Next Article in Journal
Mechanical Properties, Durability Performance, and Microstructure of CaO-Fly Ash Solidified Sludge from Northeast, China
Previous Article in Journal
Predictions of Lattice Parameters in NiTi High-Entropy Shape-Memory Alloys Using Different Machine Learning Models
Previous Article in Special Issue
Texture-Differentiated Grain Growth in Silicon Steel: Experiments and Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Giant Elastocaloric Effect and Improved Cyclic Stability in a Directionally Solidified (Ni50Mn31Ti19)99B1 Alloy

1
Key Laboratory for Anisotropy and Texture of Materials (Ministry of Education), School of Materials Science and Engineering, Northeastern University, Shenyang 110819, China
2
Western Metal Materials Co., Ltd., Xi’an 710201, China
3
Liaoning Automobile Lightweight Professional Technology Innovation Center, Tieling 112000, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(19), 4756; https://doi.org/10.3390/ma17194756
Submission received: 5 August 2024 / Revised: 24 September 2024 / Accepted: 25 September 2024 / Published: 27 September 2024

Abstract

:
Superelastic shape memory alloys with an integration of large elastocaloric response and good cyclability are crucially demanded for the advancement of solid-state elastocaloric cooling technology. In this study, we demonstrate a giant elastocaloric effect with improved cyclic stability in a <001>A textured polycrystalline (Ni50Mn31Ti19)99B1 alloy developed through directional solidification. It is shown that large adiabatic temperature variation (|ΔTad|) values more than 15 K are obtained across the temperature range from 283 K to 373 K. In particular, a giant ΔTad up to −27.2 K is achieved by unloading from a relatively low compressive stress of 412 MPa at 303 K. Moreover, persistent |ΔTad| values exceeding 8.5 K are sustained for over 12,000 cycles, exhibiting a very low attenuation behavior with a rate of 7.5 × 10−5 K per cycle. The enhanced elastocaloric properties observed in the present alloy are ascribed to the microstructure texturing as well as the introduction of a secondary phase due to boron alloying.

1. Introduction

Refrigeration systems such as air-conditioners and refrigerators are playing an increasingly important role in modern life. Vapor compression remains the predominant technology in refrigeration facilities, yet the high global-warming potential of volatile liquid refrigerants has raised extensive concerns on environmental degradation [1]. Elastocaloric refrigeration, which is based on the elastocaloric effect (eCE) of solid-state materials, offers an eco-friendly cooling alternative that circumvents potential impacts on the global environment [2,3,4]. Shape-memory alloys (SMAs) show great potential as candidate materials for elastocaloric cooling due to their ability to exhibit an elastocaloric response through the utilization of released and absorbed latent heat in association with stress-induced martensitic transformation. Consequently, very pronounced isothermal entropy change (ΔSiso) or adiabatic temperature change (ΔTad) can be expected upon the application or removal of certain uniaxial stresses [5].
Heusler-type Ni-Mn-X (e.g., X = Ga, In, Sn)-based shape-memory alloys have been shown to exhibit various functional behaviors by utilizing the martensitic transformation, such as magnetic shape memory effect [6,7], magnetocaloric effect [8,9], and elastocaloric effect [10]. However, the cycling life of elastocaloric response in these alloys suffers from strong constraint of poor mechanical properties (e.g., 15 cycles for a Ni50Mn33In15.5Cu1.5 alloy [11] and 100 cycles for a Ni50.4Mn27.3Ga22.3 alloy [12]), making them difficult to become reliable elastocaloric refrigerants [10,11,12,13]. The recently developed all-d-metal Heusler-type Ni-Mn-Ti alloys have attracted considerable attention owing to the enhanced mechanical properties that are achieved by the hybridization of d-d atomic orbitals [14]. These alloys exhibit great potential in cooling applications owing to the very large volume change in unit cell (ΔV/V0) between austenite and martensite, which substantially augments the entropy change ΔStr of martensitic transformation and consequently results in very prominent elastocaloric response [15]. In a polycrystalline (Ni50Mn31.5Ti18.5)99.8B0.2 alloy, it has been reported that removing a compressive loading of 700 MPa can yield a colossal adiabatic temperature drop of up to −31.5 K [15].
Although Ni-Mn-Ti alloys exhibit improved mechanical properties and elastocaloric effect when compared to some other typical Ni-Mn-X-based alloys [10,11,16], there remains a lack of sufficient cyclability in polycrystalline alloys, e.g., fewer than 100 cycles in a Ni50Mn31.5Ti18Cu0.5 alloy [17]. The primary cause of this dilemma stems from the constraining effect of adjacent grain boundaries in conventional polycrystalline alloys since the stress-induced martensitic transformation is very sensitive to crystallographic orientation, resulting in deformation incompatibility among grains with different crystallographic orientations. Consequently, significant stress concentration occurs at triple junctions with increasing superelastic cycles, ultimately leading to sample fracture. To enhance the deformation compatibility of neighboring grains during phase transformation, an alternative strategy is to develop a highly textured microstructure with columnar grains through directional solidification [11,18]. Moreover, it is noted that the necessary driving stress for achieving a colossal elastocaloric effect in polycrystalline Ni-Mn-Ti alloys remains too high [14,19]. Since the transformation stress can be effectively manipulated by crystallographic orientation [20,21], microstructure texturing is also expected to reduce the driving stress of martensitic transformation. In addition, alloying new elements is known to be very effective in tailoring the mechanical properties and functional stability [22,23]. Recently, the cyclic number for a (Ni50Mn30.75Ti18.25Cu1)99.8B0.2 polycrystalline alloy has been successfully increased to 650 cycles [24], owing to the substantial increase in grain boundary strength due to boron doping. It is known that the precipitate phases can also be introduced by boron doping [25], which could be exploited for the improvements in mechanical properties and functional stability.
To tackle the tough challenges of relatively low cyclability and high driving stress for the elastocaloric effect in polycrystalline Ni-Mn-Ti alloys, a strategy with the combination of microstructure texturing and boron doping was employed in this work. Considering that boron doping may increase the transformation temperatures [15], a Ni50Mn31Ti19 alloy with relatively low transformation temperature (i.e., Ms = 220 K, Mf = 210 K, As = 236 K, and Af = 243 K) was used as the starting alloy to facilitate the exploration of elastocaloric response around room temperature for the boron doped alloys. For the purpose of strengthening the matrix by utilizing the precipitates, 1% boron was introduced and a <001>A textured polycrystalline (Ni50Mn31Ti19)99B1 alloy with a coarse columnar-grained microstructure was developed by directional solidification. Large |ΔTad| values over 15 K were achieved when removing the compressive loading from 283 K to 373 K. Notably, an impressive ΔTad of −27.2 K was observed at 303 K upon eliminating a relatively low compressive stress of 412 MPa. Moreover, |ΔTad| values exceeding 8.5 K were sustained for over 12,000 cycles, showing improved cyclability owing to the synergistic effects of microstructure texturing and boron doping.

2. Materials and Methods

Polycrystalline (Ni50Mn31Ti19)99B1 (at. %) alloy was fabricated by repeatedly melting high-purity elements in a conventional arc-melting furnace. To obtain a well-oriented microstructure, directional solidification was performed by using the Bridgman method with a temperature gradient of 120 K cm−1 [26], where liquid Ga-In-Sn metal was utilized as the cooling medium. The arc-melted alloy was enveloped in a corundum crucible with an inner diameter of 10 mm and a length of 150 mm, and subsequently remelted at 1563 K. Following a holding time of 1 h, the alloy was directionally solidified by drawing the crucible into a cylinder filled with liquid Ga-In-Sn metal at a constant speed of 3 mm min−1. To eliminate the compositional inhomogeneity, the as-cast alloys were enclosed in vacuum quartz tubes and then subjected to a heat treatment at 1223 K for 48 h. Subsequently, the alloys were rapidly cooled by quenching in cold water. The composition analyses were performed using an energy dispersive spectrometer (EDS). The start and finish temperatures for the forward and reverse martensitic transformation (Ms, Mf, As, Af) were determined through differential scanning calorimetry (DSC-25, TA Instruments, New Castle, USA), using a temperature scanning rate of 10 K min−1. The microstructural features and preferred orientation were characterized using a scanning electron microscope (SEM, JEOL, JSM-IT800, Tokyo, Japan) and the attached electron backscatter diffraction (EBSD) system. To test the mechanical properties, uniaxial compression along the solidification direction (SD) was conducted using a material testing machine with a temperature control unit (AGS-X/50 kN, Shimadzu, Kyoto, Japan). The compressive axis was aligned parallel to the longer edge of a cuboid-shaped sample sized 6 mm × 4 mm × 3 mm. A K-type thermocouple with a resolution of 0.1 K was adhered to the sample surface to track the temperature variation ΔTad caused by rapid loading or unloading. The temperature of the sample was continuously recorded with an interval of 0.1 s. A 15 s dwell time was set between loading and unloading to ensure the thermal equilibrium between the sample and the ambience.

3. Results

DSC measurements demonstrate that the martensitic transformation for the directionally solidified (Ni50Mn31Ti19)99B1 alloy occurs below room temperature (Figure 1a) and the corresponding characteristic temperatures are determined to be Ms = 248 K, Mf = 232 K, As = 255 K, and Af = 269 K. In addition, the entropy change ΔStr related to martensitic transformation can be evaluated to be 66.7 Jkg−1K−1. Figure 1b plots the compressive stress–strain relation tested at 293 K with a strain rate of 2.8 × 10−4 s−1. The directionally solidified alloy can withstand a maximum compressive strength of ~1780 MPa, which is much higher than that of the arc-melted alloy (i.e., 1460 MPa, Supplementary Figure S1). In addition, the compressive strength for the present directionally solidified alloy is also higher than those of some other polycrystalline alloys, such as Ni50Mn31.6Ti18.4 alloy (i.e., 800 MPa) [27], Ni50Mn31.75Ti18.25 alloy (i.e., 1100 MPa) [14], and Ni36.5Co13.5Mn35Ti14.1Gd0.9 alloy (i.e., 1142 MPa) [28]. Achieving enhanced mechanical properties is of great significance for the functional stability of elastocaloric materials [29].
The global microstructure of the directionally solidified (Ni50Mn31Ti19)99B1 alloy is characterized by large columnar grains, with their longer axes along the SD. Figure 2a displays an EBSD orientation map colored in IPF contrast, covering the region of 5.5 mm × 2.2 mm. It is seen that the columnar grains have a homogeneous orientation with <001>A along the SD, i.e., <001>A fiber texture along the SD. As demonstrated in the corresponding inverse pole figure, the texture strength of <001>A fiber texture is 19.6. The formation of such highly textured microstructure should be attributed to the preferential growth of <001>A along the temperature gradient direction [30]. In contrast, the arc-melted alloy exhibits equiaxed grains with a very low degree of preferred orientation (Supplementary Figure S2). Thus, it can be inferred that the enhanced mechanical properties of the directionally solidified alloy should be mainly attributed to the columnar-grained microstructure with a strong <001>A preferred orientation. Such a specific microstructural feature contributes to enhancing the mechanical properties by minimizing stress concentrations at triple junctions and promoting better strain compatibility among adjacent grains. [31]. Microstructural observations also show that certain amounts of precipitated phase appear in the studied alloy, as shown in Figure 2b,c. As determined from the EDS measurements, the matrix has an average composition of Ni50.3Mn30.9Ti18.8, being close to the nominal composition, and the precipitates correspond to a boron-rich phase [32]. Furthermore, a boron-rich secondary phase is observed by TEM, as shown in Figure 2d. It can be observed that the secondary phase is in rod-like shape. The corresponding SAED indicates that the second phase exhibits a face-centered cubic structure (inset of Figure 2d). Similar observations of boron-rich precipitates have also been demonstrated in other boron-doped Ni-Mn-based alloys [13]. The introduction of precipitates has a positive effect in inhibiting crack propagation and alleviating extensive stress concentration [13,24,33,34]. Thus, the enhanced mechanical properties in the present alloy should result from the highly textured microstructure and the introduction of secondary phase due to boron alloying [35,36].
Prior to the characterizations on the superelastic and elastocaloric properties, mechanical training was conducted by performing several cycles of loading–unloading to stabilize stress-induced martensitic transformation (Supplementary Figure S3). Figure 3a presents the superelastic loop for the directionally solidified (Ni50Mn31Ti19)99B1 alloy tested at 293 K. Typical plateau-like superelastic behavior with a stress hysteresis of 105 MPa can be clearly observed. Because of the well-developed <001>A preferential orientation, a large transformation strain of 4.9% is obtained by stress-induced transformation, enabling a recoverable strain of 8.5% upon unloading.
In order to analyze the elastocaloric properties, the ΔTad values obtained by applying/releasing the compressive stress were directly measured. A typical ΔTad profile is displayed in Figure 3b, showing an adiabatic temperature rise on loading and a temperature drop upon unloading. Figure 3c depicts the variation in ΔTad values with compressive strains detected under various strain rates. The |ΔTad| values increase with increasing compressive strain owing to the gradually enhanced proportion of transformed martensite. A rapid increase in |ΔTad| values is presented until the compressive strain of 7% and then the increasing rate becomes slow. When applying a compressive strain of 8%, the ΔTad values resulting from the application and release of compressive stress at a strain rate of 2.2 × 10−2 s−1 reach 17.8 K and −17.2 K, respectively. Figure 3d presents the correlation between ΔTad and strain rate under various compressive strains. Apparently, increasing the strain rate results in a gradual rise in |ΔTad| values on account of the improved adiabatic condition. Thus, both the applied strain and strain rate strongly affect the stress-induced ΔTad values. The corresponding stress–strain curves for ΔTad measurements under various compressive strains and strain rates are displayed in Figure 4a.
It is noted that applying a higher compressive stain and strain rate also widens the stress hysteresis, resulting in enhanced energy dissipation (ΔW) that is represented by the enclosed region of the superelastic loop, as demonstrated in Figure 4. Such energy dissipation is irreversible, thereby weakening the energy conversion efficiency [37]. An important index used for evaluating the energy efficiency is the coefficient of performance of material (COPmat), which can be calculated by Equation (1) [38].
COP mat = | Δ Q / Δ W | = | ( ρ Δ T ad · C p ) / σ d ε |
where ρ denotes the density (ρ = 7278 kg m−3) and Cp is the specific heat capacity (Cp = 423 Jkg−1K−1, Supplementary Figure S4). Based on the ΔTad and ΔW mentioned above, the influence of strain level and strain rate on the COPmat was analyzed, as summarized in Figure 5. Generally, the COPmat values are gradually increased with the strain rate at a constant compressive strain. However, when the strain rate is fixed, the COPmat values are significantly reduced when increasing the compressive strain, even though there is a significant improvement in ΔTad.
In order to examine the influence of testing temperature on the superelastic response, compressive tests were conducted over a temperature range from 273 K to 383 K, by using a compressive strain of 8% at a strain rate of 2.8 × 10−4, as shown in Figure 6a. At each testing temperature, the alloy demonstrates a typical plateau-type superelasticity and a reversible strain of 8% can be clearly observed. With the increase in testing temperature, a higher level of driving stress is needed to activate the martensitic transformation owing to the increased stability of austenite. Consequently, the critical stress σcr shows a linearly increasing trend, with a rate of 6.0 MPa K−1 (inset of Figure 6a). Similar linear dependence has also been reported in some other Ni-Mn-Ti alloys [15,39]. According to the stress–strain curves obtained at different temperatures, the compressive loading-induced ΔSiso was evaluated through the Maxwell relation (Equation (2)).
Δ S iso =   v 0 0 ε σ T ε d ε
In the equation, v0 refers to the specific volume (i.e., 1.38 × 10−4 m3kg−3), as shown in Figure 6b. Under the compressive strain of 8%, a maximum ΔSiso of 38.8 Jkg−1K−1 can be achieved at 303 K.
The unloading ΔTad profiles measured at the temperature span from 273 K to 383 K when removing a compressive strain of 8% are presented in Figure 6c. Here, the unloading process was performed by using a much higher strain rate of 0.3 s−1, thereby allowing the enhanced adiabatic condition. A giant elastocaloric response, characterized by |ΔTad| values exceeding 15 K, was obtained within the testing temperature region from 283 K to 373 K. It is worth noting that the most remarkable elastocaloric response appears at 303 K, showing the maximum ΔTad up to −27.2 K. Such ΔTad is in good agreement with the value evaluated from the ΔSiso (i.e., −27.6 K) based on Equation (3).
Δ T ad cal = Δ S iso · T test / C p
Table 1 compares the elastocaloric properties between the present alloy and some other elastocaloric materials. It can be seen that the |ΔTad| value (i.e., 27.2 K) of the current alloy is higher than those of Ni-Mn-X alloys (X = Ga, In, Sn and Sb) [10,40,41,42,43,44], Cu-based alloys [45,46], and Ni-Ti alloys [46,47], but lower than those of certain Ni-Mn-Ti alloys. It should be noted that the colossal elastocaloric effect in those Ni-Mn-Ti alloys requires a quite large driving stress, indicating a high level of energy consumption. By comparison, the required external stress for the present alloy is substantially decreased, i.e., 412 MPa, thereby favoring the miniaturization and energy-saving of refrigeration equipment [48]. To quantify the elastocaloric response, the adiabatic temperature change under unit stress was determined by |ΔTad/σmax|. Notably, the maximum |ΔTad/σmax| of the present alloy reaches 78.8 K GPa−1, owing to a considerably large ΔTad of −24.9 K realized at a relatively low compressive stress of 316 MPa at 293 K. By comparison, the present |Tad/σmax| is also larger than those of some other elastocaloric materials [3,10,11,15,16,20,31,39,40,49,50,51,52,53,54,55,56], as illustrated in Figure 6d.
Figure 6. (a) Temperature-dependent superelasticity of the present alloy subjected to a strain rate of 2.8 × 10−4 s−1 by applying a maximum compressive strain of 8%. The inset displays the evolution of critical stress σcr as the variation of temperature. (b) ΔSiso induced by compressive loading under various compressive strains. (c) ΔTad profiles upon unloading measured at various temperatures. (d) Comparison of the elastocaloric properties between the present alloy and several typical elastocaloric materials [3,10,11,15,16,20,31,39,40,49,50,51,52,53,54,55,56].
Figure 6. (a) Temperature-dependent superelasticity of the present alloy subjected to a strain rate of 2.8 × 10−4 s−1 by applying a maximum compressive strain of 8%. The inset displays the evolution of critical stress σcr as the variation of temperature. (b) ΔSiso induced by compressive loading under various compressive strains. (c) ΔTad profiles upon unloading measured at various temperatures. (d) Comparison of the elastocaloric properties between the present alloy and several typical elastocaloric materials [3,10,11,15,16,20,31,39,40,49,50,51,52,53,54,55,56].
Materials 17 04756 g006
To investigate the cyclability of eCE for the present alloy, cyclic tests under a compressive strain of 4% were performed by using a strain rate of 1.1 × 10−2 s−1. Figure 7a displays the evolution of stress–strain correlation during the cyclic tests. The superelastic response demonstrates a good stability, with a small stress drop of 19 MPa over 12,000 cycles of testing. Such a stress drop could be caused by the accumulation of transformation-induced defects that stabilize the martensite phase [55,58]. It is noted that there is no indication of fracture on the sample after 12,000 cycles, which is also a reflection of enhanced mechanical properties in the present directionally solidified alloys. Figure 7b presents the representative ΔTad profiles corresponding to the various selected cycles. The present alloy exhibits persistent |ΔTad| values larger than 8.5 K for more than 12,000 cycles. The degradation rate of |ΔTad| values during such long-term tests is very low, i.e., 7.5 × 10−5 K per cycle. For a purpose of comparative analyses, the cyclic tests were also conducted in a directionally solidified Ni50Mn31Ti19 alloy without boron doping under the same testing parameters. As shown in Figure S5a, the stress–strain curves demonstrate a significant stress drop of 167 MPa after 40 mechanical cycles for the sample. Additionally, the |ΔTad| values during unloading also exhibit a sharp decrease from 8.4 K to 4.1 K (Supplementary Figure S5b). Thus, boron doping indeed yields a prominent cyclability of the elastocaloric response. As summarized in Table 2, both the cycle number and degradation rate for the present alloy are substantially improved when compared to some typical elastocaloric materials reported in the literature [12,20,39,45,46,53,57,59,60,61,62,63,64,65,66,67], thereby showing an enhanced functional stability of the elastocaloric response.
To further understand the degradation of functional behaviors, the DSC curves between the initial sample and the sample after mechanical cycles were compared, as shown in Figure 8. It is seen that mechanical cycling leads to a broadening of the endothermic/exothermic peak, which could be an indication of the accumulation of dislocations and the resulting internal stress [68]. In addition, ΔStr is also reduced after long-term testing, i.e., from 66.7 Jkg−1K−1 for the initial sample to 59.8 Jkg−1K−1 in the cycled sample, suggesting the generation of residual martensite induced by mechanical cycling.
Figure 9a shows the microstructure image for the longitudinal section of the directionally solidified (Ni50Mn31Ti19)99B1 alloy after 12,000 mechanical cycles. It is noted that some longitudinal cracks have nucleated and propagated along the loading direction. The occurrence of these cracks can be attributed to the generation of stress concentration at the grain boundary during mechanical cycling. With increasing the cycles, cracks nucleate at the grain boundary and subsequently propagate into the interior of the grains, as highlighted in red dashed squares in Figure 9a. It is worth mentioning that residual martensite can also observed in the alloy, as shown in Figure 9b, which should account for the reduced ΔStr observed in the DSC measurements after mechanical cycling. Accordingly, the reduction in the volume fraction of transforming austenite involved in the phase transition results in the attenuation of ΔTad during mechanical cycling.
To elucidate the underlying fatigue mechanisms behind long-term cyclic testing, the morphological characteristics of the fractured sample were analyzed. Figure 9c displays the morphology of the longitudinal section for the directionally solidified (Ni50Mn31Ti19)99B1 alloy after 13,000 mechanical cycles. The abundance of river patterns and cleavage steps suggests that the sample exhibits a transgranular fracture behavior, which aligns with the observed crack propagation observed in Figure 9a. Moreover, the cross-section exhibits numerous white granular precipitates that are dispersedly distributed along the grain boundary and within the grain. This provides clear evidence of enhanced cyclic stability by boron doping. These precipitates effectively mitigate deviatoric stress through plastic deformation, thereby impeding crack initiation and propagation [34].
The results presented above indicate that the present directionally solidified alloy exhibits a giant elastocaloric response and improved cyclic stability, which should be attributed to the effective microstructure tuning through directional solidification and boron doping. It is known that during mechanical cycling, dislocations and microcracks tend to preferentially form at fragile grain boundaries, resulting in premature intergranular fracture [69]. By utilizing directional solidification, a coarse-grained microstructure with <001>A preferential orientation formed in the present alloy significantly reduces the amount of grain boundaries, thereby minimizing crack initiation sites and enabling prolonged service life [70,71]. Moreover, the well-oriented microstructure is also favorable for the improvement in the deformation compatibility during stress-induced phase transformation, which thus weakens the grain boundary constraints on the lattice deformation and contributes to a giant elastocaloric response under reduced driving stress [11]. On the other hand, boron is recognized as an electron donor [72,73], and its microalloying can effectively enhance the cohesive strength of grain boundaries, thereby increasing the mechanical properties by impeding the intergranular fracture. This effect has been observed in various structural intermetallic compounds such as Ni3Al [74], Ni3Fe [75], and FeAl [76] alloys. Here, a boron-rich secondary phase is also introduced in the present alloy due to boron doping. This secondary phase is distributed within the grain and along the grain boundary (Figure 2c), which is favorable for enhancing the mechanical properties and thus extending the service life. Overall, the enhanced elastocaloric response for the present directionally solidified alloy should primarily arise from the synergy of microstructure texturing and boron doping.

4. Conclusions

In summary, the elastocaloric effect and its functional stability in a <001>A textured polycrystalline (Ni50Mn31Ti19)99B1 alloy fabricated by directional solidification were investigated. Through boron doping and microstructure control, the mechanical properties were significantly enhanced, giving rise to the maximum compressive strength of 1780 MPa. Consequently, a giant elastocaloric response represented by |ΔTad| values higher than 15 K was realized over a wide temperature window ranging from 283 K to 373 K. Because of the reduced driving stress derived from the highly textured microstructure, a very high value of specific adiabatic temperature change (|ΔTad/σmax|) up to 78.8 K GPa−1 was achieved. Moreover, persistent |ΔTad| values exceeding 8.5 K were sustained for over 12,000 cycles, showing enhanced cyclability when compared to some other elastocaloric materials. It is demonstrated that the combination of microstructure texturing and boron doping can be exploited as an effective approach in enhancing the cyclic stability of the elastocaloric effect.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17194756/s1, Figure S1: Compressive stress–strain curve for the arc-melted (Ni50Mn31Ti19)99B1 alloy measured at 293 K; Figure S2: EBSD orientation micrograph for the arc-melted (Ni50Mn31Ti19)99B1 alloy and corresponding inverse pole figure; Figure S3: Compressive stress–strain curves for the directionally solidified (Ni50Mn31Ti19)99B1 alloy during 10 cycles of superelastic training; Figure S4: Temperature dependence of the specific heat capacity Cp measured on heating for the directionally solidified alloy; Figure S5: Cyclic tests for the directionally solidified Ni50Mn31Ti19 alloy.

Author Contributions

Conceptualization, H.W. and Z.L.; validation, B.Y. and H.Y.; investigation, H.W.; data curation, Y.W., G.Z., and J.Y.; writing—original draft preparation, H.W.; writing—review and editing, Z.L. and L.Z.; supervision, J.L.; project administration, Z.L.; funding acquisition, L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant Nos. 52371006, 52171005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

Author Jiajing Yang was employed by the Western Metal Materials Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships.

References

  1. Lloveras, P.; Stern-Taulats, E.; Barrio, M.; Tamarit, J.L.; Crossley, S.; Li, W.; Pomjakushin, V.; Planes, A.; Mañosa, L.; Mathur, N.D.; et al. Giant barocaloric effects at low pressure in ferrielectric ammonium sulphate. Nat. Commun. 2015, 6, 8801. [Google Scholar] [CrossRef] [PubMed]
  2. Mañosa, L.; Planes, A.; Acet, M. Advanced materials for solid-state refrigeration. J. Mater. Chem. A 2013, 1, 4925–4936. [Google Scholar] [CrossRef]
  3. Tušek, J.; Engelbrecht, K.; Millán-Solsona, R.; Mañosa, L.; Vives, E.; Mikkelsen, L.P.; Pryds, N. The elastocaloric effect: A way to cool efficiently. Adv. Energy Mater. 2015, 5, 1500361. [Google Scholar] [CrossRef]
  4. Mañosa, L.; Planes, A. Materials with giant mechanocaloric effects: Cooling by strength. Adv. Mater. 2017, 29, 1603607. [Google Scholar] [CrossRef]
  5. Yang, J.; Wang, H.; Li, Z.; Zou, N.; Yan, H.; Yang, B.; Zuo, L. Crystallography of stress-induced martensitic transformation and giant elastocaloric effect in a <001>A textured Ni27Cu21Mn46Sn6 shape memory alloy. Acta Mater. 2024, 263, 119546. [Google Scholar]
  6. Heczko, O.; Cejpek, P.; Drahokoupil, J.; Holý, V. Structure and microstructure of Ni-Mn-Ga single crystal exhibiting magnetic shape memory effect analysed by high resolution X-ray diffraction. Acta Mater. 2016, 115, 250–258. [Google Scholar] [CrossRef]
  7. Heczko, O.; Kopeček, J.; Straka, L.; Seiner, H. Differently mobile twin boundaries and magnetic shape memory effect in 10M martensite of Ni–Mn–Ga. Mater. Res. Bull. 2013, 48, 5105–5109. [Google Scholar] [CrossRef]
  8. Liu, J.; Gottschall, T.; Skokov, K.P.; Moore, J.D.; Gutfleisch, O. Giant magnetocaloric effect driven by structural transitions. Nat. Mater. 2012, 11, 620–626. [Google Scholar] [CrossRef]
  9. Gottschall, T.; Skokov, K.P.; Frincu, B.; Gutfleisch, O. Large reversible magnetocaloric effect in Ni-Mn-In-Co. Appl. Phys. Lett. 2015, 106, 118694. [Google Scholar] [CrossRef]
  10. Zhang, G.Y.; Li, Z.B.; Yang, J.J.; Yang, B.; Wang, D.H.; Zhang, Y.D.; Esling, C.; Hou, L.; Li, X.; Zhao, X.; et al. Giant elastocaloric effect in a Mn-rich Ni44Mn46Sn10 directionally solidified alloy. Appl. Phys. Lett. 2020, 116, 023902. [Google Scholar] [CrossRef]
  11. Wang, H.; Li, D.; Zhang, G.; Li, Z.; Yang, B.; Yan, H.; Cong, D.; Esling, C.; Zhao, X.; Zuo, L. Highly sensitive elastocaloric response in a directionally solidified Ni50Mn33In15.5Cu1.5 alloy with strong <001>A preferred orientation. Intermetallics 2022, 140, 107379. [Google Scholar]
  12. Wei, L.; Zhang, X.; Liu, J.; Geng, L. Orientation dependent cyclic stability of the elastocaloric effect in textured Ni-Mn-Ga alloys. AIP Adv. 2018, 8, 055312. [Google Scholar] [CrossRef]
  13. Yang, Z.; Cong, D.Y.; Sun, X.M.; Nie, Z.H.; Wang, Y.D. Enhanced cyclability of elastocaloric effect in boron-microalloyed Ni-Mn-In magnetic shape memory alloys. Acta Mater. 2017, 127, 33–42. [Google Scholar] [CrossRef]
  14. Yan, H.L.; Wang, L.D.; Liu, H.X.; Huang, X.M.; Jia, N.; Li, Z.B.; Yang, B.; Zhang, Y.D.; Esling, C.; Zhao, X.; et al. Giant elastocaloric effect and exceptional mechanical properties in an all-d-metal Ni–Mn–Ti alloy: Experimental and ab-initio studies. Mater. Des. 2019, 184, 108180. [Google Scholar] [CrossRef]
  15. Cong, D.Y.; Xiong, W.X.; Planes, A.; Ren, Y.; Manosa, L.; Cao, P.Y.; Nie, Z.H.; Sun, X.M.; Yang, Z.; Hong, X.F.; et al. Colossal elastocaloric effect in ferroelastic Ni-Mn-Ti alloys. Phys. Rev. Lett. 2019, 122, 255703. [Google Scholar] [CrossRef] [PubMed]
  16. Li, Y.; Sun, W.; Zhao, D.; Xu, H.; Liu, J. An 8 K elastocaloric temperature change induced by 1.3% transformation strain in Ni44Mn45−xSn11Cux alloys. Scr. Mater. 2017, 130, 278–282. [Google Scholar] [CrossRef]
  17. Ma, G.; Li, C.; Chen, M.; Zong, S.; Zhang, Y.; Zhao, S.; Chen, F.; Xuan, H. Elastocaloric effect and magnetic properties of Ni50Mn31.5Ti18Cu0.5 shape memory alloy. J. Supercond. Novel Magn. 2022, 35, 1669–1676. [Google Scholar] [CrossRef]
  18. Fornell, J.; Tuncer, N.; Schuh, C.A. Orientation dependence in superelastic Cu-Al-Mn-Ni micropillars. J. Alloys Compd. 2017, 693, 1205–1213. [Google Scholar] [CrossRef]
  19. Villa, F.; Villa, E.; Righi, L.; Ruggieri, P.; Bennato, N.; Battiston, S.; Passaretti, F.; Casati, R. Effect of the thermal processing on the microstructural, functional and mechanical properties of cast polycrystalline NiMnTi alloys. J. Alloys Compd. 2024, 1000, 175099. [Google Scholar] [CrossRef]
  20. Xiao, F.; Jin, M.J.; Liu, J.; Jin, X.J. Elastocaloric effect in Ni50Fe19Ga27Co4 single crystals. Acta Mater. 2015, 96, 292–300. [Google Scholar] [CrossRef]
  21. Fazeli, S.; Izadifar, M.; Dolado, J.S.; Ramazani, A.; Sadrnezhaad, S.K. Atomistic study of the effect of crystallographic orientation on the twinning and detwinning behavior of NiTi shape memory alloys. Comput. Mater. Sci. 2022, 203, 111080. [Google Scholar] [CrossRef]
  22. Hsu, Y.T.; Wu, C.T.; Chen, C.H. Nanoscale-precipitate-strengthened (Ni,Cu)-rich TiNiCu shape memory alloy with stable superelasticity and elastocaloric performance. J. Alloys Compd. 2024, 997, 174937. [Google Scholar] [CrossRef]
  23. Meng, J.; Xie, L.; Yu, Q.; Wang, J.; Jiang, C. Toughening the grain boundaries by introducing a small amount of the second phase: Ni-Cu-Mn-Ga shape memory alloys as an example. Acta Mater. 2024, 263, 119469. [Google Scholar] [CrossRef]
  24. Guan, Z.; Bai, J.; Zhang, Y.; Sun, S.; Gu, J.; Liang, X.; Zhang, Y.; Esling, C.; Zhao, X.; Zuo, L. Achieved good mechanical properties and large elastocaloric effect in Ni-Mn-Ti-Cu-B alloy: Experiments and first-principles calculations. J. Alloys Compd. 2023, 930, 167477. [Google Scholar] [CrossRef]
  25. Zhang, Y.; Yang, S.; Wang, L.; Pan, S.; Zhang, J.; Liu, X.; Wang, C. Development of boron-microalloyed Co–V–Al–Fe shape memory alloys. Intermetallics 2023, 157, 107889. [Google Scholar] [CrossRef]
  26. Li, X.; Fautrelle, Y.; Gagnoud, A.; Du, D.; Wang, J.; Ren, Z.; Nguyen-Thi, H.; Mangelinck-Noel, N. Effect of a weak transverse magnetic field on solidification structure during directional solidification. Acta Mater. 2014, 64, 367–381. [Google Scholar] [CrossRef]
  27. Li, B.; Li, S.M.; Yang, B.; Zhong, H.; Liu, Z.P.; Li, D. Enhancing the elastocaloric effect and thermal cycling stability in dendritic-like Ni50Mn31.6Ti18.4 single crystal. J. Alloys Compd. 2023, 936, 168310. [Google Scholar] [CrossRef]
  28. Xu, F.; Zhu, C.; Wang, J.; Luo, F.; Zhu, X.; Xu, J.; Chen, S.; Wang, J.; Ma, G.; Chen, F.; et al. Enhanced elastocaloric effect and mechanical properties of Gd-doped Ni-Co-Mn-Ti-Gd metamagnetic shape memory alloys. J. Alloys Compd. 2023, 960, 170768. [Google Scholar] [CrossRef]
  29. Li, B.; Zheng, L.J.; Zhang, H. Microstructure-property relationship in Zr-alloyed Ni-rich NiTi alloys: Enhancements in high-temperature stability and superelasticity. Mater. Charact. 2024, 208, 113649. [Google Scholar] [CrossRef]
  30. Zhao, Y.; Hou, L.; Li, X.; Su, H.; Zhang, J. On the formation of gradient-distributed dendrites in a single crystal nickel-based superalloy directionally solidified under transverse static magnetic field. Mater. Charact. 2024, 207, 113508. [Google Scholar] [CrossRef]
  31. Niu, Y.; Chen, H.; Zhang, X.; Li, S.; Cong, D.; Ma, T.; Li, S.; Lin, J.; Wang, Y.D. Achieving excellent superelasticity and extraordinary elastocaloric effect in a directionally solidified Co-V-Ga alloy. Scr. Mater. 2021, 204, 114123. [Google Scholar] [CrossRef]
  32. Zhao, Y.; Ming, F.; Jia, N.; Chen, J.; Ren, S.; Xu, W.; Qu, X. High-strength superelastic as-cast Ni50.9Ti49.1-TiB2 in-situ composites. Mater. Sci. Eng. A 2021, 818, 141451. [Google Scholar] [CrossRef]
  33. Pfeuffer, L.; Lemke, J.; Shayanfar, N.; Riegg, S.; Koch, D.; Taubel, A.; Scheibel, F.; Kani, N.A.; Adabifiroozjaei, E.; Molina-Luna, L.; et al. Microstructure engineering of metamagnetic Ni-Mn-based Heusler compounds by Fe-doping: A roadmap towards excellent cyclic stability combined with large elastocaloric and magnetocaloric effects. Acta Mater. 2021, 221, 117390. [Google Scholar] [CrossRef]
  34. Zhang, X.; Chen, H.; Li, S.; Niu, Y.; Yin, T.; Song, C.; Lang, R.; Cong, D.; Li, S.; Wang, Y.D. Enhanced cyclability of superelasticity and elastocaloric effect in Cu and B co-doped Co-Ni-Ga shape memory alloys. J. Alloys Compd. 2022, 918, 165633. [Google Scholar] [CrossRef]
  35. Zhang, X.; Zhang, M.; Cui, T.; Li, J.; Liu, Q.; Wang, H. The enhancement of the mechanical properties and the shape memory effect for the Cu-13.0Al-4.0Ni alloy by boron addition. J. Alloys Compd. 2019, 776, 326–333. [Google Scholar] [CrossRef]
  36. Li, X.W.; Yang, M.H.; Sun, C.H.; Ruan, Y.; Wei, B. Enhanced tensile performance of ternary Fe-Ni-Ti peri-eutectic alloy through directional solidification associated with thermal processing. Mater. Lett. 2024, 369, 136743. [Google Scholar] [CrossRef]
  37. Qian, S.X.; Ling, J.Z.; Hwang, Y.; Radermacher, R.; Takeuchi, I. Thermodynamics cycle analysis and numerical modeling of thermoelastic cooling systems. Int. J. Refrig. 2015, 56, 65–80. [Google Scholar] [CrossRef]
  38. Moya, X.; Defay, E.; Heine, V.; Mathur, N.D. Too cool to work. Nat. Phys. 2015, 11, 202–205. [Google Scholar] [CrossRef]
  39. Zhang, G.; Wang, H.; Li, Z.; Yang, B.; Yan, H.; Zuo, L. Giant elastocaloric effect covering a wide temperature region in a directionally solidified Ni50Mn30Ti20 alloy. Scr. Mater. 2023, 237, 115725. [Google Scholar] [CrossRef]
  40. Li, D.; Li, Z.; Yang, J.; Li, Z.; Yang, B.; Yan, H.; Wang, D.; Hou, L.; Li, X.; Zhang, Y.; et al. Large elastocaloric effect driven by stress-induced two-step structural transformation in a directionally solidified Ni55Mn18Ga27 alloy. Scr. Mater. 2019, 163, 116–120. [Google Scholar] [CrossRef]
  41. Zhao, D.; Castán, T.; Planes, A.; Li, Z.; Sun, W.; Liu, J. Enhanced caloric effect induced by magnetoelastic coupling in NiMnGaCu Heusler alloys: Experimental study and theoretical analysis. Phys. Rev. B 2017, 96, 224105. [Google Scholar] [CrossRef]
  42. Huang, X.M.; Zhao, Y.; Yan, H.L.; Tang, S.; Yang, Y.; Jia, N.; Yang, B.; Li, Z.; Zhang, Y.; Esling, C.; et al. A first-principle assisted framework for designing high elastocaloric Ni–Mn-based magnetic shape memory alloy. J. Mater. Sci. Technol. 2023, 134, 151–162. [Google Scholar] [CrossRef]
  43. Sun, W.; Liu, J.; Zhao, D.; Zhang, M. Directional solidification and elastocaloric effect in a Ni45Mn44Sn11 magnetic shape memory alloy. J. Phys. D Appl. Phys. 2017, 50, 444001. [Google Scholar] [CrossRef]
  44. Li, Z.; Li, Z.; Yang, J.; Li, D.; Yang, B.; Yan, H.; Nie, Z.; Hou, L.; Li, X.; Zhang, Y.; et al. Large elastocaloric effect in a polycrystalline Ni45.7Co4.2Mn37.3Sb12.8 alloy with low transformation strain. Scr. Mater. 2019, 162, 486–491. [Google Scholar] [CrossRef]
  45. Wang, Y.; Liu, C.; Wang, H.; Li, Z.; Li, J.; Yang, B.; Yan, H.; Zuo, L. Orientation dependence of elastocaloric effect in a Cu71.3Al17.5Mn11.2 single crystal. J. Alloys Compd. 2023, 969, 172392. [Google Scholar] [CrossRef]
  46. Wu, Y.; Ertekin, E.; Sehitoglu, H. Elastocaloric cooling capacity of shape memory alloys—Role of deformation temperatures, mechanical cycling, stress hysteresis and inhomogeneity of transformation. Acta Mater. 2017, 135, 158–176. [Google Scholar] [CrossRef]
  47. Tušek, J.; Engelbrecht, K.; Mikkelsen, L.P.; Pryds, N. Elastocaloric effect of Ni-Ti wire for application in a cooling device. J. Appl. Phys. 2015, 117, 124901. [Google Scholar] [CrossRef]
  48. Qian, S.X.; Geng, Y.L.; Wang, Y.; Muehlbauer, J.; Ling, J.Z.; Hwang, Y.; Radermacher, R.; Takeuchi, I. Design of a hydraulically driven compressive elastocaloric cooling system. Sci. Technol. Built Environ. 2016, 22, 500–506. [Google Scholar] [CrossRef]
  49. Wang, J.; Yu, Q.; Xu, K.; Zhang, C.; Wu, Y.; Jiang, C. Large room-temperature elastocaloric effect of Ni57Mn18Ga21In4 alloy undergoing a magnetostructural coupling transition. Scr. Mater. 2017, 130, 148–151. [Google Scholar]
  50. Masdeu, F.; Pons, J.; Torrens-Serra, J.; Chumlyakov, Y.; Cesari, E. Superelastic behavior and elastocaloric effect in a Ni51.5Fe21.5Ga27.0 ferromagnetic shape memory single crystal under compression. Mater. Sci. Eng. A 2022, 833, 142362. [Google Scholar] [CrossRef]
  51. Li, Z.Z.; Li, Z.B.; Li, D.; Yang, J.J.; Yang, B.; Hu, Y.; Wang, D.H.; Zhang, Y.D.; Esling, C.; Zhao, X.; et al. Achieving a broad refrigeration temperature region through the combination of successive caloric effects in a multiferroic Ni50Mn35In15 alloy. Acta Mater. 2020, 192, 52–59. [Google Scholar] [CrossRef]
  52. Zhao, D.W.; Liu, J.; Chen, X.; Sun, W.; Li, Y.; Zhang, M.X.; Shao, Y.Y.; Zhang, H.; Yan, A. Giant caloric effect of low-hysteresis metamagnetic shape memory alloys with exceptional cyclic functionality. Acta Mater. 2017, 133, 217–223. [Google Scholar] [CrossRef]
  53. Yuan, B.; Zhong, S.; Qian, M.; Zhang, X.; Geng, L. Elastocaloric effect in bamboo-grained Cu71.1Al17.2Mn11.7 microwires. J. Alloys Compd. 2021, 850, 156612. [Google Scholar] [CrossRef]
  54. Xu, S.; Huang, H.; Xie, J.; Takekawa, S.; Xu, X.; Omori, T.; Kainuma, R. Giant elastocaloric effect covering wide temperature range in columnar-grained Cu71.5Al17.5Mn11 shape memory alloy. APL Mater. 2016, 4, 106106. [Google Scholar] [CrossRef]
  55. Liu, C.; Li, Z.B.; Wang, H.L.; Wang, Y.P.; Yang, B.; Yan, H.L.; Cong, D.Y.; Zhao, X.; Zuo, L. Long-term stable elastocaloric effect in a Heusler-type Co51V33Ga16 polycrystalline alloy. ACS Appl. Energy Mater. 2022, 5, 12953–12965. [Google Scholar] [CrossRef]
  56. Zhou, M.; Li, Y.; Zhang, C.; Li, L. Elastocaloric effect and mechanical behavior for NiTi shape memory alloys. Chin. Phys. B 2018, 27, 106501. [Google Scholar] [CrossRef]
  57. Liu, S.; Xuan, H.; Cao, T.; Wang, L.; Xie, Z.; Liang, X.; Li, H.; Feng, L.; Chen, F.; Han, P. Magnetocaloric and elastocaloric effects in all-d-metal Ni37Co9Fe4Mn35Ti15 magnetic shape memory alloy. Phys. Status Solidi 2019, 216, 1900563. [Google Scholar] [CrossRef]
  58. Sidharth, R.; Abuzaid, W.; Vollmer, M.; Niendorf, T.; Sehitoglu, H. Fatigue crack initiation in the iron-based shape memory alloy FeMnAlNiTi. Shape Mem. Superelasticity 2020, 6, 323–331. [Google Scholar] [CrossRef]
  59. Xu, Y.; Lu, B.; Sun, W.; Yan, A.; Liu, J. Large and reversible elastocaloric effect in dual-phase Ni54Fe19Ga27 superelastic alloys. Appl. Phys. Lett. 2015, 106, 201903. [Google Scholar] [CrossRef]
  60. Imran, M.; Zhang, X.; Qian, M.; Geng, L. Enhanced working stability of elastocaloric effects in polycrystalline Ni-Fe-Ga dual phase alloy. Intermetallics 2021, 136, 107255. [Google Scholar] [CrossRef]
  61. Yuan, B.; Qian, M.; Zhang, X.; Imran, M.; Geng, L. Enhanced cyclic stability of elastocaloric effect in oligocrystalline Cu–Al–Mn microwires via cold-drawing. Int. J. Refrig 2020, 114, 54–61. [Google Scholar] [CrossRef]
  62. Liu, K.; Yuan, Y.; Ma, S.; Feng, G.; Wan, D.; Wang, S.; Chen, C.; Luo, X.; Zhong, Z. Large elastocaloric effect around room temperature in directionally solidified Co49Fe3V33Ga15 superelastic alloy. J. Alloys Compd. 2021, 884, 161094. [Google Scholar] [CrossRef]
  63. Liu, C.; Li, D.; Li, Z.; Yang, B.; Yan, H.; Li, J.; Li, Z.; Zhao, X.; Zuo, L. Large elastocaloric effect in a Heusler-type Co50V35Ga14Ni1 polycrystalline alloy. Appl. Phys. Lett. 2021, 118, 103904. [Google Scholar] [CrossRef]
  64. Yang, Z.; Cong, D.; Yuan, Y.; Li, R.; Zheng, H.; Sun, X.; Nie, Z.; Ren, Y.; Wang, Y. Large room-temperature elastocaloric effect in a bulk polycrystalline Ni-Ti-Cu-Co alloy with low isothermal stress hysteresis. Appl. Mater. Today 2020, 21, 100844. [Google Scholar] [CrossRef]
  65. Bechtold, C.; Chluba, C.; Lima de Miranda, R.; Quandt, E. High cyclic stability of the elastocaloric effect in sputtered TiNiCu shape memory films. Appl. Phys. Lett. 2012, 101, 091903. [Google Scholar] [CrossRef]
  66. Wei, L.; Zhang, X.; Gan, W.; Ding, C.; Geng, L. Hot extrusion approach to enhance the cyclic stability of elastocaloric effect in polycrystalline Ni-Mn-Ga alloys. Scr. Mater. 2019, 168, 28–32. [Google Scholar] [CrossRef]
  67. Eftifeeva, A.; Panchenko, E.; Yanushonite, E.; Kurlevskaya, I.; Timofeeva, E.; Tokhmetova, A.; Surikov, N.; Tagiltsev, A.; Chumlyakov, Y. Superelasticity and elastocaloric cooling capacity in stress-induced martensite aged [001]А-oriented Ni54Fe19Ga27 single crystals. Mater. Sci. Eng. A 2022, 855, 143855. [Google Scholar] [CrossRef]
  68. Liang, X.; Xiao, F.; Jin, X.; Fukuda, T.; Kakeshita, T. Suppression of martensitic transformation in Co2Cr(Ga,Si) Heusler alloys by thermal cycling. Metall. Mater. Trans. A 2017, 48, 2105–2108. [Google Scholar] [CrossRef]
  69. Sidharth, R.; Celebi, T.B.; Sehitoglu, H. Origins of functional fatigue and reversible transformation of precipitates in NiTi shape memory alloy. Acta Mater. 2024, 274, 119990. [Google Scholar] [CrossRef]
  70. Dadda, J.; Maier, H.J.; Niklasch, D.; Karaman, I.; Karaca, H.E.; Chumlyakov, Y.I. Pseudoelasticity and cyclic stability in Co49Ni21Ga30 shape-memory alloy single crystals at ambient temperature. Metall. Mater. Trans. A 2008, 39, 2026–2039. [Google Scholar] [CrossRef]
  71. Fu, H.; Xu, S.; Zhao, H.; Dong, H.; Xie, J. Cyclic stress-strain response of directionally solidified polycrystalline Cu-Al-Ni shape memory alloys. J. Alloys Compd. 2017, 714, 154–159. [Google Scholar] [CrossRef]
  72. Briant, C.L.; Messmer, R.P. Electronic effects of sulphur in nickel. Philos. Mag. B 1980, 42, 569–576. [Google Scholar] [CrossRef]
  73. Messmer, R.P.; Briant, C.L. The role of chemical bonding in grain boundary embrittlement. Acta Metall. 1982, 30, 457–467. [Google Scholar] [CrossRef]
  74. Schulson, E.M.; Weihs, T.P.; Baker, I.; Frost, H.J.; Horton, J.A. Grain boundary accommodation of slip in Ni3Al containing boron. Acta Metall. 1986, 34, 1395–1399. [Google Scholar] [CrossRef]
  75. Liu, Y.; Liu, C.T.; Heatherly, L.; George, E.P. Effect of boron on the fracture behavior and grain boundary chemistry of Ni3Fe. Scr. Mater. 2011, 64, 303–306. [Google Scholar] [CrossRef]
  76. Liu, C.T.; George, E.P. Environmental embrittlement in boron-free and boron-doped FeAl (40 at. % Al) alloys. Scr. Metall. Mater. 1990, 24, 1285–1290. [Google Scholar] [CrossRef]
Figure 1. (a) DSC curves for the directionally solidified (Ni50Mn31Ti19)99B1 alloy; (b) compressive stress–strain curve measured at 293 K.
Figure 1. (a) DSC curves for the directionally solidified (Ni50Mn31Ti19)99B1 alloy; (b) compressive stress–strain curve measured at 293 K.
Materials 17 04756 g001
Figure 2. (a) EBSD orientation micrograph (IPF contrast) covering the longitudinal section for the (Ni50Mn31Ti19)99B1 alloy and the corresponding inverse pole figure related to the SD; (b) backscattered electron (BSE) image for the directionally solidified alloy; (c) BSE image corresponding to the squared region of (b); (d) TEM bright field image and the corresponding selected-area electron diffraction (SAED) for the boron-rich secondary phase along the [ 1 - 11] axis.
Figure 2. (a) EBSD orientation micrograph (IPF contrast) covering the longitudinal section for the (Ni50Mn31Ti19)99B1 alloy and the corresponding inverse pole figure related to the SD; (b) backscattered electron (BSE) image for the directionally solidified alloy; (c) BSE image corresponding to the squared region of (b); (d) TEM bright field image and the corresponding selected-area electron diffraction (SAED) for the boron-rich secondary phase along the [ 1 - 11] axis.
Materials 17 04756 g002
Figure 3. (a) Superelastic loop of the directionally solidified alloy measured under compression at 293 K; (b) typical ΔTad profile under a compressive strain of 8% by applying and removing the compressive loading of 510 MPa at 293 K; (c) correlation between ΔTad and compressive strain under various strain rates tested at 293 K; (d) correlation between ΔTad and strain rate under various compressive strains tested at 293 K.
Figure 3. (a) Superelastic loop of the directionally solidified alloy measured under compression at 293 K; (b) typical ΔTad profile under a compressive strain of 8% by applying and removing the compressive loading of 510 MPa at 293 K; (c) correlation between ΔTad and compressive strain under various strain rates tested at 293 K; (d) correlation between ΔTad and strain rate under various compressive strains tested at 293 K.
Materials 17 04756 g003
Figure 4. (a) Influence of strain rate on the superelastic response at various compressive strains tested at 293 K, by applying the maximum compressive loading of 350 MPa, 394 MPa, 415 MPa, 417 MPa, 445 MPa, and 506 MPa for the compressive strains of 3%, 4%, 5%, 6%, 7%, and 8%, respectively; (b) correlation between energy dissipation ΔW and compressive strain under various strain rates.
Figure 4. (a) Influence of strain rate on the superelastic response at various compressive strains tested at 293 K, by applying the maximum compressive loading of 350 MPa, 394 MPa, 415 MPa, 417 MPa, 445 MPa, and 506 MPa for the compressive strains of 3%, 4%, 5%, 6%, 7%, and 8%, respectively; (b) correlation between energy dissipation ΔW and compressive strain under various strain rates.
Materials 17 04756 g004
Figure 5. Graphical representation of COPmat as a function of ΔTad.
Figure 5. Graphical representation of COPmat as a function of ΔTad.
Materials 17 04756 g005
Figure 7. (a) Selective stress–strain curves during cyclic tests; (b) ΔTad profiles induced by loading and unloading for some selected cycles.
Figure 7. (a) Selective stress–strain curves during cyclic tests; (b) ΔTad profiles induced by loading and unloading for some selected cycles.
Materials 17 04756 g007
Figure 8. Comparison on the DSC curves for the samples before and after 12,000 mechanical cycles.
Figure 8. Comparison on the DSC curves for the samples before and after 12,000 mechanical cycles.
Materials 17 04756 g008
Figure 9. (a) Morphology of the longitudinal section after 12,000 mechanical cycles; (b) local BSE image showing the appearance of residual martensite after mechanical cycling; (c) fractured surface morphology after 13,000 cycles of mechanical tests.
Figure 9. (a) Morphology of the longitudinal section after 12,000 mechanical cycles; (b) local BSE image showing the appearance of residual martensite after mechanical cycling; (c) fractured surface morphology after 13,000 cycles of mechanical tests.
Materials 17 04756 g009
Table 1. Comparison of the elastocaloric properties between the present alloy and some other elastocaloric materials.
Table 1. Comparison of the elastocaloric properties between the present alloy and some other elastocaloric materials.
AlloySample StatusTtest (K)Tad| (K)σmax (MPa)Strain Rate (s−1)Ref.
(Ni50Mn31Ti19)99B1Polycrystal (textured)30327.24123.0 × 10−1This work
Ni50Mn31.75Ti18.25Polycrystal (textured)29320.49132.0[14]
Ni50Mn35Ti15Polycrystal (untextured)2934.54502.8 × 10−2[19]
Ni50Mn31.6Ti18.4Single crystal29329.08302.0[27]
Ni37Co9Fe4Mn35Ti15Polycrystal (untextured)2936.34001.4 × 10−1[57]
(Ni50Mn31.5Ti18.5)99.8B0.2Polycrystal (untextured)30831.57005.33[15]
Ni55Mn18Ga27Polycrystal (textured)30010.73502.0 × 10−1[40]
Ni50Mn18.5Ga25Cu6.5Polycrystal (textured)3158.11004.2 × 10−2[41]
Ni50(Mn28.5Cu4.5)(In14Ga3)Polycrystal (textured)29319.07102.0[42]
Ni44Mn46Sn10Polycrystal (textured)32018.03903.0 × 10−1[10]
Ni45Mn44Sn11Polycrystal (textured)29810.03103.0 × 10−2[43]
Ni47.5Co4.2Mn37.3Sb12.8Polycrystal (textured)3038.74005.0 × 10−2[44]
Cu71.3Al17.5Mn11.2Single crystal29311.91201.4[45]
Cu59.1Zn27Al13.8Zr0.1Single crystal34314.22002.0 × 10−1[46]
Ni48.9Ti51.1Wire32221.09002.0 × 10−1[47]
Ni50.8Ti49.2Polycrystal (textured)32317.99132.0 × 10−1[46]
Table 2. Cyclability and degradation rate of the elastocaloric effect for the present alloy and some elastocaloric materials.
Table 2. Cyclability and degradation rate of the elastocaloric effect for the present alloy and some elastocaloric materials.
AlloySample StatusNumber of CyclesDegradation Rate of ΔTad (K per Cycle)Ref.
(Ni50Mn31Ti19)99B1Polycrystal (textured)12,0007.5 × 10−5This work
Ni50Mn30Ti20Polycrystal (textured)20005.0 × 10−4[39]
Ni37Co9Fe4Mn35Ti15Polycrystal (untextured)10002.0 × 10−4[57]
Ni50.4Mn27.3Ga22.3Polycrystal (textured)2501.2 × 10−3[66]
Ni50.4Mn27.3Ga22.3Polycrystal (textured)1002.0 × 10−3[12]
Ni50Fe19Ga27Co4Single crystal30009.3 × 10−4[20]
Ni54Fe19Ga27Single crystal1001.0 × 10−3[67]
Ni54Fe19Ga27Polycrystal (untextured)1009.0 × 10−3[59]
Ni53.2Fe19.4Ga27.4Polycrystal (untextured)4802.3 × 10−3[60]
Cu71.3Al17.5Mn11.2Single crystal504.0 × 10−3[45]
Cu71Al18Mn11Polycrystal (textured)2751.5 × 10−3[61]
Cu71.1Al17.2Mn11.7Polycrystal (textured)2001.0 × 10−3[53]
Cu59.1Zn27Al13.8Polycrystal (untextured)10,0001.4 × 10−4[46]
Co49Fe3V33Ga15Polycrystal (textured)2001.5 × 10−3[62]
Co50V35Ga14Ni1Polycrystal (untextured)40008.5 × 10−4[63]
(Ni42.5Ti50Cu7.5)99Co1Polycrystal (textured)2002.2 × 10−2[64]
Ti54.9Ni32.5Cu12.6Polycrystal (untextured)15022.0 × 10−4[65]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Wang, Y.; Zhang, G.; Li, Z.; Yang, J.; Li, J.; Yang, B.; Yan, H.; Zuo, L. Giant Elastocaloric Effect and Improved Cyclic Stability in a Directionally Solidified (Ni50Mn31Ti19)99B1 Alloy. Materials 2024, 17, 4756. https://doi.org/10.3390/ma17194756

AMA Style

Wang H, Wang Y, Zhang G, Li Z, Yang J, Li J, Yang B, Yan H, Zuo L. Giant Elastocaloric Effect and Improved Cyclic Stability in a Directionally Solidified (Ni50Mn31Ti19)99B1 Alloy. Materials. 2024; 17(19):4756. https://doi.org/10.3390/ma17194756

Chicago/Turabian Style

Wang, Honglin, Yueping Wang, Guoyao Zhang, Zongbin Li, Jiajing Yang, Jinwei Li, Bo Yang, Haile Yan, and Liang Zuo. 2024. "Giant Elastocaloric Effect and Improved Cyclic Stability in a Directionally Solidified (Ni50Mn31Ti19)99B1 Alloy" Materials 17, no. 19: 4756. https://doi.org/10.3390/ma17194756

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