Introduction

Shape-memory alloys (SMAs) are alloys that undergo the diffusionless solid–solid phase transformation known as martensitic transformation (MT). Their functionalities—the shape-memory effect and superelasticity—have long stimulated considerable interest in the field of materials science and engineering. The superelasticity enables a large recoverable strain to be attained using the stress-induced forward/reverse MTs via the application/release of uniaxial stress.1 The behavior is similar to that of elastomers; the stress–strain (σɛ) curve exhibits plateaus but not linearity as a superelastic response.1 This functionality has resulted in many important applications, such as guide wires and dam** components.1, 2, 3 In addition, refrigeration using the latent heat yielded by the stress-induced MT has recently received considerable attention as an environmentally friendly refrigeration technique called elastocaloric cooling (EC),4 which may be a promising alternative to conventional vapor-compression techniques. This EC effect (previously termed the piezocaloric effect5) exploits the transformation entropy change ΔS as a cooling source under adiabatic conditions. ΔS is related to the equilibrium stress σ0 and transformation strain ɛSE by the Clausius–Clapeyron equation:6

where Vm represents molar volume. Therefore, investigating the temperature dependence of superelastic behaviors can be used as an indirect approach to evaluate the EC potential of SMAs.

In the design of practical SMAs, minimizing the transformation hysteresis without deteriorating the shape-memory and/or superelastic property is a general strategy because it leads to reproducibility of these functionalities and enhancement of their fatigue properties.7 However, simultaneously achieving small σhys and large ɛSE (or ΔS) remains challenging. Even for the R-phase (trigonal-symmetric martensite) transformation in Ti–Ni, which exhibits a small temperature hysteresis of less than a few degrees Kelvin, the obtainable ɛSE is <1%;8 therefore, ΔS is also small. In contrast, σhys in SMAs with large ɛSE, for example, in FeNiCoAlTaB,9 is generally large. Recently, an extremely small temperature hysteresis and thereby excellent fatigue properties were achieved in certain SMAs10, 11, 12 because of the very small lattice mismatch between the parent and martensite phases10, 11 or among them and precipitates12 arising from special relationships intrinsic to their lattice correspondence10, 11, 12, 13 (termed ‘supercompatibility’ by James14). Another, more conventional approach involves tailoring the microstructure to reduce intergranular constraints; the fatigue properties of Cu–Al–Mn SMA in the forms of a single crystal, bamboo-like polycrystals, and columnar-oriented polycrystals have been successfully improved while realizing small σhys and large ɛSE.15,

Materials and methods

Ingots of Cu–17Al–15Mn and Cu–17Al–12Mn (at.%) were prepared by induction melting in an Ar atmosphere. The ingots were hot-rolled at 1143 K to reduce their thickness to ~3 mm. The obtained plates were subjected to cyclic heat treatment (see Supplementary Figure S1a) to obtain single crystals by inducing abnormal grain growth.26 The plates (shown in Supplementary Figure S1b) were subsequently annealed at 473 K for 15 min to stabilize the MT behavior between the cubic (L21) parent and monoclinic (18R, now often termed 6M) martensite phases.27, 28 As shown in Supplementary Figure S2, it was confirmed that the obtained Cu–17Al–15Mn does not undergo the thermally induced MTs in the temperature range 4–300 K.

A series of tensile tests of Cu–17Al–15Mn single crystals was performed at a strain rate of 5 × 10−5 s−1 in order from the highest to lowest test temperatures. To account for the training effect, the tests were conducted at least twice at every temperature, and the σɛ curve obtained in the final test was used. The tensile orientations in the tested single crystals were determined using the electron backscattered diffraction technique.

Results and discussion

The tensile σɛ curves of the Cu–Al–Mn single-crystal specimen (#1) in the low-strain region at temperatures of 4.2–160 K are presented in Figure 2. The series of σɛ curves indicates that the superelastic property was successfully achieved at all the temperatures tested. The two insets show the full superelastic curve at 4.2 K and the crystallographic orientation of the tested crystal on an inverse pole figure. Those for another crystal (#2) with a different orientation are provided in Supplementary Figure S3. The ɛSE values of 7.1% and 5.3% for specimens #1 and #2 at 4.2 K are less than the theoretical values of 9.1% and 7.6%, respectively, estimated from the crystallographic relationship between the cubic (L21) parent and monoclinic (18R) martensite phases in a Cu–17Al–10Mn SMA at ambient temperature.29 This discrepancy may have been caused by a change in the lattice parameters resulting from the different Mn concentration and temperature. The critical stresses, σMs and σAf, of the stress-induced forward/reverse MTs, as depicted in Figure 2, are plotted in Figure 3a as a function of temperature. Those reported for bamboo-textured30 and columnar-grained27 high machinability,27 improved fatigue life,16 grain size tunability,26 and transformation temperature tunability.28 It has been reported that the near <100>-oriented single crystal exhibits very large ɛSE approaching 10%.31 If we assume ɛSE=10%, σMs=50 MPa at 4.2 K, temperature-independent σhys=20 MPa, and critical slip stress=500 MPa in the near <100>-oriented single crystal, operations of superelasticity at temperatures up to 366 K and of the EC effect at temperatures as low as TCO=24 K may be feasible; one specimen can exhibit both superelasticity and the EC effect over a wide temperature window of ~340 K. As previously mentioned, the crystallographic compatibility between the parent and martensite phases10, 11, 12, 13 was considered to be a factor determining the magnitude of the MT temperature hysteresis. Although experimental investigations have not been attempted, this guidance appears to be valid for suppressing σhys and temperature hysteresis; thus, the existing COPmat metrics (Figure 5) may be revised for such supercompatible superelastic alloys. However, it is concerning that these conditions can be easily disrupted by changing temperature and composition. It remains unknown how thermal activation develops as temperature decreases because the extremely small hystereses are all observed at ambient temperatures at which the athermal nature is dominant. Therefore, the novel feature of the present Cu–Al–Mn alloy, namely, the temperature-independent small σhys and persistent EC performance over a very wide temperature range, will not be lost even after the supercompatible strategy is applied to design a superelastic alloy.

As shown in Figure 6, the present Cu–Al–Mn alloy exclusively covers the application fields of space engineering, superconducting technologies, and liquefied-gas technologies, for which it has been almost impossible to implement existing SMAs. This material will contribute to innovative developments of these cryogenic technologies as well as the downsizing or simplification of various cryogenic systems.

Figure 6
figure 6

Existing and potential application fields and operational temperature windows of shape-memory materials.