Elsevier

Journal of Alloys and Compounds

Volume 741, 15 April 2018, Pages 148-154
Journal of Alloys and Compounds

Ni-Mn-Ga high temperature shape memory alloys: Function stability in β and β + γ regions

https://doi.org/10.1016/j.jallcom.2018.01.068Get rights and content

Highlights

  • Ni-Mn-Ga HTSMAs are innovative actuators materials for T = 400–500 °C.

  • Thermo-mechanical training stabilize the functional properties of Ni-Mn-Ga HTSMAs.

  • Training induces a defect structure driving a reproducible twin pattern.

  • Microstructure, heat treatments, and phase structure control the transformation.

Abstract

Record-breaking values of the tensile superelastic strain (about 12%) have been found previously in Ni-Mn-Ga single crystalline alloys at 400 °C which placed such materials ahead of known high temperature shape memory alloys (HTSMAs) promising in the automotive or aerospace industries operating in the range of 400–500 °C and above. This paper addresses two main issues that commonly affect Ni-Mn-Ga HTSMAs and limit their application, namely cycling stability of the transformation temperatures and thermomechanical actuation. These issues have been studied systematically by using six different Ni-Mn-Ga HTSMAs. The results show initial transformation temperatures up to 500 °C, which evolve, together with transformation strains, during more than 300 thermal cycles with and without mechanical loading. The specific evolution of a given sample depends on the microstructure, heat treatment prior to the cycling and whether the initial state of austenite is a single (β) or a dual phase (β + γ). The cycling protocol employed can be considered as an innovative training procedure to achieve the stabilization of the functionality and longer lifetime of these materials.

Introduction

Over the past decades, shape memory alloys (SMAs) have received considerable attention as functional materials due to their ability to recover the shape by heating and/or to exhibit superelasticity; both effects are associated with the thermoelastic martensitic transformation (MT) [1], [2]. NiTi- and Cu-based alloys are the most commonly used SMA materials [2]. These materials show MT temperatures limited to about 100 °C, whereby preventing their application at higher temperatures [3], [4], [5]. NiTi-based alloys are widely used in biomedical applications thanks to their good biocompatibility and the fact that in most of these applications the material needs to be deformed only once, as is the case of stents [5], [6], [7].

A growing interest has recently arisen in the development of high temperature SMAs (HTSMAs), working up to 550 °C, as a response to the demand from the different highly technological areas, such as automotive or aerospace industries [3], [4]. Currently, the FeMnSi-, CuAlNi-, NiMn-, NiAl-, Ti(Pt, Pd, Au)- and NiTi-based alloys systems are under intense research as HTSMAs candidates [4], [8], [9], [10], [11], [12]. While many of these systems exhibit a high temperature MT, no real breakthrough has been made yet in their actuation capability at high temperatures, due to issues like martensite stabilization [2], low strain output [7], thermal and thermomechanical instability, etc. [13], [14]. In order to meet the industrial requirements, new low cost alloys with higher operational actuation range, larger work output and stable properties are highly desirable [2], [7], [8], [9]. Among all known today HTSMAs, suitable for applications in the range of 400–500 °C, only Ti-Pd-Ni alloys with an actuation strain of about 4% seem promising, although their high cost constitutes a big disadvantage for practical implementation (see Ref. [4] and references therein).

Ni-Mn-Ga alloys have been widely studied as magnetic and non-magnetic shape memory materials [15], [16], [17], [18]. The composition-sensitive dependence of their properties [17], [19], [20], [21] makes them easily tunable materials, which can exhibit high-temperature, low-hysteresis thermoelastic martensitic transformations [22], [23], with over 9% recoverable strains [24] and good compressive and tensile superelastic behavior [25], [26]. Should these properties being combined in a single alloy, they would make it a real candidate for a highly efficient low-cost HTSMA.

The design of new Ni-Mn-Ga HTSMAs should not only focus on the increase of the MT temperature [23], [27], [28] but also on the improvement of their properties' stability [29], [30]. In the Ni-Mn-Ga phase diagram there are two regions of single phase: β (B2-ordered body-centered cubic (bcc)) and γ (Ni-rich and Ga-depleted disordered face-centered cubic (fcc)), as well as a region of a β + γ dual phase [29]. Mechanical properties [31] and thermal stability [13], [31], [32] of the alloy may strongly depend on its phase content.

Research on Ni-Mn-Ga HTSMAs has been mostly focused on improving ductility by the precipitation of second phases [33], [34], increasing MT [23], [27], [28] and studying aging effects [13], [31], [35], [36]. However, little has been done on understanding the underlying physics behind the martensite stabilization [37] or evaluating the thermal [38], [39] and thermomechanical stability [40], [41] of these alloys. Indeed, experiments on cyclic stability would be essential in order to disclose the mechanisms for aging and stable actuation at high temperature for single and dual phase samples. Furthermore, one should also emphasize that thermal treatments of the samples prior to the thermomechanical cycling, can strongly modify the physical and transformation characteristics of SMAs by changing the stabilization processes [35], [42].

The aim of this work was to elaborate Ni-Mn-Ga HTSMAs with MT higher than 400 °C exhibiting a stabilized actuation response of nearly 4% strain, after a cyclic training. Our approach to reach this goal was to explore different compositions of these alloys that show either a single β phase or a dual β + γ phase, and to study the mechanisms leading to the cycling stability (thermal and thermomechanical) in single crystal and polycrystalline forms. The influence of annealing (prior to a thermomechanical cycling) on the transformation characteristics of the single crystalline samples was also examined.

Section snippets

Experimental

Polycrystalline Ni-Mn-Ga alloys were prepared from high purity elements (Alfa-Aesar Ga 99.999% Ni 99.95+% and Mn 99.95%) by melting in an Induret compact Reitel induction furnace under argon atmosphere. The furnace chamber was evacuated and refilled with argon gas several times before melting to reduce the amount of oxygen present. The molten alloy was cast into a cold copper mold. The final compositions of the alloys were measured by EDX using a Bruker Quantax 70 detector in a Hitachi TM3000

Results and discussion

Two groups of alloys were selected in the ternary phase diagram (Fig. 1) with the aim of finding compositions with the highest possible transformation temperatures. One group of the alloys should exhibit, at room temperature a non-modulated tetragonal martensite (2M) originated from a single β phase. The second group, exhibiting a dual β + γ phase structure, should show at room temperature a mixture of the 2M martensite, resulting from the transformation of the austenite, with the precipitates

Conclusions

We have studied the cyclic behavior of four polycrystalline Ni-Mn-Ga HTSMAs. Then, we have grown two single crystals with composition pertaining to a single (β) and a dual (β + γ) phase regions on the phase diagram, which show thermo-mechanical actuation in the temperature range of 400–500 °C.

Two very distinct behaviors are observed in the thermal and thermomechanical cycling behavior of the samples. Under thermal cycling: (i) the samples from the dual (β + γ) phase regions show a decrease of

Acknowledgements

Authors would like to acknowledge the financial support from Alstom Inc. (project ANS GTA-2014-0410) and from Spanish Ministry of Economy and Competitiveness (project MAT2014-56116-C4-3-4-R). A. Pérez-Checa acknowledges a PhD grant from the Basque Government. Technical and human support provided by SGIker (UPV/EHU, MINECO, GV/EJ, ERDF and ESF), and in particular by A. Larrañaga, is also gratefully acknowledged.

References (46)

  • V.A. Chernenko et al.

    Giant two-way shape memory effect in high-temperature Ni–Mn–Ga single crystal

    Physics Procedia

    (2010)
  • V.A. L’vov et al.

    Destabilization of Ni–Mn–Ga martensite: experiment and theory

    Acta Mater.

    (2012)
  • C.B. Jiang et al.

    Effect of Ni excess on phase transformation temperatures of NiMnGa alloys

    Mater. Sci. Eng.: A

    (2003)
  • R.K. Singh et al.

    Magnetic and structural transformation in off-stoichiometric NiMnGa alloys

    Mater. Sci. Eng.: A.

    (2008)
  • S. Yang et al.

    Phase equilibria and composition dependence of martensitic transformation in Ni–Mn–Ga ternary system

    Intermetallics

    (2012)
  • Y. Xin et al.

    Microstructure, mechanical and shape memory properties of polycrystalline Ni–Mn–Ga High temperature shape memory alloys

    Mater. Sci. Eng.: A.

    (2016)
  • S. Yang et al.

    Microstructure and shape-memory characteristics of Ni 56 Mn 25−x Co x Ga 19 (x= 4, 8) high-temperature shape-memory alloys

    Intermetallics

    (2011)
  • S. Yang et al.

    Simultaneous improvement of ductility and shape memory effect of two-phase Ni 53 Mn 22 Co 6 Ga 19 high-temperature shape memory alloy with the addition of Ta

    Mater. Lett.

    (2012)
  • M. Kök et al.

    Effect of heat treatment on the physical properties of Ni–Mn–Ga alloy

    Thermochim. Acta

    (2012)
  • E. Cesari et al.

    Thermal stability of high-temperature Ni–Mn–Ga alloys

    Scripta Mater.

    (2008)
  • C. Segui et al.

    Martensite stabilization in a high temperature Ni–Mn–Ga alloy

    Scripta Mater.

    (2005)
  • Y.Q. Ma et al.

    Thermal stability of the Ni 54 Mn 25 Ga 21 Heusler alloy with high temperature transformation

    Scripta Mater.

    (2003)
  • S. Yang et al.

    Martensite stabilization and thermal cycling stability of two-phase NiMnGa-based high-temperature shape memory alloys

    Acta Mater.

    (2012)
  • Cited by (0)

    View full text