Next Article in Journal
Molecular Dynamics Study on the Impact of Cu Clusters at the BCC-Fe Grain Boundary on the Tensile Properties of Crystal
Next Article in Special Issue
Control of Density and Grain Structure of a Laser Powder Bed-Fused Superelastic Ti-18Zr-14Nb Alloy: Simulation-Driven Process Mapping
Previous Article in Journal
Microstructure Evolution and Mechanical Properties of Refractory Mo-Nb-V-W-Ti High-Entropy Alloys
Previous Article in Special Issue
Softening of Shear Elastic Coefficients in Shape Memory Alloys Near the Martensitic Transition: A Study by Laser-Based Resonant Ultrasound Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

TiPd- and TiPt-Based High-Temperature Shape Memory Alloys: A Review on Recent Advances

by
Yoko Yamabe-Mitarai
1,2
1
Department of Advanced Materials Science, Graduate School of Frontier Sciences, The University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa-shi, Chiba 277-8561, Japan
2
National Institute for Materials Science, 1-2-1 Sengen, Tsukuba, Ibaraki 305-0047, Japan
Metals 2020, 10(11), 1531; https://doi.org/10.3390/met10111531
Submission received: 30 September 2020 / Revised: 10 November 2020 / Accepted: 12 November 2020 / Published: 18 November 2020
(This article belongs to the Special Issue Shape Memory Alloys 2020)

Abstract

:
In this paper high-temperature shape memory alloys based on TiPd and TiPt are reviewed. The effect of the alloying elements in ternary TiPd and TiPt alloys on phase transformation and strain recovery is also discussed. Generally, the addition of alloying elements decreases the martensitic transformation temperature and improves the strength of the martensite and austenite phases. Additionally, it also decreases irrecoverable strain, but without perfect recovery due to plastic deformation. With the aim to improve the strength of high-temperature shape memory alloys, multi-component alloys, including medium- and high-entropy alloys, have been investigated and proposed as new structural materials. Notably, it was discovered that the martensitic transformation temperature could be controlled through a combination of the constituent elements and alloys with high austenite finish temperatures above 500 °C. The irrecoverable strain decreased in the multi-component alloys compared with the ternary alloys. The repeated thermal cyclic test was effective toward obtaining perfect strain recoveries in multi-component alloys, which could be good candidates for high-temperature shape memory alloys.

1. Introduction

Shape recovery in shape memory alloys (SMAs) occurs during a reverse martensitic transformation from martensite to austenite phases. Thereafter, the SMA operating temperature is related to the martensitic transformation temperature (MTT). High-temperature shape memory alloys (HT-SMAs) are defined as SMAs that can recover their shapes at temperatures above 100 °C. Several applications of HT-SMAs have been proposed. For example, Ni30Pt20Ti50, whose MTTs include austenite start temperature, As: 262 °C; austenite finish temperature, Af: 275 °C; martensite start temperature, Ms: 265 °C; and martensite finish temperature, Mf: 240 °C, was applied for active clearance control actuation in the high-pressure turbine section of a turbofan engine [1]. This indicates that the design can offer a small and lightweight package without requiring motion amplifiers that cause efficiency losses and introduce an additional failure mode [1]. Another example is the helical actuators for surge-control applications in helicopter engine compressors [2]. In this application, Ni19.5Ti50.5Pd25Pt5, whose MTTs comprise As: 243 °C, Af: 259 °C, Ms: 247 °C, and Mf: 228 °C, was applied because the alloy exhibited good work capabilities, a 2.5% recoverable strain, and a work output of 9.45 J/cm3 at 400 MPa [2]. Several SMA applications, such as the active jet engine chevron, springs and wires for a general class of high-temperature actuators, oxygen mask deployment latch, SMA-activated thermal switch for lunar surface applications, variable geometry chevrons, and gas turbine variable area nozzles, have also been proposed [3]. Here, HT-SMAs, Ni19.5Ti50.5Pd25Pt5 or Ni50.3Ti29.7Hf20 are used only in springs and wires for a general class of high-temperature actuator. Furthermore, NiTi-based SMAs that can actuate in the temperature range of 70–90 °C are used in other applications.
Raising MTTs is necessary for the development of HT-SMAs. In addition, improving SMA strength is also important because plastic deformation easily occurs at high temperatures, thereby resulting in incomplete shape recovery. Several studies have been conducted to increase MTTs by adding alloying elements such as Hf, Zr, Pd, Pt, and Au, to NiTi [4,5,6,7,8,9,10]. Their MTTs successfully increased by adding an alloying element, but a perfect shape recovery was not obtained. Recently, research of NiTi alloys has shifted to Ni50.3Ti29.7Hf20, which is strengthened by nano-size precipitates called the “H phase” [11,12,13,14,15,16,17,18,19,20,21,22,23,24]. The austenite finishing temperature Af of Ni50.3Ti29.7Hf20 is typically 166 °C under unloading conditions [13], but it rises to 270 °C under tensile loading conditions at 500 MPa [16]. Furthermore, ageing increased the work output due to the higher transformation strain and the work output under 500 MPa was 16.45 J/cm3 at Af of 270 °C [16]. High strength Ni-rich Ni51.2Ti28.8Hf20 was also developed and its work output was 23 J/cm3 under 1700 MPa at Af of approximately 100 °C, 27 J/cm3 under 1500 MPa at Af of approximately 220 °C, and 15 J/cm3 under 1000 MPa at As of 151 °C (Af was not clearly shown) [19]. The effect of 2000 thermal training cycles under 300 MPa of Ni50.3Ti29.7Hf20 was also investigated and it was found that the stable cyclic strain recovery with the almost constant transformation strain [24]. The work output under 300 MPa was approximately 7.5 J/cm3 at Af of approximately 220 °C [24].
Another approach to increasing MTTs is using other alloys with MTTs higher than those of NiTi alloys. Therefore, TiPd, TiAu, and TiPt have been studied because they exhibit a martensitic transformation from a B2 to a B19 structure, and their MTT values are higher than 500 °C [25,26]. For example, typical martensitic twin structures were observed in TiPt, whose high potential as an HT-SMA has been established [27,28]. The first investigation on strain recovery at high temperatures was performed for TiPd [29]. A binary TiPd sample was deformed at 500 °C, and the change in its length after it was heated above the Af was investigated to measure strain recovery [29]. However, only a 10% strain was recovered owing to plastic deformation at 500 °C [29]. The shape recovery behavior of the TiAu alloy was investigated through thermomechanical analysis measurements after compressive deformation at 500 °C [30]. It was found that an 80% strain recovery occurred after a 5% deformation [30]. The effects of both Zr [31] and Ag addition [32] into TiAu on the martensitic transformation and strain recovery was investigated; the MTT decreased by 10% with the addition of Zr and Ag. For example, Af decreased from 624 °C for TiAu to 511 °C and 594 °C for Ti–50Au–10Zr and Ti–40Au–10Ag, respectively [31,32]. Compressive deformation was applied at a test temperature of 50 °C below Mf, and the deformed samples were heated above Af to measure the strain recovery. The strain recovery ratios of the Ti–50Au, Ti–50Au–10Zr, and Ti–40Au–10Ag alloys were 68%, 82%, and 76%, respectively [32]. The addition of Zr and Ag was effective in improving the strain recovery of TiAu.
The effects of an alloying element on MTTs, strain recovery, as well as strength of the martensite and austenite phases in TiPd and TiPt alloys, have been investigated by my group [33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50] and are reviewed herein. In addition to the TiPd and TiPt alloys, high- and medium- entropy SMAs (HEAs or MEAs) are also appraised because HEAs and MEAs have been attracting considerable attention as new SMAs. Notably, HEAs and MEAs are multi-component equiatomic or near-equiatomic alloys, which have garnered much interest as new generation structural materials because their high-entropy effects, such as severe lattice distortion and sluggish diffusion, are expected to improve the high-temperature strength of alloys [51,52]. As already shown, it is difficult to achieve perfect strain recovery in HT-SMAs because of the easy introduction of plastic deformation at high temperatures. Furthermore, improvement of strength of SMAs is a key issue for HT-SMAs. Application of HEAs and MEAs to HT-SMAs is expected in this area for which results on multi-component alloys, in particularly MEAs and HEAs, are presented in this paper.

2. Martensitic Transformation Temperature

The MTTs of the TiPd and TiPt alloys measured through differential scanning calorimetry (DSC) (DSC3200, NETZSCH Japan, Yokohama, Japan) are shown in Table 1. The temperature hysteresis, AfMs, is also summarized in Table 1. Data without reference are from my group’s unpublished work. The addition of a third element, such as Hf, Zr, V, Nb, Ta, Cr, Mo, W, Ru, Ir, and Co, was performed for TiPd, while elements, including Ir, Ru, Co, Zr, and Hf, were added for TiPt. In most cases, the MTT decreased with the addition of an alloying element. In Figure 1a, Af changes with increasing alloying element concentration are shown for some of the ternary alloys. Notably, Af decreased rather linearly with increasing amounts of alloying elements. The same trend appeared for As, Ms, and Mf. The average decrease in Ms per 1 at% of an alloying element up to 5 at% addition is shown in Table 1. Ta and W did not melt well into TiPd, and their particles remained in the alloy, thereby making the matrix composition of Ti45Pd50W5 and Ti45Pd50Ta5 approximately close to that of TiPd. Consequently, the decrease in the MTT by W and Ta was small. For other alloying elements, the decrease in MTT was between 15–37 °C. In Figure 1b, the MTT change with respect to Pd concentration is observed for Ti(95 − x)PdxZr5; the MTT increased with an increase in the Pd concentration.
The MTT of TiPt was very high; for example, As and Af were 1000 °C and 1057 °C, respectively, as shown in Table 1. Although the addition of Ru, Co, Zr, and Hf decreased the MTT of TiPt, the addition of Ir increased it.
The MTTs of multi-component alloys are summarized in Table 2. Data without reference are from my group’s unpublished work. The multi-component alloys were designed based on TiPd as follows: Pd, a group 10 element in the periodic table, was replaced by Co and Ir, which are group 9 elements; by Ni and Pt which are group 10 elements; and by Au, a group 11 element. Furthermore, Ti, a group 4 element, was replaced by Zr and Hf, which are also group 4 elements. Evident DSC peaks were not observed in the three alloys, Ti40Zr10Pd25Pt25, Ti45Zr5Pd37Ni13, and Ti45Zr5Pd25Pt20Ni5, and their MTTs were estimated from the strain–temperature (S–T) curves during the thermal cyclic test under constant stress.
Multi-component alloys are classified according to the mixing entropy, ΔSmix, using the following equations [51].
Here, ΔSmix is defined by the following equation:
Δ S mix   =   R i = 1 n x i l n x i
where xi is the mole fraction of component i, n is the number of constituent elements, and R is the gas constant (8.314 J/Kmol).
HEA:
Δ S mix     1.5 R
Medium-entropy alloy (MEA):
1.0 R     Δ S mix     1.5 R
Low-entropy alloy (LEA, conventional solid-solution alloy):
Δ S mix   1.0 R
The calculated mixing entropies are presented in Table 2. Based on the mixing entropies, the alloy classification is also shown in Table 2. Most quaternary alloys are classified as MEAs, while some are classified as LEAs, i.e., conventional solid-solution alloys. Notably, HEAs have been identified in some multi-component alloys with five or six constituent elements.
The MTTs of the five quaternary alloys, TiZrPdIr, TiZrPdPt, TiZrPdNi, TiZrPdCo, and TiZrVPd, are shown in Table 2. The combination of TiZrPdIr, TiZrPdNi, and TiZrPdCo decreased the MTTs compared with those of Ti45Zr5Pd50. This is because the phase transformations of TiIr, TiNi, and TiCo are different from those of TiPd. For example, TiIr undergoes a two-step phase transformation from the B2 structure at high temperature to a tetragonal structure (distorted B2 structure) at middle temperature, and finally to an orthorhombic structure with space group 65 at low temperature [56]. The MTT depends on Ir concentration. Notably, TiNi undergoes martensitic transformation from the B2 to the B19′ structure [4]. The B2 structure is very stable in TiCo from room temperature to the melting temperature [25]; the addition of these elements is considered to stabilize the B2 structure. However, TiZrPdPt increased the MTT relative to Ti45Zr5Pd50. This is because the same martensitic transformation from B2 to B19 structures, as well as TiPd, occur in TiPt, whose MTT is very high, approximately 1000 °C, as shown in Table 1. Furthermore, Pt and Pd exhibit perfect solubility in both austenite and martensite phases. Consequently, MTT gradually increased with Pt content. In the case of TiZrVPd, the total amounts of Zr and V were maintained at 5 at%, and the values of As and Af were approximately the same as those of Ti45Zr5Pd50, while the Ms and Mf values of TiZrVPd were lower than those of Ti45Zr5Pd50.
Although the MTTs of multi-component alloys depend on the combination of elements, there are some trends whereby those of TiZrPdPtNi and TiZrPdPtAu are relatively high, whereas that of TiZrPdNiCo is relatively low.
The potential of HEAs as SMAs was presented for the first time by Fistov et al., wherein Ti16.7Zr16.7Hf16.7Ni25Cu25 was investigated and its As and Af were 184 and 338 °C, respectively [53]. The addition of Co to Ti16.7Zr16.7Hf16.7Ni25Cu25 was investigated, but its MTT drastically decreased and Af was –23 °C [54]. The MTTs of these alloys are summarized in Table 2. The potential of HEAs as HT-SMAs was first shown by Canadinc [55]. The MTT of their alloys was obtained from the DSC curves in reference [55] and is summarized in Table 2. The As and Af of Ti16.7Zr16.7Hf16.7Pd25Ni25 were 740 and 780 °C, respectively. In my group’s experiments, HEAs and some MEAs did not show martensitic transformation up to 700 °C during DSC measurement. However, martensitic transformation may be possible for these alloys at temperatures higher than 700 °C. The MTT measurement of my group’s HEAs is presently undergoing ultra-high temperature DSC. We found that some alloys have a high MTT, approximately close to 1000 °C in multi-component alloys.
Notably, the temperature hysteresis of the multi-component alloys becomes larger relative to that of ternary alloys. The number of alloys with temperature hysteresis exceeding 100 °C and their ratios are summarized in Table 3. The ratio of alloys with temperature hysteresis exceeding 100 °C in ternary alloys is 12% and it increases to 28% in LEAs and to 40% in MEAs and HEAs. It indicates that the ratio of alloys with temperature hysteresis exceeding 100 °C increased with increase of amount of constituent elements in alloys. In multi-component alloys, in particularly in MEAs and HEAs, the amount of constituent elements is equivalent or near-equivalent; therefore, it is expected that severe lattice distortion, which is considered to obstruct martensitic transformation, will occur. In many MEA cases, a drastic decrease in Ms was observed compared with Af. When SMAs are used as actuators, a smaller temperature hysteresis is necessary to quickly respond to the surrounding environment.

3. Strain Recovery Determined by Compression Test

In the early stage of my group’s research [33,34,35,37,38,39,40,41,42,43,44], the recovery strain was estimated by measuring the sample length before and after the compression test using a strain rate of 3 × 10−3/s, and after heat treatment above Af. A sample was deformed by approximately 5% at the test temperature. The deformed sample length (L’) was measured after cooling to room temperature. Thereafter, the sample was re-heated over Af, cooled to room temperature, and the recovered sample length (L”) was measured. When the initial sample length is L0, the applied strain εa is defined by the following equation:
εa = 100 × ((L0 − L’)/L0)
Recoverable strain εr is defined according to the following equation:
εr = 100 × ((L” − L’)/L0)
Strain recovery ratio, i.e., shape memory effect (SME) was evaluated using the following equation:
SME (%) = εra × 100
The applied and recoverable strains, as well as the strain recovery ratios of the ternary TiPd and TiPt alloys, are summarized in Table 4. Data without reference are from my group’s unpublished work. For the quaternary alloys, this method was used only for TiZrPdIr, and the results are summarized in Table 5. Although it is difficult to compare the effect of alloying elements due to different applied strains, some alloys have a high strain recovery ratio of above 80%. The quaternary alloys also exhibited a high strain recovery ratio of above 80%, as shown in Table 5. In Figure 2, the strain recovery ratios of Ti45Pd50 × 5 are plotted for the periodic table family; it can be observed that the addition of Zr and Hf is effective toward improving strain recovery. Since the atomic size (metallic) of Zr and Hf are 160 and 159 pm, respectively, and larger than those of Ti (147 pm) and Pd (137 pm) [57], the solid-solution hardening effects of Zr and Hf are considered to exceed those of the other alloying elements. Thereafter, the 0.2% flow stress of the austenite and martensite phases were investigated by compression testing using samples with dimensions of 2.5 × 2.5 × 5 mm3. The 0.2% flow stresses of ternary TiPd and TiPt alloys are summarized in Table 6. Data without reference are from my group’s unpublished work. When the martensite phase is deformed, a double yielding behavior often occurs. The first yielding behavior represents the rearrangement of the martensite variant and is referred to as the detwinning stress. The second yielding behavior represents the yield stress of martensite. In Table 6, the detwinning stresses of the martensite phase are also shown. In Figure 3, the 0.2% flow stress of the austenite and martensite phases of Ti45Pd50X5 (X in group 4–6 elements) and Ti50Pd46Y4 (Y in group 8–10 elements) are plotted for the periodic table family. Most alloying elements improved the strength of the austenite and martensite phases compared with those of Ti50Pd50. The strengthening behavior of the alloying element on the austenite and martensite phases was similar. Comparing Figure 2 and Figure 3, it is evident that high-strength alloys, such as Ti45Pd50V5, have a lower strain recovery than Ti45Pd50Zr5 and Ti45Pd50Hf5 with a smaller strengthening effect than Ti45Pd50V5. It is difficult to establish the correlation between strength and strain recovery ratio.

4. Strain Recovery Determined by Thermal Cyclic Test under Constant Stress

Recently, strain recovery was investigated through a thermal cyclic compression test under a constant load (Shimadzu AG-X, Shimadzu, Kyoto, Japan). The specimens were heated to 30 °C above Af and then cooled to a temperature lower than Mf under a constant load of 15–200 MPa. Five thermal cycles under loads of 15, 50, 100, 150, and 200 MPa were applied for the same samples in the order given with heating and cooling rates of 50 °C/min. The work output (work per volume) was obtained from the product of the recovery strain and the applied stress. The strain temperature (S–T) curves of Ti45Pd50X5 (X = Zr, Hf, V, and Nb) are plotted in Figure 4. The transformation strain increased with an increase in the applied stress. The transformation strains of Ti45Pd50Zr5 and Ti45Pd50Hf5 exceed 2% above 50 MPa, as shown in Figure 4a,b. Nonetheless, the transformation strains of Ti45Pd50V5 are less than 1% at an applied stress of 200 MPa. Phase transformation was not evident in TiPdNb, even at 200 MPa. The perfect strain recovery was obtained under a small applied stress of less than 100 MPa in Ti45Pd50V5, but the irrecoverable strain appeared at 150 MPa and increased at 200 MPa. In Ti45Pd50Zr5 and Ti45Pd50Hf5, the small irrecoverable strain was observed, even at 15 MPa, which increased with the applied stress. At 150 and 200 MPa in Ti45Pd50Hf5, trumpet-shaped S–T curves were obtained, as shown by arrows in Figure 4b. This is because the compressive plastic deformation is larger than the expansion of the sample during heating above Af.
The recoverable and irrecoverable strains, as well as the work output of Ti45Pd50X5 (X = Zr, Hf, and V) are plotted as a function of the applied stress, as shown in Figure 5; for reference, those of Ti50Pd50 at 50 MPa are also plotted [58]. The recoverable strain increased with the applied stress; however, by increasing the irrecoverable strain, the recoverable strain started to decrease after the peak recoverable strain. Although the recoverable strain of Ti50Pd50 is similar to that of Ti45Pd50X5 (X = Zr and Hf), the irrecoverable strain is very large compared to that of the ternary alloys. The irrecoverable strain is expected to increase drastically with an increase in the applied stress. It also indicates that the addition of an alloying element is effective toward decreasing the irrecoverable strain. The work output increased with applied stress. Although an irrecoverable strain appeared, a large work output of 7.6 J/cm3 was obtained for Ti45Pd50Zr5.

5. Effects of Training

It is known that repeated thermal cyclic tests, which is referred to as “training,” reduces the irrecoverable strain, and a perfect recovery is finally obtained [8,59]. For example, the irrecoverable strain of TiNi-based alloys became approximately 0 after 40 cycles under a load of 80 MPa. The training was effective for the TiNi alloys strengthened by the addition of Pd or Sc. Therefore, training was performed for Ti45Pd50X5 (X = Zr and Hf) under a 50-MPa load. The irrecoverable strain vs. the number of thermal cycles is plotted in Figure 6. The irrecoverable strain of Ti45Pd50Zr5 was saturated after 10 cycles, and it remained at approximately 0.1%. The irrecoverable strain of Ti45Pd50Hf5 was approximately double that of Ti45Pd50Zr5, and it was unsaturated. This indicates that the strain recovery of ternary alloys is unstable for repeated cycles. The effect of Zr content on the training effect was investigated for Ti45Pd50Zr7 and Ti45Pd50Zr10 [36]. By increasing Zr addition, the irrecoverable strain finally disappeared during training and perfect strain recovery was achieved [36]. This outcome can be attributed to two reasons: (1) the solid-solution hardening effect is larger at high Zr contents, and (2) MTT reduction advantageously affects the suppression of plastic deformation.

6. Strain Recovery of Multi-Component Alloys

A thermal cyclic test was performed on the multi-component alloys to investigate strain recovery. Some results have already been published [45,46,47,48,49,50]. The recoverable and irrecoverable strains, as well as the work output of some of the multi-component alloys are shown in Figure 7; the TiZrPdNi and TiZrPdPt alloys are shown in Figure 7a,c,e, and in Figure 7b,d,f, respectively, while Ti45Pd50Zr5 is shown as a standard sample in all the diagrams. Furthermore, Ti45Zr5Pd45Ni5, Ti45Zr5Pd37Ni13, Ti40Zr10Pd25Pt25, and Ti45Zr5Pd25Pt20Ni5 are the original data source in this study. In Figure 7a,c,e, the concentrations of Ti and Zr were maintained at 45 and 5 at%, respectively, in TiZrPdNi, and only the concentrations of Pd and Ni were changed. Among the tested alloys in Figure 7a,c,e, TiZrPdNi alloys exhibited a relatively high recoverable strain, although the recoverable strain of Ti45Zr5Pd40Ni10 [48] was similar to that of ternary Ti45Pd50Zr5. Moreover, Ni addition seems to increase the recoverable strain of Ti45Pd50Zr5. The recoverable strain of Ti45Zr5Pd40Co10 [48] was very small, thereby suggesting that Co addition drastically decreased the recovery strain. However, the recoverable strain of the multi-component alloys of TiZrPdNiCo was larger than that of Ti45Zr5Pd40Ni10 [48]. This indicated that the recoverable strain was increased by the effect of the multi-component alloy, i.e., the high-entropy effect. In Figure 7c, the irrecoverable strains of most of the tested alloys were smaller than 0.2%. Only the ternary Ti45Pd50Zr5 and quaternary alloy of Ti45Zr5Pd45Ni5, which were defined as LEA, represented a large irrecoverable strain exceeding 0.2%. A decrease in the irrecoverable strain of MEAs is also considered to be consequent of the high-entropy effect. As a result of the large recoverable strain, the work output of TiZrPdNi-type multi-component alloys also becomes large, as shown in Figure 7e. The work outputs of some alloys exceeded 10 J/cm3. The repeated thermal cycling test, i.e., training was also applied for some alloys. For example, training under 300 MPa was performed for Ti45Zr5Pd40Ni10 for 100 cycles [48]. Although the transformation strain decreased, the irrecoverable strain decreased, and a perfect recovery was finally achieved during the thermal cyclic test. The final recovery strain was approximately 3.6% under 300 MPa. Thereafter, the work output was 10.8 J/cm3 at the Af of 258 °C. For Ti45Zr5Pd40Co10, training was performed under 700 MPa. After nine cycles, the irrecoverable strain disappeared, and a perfect recovery was achieved. The final recovery strain was approximately 1%, and the work output was 7 J/cm3 at the Af of 361 °C.
In Figure 7b,d,f, TiZrPdPt alloys are compared. Again, the concentrations of Ti and Zr were maintained at 45 and 5 at%, respectively, in most of the alloys. Compared with the recoverable strain in TiZrPdNi alloys, as shown in Figure 7a and in TiZrPdPt alloys, as presented in Figure 7b, those of TiZrPdPt alloys were smaller than 3% and less than those of TiZrPdNi alloys. This indicates that the addition of Pt decreases the recoverable strain. The same trend can be observed in quaternary alloys by comparing Ti45Zr5Pd25Pt25 and Ti45Zr5Pd45Pt5. A small recoverable strain was achieved in the alloy with high Pt content. The effect of Zr on the recoverable strain in quaternary alloys could be understood by comparing Ti40Zr10Pd25Pt25 and Ti45Zr5Pd25Pt25, and it was found that Zr addition increased the recoverable strain. In the multi-component alloys, it was found that high Pt addition decreased recoverable strain, as shown in Figure 7b. In TiZrPdPt alloys, it was found that the irrecoverable strain decreased in the multi-component alloys including the MEAs and HEAs, except for the LEA, Ti45Zr5Pd45Pt5, as shown in Figure 7d. The irrecoverable strain of Ti45Zr5Pd25Pt25 is also small, which may be due to high MTT. The work outputs of the TiZrPdPt alloys are shown in Figure 7f; they are all between 2–6 J/cm3 and smaller than that of Ti45Pd50Zr5.
Training was performed for Ti45Zr5Pd20Pt25Ni5 and Ti45Zr5Pd20Pt20Ni10 [49], and perfect recovery was achieved during the thermal cyclic test for approximately 100 cycles under loads of 200, 300, and 400 MPa, whereby the final work outputs were approximately 3.5, 3, and 2 J/cm3, respectively, for Ti45Zr5Pd20Pt25Ni5. For Ti45Zr5Pd20Pt20Ni10, final work outputs of 3 and 1.5 J/cm3 under loads of 200 and 300 MPa were obtained, respectively. The small work output for the large applied stress represents a drastic decrease in the transformation strain. It is necessary to keep the transformation strain of the multi-component alloys during the thermal cyclic test. The thermal cyclic test is considered as a kind of thermal fatigue test and the stable cyclic strain recovery with the constant transformation strain indicates the thermal fatigue life of SMAs and stability as SMA actuators.
The strength of the austenite and martensite phases of some of the multi-component alloys are investigated, and the results are summarized in Table 7. The strength of the austenite phases in the tested alloys was between 200 and 300 MPa, and they were similar to those of the ternary alloys. This is because the MTTs of the multi-component alloys are relatively high. Therefore, it is difficult to correlate the strength and strain recovery directly. Thereafter, the temperature dependence of the strength of the martensite and austenite phases was investigated for ternary and multi-component alloys [49]. The strength of the multi-component alloys was higher in both martensite and austenite phases when compared at the same temperature. The solid-solution strengthening effect of the multi-component alloys was more evident when the strength of the austenite phase was compared at the same test temperature of 700 °C. The strength of the multi-component alloys was higher than that of the ternary alloys or LEAs.

7. Change in Microstructure before and after Training

The microstructural changes during training were also investigated in Ti43Zr7Pd50 and Ti40Zr10Pd50 [36], Ti45Zr5Pd35Pt15 [45], as well as Ti50Pd50 [58]. A typical twin structure with multiple variants was observed in the heat-treated sample, but variant selection occurred during the thermal cyclic test, and the major variant was predominantly the [010] axis parallel to the compression axis in all observed alloys. It is suggested that a martensite variant is selected to obtain the largest constriction of the lattice during the compression test. The crystal orientation relationship between the B2 and B19 structures is shown as follows:
[ 110 ] B 2 / / [ 100 ] B 19 ,   [ 010 ] B 2 / / [ 010 ] B 19 ,   [ 1 ¯ 10 ] B 2 / / [ 001 ] B 19
The lattice parameters of the orthorhombic B19 structure of TiPd were a = 0.459, b = 0.28, and c = 0.484 nm [60]. The lattice parameter of b in the B19 structure was the smallest, and the reorientation of the martensite variant along the [010] direction is, therefore, considered to relax the compressive strain under compressive stress.

8. Potential of High-Temperature Shape Memory Alloys (HT-SMAs)

The work outputs obtained by the thermal cyclic test are plotted for Af, as shown in Figure 8. The open symbols indicate the alloys with imperfect recovery after a single thermal cycle, while solid symbols indicate perfect recovery after training. The green square indicates the work output of Ni50.3Ti29.7Hf20 with nano-sized precipitation, which is developed in the USA and is the most expected HT-SMA [16,19]. It is said that the work output of the commercially used TiNi is between 12–18 J/cm3 for operating temperatures under 100 °C [61]. The work output of Ni50.3Ti29.7Hf20 is equivalent to that of TiNi alloys at higher Af (220 °C) than TiNi. Although the work outputs of some alloys in my group’s research exceed 12 J/cm3 at approximately 200 °C [48], most of them underwent decreasing work outputs with increasing Af. Nevertheless, it is notable that perfect recovery is achieved at temperatures between 200–600 °C because it is difficult to obtain above 200 °C. The work output obtained at 600 °C was approximately 3 J/cm3.

9. Conclusions

This paper reviewed TiPd- and TiPt-based HT-SMAs. The effects of alloying elements on phase transformation, strain recovery, and strength are shown in ternary TiPd and TiPt alloys. In most cases, MTT decreased through the addition of alloying elements, as well as an increase in the alloying element content. The alloying element improved the strength of the martensite and austenite phases by the solid-solution strengthening effect. However, it is difficult to correlate with solid-solution strengthening and strain recovery. A thermal cyclic test was performed to investigate the strain recovery. The irrecoverable strain decreased with the addition of an alloying element, but it was difficult to achieve perfect recovery, even after repeated thermal cyclic tests (training). Multi-component alloys were also investigated. Notably, MTT can be controlled by a combination of the constituent elements. For example, Zr, Ni, and Co decreased MTT, whereas Pt increased MTT. The effect of alloying elements on the recoverable strain differed from that of the MTT. For example, Ni increased the recoverable strain, while Pt and Co decreased it. The extra solid-solution hardening effect relative to ternary alloys was found in the multi-component alloys. Consequently, the irrecoverable strains were smaller in the multi-component alloys than those in the ternary alloys. In many cases, training was effective in achieving perfect recovery in the multi-component alloys; however, a decrease in the transformation strain also occurred. Multi-component alloys can be good candidates for HT-SMAs; the limitations that need to be overcome entail the suppression of the transformation strain reduction and temperature hysteresis increment.

Funding

The study was partly supported by “Precious Metals Research Grant of TANAKA Memorial Foundation”, for which the author expresses her thanks.

Acknowledgments

This work was performed in cooperation with M. Kawakita, A. Wadood, and R. Arockiakumar as a part of their Post-doctoral research; W. Tasaki and W. Takebe as a part of their Bachelors research; H. Sato and H. Matsuda as a part of their Bachelors and Masters research; and B. Ohl, K. Bogdanowicz, E. Muszalska, D. Kuczyska, A. Wierzchowska, A. Chmielewska, K. Aimizu, and M. Yamamoto as a part of their internship program.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. DeCastro, J.A.; Melcher, K.J.; Noebe, R.D. NASA/TM-2005-213834, AIAA-2005-3988. Available online: https://ntrs.nasa.gov/citations/20050203852 (accessed on 30 September 2020).
  2. Stebner, A.; Padula, S., II; Noebe, R.; Lerch, B.; Quinn, D. Development, Characterization, and Design Considerations of Ni19.5Ti50.5Pd25Pt5 High-temperature Shape Memory Alloy Helical Actuators. J. Intell. Mater. Syst. Struct. 2009, 20, 2107–2126. [Google Scholar] [CrossRef]
  3. Benefan, O.; Brown, J.; Calkins, F.T.; Kumar, P.; Stebner, A.P.; Turner, T.L.; Vaidyanathan, R.; Webster, J.; Young, M.L. Shape memory alloy actuator design: CASMART collaborative best practices and case studies. Int. J. Mech. Mater. Des. 2014, 10, 1–42. [Google Scholar] [CrossRef]
  4. Ma, J.; Karaman, I.; Noebe, R.D. High temperature shape memory alloys. Int. Mater. Rev. 2010, 55, 257–315. [Google Scholar] [CrossRef]
  5. Bigelow, G.S.; Padula, S.A., II; Garg, A.; Gaydosh, D.; Noebe, R.D. Characterization of ternary NiTiPd high-temperature shape-memory alloys under load-biased thermal cycling. Metall. Mater. Trans. A 2010, 41, 3065–3079. [Google Scholar] [CrossRef] [Green Version]
  6. “Development of a Numerical Model for High-Temperature Shape Memory Alloys” NASA Technical Reports Server, 2013. Available online: https://ntrs.nasa.gov/citations/20060028493 (accessed on 30 September 2020).
  7. Atli, K.C.; Karaman, I.; Noebe, R.D.; Garg, A.; Chumlyakov, Y.I.; Kireeva, I.V. Improvement in the Shape Memory Response of Ti50.5Ni24.5Pd25 High-Temperature Shape Memory Alloy with Scandium Microalloying. Metall. Mater. Trans. A 2010, 41, 2485–2497. [Google Scholar] [CrossRef]
  8. Atli, K.C.; Karaman, I.; Noebe, R.D.; Maier, H.J. Comparative analysis of the effects of severe plastic deformation and thermomechanical training on the functional stability of Ti50.5Ni24.5Pd25 high-temperature shape memory alloy. Scr. Mater. 2011, 64, 315–318. [Google Scholar] [CrossRef]
  9. Atli, K.C.; Karaman, I.; Noebe, R.D. Work output of the two-way shape memory effect in Ti50.5Ni24.5Pd25 high-temperature shape memory alloy. Scr. Mater. 2011, 65, 903–906. [Google Scholar] [CrossRef]
  10. Atli, K.C.; Karaman, I.; Noebe, R.D.; Garg, A.; Chumlyakov, Y.I.; Kireeva, I.V. Shape memory characteristics of Ti50.5Ni24.5Pd25Sc0.5 high-temperature shape memory alloy after severe plastic deformation. Acta Mater. 2011, 59, 4747–4760. [Google Scholar] [CrossRef]
  11. Yang, F.; Coughlin, D.R.; Phillips, P.J.; Yang, L.; Devaraj, A.; Kovarik, L. Structure analysis of a precipitate phase in an Ni-rich high-temperature NiTiHf shape memory alloy. Acta Mater. 2013, 61, 3335–3346. [Google Scholar] [CrossRef]
  12. Santamarta, R.; Arroyave, R.R.; Pons, J.; Evrigen, A.; Karaman, I.; Karacka, H.E.; Noebe, R.D. TEM study of structural and microstructural characteristics of a precipitate phase in NI-rich Ni-Ti-Hf and Ni-Ti-Zr shape memory alloys. Acta Mater. 2013, 61, 6191–6206. [Google Scholar] [CrossRef]
  13. Bigelow, G.S.; Garg, A.; Padula, S.A., II; Gaydosh, D.J.; Noebe, R.D. Load-biased shape-memory and superelastic properties of a precipitation strengthened high-temperature Ni50.3Ti29.7Hf20 alloy. Scr. Mater. 2011, 64, 725–728. [Google Scholar] [CrossRef]
  14. Coughlin, D.R.; Phillips, P.J.; Bigelow, G.S.; Garg, A.; Noebe, R.D.; Mills, M.J. Characterization of the microstructure and mechanical properties of a 50.3Ni-29.7Ti-20Hf shape memory alloy. Scr. Mater. 2012, 67, 112–115. [Google Scholar] [CrossRef]
  15. Evirgen, A.; Basner, F.; Karaman, I.; Noeve, R.D.; Pons, J.; Santamarta, R. Effect of Aging on the martensitic transformation characteristics of a Ni-rich NiTiHf high temperature shape meory alloys. Funct. Mater. Lett. 2012, 5, 1250038. [Google Scholar] [CrossRef]
  16. Karaca, H.E.; Saghaian, S.M.; Ded, G.; Tobe, H.; Basaran, B.; Maier, H.J.; Noebe, R.D.; Chumlyakov, Y.I. Effects of nanoprecipitation on the shape memory and material properties of an Ni-rich NiTiHf high temperature shape memory alloy. Acta Mater. 2013, 61, 7422–7431. [Google Scholar] [CrossRef]
  17. Benafan, O.; Garg, A.; Noebe, R.D.; Bigelow, G.S.; Padula, S.A.; Gaydosh, D.J.; Schell, N.; Mabe, J.H.; Vaidyanathan, R. Mechanical and functional behavior of Ni-rich Ni50.3Ti29.7Hf20 high temperature shape memory alloy. Intermet 2014, 50, 94–107. [Google Scholar] [CrossRef] [Green Version]
  18. Evirgen, A.; Karaman, I.; Santamarta, R.; Pons, J.; Hayrettin, C.; Noebe, R.D. Relationship between crystallographic compatibility and thermal hysteresis in Ni-rich NiTiHf and NiTiZr high temperature shape memory alloys. Acta Mater. 2016, 121, 374–383. [Google Scholar] [CrossRef] [Green Version]
  19. Saghaian, S.M.; Karaca, H.E.; Tobe, H.; Turabi, A.S.; Saedi, S.; Saghaian, S.E.; Chumlyakov, Y.I.; Noebe, R.D. High strength NiTiHf shape memory alloys with tailorable properties. Acta Mater. 2017, 134, 211–220. [Google Scholar] [CrossRef]
  20. Amin-Ahmadi, B.; Gallmeyer, T.; Pauza, J.G.; Duerig, T.W.; Noebe, R.D.; Stebner, A.P. Effect of a pre-aging treatment on the mechanical behaviors of Ni50.3Ti49.7−xHfx(x < 9 at.%) Shape memory alloys. Scr. Mater. 2018, 147, 11–15. [Google Scholar]
  21. Amin-Ahmadi, B.; Pauza, J.G.; Shamimi, A.; Duerig, T.W.; Noebe, R.D.; Stebner, A.P. Coherency strains of H-phase precipitates and their influence on functional properties of nickel-titanium-hafnium shape memory alloys. Scr. Mater. 2018, 147, 83–87. [Google Scholar] [CrossRef]
  22. Evirgen, A.; Pons, J.; Karaman, I.; Santamarta, R.; Noebe, R.D. H-Phase Precipitation and Martensitic Transformation in Ni-rich Ni-Ti-Hf and Ni-Ti-Zr High-Temperature Shape Memory Alloys. Shap. Mem. Superelasticity 2018, 4, 85–92. [Google Scholar] [CrossRef] [Green Version]
  23. Amin-Ahmadi, B.; Noebe, R.D.; Stebner, A.P. Crack propagation mechanisms of an aged nickel-titanium-hafnium shape memory alloy. Scr. Mater. 2019, 159, 85–88. [Google Scholar] [CrossRef]
  24. Karakoc, O.; Atli, K.C.; Evirgen, A.; Pons, J.; Santamarta, R.; Benafan, O.; Noebe, R.D.; Karaman, I. Effects of training on the thermomechanical behavior of NiTiHf and NiTiZr high temperature shape memory alloys. Mater. Sci. Eng. A 2020, 794, 139857. [Google Scholar] [CrossRef]
  25. ASM Alloy Phase Diagrams Center; Villars, P.; Okamoto, H.; Cenzual, K. (Eds.) ASM International, Materials Park, OH, 2006. Available online: http://www1.asminternational.org/AsmEnterprise/APD (accessed on 4 September 2020).
  26. Donkersloot, H.C.; van Vocht, J.H.N. Martensitic transformations in gold-titanium, palladium-titanium and platinum-titanium alloys near the equiatomic composition. J. Less Common Met. 1970, 20, 83–91. [Google Scholar] [CrossRef]
  27. Biggs, T.; Witcomb, M.J.; Cornish, L.A. Martensite-type transformations in platinum alloys. Mater. Sci. Eng. A 1999, 273–278, 204–207. [Google Scholar] [CrossRef]
  28. Biggs, T.; Corite, M.B.; Witcomb, M.J.; Cornish, L.A. Martensitic Transformations, Microstructure, and Mechanical Workability of TiPt. Metall. Trans. A 2003, 32A, 2267–2272. [Google Scholar] [CrossRef]
  29. Otsuka, K.; Oda, K.; Ueno, Y.; Piao, M.; Ueki, T.; Horikawa, H. The shape memory effect in a Ti50Pd50 alloy. Scr. Met. Mater. 1993, 29, 1355–1358. [Google Scholar] [CrossRef]
  30. Declairieux, C.; Denquin, A.; Ochin, P.; Poriter, R.; Vermaut, P. On the potential of Ti50Au50compound as a high temperature shape memory alloy. Intermetallics 2011, 19, 1461–1465. [Google Scholar] [CrossRef]
  31. Wadood, A.; Hosoda, H.; Yamabe-Mitarai, Y. Phase transformation, osidation and shape memory properties of Ti-50Au-10Zr alloy for high temperature applications. J. Alloys Compd. 2014, 595, 200–205. [Google Scholar] [CrossRef]
  32. Wadood, A.; Yamabe-Mitarai, Y. Silver- and Zirconium-added ternary and quaternary TiAu based high temperature shape memory alloys. J. Alloys Compd. 2015, 646, 1172–1177. [Google Scholar] [CrossRef]
  33. Kawakita, M.; Takahashi, M.; Takahashi, S.; Yamabe-Mitarai, Y. Effect of Zr on phase transformation and high-temperature shape memory effect in TiPd alloys. Mater. Lett. 2012, 89, 336–338. [Google Scholar] [CrossRef]
  34. Yamabe-Mitarai, Y. Development of High-Temperature Shape Memory Alloys above 673 K. Mater. Sci. Forum 2017, 879, 107–112. [Google Scholar] [CrossRef]
  35. Arockiakumar, R.; Takahashi, M.; Takahashi, S.; Yamabe-Mitarai, Y. Microstructure, mechanical and shape memory properties of Ti-55Pd-5x (x = Zr, Hf, V, Nb) alloys. Mater. Sci. Eng. A 2013, 585, 86–93. [Google Scholar] [CrossRef]
  36. Sato, H.; Kim, H.Y.; Shimojo, M.; Yamabe-Mitarai, Y. Training effect on microstructure and shape recovery in Ti-Pd-Zr alloys. Mater. Trans. 2017, 10, 1479–1486. [Google Scholar] [CrossRef] [Green Version]
  37. Yamabe-Mitarai, Y.; Wadood, A.; Arockiakumar, R.; Hara, T.; Takahashi, M.; Takahashi, S.; Hosoda, H. High-Temperature Shape Memory Alloys Based on Ti-Platinum Group Metals compounds. Mater. Sci. Forum 2014, 783–786, 2541–2545. [Google Scholar] [CrossRef]
  38. Yamabe-Mitarai, Y.; Hara, T.; Miura, S.; Hosoda, H. Mechanical properties of Ti-50(Pt, Ir) high-temperature shape memory alloys. Mater. Trans. 2006, 47, 650–657. [Google Scholar] [CrossRef] [Green Version]
  39. Yamabe-Mitarai, Y.; Hara, T.; Miura, S.; Hosoda, H. Phase transformation and shape memory effect of Ti(Pt, Ir). Metall. Trans. A 2012, 43A, 2901–2911. [Google Scholar] [CrossRef]
  40. Yamabe-Mitarai, Y.; Arockiakumar, R.; Wadood, A.; Suresh, K.S.; Kitashima, T.; Hara, T.; Shimojo, M.; Tasaki, W.; Takahashi, M.; Takahashi, S.; et al. Ti(Pt, Pd, Au) based high temperature shape memory alloys. Mater. Today 2015, 2S, S517–S522. [Google Scholar] [CrossRef]
  41. Yamabe-Mitarai, Y.; Hara, T.; Kitashima, T.; Miura, S.; Hosoda, H. Composition dependence of phase transformation behavior and shape memory effect of Ti(Pt, Ir). J. Alloys Compd. 2013, 577S, S399–S403. [Google Scholar] [CrossRef] [Green Version]
  42. Wadood, A.; Takahashi, M.; Takahashi, S.; Hosoda, H.; Yamabe-Mitarai, Y. High-temperature mechanical and shape memory properties of TiPt-Zr and TiPt-Ru alloys. Mater. Sci. Eng. A 2013, 564, 34–41. [Google Scholar] [CrossRef]
  43. Wadood, A.; Yamabe-Mitarai, Y. TiPt-Co and TiPt-Ru high temperature shape memory alloys. Mater. Sci. Eng. A 2014, 610, 106–110. [Google Scholar] [CrossRef]
  44. Wadood, A.; Yamabe-Mitarai, Y. Recent research and developments related to near-equiatomic titanium-platinum alloys for high-temperature applications. Platinum Metals Rev. 2014, 2, 61–67. [Google Scholar] [CrossRef]
  45. Tasaki, W.; Shimojo, M.; Yamabe-Mitarai, Y. Thermal cyclic properties of Ti-Pd-Pt-Zr high-temperature shape memory alloys. Crystals 2019, 9, 595. [Google Scholar] [CrossRef] [Green Version]
  46. Yamabe-Mitarai, Y.; Takebe, W.; Shimojo, M. Phase transformation and shape memory effect of Ti-Pd-Pt-Zr high-temperature shape memory alloys. Shap. Mem. Superelasticity 2017, 3, 381–391. [Google Scholar] [CrossRef]
  47. Yamabe-Mitarai, Y.; Tasaki, W.; Ohl, B.; Sato, H. Effect of Ni, Co, and Pt on Phase transformation and shape recovery of TiPd high temperature shape memory alloys. In Proceedings of the International Conference on Processing & Manufacturing of Advanced Materials, Paris, France, 9–13 July 2018. [Google Scholar]
  48. Yamabe-Mitarai, Y.; Ohl, B.; Bogdanowicz, K.; Muszalska, E. Effect of Ni and Co on Phase Transformation and Shape Memory Effect of Ti-Pd-Zr alloys. Shap. Mem. Superelasticity 2020, 6, 170–180. [Google Scholar] [CrossRef]
  49. Matsuda, H.; Sato, H.; Shimojo, M.; Yamabe-Mitarai, Y. Improvement of high-temperature shap-memory effect by multi-component alloyin for TiPd alloys. Mater. Trans. 2019, 11, 2282–2291. [Google Scholar] [CrossRef]
  50. Matsuda, H.; Shimojo, M.; Murakami, H.; Yamabe-Mitarai, Y. Martensitic Transformation of High-Entropy and Medium-Entropy Shape Memory Alloys. Mater. Sci. Forum 2021, 1016, 1802–1810. [Google Scholar]
  51. Yeh, J.W.; Chen, S.K.; Lin, S.J.; Gan, J.Y.; Chin, T.S.; Shun, T.T.; Tsau, C.H.; Chang, S.Y. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  52. Canter, B.; Chang, I.T.H.; Knight, P.; Vincent, A.J.B. Microstructural development in equiatomic multicomponent alloys. Mater. Sci. Eng. A 2004, 375–377, 213–218. [Google Scholar] [CrossRef]
  53. Firstov, G.S.; Kosorukova, T.A.; Koval, Y.N.; Odnosum, V.V. High entropy shape memory alloys. Mater. Today 2015, 2S, S499–S504. [Google Scholar] [CrossRef]
  54. Chen, C.-H.; Chen, Y.-J. Shape memory characteristics of (TiZrHf)50Ni25Co10Cu15 high entropy shape memory alloy. Scr. Mater. 2019, 162, 185–189. [Google Scholar] [CrossRef]
  55. Canadinc, D.; Trehern, W.; Ma, J.; Karaman, I.; Sun, F.; Chaudhry, Z. Ultra-high temperature multi-component shape memory alloys. Scr. Mater. 2019, 158, 83–87. [Google Scholar] [CrossRef]
  56. Yamabe-Mitarai, Y.; Hara, T.; Phasha, M.; Ngoepe, P.; Chikwanda, H. Phase transformation and crystal structure of IrTi. Intermetallics 2012, 31, 26–33. [Google Scholar] [CrossRef]
  57. Available online: https://en.wikipedia.org/wiki/Atomic_radii_of_the_elements_(data_page) (accessed on 1 September 2020).
  58. Hisada, S.; Matsuda, M.; Yamabe-Mitarai, Y. Shape Change and Crystal Orientation of B19 Martensite in Equiatomic TiPd Alloy by Isobaric Test. Metals 2020, 10, 375. [Google Scholar] [CrossRef] [Green Version]
  59. Atli, K.C.; Franco, E.B.; Karaman, I.; Gaydosh, D.; Noebe, R.D. Influence of crystallographic compatibility on residual strain of TiNi based shape memory alloys during thermo-mechanical cycling. Mater. Sci. Eng. A 2013, 574, 9–16. [Google Scholar] [CrossRef]
  60. Pearson’s Crystal Data—Crystal Structure Database for Inorganic Compounds (on CD-ROM); Villars, P.; Cenzual, K. (Eds.) ASM International: Materials Park, OH, USA, 2010. [Google Scholar]
  61. Karaca, H.E.; Acar, E.; Tobe, H.; Saghaian, S.M. NiTiHf-based shape memory alloys. Mater. Sci. Technol. 2014, 30, 1530–1544. [Google Scholar] [CrossRef]
Figure 1. (a) Austenite finish temperature change for concentration of alloying element of TiPd [33], TiPd-Zr [33,36], TiPd-Hf [34], TiPd-Ru [37], TiPd-Ir [37], and TiPd-Co [37], and (b) martensite transformation temperature for concentration of Pd of Ti50Pd45Zr5, Ti47Pd48Zr5, Ti45Pd50Zr5 [33], and Ti40Pd55Zr5 [35].
Figure 1. (a) Austenite finish temperature change for concentration of alloying element of TiPd [33], TiPd-Zr [33,36], TiPd-Hf [34], TiPd-Ru [37], TiPd-Ir [37], and TiPd-Co [37], and (b) martensite transformation temperature for concentration of Pd of Ti50Pd45Zr5, Ti47Pd48Zr5, Ti45Pd50Zr5 [33], and Ti40Pd55Zr5 [35].
Metals 10 01531 g001
Figure 2. Strain recovery ratio of TiPd [33], Ti45Pd50X5 (X = Zr [33], Hf [34], V [34], Nb [34], Cr [34], and Mo [34]) and Ti50Pd46Y4 (Y = Ru, Co, and Ir) [37] for the periodic table family.
Figure 2. Strain recovery ratio of TiPd [33], Ti45Pd50X5 (X = Zr [33], Hf [34], V [34], Nb [34], Cr [34], and Mo [34]) and Ti50Pd46Y4 (Y = Ru, Co, and Ir) [37] for the periodic table family.
Metals 10 01531 g002
Figure 3. The 0.2% flow stress of the austenite and martensite phases of TiPd [33]: (a) Ti45Pd50X5 (X = Zr [33], Hf [34], V [34], Nb [34], Cr [34], and Mo [34]) and (b) Ti50Pd46Y4 (Y = Ru, Co, and Ir) [37] for the periodic table family.
Figure 3. The 0.2% flow stress of the austenite and martensite phases of TiPd [33]: (a) Ti45Pd50X5 (X = Zr [33], Hf [34], V [34], Nb [34], Cr [34], and Mo [34]) and (b) Ti50Pd46Y4 (Y = Ru, Co, and Ir) [37] for the periodic table family.
Metals 10 01531 g003
Figure 4. Strain–temperature curves of thermal cycle tests of between 15 and 200 MPa for (a) Ti45Zr5Pd50 [45], (b) Ti45Hf5Pd50, (c) Ti45V5Pd50, and (d) Ti45Nb5Pd50.
Figure 4. Strain–temperature curves of thermal cycle tests of between 15 and 200 MPa for (a) Ti45Zr5Pd50 [45], (b) Ti45Hf5Pd50, (c) Ti45V5Pd50, and (d) Ti45Nb5Pd50.
Metals 10 01531 g004
Figure 5. Changes in (a) recoverable strain and irrecoverable strain, and (b) work output for the applied stress of Ti50Pd50 [58], Ti45Pd50Zr5 [45], Ti45Pd50Hf5, and Ti45Pd50V5 derived from the strain–temperature curves shown in Figure 4.
Figure 5. Changes in (a) recoverable strain and irrecoverable strain, and (b) work output for the applied stress of Ti50Pd50 [58], Ti45Pd50Zr5 [45], Ti45Pd50Hf5, and Ti45Pd50V5 derived from the strain–temperature curves shown in Figure 4.
Metals 10 01531 g005
Figure 6. Irrecoverable strain of Ti45Pd50Zr5 and Ti45Pd50Hf5 as a function of number of thermal cycles under 50 MPa.
Figure 6. Irrecoverable strain of Ti45Pd50Zr5 and Ti45Pd50Hf5 as a function of number of thermal cycles under 50 MPa.
Metals 10 01531 g006
Figure 7. (a,b) Recoverable strain, (c,d) irrecoverable strain, and (e,f) work output obtained from strain–temperature curves of thermal cycle tests of between 15 and 200 MPa for Ti45Zr5 Pd50 [45] and multi-component alloys. (a,c,d) TiZrPdNi alloys [47,48], and (b,d,f) TiZrPdPt alloys [45,47,49].
Figure 7. (a,b) Recoverable strain, (c,d) irrecoverable strain, and (e,f) work output obtained from strain–temperature curves of thermal cycle tests of between 15 and 200 MPa for Ti45Zr5 Pd50 [45] and multi-component alloys. (a,c,d) TiZrPdNi alloys [47,48], and (b,d,f) TiZrPdPt alloys [45,47,49].
Metals 10 01531 g007
Figure 8. Work output vs. Af. The open and solid symbols represent the work output of the alloys with imperfect recovery and the work output of the alloys with perfect recovery.
Figure 8. Work output vs. Af. The open and solid symbols represent the work output of the alloys with imperfect recovery and the work output of the alloys with perfect recovery.
Metals 10 01531 g008
Table 1. Martensitic transformation temperature of ternary alloys (°C).
Table 1. Martensitic transformation temperature of ternary alloys (°C).
AlloyAustenite Start Temperature
As
Austenite Finish Temperature
Af
Martensite Start Temperature
Ms
Martensite Finish Temperature
Mf
Temperature Hysteresis
AfMs
Average Decrease in Ms per 1 at%
ΔMs/at%
Ref.
Ti50Pd5056858752751560-[33]
Ti48Pd50Hf253857049447276-
Ti45Pd50Hf548150546042945−15[34]
Ti40Pd50Hf1022730627618830-
Ti40Pd55Hf552855147148780-[35]
Ti47Pd50Zr352053349147042-
Ti45Pd50Zr549250846744541−15[33]
Ti43Pd50Zr749150646444342-[36]
Ti40Pd50Zr1025630822120487-[33,36]
Ti50Pd45Zr532437434431230-
Ti47Pd48Zr542945037033180-
Ti43Pd52Zr5482529393366136-
Ti40Pd55Zr5530593451438142-[35]
Ti45Pd50V550852346244061−15[34]
Ti40Pd55V5437559442398117-[35]
Ti45Pd50Nb548650843140877−20[34]
Ti40Pd55Nb5487535409360126-[35]
Ti45Pd50Ta5 *56157951248967−4[34]
Ti45Pd50Cr52683563472939−37[34]
Ti45Pd50Mo5267383401268−18−27[34]
Ti45Pd50W5 *55857651349263−4[34]
Ti50Pd48Ru256157651651060-[37]
Ti50Pd46Ru444048243639346−23[37]
Ti50Pd42Ru829835433327522-[37]
Ti50Pd37.5Ru12.5------[37]
Ti50Pd25Ru25------[37]
Ti50Pd48Ir255658150850073-[37]
Ti50Pd46Ir452555251048442−4[37]
Ti50Pd42Ir847852147943943-[37]
Ti50Pd38Ir1247150845239556-
Ti50Pd37.5Ir12.547849245243339-[37]
Ti50Pd25Ir25------[37]
Ti50Pd12.5Ir37.5------[37]
Ti50Pd48Co252354547846467-[37]
Ti50Pd46Co447349345542738−18[37]
Ti50Pd42Co837539825932734-[37]
Ti50Pt501000105798996368-[38]
Ti50Pt45Ir5105010811030100851-[39,40]
Ti50Pt37.5Ir12.5103611031000978103-[38]
Ti50Pt25Ir25112111901110106880-[38]
Ti50Pt20Ir30115911891145112744-[39]
Ti50Pt12.5Ir37.5117512181184116934-[38]
Ti55Pt35Ir10931959----[41]
Ti50Pt45Ru592597591385662−20[40,42]
Ti45Pt50Zr593998589784088−23[40,42]
Ti50Pt45Co5959100391385590−15[40,43]
Ti45Pt50Hf595299690585591−17[40,44]
*: melting was not perfect.
Table 2. Martensitic transformation temperature of multi-element alloys.
Table 2. Martensitic transformation temperature of multi-element alloys.
AlloyAsAfMsMfAf–MsΔSmixAlloy ClassificationRef.
Ti45Zr5Pd45Ir5457475422398531.0RLow-entropy alloy
LEA
-
Ti45Zr5Pd35Ir15313348295256531.2RMedium-entropy alloy
MEA
-
Ti45Zr5Pd25Ir25-----1.2RMEA-
Ti45Zr5Pd45Pt5497511464439471.0RLEA[45]
Ti45Zr5Pd35Pt15539557485471721.2RMEA[45]
Ti45Zr5Pd25Pt25628648571539771.2RMEA[45]
Ti45Zr5Pd15Pt35725804713590911.2RMEA[46]
Ti45Zr5Pd5Pt458969357747131121.0RLEA[46]
Ti40Zr10Pd25Pt253904302452202881.3RMEA[47] ST
Ti45Zr4V1Pd50497508409384990.9RLEA[36]
Ti45Zr2.5V2.5Pd505235463933681530.9RLEA[36]
Ti45Zr1V4Pd50496518441424770.9RLEA[36]
Ti45Zr5Pd45Ni5348378328297501.0RLEA[47]
Ti45Zr5Pd40Ni10216258206177521.1RMEA[48]
Ti45Zr5Pd37Ni13114230206108241.2RMEA[47] ST
Ti42Zr8Pd40Ni10-----1.2RMEA[47]
Ti45Zr5Pd30Ni20-----1.2RMEA[48]
Ti45Zr5Pd40Co10303361339298221.1RMEA[48]
Ti45Zr5Pd40Ni8Co2201273212199611.2RMEA[48]
Ti45Zr5Pd40Ni5Co5-----1.2RMEA[48]
Ti45Zr5Pd40Ni2Co8143189135108541.2RMEA[48]
Ti35Zr15Pd20Pt15Ni15-----1.5RHigh-entropy alloy
HEA
[49]
Ti40Zr10Pd20Pt15Ni15-----1.5RHEA[49]
Ti45Zr5Pd20Pt25Ni5559598502432961.3RMEA[49]
Ti45Zr5Pd25Pt20Ni53904302452201851.3RMEA[47] ST
Ti45Zr5Pd20Pt20Ni103744423372561051.3RMEA[49]
Ti35Zr15Pd20Pt15Au15-----1.5RHEA[50]
Ti35Zr15Pd20Pt15Co15-----1.5RHEA[50]
Ti45Zr5Pd25Pt20Au5520590515453751.3RMEA[50]
Ti45Zr5Pd25Pt20Co54195384313241071.3RMEA[50]
Ti16.7Zr16.7Hf16.7Ni25Cu251843382261121121.6RHEA[53]
Ti16.7Zr16.7Hf16.7Ni25Co10Cu15−237136−80351.8RHEA[54]
Ti30Hf20Pd15Ni355376865254791611.3RMEA[55]
Ti25Hf25Pd25Ni256807206205801001.4RMEA[55]
Ti16.7Zr16.7Hf16.7Pd25Ni257407806606201201.6RHEA[55]
Table 3. Ratio of alloys with temperature hysteresis over 100 °C (%).
Table 3. Ratio of alloys with temperature hysteresis over 100 °C (%).
AlloysNumber of Alloys with Temperature Hysteresis over 100 °CNumber of Tested AlloysRatio of Alloys with Temperature Hysteresis over 100 °C (%)
Ternary alloy in Table 154212
LEA in Table 22728
MEA and HEA in Table 282040
Table 4. Strain recovery of ternary alloys.
Table 4. Strain recovery of ternary alloys.
AlloyTest Temp., °CThe Difference from Martensite Transformation Temperature, °CApplied Strain, %Recoverable Strain, %Strain Recovery Ratio, %Ref.
Ti50Pd50538-4.11.640[33]
Ti50Pd50485Mf − 303.82.567[34]
Ti50Pd50380-3.70.4613[33]
Ti50Pd50320-1.240.5948
Ti48Pd50Hf2380-2.80.414
Ti45Pd50Hf5456As − 302.92.174
Ti45Pd50Hf5440-2.72.380[34]
Ti45Pd50Hf5380Mf − 501.91.580
Ti45Pd50Hf5200-5.254.178
Ti40Pd50Hf10380-3.02.480
Ti40Pd55Hf5457Mf − 303.71.745[35]
Ti50Pd45Zr5210-4.280.717
Ti50Pd45Zr5295-3.10.5518
Ti50Pd45Zr5380-3.21.4746
Ti47Pd48Zr5301-4.23.685
Ti47Pd48Zr5399-8.35.769
Ti47Pd48Zr5380-2.21.150
Ti48Pd50Zr2400-6.53.453
Ti48Pd50Zr2380-3.11.343
Ti47Pd50Zr3490As − 303.72.670
Ti47Pd50Zr3415Mf − 303.32.988
Ti47Pd50Zr3380-3.22.887
Ti45Pd50Zr5415Mf − 304.43.581[34]
Ti45Pd50Zr5426As − 303.12.5884[33]
Ti45Pd50Zr5460-85.265
Ti45Pd50Zr5380-3.52,262
Ti45Pd50Zr5380-4.13.994[33]
Ti43Pd52Zr5380-2.21.7581
Ti43Pd52Zr5451As − 304.64.190
Ti40Pd55Zr5490-7.22.027
Ti40Pd55Zr5450-1.20.3934
Ti40Pd55Zr5380-1.90.422
Ti43Pd50Zr7461As − 301.41.284
Ti43Pd50Zr7413Mf − 303.63.6100
Ti43Pd50Zr7380-3.22.990
Ti40Pd50Zr10225-7.74.863[33]
Ti40Pd50Zr10380-1.41.4100
Ti40Pd55Zr5408Mf − 3041.230[35]
Ti45Pd50V5413Mf − 302.41.260[34]
Ti40Pd55V5368Mf − 3030.310[35]
Ti45Pd50Nb5378Mf − 303.11.137[34]
Ti40Pd55Nb5330Mf − 3030.310[35]
Ti45Pd50Ta5 *483Mf − 303.01.135[34]
Ti45Pd50Cr5228Mf − 303.30.412[34]
Ti45Pd50Mo5256Mf − 302.80.414[34]
Ti45Pd50W5 *474Mf − 302.30.522[34]
Ti50Pd46Ru2320---40[37]
Ti50Pd46Ru4320-3.72.052
Ti50Pd42Ru8320-3.42.572
Ti50Pd48Ir2320-2.61.2247[37]
Ti50Pd46Ir4320-2.91.4951[37]
Ti50Pd46Ir4454Mf − 301.01.0100
Ti50Pd46Ir4427Mf − 602.90.5519
Ti50Pd42Ir8320-2.171.2557[37]
Ti50Pd42Ir8408Mf − 301.51.387
Ti50Pd42Ir8427Mf − 606.362.0532
Ti50Pd38Ir12401Mf − 301.41.0172
Ti50Pd48Co2320-2.351.3658[37]
Ti50Pd46Co4320-3.21.447[37]
Ti50Pd42Co8320-3.31.031[37]
Ti50Pt50910Mf − 50600[38]
Ti50Pt50850---11[39]
Ti50Pt45Ir5850---10[39,40]
Ti50Pt37.5Ir12.5928Mf − 5017212[38]
Ti50Pt37.5Ir12.5850---51[39]
Ti50Pt25Ir251018Mf − 5017423[38]
Ti50Pt25Ir25850---57[39]
Ti50Pt20Ir30850---36[39]
Ti50Pt12.5Ir37.51119Mf − 501500[38]
Ti50Pt12.5Ir37.5850---21[39]
Ti55Pt35Ir10850-4.8120[41]
Ti56Pt22Ir22850-2.51.456[41]
Ti58Pt10Ir32850-1.91.473[41]
Ti50Pt45Ru5802Mf − 502.91.345[40,42]
Ti45Pt50Zr5790Mf − 502.11.258[40,42]
Ti50Pt45Co5801Mf − 502.50.936[40,43]
Ti45Pt50Hf5806Mf − 502.41.458[40,44]
*: melting was not perfect.
Table 5. Strain recovery of quaternary alloys.
Table 5. Strain recovery of quaternary alloys.
AlloyTest Temp., °CThe Difference from Martensite Transformation Temperature, °CApplied Strain, %Recoverable Strain, %Strain Recovery Ratio, %
Ti45Pd45Ir5Zr5368Mf − 302.82.796
Ti45Pd35Ir15Zr5226Mf − 302.62.285
Table 6. Strength of the martensite and austenite phases of the binary and ternary alloys.
Table 6. Strength of the martensite and austenite phases of the binary and ternary alloys.
AlloyTest Temp., °CThe Difference from Martensite Transformation Temperature, °CDetwining Stress, MPa0.2% Flow Stress, MPaRef.
Ti50Pd50617Af + 30-82[33]
Ti50Pd50538As − 30231293[33]
Ti50Pd50485Mf − 30249617
Ti45Pd50Hf5534Af + 30-267[34]
Ti45Pd50Hf5456As − 30152794[34]
Ti40Pd55Hf5581Af + 30-341[35]
Ti40Pd55Hf5441Mf − 307251300[35]
Ti47Pd48Zr5301Mf − 303011163
Ti47Pd48Zr5550Af + 30-628
Ti47Pd48Zr5399As − 303021449
Ti48Pd50Zr2400As − 306311504
Ti48Pd50Zr2563Af + 30-467
Ti47Pd50Zr3563Af + 30-183
Ti47Pd50Zr3490As − 30219666
Ti47Pd50Zr3439Mf − 30234863
Ti45Pd50Zr5538Af + 30-274[33]
Ti45Pd50Zr5426As − 30161722[33]
Ti45Pd50Zr5415Mf − 30298937
Ti43Pd50Zr7536Af + 30-245
Ti43Pd50Zr7461As − 30255733
Ti43Pd50Zr7413Mf − 30227954
Ti40Pd50Zr10337Af + 30-386[33]
Ti40Pd50Zr10225As − 30331894[33]
Ti43Pd52Zr5451As − 303711080
Ti43Pd52Zr5337Mf − 305681378
Ti40Pd55Zr5623Af + 30-386[35]
Ti40Pd55Zr5408Mf − 3010301492[35]
Ti45Pd50V5591Af + 30-644[34]
Ti45Pd50V5413Mf − 305211252[34]
Ti40Pd55V5589Af + 30643386[35]
Ti40Pd55V5368Mf − 30744386[35]
Ti45Pd50Nb5625Af + 30-389[34]
Ti45Pd50Nb5378Mf − 30329836[34]
Ti40Pd55Nb5565Af + 30601386[35]
Ti40Pd55Nb5330Mf − 30651386[35]
Ti45Pd50Ta5 *625Af + 30-287[34]
Ti45Pd50Ta5 *483Mf − 30462941[34]
Ti45Pd50Cr5333Af + 30-273[34]
Ti45Pd50Cr5228Mf − 303171000[34]
Ti45Pd50Mo5503Af + 30-364[34]
Ti45Pd50Mo5256Mf − 305121239[34]
Ti45Pd50W5 *615Af + 30-232[34]
Ti45Pd50W5 *474Mf − 305911008[34]
Ti50Pd48Ru2320--301
Ti50Pd46Ru4520Af + 30-150
Ti50Pd46Ru4320-196598
Ti50Pd46Ir4580Af + 30-100
Ti50Pd46Ir4454Mf − 30260740
Ti50Pd46Ir4320-280800
Ti50Pd42Ir8551Af + 3065150
Ti50Pd42Ir8408Mf − 30320720
Ti50Pd38Ir12521-215744
Ti50Pd38Ir12401Mf − 30360850
Ti50Pd46Co4520Af + 30-65
Ti50Pd46Co4320-200710
Ti50Pt501100Af + 50-27[38]
Ti50Pt50910Mf − 50-320[38]
Ti50Pt37.5Ir12.51153Af + 50-42[38]
Ti50Pt37.5Ir12.5928Mf − 50300640[38]
Ti50Pt25Ir251240Af + 50-35[38]
Ti50Pt25Ir251018Mf − 50230450[38]
Ti50Pt12.5Ir37.51268Af + 50-170[38]
Ti50Pt12.5Ir37.51119Mf − 50285435[38]
Ti50Pt45Ru5-Af + 50-100[40,42]
Ti50Pt45Ru5800Mf − 50355712[42]
Ti45Pt50Zr5-Af + 50-111[40,42]
Ti45Pt50Zr5790Mf − 506001468[40,42]
Ti50Pt45Co5-Af + 50-72[40,43]
Ti50Pt45Co5-Mf − 50407523[40,43]
Ti45Pt50Hf5-Af + 50-149[40,44]
Ti45Pt50Hf5-Mf − 505651113[40,44]
*: melting was not perfect.
Table 7. Strength of the martensite and austenite phases of the multi-element alloys.
Table 7. Strength of the martensite and austenite phases of the multi-element alloys.
AlloyTest Temp., °CThe Difference from Martensite Transformation Temperature, °CDetwining Stress, MPa0.2% Flow Stress, MPaRef.
Ti45Pd45Pt5Zr5-Af + 30-281[45]
Ti45Pd45Pt5Zr5425Mf − 30246877[45]
Ti45Pd35Pt15Zr5-Af + 30-315[45]
Ti45Pd35Pt15Zr5442Mf − 303871080[45]
Ti45Pd25Pt25Zr5-Af + 30-231[45]
Ti45Pd25Pt25Zr5509Mf − 304361205[45]
Ti45Zr5Pd20Ni5Pt25628Af + 30-322[49]
Ti45Zr5Pd20Ni5Pt25402Mf − 30-1266[49]
Ti45Zr5Pd20Ni10Pt20472Af + 30-610[49]
Ti45Zr5Pd20Ni10Pt20226Mf − 30-1615[49]
Ti45Zr5Pd25Pt20Au5620Af + 30-267[50]
Ti45Zr5Pd25Pt20Au5423Mf − 30-1169[50]
Ti45Zr5Pd25Pt20Co5568Af + 30-269[50]
Ti45Zr5Pd25Pt20Co5294Mf − 30-741[50]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yamabe-Mitarai, Y. TiPd- and TiPt-Based High-Temperature Shape Memory Alloys: A Review on Recent Advances. Metals 2020, 10, 1531. https://doi.org/10.3390/met10111531

AMA Style

Yamabe-Mitarai Y. TiPd- and TiPt-Based High-Temperature Shape Memory Alloys: A Review on Recent Advances. Metals. 2020; 10(11):1531. https://doi.org/10.3390/met10111531

Chicago/Turabian Style

Yamabe-Mitarai, Yoko. 2020. "TiPd- and TiPt-Based High-Temperature Shape Memory Alloys: A Review on Recent Advances" Metals 10, no. 11: 1531. https://doi.org/10.3390/met10111531

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop