Next Article in Journal
Plasma Temperatures in the Martian Subsolar Magnetosheath: MAVEN Observations
Next Article in Special Issue
Effect of Substrate on One-Dimensional Multiferroic Properties
Previous Article in Journal
Synthesis and Single Crystal Growth by Floating Zone Technique of FeCr2O4 Multiferroic Spinel: Its Structure, Composition, and Magnetic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Large Elastocaloric Effect Driven by Low Stress Induced in [001]-Oriented Polycrystalline Co51.6V31.4Ga17 Alloy

1
Center for Advanced Solidification Technology, School of Materials Science and Engineering, Shanghai University, Shanghai 200444, China
2
Jiangxi Key Laboratory for Rare Earth Magnetic Materials and Devices, Faculty of Materials Metallurgy and Chemistry, Jiangxi University of Science and Technology, Ganzhou 341000, China
3
Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
*
Author to whom correspondence should be addressed.
Magnetochemistry 2022, 8(8), 87; https://doi.org/10.3390/magnetochemistry8080087
Submission received: 21 July 2022 / Revised: 1 August 2022 / Accepted: 6 August 2022 / Published: 9 August 2022
(This article belongs to the Special Issue Phase Change Material and Magnetic Research)

Abstract

:
In this work, we have studied the elastocaloric effect in directionally solidified Co51.6V31.4Ga17 alloys with a strong [001] preferred orientation. The entropy change for thermal-induced martensitic transformation is determined as 19.6 J kg−1 K−1. The sample exhibits stress-induced martensitic transformation with a hysteresis of 46 MPa, and the superelasticity is also verified by the in situ X-ray diffraction method. According to the elastocaloric effect tests, a noticeable change in adiabatic temperature up to 12.2 K has been achieved at the strain of 6%. The specific temperature change upon the critical stress loading can be attained as 132 K MPa−1. In addition, the difference in the loading–unloading temperature change can be ascribed to the imperfect adiabatic environment.

1. Introduction

Considerable attention among researchers has been paid to the development of advanced refrigeration technology in recent years. This is because of the expected enhanced efficiency of traditional gas compression refrigeration and the greenhouse gas emissions of the refrigerant. By searching 600 million organic compounds in the PubChem database, it was concluded that there are currently no unknown compound refrigerants [1]. Being one of the most promising new refrigeration technologies, solid-state refrigeration has aroused great interest as an environmentally friendly and energy-saving technique. In particular, elastocaloric refrigeration based on the elastocaloric effect (eCE) has recently become a popular topic due to its simple uniaxial loading style and wide refrigeration temperature range. At present, the reported elastocaloric materials include shape memory alloys (SMAs), oxide ceramics [2], natural rubber [3], and graphene [4]. Specifically, the eCE in SMAs has been studied extensively and in depth over the past few decades. Among the mainstream elastocaloric alloys, there are Ni-Ti alloys [5,6,7], Cu-based alloys [8,9], Heusler-type alloys [7,10,11], Fe-based alloys [12,13], and so on.
The external factors, including applied stress field, working temperature, and adiabatic environment, can directly affect the eCE of materials [14,15,16]. Moreover, the chemical composition, crystal structure, and grain orientation of the alloy may impact the eCE associated with the stress-induced martensitic transformation [17,18,19,20,21,22]. For anisotropic SMAs, the mechanical properties, transformation strain, and critical stress may vary with grain orientation. In particular, a large amount of research has been performed on the eCE of polycrystalline, single-crystal, and textured polycrystalline Ni-Mn-based Heusler alloys [12,17,18,19,20,21,22,23,24]. According to various reports, the eCE properties of Ni-Mn based alloys are directly related to the temperature sensitivity of the transformation strain and critical stress, and strongly refer to the crystallographic orientation. For example, the adiabatic temperature change values of the [001] oriented Ni50Mn32In16Cr2 alloys are 1.9 K and 2.6 K, respectively, larger than those of the [111] and [110] oriented Ni50Mn32In16Cr2 alloys [16]. For polycrystalline Ni50.4Mn27.3Ga22.3 alloys with different orientations, the adiabatic temperature changes depend on whether the loading direction is parallel or perpendicular to the [001] direction [14]. In Ni45Mn44Sn11 alloy, the thermal responses of samples with [111] orientation vary from those corresponding to other orientations under the same stress conditions [15].
Therefore, the grain orientation of the alloy has a significant influence on its eCE. Apart from the aforementioned Ni-Mn based alloys, Co-Ni-Ga alloys, as a representative of Co-based Heusler alloys, have also attracted much attention in recent years. It is noteworthy that in the case of Co-based alloys with grain orientation, most researchers mainly focus on the microstructure, magnetic properties, and martensitic transformation of Co-Ni-Ga textured polycrystalline alloys prepared by directional solidification or single-crystal growth [25,26,27,28]. Meanwhile, the elastocaloric effect in textured polycrystalline alloys has less been studied systematically. A pronounced eCE has been found in Co-V-Ga alloys undergoing a stress-induced phase transformation from the L21 atomic ordered austenite to the DO22 tetragonal martensite [29,30,31,32]. However, the reported eCE was observed in texture-free polycrystalline alloys. Therefore, this study aims to investigate the eCE in textured polycrystalline Co-V-Ga alloys.

2. Materials and Methods

Co51.6V31.4Ga17 polycrystalline alloy ingots were prepared by arc-melting pure Co (99.99%), V (99.99%), and Ga (99.99%) in an argon atmosphere, and then casting them into a 7-mm-diameter cold copper mold. The rods were subsequently re-melted at 1773 K and grown in alumina crucibles via the liquid–metal cooling directional solidification method, with a pulling rate of 150 μm s−1. The compressive test specimens were cut from the steady growing zone of the solidified sample into rectangular parallelepipeds with nominal dimensions of 3 mm × 3 mm × 6 mm, so that their long axis was along the growth direction. The specimens were afterward annealed at 1473 K for 24 h in an Ar atmosphere and quenched in ice water.
The temperature dependence of magnetization M-T and magnetic field dependence of magnetization M-H were measured by a vibrating sample magnetometer attached to a superconducting quantum interference device (SQUID-VSM, Quantum Design, San Diego, CA, USA). The calorimetric measurements were performed using differential scanning calorimetry (DSC, NETZSCH DSC 214) at the heating/cooling rate of 10 K min−1. The texture analysis was conducted via electron backscattering diffraction (EBSD, OXFORD Nordlys Max2). Prior to the EBSD experiments, specimens were exposed to ion beam polishing (LEICA EM TIC3X) to reduce surface residual stress. Crystal structure testing was implemented on a two-dimensional X-ray diffraction system (VANTEC500 XRD4). Before the XRD analysis, the samples were compressed with a self-designed micro-compressive device to maintain their compression state. The room-temperature compressive experiments were conducted on a universal compression testing machine (SUNS-UTM5000) at 303 K. The surface temperature changes were monitored via infrared (IR) thermography on samples pre-sprayed with black matte paint. For compressive tests, the loading was applied along the direction of solidification growth. For the eCE measurements, the samples were rapidly loaded–unloaded at a strain rate of 2.58% s−1 for each curve. The high strain rate provided a quasi-adiabatic environment, and a holding step of 30 s was applied after loading to ensure that the sample returned to the thermal equilibrium state before unloading started.

3. Results and Discussion

Figure 1a displays the DSC and M-T curves acquired under a magnetic field of 5 T during heating and cooling. Applying the tangent extrapolation method to the DSC curves enabled us to establish the start and finish temperatures of the forward martensitic transformation (Ms and Mf) and the reverse transformation (As and Af) as 270 K, 255 K, 272 K, and 285 K, respectively. The transformation equilibrium temperature T0, determined as a function of T0 = (Af + Ms)/2 [33], was 271 K. The average transformation enthalpy change ΔHtr found from the area under the exothermic and endothermic peaks was 5.3 J g−1. The transformation entropy change ΔStr associated with the martensitic transformation was determined to be 19.6 J kg−1 K−1 according to the following formula: ΔStr = ΔHtr/T0. In addition, the characteristic martensitic transformation temperatures confirmed from the M-T curve were almost consistent with those obtained from the DSC curve. The magnetization difference ΔM in the two phases crossing the martensitic transformation was only 1.9 emu g−1 under the magnetic field of 5 T (Figure 1a), indicating that it is difficult to drive the phase transition by applying a magnetic field.
Figure 1b depicts the EBSD orientation micrograph of the longitudinal section for the Co51.6V31.4Ga17 alloy at room temperature, where the alloy was in the austenite phase state. The EBSD phase calibration was performed assuming the lattice parameter a of 5.80 Å, the space group Fm3m (225), and the calibration rate of 93.7%. The inverse pole figure (see the inset) reveals that the sample had a strong [001]A preferred orientation along the solidification growth direction. Generally, the solid–liquid interface energy required for the [001] orientation growth of the cubic phase was lower than that in other orientations, so the [001] orientation tended to be the most common growth direction. Such a microstructural feature in the directionally solidified alloy also plays a significant role in decreasing the transformation hysteresis due to the reduced amount of grain boundaries [34,35]. Hence, the [001] orientation with low critical stress is the preferred uniaxial stress loading direction, leading to a maximum transformation strain [36,37], being similar to previously reported SMAs, including Cu-Al-Mn [38], Ni-Mn-In-Co [39], Ni-Mn-Ga [40], and Ni-Co-Mn-Ti [41].
Figure 2a presents the isothermal compressive stress–strain curve for the textured Co51.6V31.4Ga17 alloy at 303 K, which is above Af = 285 K. The strain rate of 0.03% s−1 was applied to provide sufficient heat exchange between the sample and the surrounding environment. The sample was then loaded slowly in an isothermal process so that the strain was further increased up to 5.0%. It should be mentioned that the critical stress (σcr) required for the alloy is approximately 109 MPa. The critical stress is an important parameter closely related to the eCE of the material, and is usually affected by the testing temperature, microstructure, and deformation conditions. As shown in Figure 2a, once the applied uniaxial stress reached 109 MPa, the alloy started to undergo martensitic transformation, indicating that the critical stress of this alloy was significantly lower than that of the texture-free polycrystalline alloy with the same composition [29]. Moreover, it was inferior to that of the Ni-Ti based alloy system [6,7,42,43,44] and several Ni-Mn based alloys [14,16,45,46]. In addition, the alloy under consideration exhibited very good superelastic behavior, where a nearly reversible strain up to 5.0% could be obtained through the reversible stress-induced martensitic transformation under a relatively low stress of 188 MPa. Under the same stress conditions, the stress-induced transformation fraction of the textured polycrystalline alloy was significantly larger than that of the texture-free polycrystalline alloy [29]. The stress hysteresis of the shape memory alloy is defined as: Δσhy = σMsσAf. For our textured alloy, the calculated Δσhy hysteresis was only 45 MPa, being lower than those of some Ni-Mn based alloys. The low stress hysteresis reduces the energy dissipation and enhances the repetitive stability of stress cycles [15,47].
A self-designed micro-compressive device was utilized to compress the textured alloy at room temperature and to hold the sample in a compressed state for further testing of its crystal structure. Figure 2b displays the one-dimensional XRD patterns within the angular range of 40–50° under various strains. It was evident that the uncompressed alloy was in the austenite phase. As the amount of applied strain increased, the main (220)A peak began to split into the peaks of the martensitic phase, indicating the stress-induced martensitic transformation. The diffraction peak at 43° corresponds to the (112)M. Once the applied strain increased to 3.33%, the alloy transformed into the state of martensite–austenite phase coexistence, which was consistent with the stress–strain curve shown in Figure 2a.
The eCE of the [001] oriented Co51.6V31.4Ga17 polycrystal was studied at room temperature, and the temperature change was characterized by infrared thermography (IR). According to Figure 3a, the temperature change value gradually increased with the applied strain. The maximum adiabatic temperature change of 12.2 K was achieved under the applied strain of 6%. Taking the compressive curve and the corresponding infrared spectrum at εappl = 5.5% as an example, as shown in Figure 3a, at the critical stress of 91 MPa, the corresponding average temperature changes during loading and unloading, obtained from the rectangular area in the center of the sample (highlighted in Figure 3b), were 12.0 K and 10.9 K, respectively. Moreover, as seen in Figure 3b, the temperature at the edge of the sample was slightly lower due to part of the heat loss originating from the heat exchange between the sample and the clamps [48]. Thus, the IR images revealed that the temperature space evolution of the samples was homogeneous.
Figure 3c depicts the loading–unloading temperature change values under different applied strains. It can be seen that when the strain amount was below 4.5%, the loading temperature changed more slowly than the unloading one. However, with the gradual increase in the applied strain, the values of ΔTloading and ΔTunloading gradually decreased. This temperature asymmetry might be related to the imperfect adiabatic environment. In other words, during the initial loading stage, the contact of the compressive machine with the top of the sample was incomplete, resulting in heat leakage. Meanwhile, with the increase in testing time, the loading temperature change gradually exceeded the unloading one. This could be attributed to the fact that the temperature between the compressive machine and the sample tended to be closer, which resulted in the improvement of the adiabaticity, and consequently, the phenomenon of temperature change was reduced due to heat leakage during loading. In addition, according to the stress–strain curve in Figure 3d, as the residual strain after the rapid loading–unloading cycle was approximately 1.4%, we might exclude the possibility of temperature asymmetry being responsible for the generation of the unrecovered martensite phase upon plastic deformation. Instead, the temporary residual strain in the texture-free Co51.6V31.4Ga17 polycrystal [29] was not observed in the present work.
The temperature interval between the testing temperature Ttest and Af has a great influence on the eCE of the alloy. In view of this, the typical elastocaloric alloy systems with |TtestAf| approximately 20 K were selected for comparison with our textured Co51.6V31.4Ga17 alloy (the relevant data taken at εappl = 5.5% are listed in Table 1). It can be seen that the adiabatic temperature change of the textured alloy under consideration was at a relatively high level, whereas the testing was also near room temperature. In turn, the critical stress exceeded those of Ni-Mn based alloys [47,48,49,50,51], but was significantly lower than that of Ni-Ti based alloys [52,53]. It is worth noting that our textured alloy exhibited a relatively high | Δ T ad / Δ σ cr | 132 K GPa−1, which was lower than other elastocaloric SMAs and higher than that of other components in the Co-V-Ga system [31,32]. The low specific adiabatic temperature change per stress would be beneficial for saving the driven component space and minimizing the size of the elastocaloric cooling device. This indicates that polycrystalline Co51.6V31.4Ga17 with the [001] orientation is a promising elastocaloric material with high efficiency and great energy-saving potential for next-generation refrigeration technology.

4. Conclusions

The Co51.6V31.4Ga17 polycrystalline alloy prepared by directional solidification exhibited a strong [001] preferred orientation and a large entropy change of 19.6 J kg−1 K−1. The alloy demonstrated good superelasticity. The stress-induced martensitic transformation in the alloy was confirmed from the XRD patterns under stressing, revealing an obvious austenite–martensite coexistence state under the strain of 3.33%. Moreover, a large ΔTad of 12.0 K was achieved under the applied strain of 5.5%, resulting in the large value of |ΔTad/σcr| ≈ 132 K MPa−1. Thus, the [001]-oriented Co51.6V31.4Ga17 polycrystalline alloy has application potential in the field of solid-state refrigeration based on eCE.

Author Contributions

J.L.: Conceptualization, Methodology, Investigation, Writing—Original Draft, Writing—Review and Editing. J.H.: Methodology, Investigation, Writing—Review and Editing. S.M.: Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ningbo Science and Technology Innovation 2025 Major Project (2020Z063), National Natural Science Foundation of China (51971232), and Zhejiang Provincial Natural Science Foundation of China (Grant No. LD21E010001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mclinden, M.O.; Brown, J.S.; Brignoli, R.; Kazakov, A.F.; Domanski, P.A. Limited options for low-global-warming-potential refrigerants. Nat. Commun. 2017, 8, 14476. [Google Scholar] [CrossRef] [PubMed]
  2. Chauhan, A.; Patel, S.; Vaish, R. Elastocaloric effect in ferroelectric ceramics. Appl. Phys. Lett. 2015, 106, 17290110. [Google Scholar] [CrossRef]
  3. Guyomar, D.; Li, Y.; Sebald, G.; Cottinet, P.-J.; Ducharne, B.; Capsal, J.-F. Elastocaloric modeling of natural rubber. Appl. Therm. Eng. 2013, 57, 33–38. [Google Scholar] [CrossRef]
  4. Meng, L.I.; Guo, Z.R.; Chang, T.C. Adhesion and Stress Enhanced Elastocaloric Effect in Graphene. Sci. China Technol. Sci. 2018, 59, 1674–7321. [Google Scholar]
  5. Cui, J.; Wu, Y.M.; Muehlbauer, J.; Hwang, Y.; Radermacher, R.; Fackler, S.; Wuttig, M.; Takeuchi, I. Demonstration of high efficiency elastocaloric cooling with large ΔT using NiTi wires. Appl. Phys. Lett. 2012, 101, 073904. [Google Scholar] [CrossRef] [Green Version]
  6. 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] [Green Version]
  7. Pataky, G.J.; Ertekin, E.; Sehitoglu, H. Elastocaloric cooling potential of NiTi, Ni2FeGa, and CoNiAl. Acta Mater. 2015, 96, 420–427. [Google Scholar] [CrossRef]
  8. Brown, L.C. The thermal effect in pseudoelastic single crystals of β-CuZnSn. Metal. Mater. Trans. A 1981, 12, 1491–1494. [Google Scholar] [CrossRef]
  9. Niitsu, K.; Kimura, Y.; Omori, T.; Kainuma, R. Cryogenic superelasticity with large elastocaloric effect. NPG Asia Mater. 2018, 10, 457. [Google Scholar] [CrossRef] [Green Version]
  10. Xiao, F.; Fukuda, T.; Kakeshita, T.; Jin, X. Elastocaloric effect by a weak first-order transformation associated with lattice softening in an Fe-31.2Pd (at. %) alloy. Acta Mater. 2015, 87, 8–14. [Google Scholar] [CrossRef]
  11. Zhao, D.W.; Xiao, F.; Nie, Z.H.; Cong, D.Y.; Sun, W.; Liu, J. Burst-like superelasticity and elastocaloric effect in [011] oriented Ni50Fe19Ga27Co4 single crystals. Scri. Mater. 2018, 149, 6–10. [Google Scholar] [CrossRef]
  12. 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. Scri. Mater. 2017, 130, 148–151. [Google Scholar] [CrossRef]
  13. Nikitin, S.A.; Myalikgulyev, G.; Annaorazov, M.P.; Tyurin, A.L.; Myndyev, R.W.; Akopyan, S.A. Giant elastocaloric effect in FeRh alloy. Phys. Lett. A 1992, 171, 234–236. [Google Scholar] [CrossRef]
  14. Lu, B.F.; Xiao, F.; Yan, A.R.; Liu, J. Elastocaloric effect in a textured polycrystalline Ni-Mn-In-Co metamagnetic shape memory alloy. Appl. Phys. Lett. 2014, 105, 161905. [Google Scholar] [CrossRef]
  15. Guo, J.P.; Wei, Z.Y.; Shen, Y.; Zhang, Y.F.; Li, J.Y.; Huo, X.L.; Liu, J. Low-temperature superelasticity and elastocaloric effect in textured Ni–Mn–Ga–Cu shape memory alloys. Scri. Mater. 2020, 185, 56–60. [Google Scholar] [CrossRef]
  16. Mañosa, L.; Planes, A. Mechanocaloric effects in shape memory alloys. Philo. Trans. R. Soc. A 2016, 374, 0150310. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, J.; Woodcock, T.G.; Scheerbaum, N.; Gutfleisch, O. Influence of annealing on magnetic field-induced structural transformation and magnetocaloric effect in Ni-Mn-In-Co ribbons. Acta Mater. 2009, 57, 4911–4920. [Google Scholar] [CrossRef]
  18. 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 alloy. Acta Mater. 2017, 127, 33–42. [Google Scholar] [CrossRef]
  19. Hernández-Navarro, F.; Camarillo-Garcia, J.-P.; Aguilar-Ortiz, C.-O.; Flores-Zúñiga, H.; Ríos, D.; González, J.-G.; Álvarez-Alonso, P. The influence of texture on the reversible elastocaloric effect of a polycrystalline Ni50Mn32In16Cr2 alloy. Appl. Phys. Lett. 2018, 112, 164101. [Google Scholar] [CrossRef]
  20. Wei, L.S.; Zhang, X.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] [Green Version]
  21. Shen, Y.; Sun, W.; Wei, Z.Y.; Shen, Q.; Zhang, Y.F.; Liu, J. Orientation dependent elastocaloric effect in directionally solidified Ni-Mn-Sn alloys. Scri. Mater. 2019, 163, 14–18. [Google Scholar] [CrossRef]
  22. Wei, Z.Y.; Sun, W.; Shen, Q.; Shen, Y.; Zhang, Y.F.; Liu, E.K.; Liu, J. Elastocaloric effect of all-d-metal Heusler NiMnTi(Co) magnetic shape memory alloys by digital image correlation and infrared thermography. Appl. Phys. Lett. 2019, 114, 101903. [Google Scholar] [CrossRef]
  23. Karaca, H.; Karaman, I.; Basaran, B.; Ren, Y.; Chumlyakov, Y.; Maier, H. Magnetic field-induced phase transformation in NiMnCoIn magnetic shape-memory alloys—A new actuation mechanism with large work output. Adv. Funct. Mater. 2009, 19, 938. [Google Scholar] [CrossRef]
  24. Pfeuffer, L.; Gracia-Condal, A.; Gottschall, T.; Koch, D.; Faske, T.; Bruder, E.; Lemke, J.; Taubel, A.; Ener, S.; Scheibel, F.; et al. Influence of microstructure on the application of Ni-Mn-In Heusler compounds for multicaloric cooling using magnetic field and uniaxial stress. Acta Mater. 2021, 217, 117157. [Google Scholar] [CrossRef]
  25. Xie, H.; Liu, J.; Huo, Y.Q.; Liu, J. Preferred orientation and microstructural evolution of directionally solidified Co-Ni-Ga ferromagnetic shape memory alloys. Acta Mater. Sin. 2007, 43, 417–421. [Google Scholar]
  26. Liu, J.; Li, J.G. Magnetic force microscopy observations of Co–Ni–Ga and Co–Ni–Al alloys with two-phase structures. Scri. Mater. 2006, 55, 755–758. [Google Scholar] [CrossRef]
  27. Liu, J.; Li, J.G.; Xie, H. Crystal growth of Co46Ni27Ga27 ferromagnetic shape memory alloys with large single-variant set. J. Cryst. Growth 2007, 300, 272–276. [Google Scholar] [CrossRef]
  28. Shen, A.; Zhao, D.W.; Sun, W.; Liu, J.; Li, C. Elastocaloric effect in a Co50Ni20Ga30 single crystal. Scri. Mater. 2017, 127, 1–5. [Google Scholar] [CrossRef]
  29. He, J.; Wei, Z.Y.; Sun, W.; Lu, X.; Ma, S.C.; Liu, J. Martensitic transformation and elastocaloric effect of Co51.5+xV31.5−xGa17 (x = 0.1, 0.2, 0.3) alloys. Intermetallics 2021, 139, 107348. [Google Scholar] [CrossRef]
  30. Su, J.C.; Xuan, H.C.; Wang, L.B.; Zhu, Y.D.; Xie, Z.G.; Chen, F.H.; Zhang, K.W.; Han, P.D. Large Elastocaloric Effect in Fe-Doped Co-Fe-V-Ga Shape Memory Alloys. Phys. Status Solidi A-Appl. Mater. Sci. 2022, 13, 6666–6675. [Google Scholar] [CrossRef]
  31. Niu, Y.R.; Chen, H.Y.; Zhang, X.Y.; Li, X.W.; Cong, D.Y.; Ma, T.Y.; Li, S.H.; Lin, J.P.; Wang, Y.-D. Achieving excellent superelasticity and extraordinary elastocaloric effect in a directionally solidified Co-V-Ga alloy. Scri. Mater. 2021, 204, 114123. [Google Scholar] [CrossRef]
  32. Liu, C.; Li, D.; Li, Z.B.; Yang, B.; Zuo, L. Large elastocaloric effect in a Heusler-type Co50V35Ga14Ni1 polycrystalline alloy. Appl. Phys. Lett. 2021, 118, 103904. [Google Scholar] [CrossRef]
  33. Ortin, J.; Planes, A. Thermodynamic analysis of thermal measurements in thermoelastic martensitic transformations. Acta Metall. 1988, 36, 1873. [Google Scholar] [CrossRef]
  34. Stupkiewicz, S.; Petryk, H. Grain-size effect in micromechanical modelling of hysteresis in shape memory alloys. ZAMM-J. Apll. Math. Mech. 2010, 90, 783–795. [Google Scholar] [CrossRef]
  35. Huang, X.M.; Wang, L.D.; Liu, H.X.; Yan, H.L.; Jia, N.; Yang, B.; Li, Z.B.; Zhang, Y.D.; Esling, C.; Zhao, X.; et al. Correlation between microstructure and martensitic transformation, mechanical properties and elastocaloric effect in Ni–Mn-based alloy. Intermetallics 2019, 113, 106579. [Google Scholar] [CrossRef]
  36. Li, J.Z.; Li, J.G. Effects of solidification rate and temperature gradient on microstructure and crystal orientation of Co32Ni40Al28 alloys. Mater. Lett. 2012, 68, 40–43. [Google Scholar] [CrossRef]
  37. Lu, B.F.; Liu, J. Mechanocaloric materials for solid-state cooling. Sci. Bull. 2015, 60, 1638–1643. [Google Scholar] [CrossRef]
  38. Xu, S.; Huang, H.Y.; Xie, J.X.; Takekawa, S.; 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] [Green Version]
  39. Wang, J.M.; Jiang, C.B. Single-crystal growth of NiMnGa magnetic shape memory alloys. J. Cryst. Growth 2008, 310, 865–869. [Google Scholar] [CrossRef]
  40. He, X.J.; Kang, Y.R.; Wei, S.X.; Zhang, Y.L.; Cao, Y.M.; Xu, K.; Li, Z.; Jing, C.; Li, Z.B. A large barocaloric effect and its reversible behavior with an enhanced relative volume change for Ni42.3Co7.9Mn38.8Sn11 Heusler alloy. J. Alloys Compd. 2018, 741, 821–825. [Google Scholar] [CrossRef]
  41. Shen, Y.; Wei, Z.Y.; Sun, W.; Zhang, Y.F.; Liu, J. Large elastocaloric effect in directionally solidified all-d-metal Heusler metamagnetic shape memory alloys. Acta Mater. 2020, 188, 677–685. [Google Scholar] [CrossRef]
  42. Waitz, T.; Karnthaler, H.P. Martensitic transformation of NiTi nanocrystals embedded in an amorphous matrix. Acta Mater. 2004, 52, 5461–5469. [Google Scholar] [CrossRef]
  43. Waitz, T.; Antretter, T.; Fischer, F.D.; Simha, N.K.; Karnthaler, H.P. Size effects on the martensitic phase transformation of NiTi nanograins. J. Mech. Phys. Solids 2007, 55, 419–444. [Google Scholar] [CrossRef]
  44. 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]
  45. Li, Z.Z.; Li, Z.B.; Li, D.; Yang, J.; Zuo, L. 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]
  46. Cong, D.Y.; Xiong, W.X.; Planes, A.; Ren, Y.; Mañosa, L.; Cao, P.; 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] [Green Version]
  47. Li, D.; Li, Z.B.; Yang, J.J.; Li, Z.Z.; Yang, B.; Yan, H.L.; Wang, D.H.; Hou, L.; Li, X.; Zhang, Y.D.; et al. Large elastocaloric effect driven by stress-induced two-step structural transformation in a directionally solidified Ni55Mn18Ga27 alloy. Scri. Mater. 2019, 163, 116–120. [Google Scholar] [CrossRef]
  48. 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]
  49. Lu, B.F.; Liu, J. Elastocaloric effect and superelastic stability in Ni–Mn–In–Co polycrystalline Heusler alloys: Hysteresis and strain-rate effects. Sci. Rep. 2017, 7, 2084. [Google Scholar] [CrossRef] [Green Version]
  50. Li, D.; Li, Z.B.; Zhang, X.L.; Yang, B.; Wang, D.H.; Zhao, X.; Zuo, L. Enhanced cyclability of elastocaloric effect in a directionally solidified Ni55Mn18Ga26Ti1 alloy with low hysteresis. Scri. Mater. 2020, 189, 78–83. [Google Scholar] [CrossRef]
  51. Sun, W.; Liu, J.; Zhao, D.W.; Zhang, M.X. Directional solidification and elastocaloric effect in a Ni45Mn44Sn11 magnetic shape memory alloy. J. Phys. D Appl. Phys. 2017, 50, 444001. [Google Scholar] [CrossRef]
  52. Cao, Y.; Zhou, X.; Cong, D.Y.; Zheng, H.; Cao, Y.; Nie, Z.; Chen, Z.; Li, S.H.; Xu, N.; Gao, Z.Y.; et al. Large tunable elastocaloric effect in additively manufactured Ni–Ti shape memory alloys. Acta Mater. 2020, 194, 178–189. [Google Scholar] [CrossRef]
  53. Yang, Z.; Cong, D.Y.; Yuan, Y.; Li, R.G.; Zheng, H.X.; Sun, X.M.; Nie, Z.; Ren, Y.; Wang, Y.D. 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]
Figure 1. (a) DSC without magnetic field and M-T curves at 5 T for the directionally solidified Co51.6V31.4Ga17 alloy. (b) EBSD orientation micrograph (IPF mode) of the longitudinal section for the directionally solidified Co51.6V31.4Ga17 alloy at room temperature. The inset shows the inverse pole figure.
Figure 1. (a) DSC without magnetic field and M-T curves at 5 T for the directionally solidified Co51.6V31.4Ga17 alloy. (b) EBSD orientation micrograph (IPF mode) of the longitudinal section for the directionally solidified Co51.6V31.4Ga17 alloy at room temperature. The inset shows the inverse pole figure.
Magnetochemistry 08 00087 g001
Figure 2. (a) The isothermal compressive stress–strain curve for the textured Co51.6V31.4Ga17 alloy at room temperature (the strain rate is 0.03% s−1). (b) XRD patterns of the textured alloy under uniaxial stress at room temperature.
Figure 2. (a) The isothermal compressive stress–strain curve for the textured Co51.6V31.4Ga17 alloy at room temperature (the strain rate is 0.03% s−1). (b) XRD patterns of the textured alloy under uniaxial stress at room temperature.
Magnetochemistry 08 00087 g002
Figure 3. (a) ΔTad parameter as a function of time under different applied strains with a strain rate of 2.58% s−1 at 303 K. The temperature–time profile is obtained from the dashed frame areas in the IR spectrograms. (b) The IR images at the applied strain of 6.0%. (c) The average ΔTad values at various compressive strains during loading and unloading. (d) The stress–strain curve under applied strain of 5.5%.
Figure 3. (a) ΔTad parameter as a function of time under different applied strains with a strain rate of 2.58% s−1 at 303 K. The temperature–time profile is obtained from the dashed frame areas in the IR spectrograms. (b) The IR images at the applied strain of 6.0%. (c) The average ΔTad values at various compressive strains during loading and unloading. (d) The stress–strain curve under applied strain of 5.5%.
Magnetochemistry 08 00087 g003
Table 1. The values of | Δ T ad / Δ σ cr | for the present textured Co51.6V31.4Ga17 alloy and some typical elastocaloric alloys.
Table 1. The values of | Δ T ad / Δ σ cr | for the present textured Co51.6V31.4Ga17 alloy and some typical elastocaloric alloys.
AlloysAf (K)Ttest (K) Δ Tad (K) σcr (MPa)| Δ Tad/σcr| (K GPa−1)
This work28530312.091132
Co50.7V33.3Ga16 [31]24927317.318096
Co50V35Ga14Ni1 [32]28030012.112994
Co50Ni20Ga30 [28]2642913.77549
Ni45Mn36.4In13.6Co5 [49]2682932.49625
Ni55Mn18Ga26Ti1 [50]2652953.18536
Ni50Mn31.5In16Cu2.5 [48]2803006.014541
Ni55Mn18Ga27 [47]28129810.715071
Ni45Mn44Sn11 [51]2732987.525030
Ni51.4Ti48.6 [52]31833323.250046
(Ni42.5Ti50Cu7.5)99Co1 [53]28630414.432045
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; He, J.; Ma, S. Large Elastocaloric Effect Driven by Low Stress Induced in [001]-Oriented Polycrystalline Co51.6V31.4Ga17 Alloy. Magnetochemistry 2022, 8, 87. https://doi.org/10.3390/magnetochemistry8080087

AMA Style

Liu J, He J, Ma S. Large Elastocaloric Effect Driven by Low Stress Induced in [001]-Oriented Polycrystalline Co51.6V31.4Ga17 Alloy. Magnetochemistry. 2022; 8(8):87. https://doi.org/10.3390/magnetochemistry8080087

Chicago/Turabian Style

Liu, Jian, Jing He, and Shengcan Ma. 2022. "Large Elastocaloric Effect Driven by Low Stress Induced in [001]-Oriented Polycrystalline Co51.6V31.4Ga17 Alloy" Magnetochemistry 8, no. 8: 87. https://doi.org/10.3390/magnetochemistry8080087

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