Download PDF
Research Article  |  Open Access  |  27 Apr 2023

Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in Mn3-xPt1+x

Views: 545 |  Downloads: 610 |  Cited:   1
Microstructures 2023;3:2023022.
10.20517/microstructures.2022.46 |  © The Author(s) 2023.
Author Information
Article Notes
Cite This Article

Abstract

A large driving pressure is required for barocaloric effects (BCEs) in intermetallics, usually above 100 MPa. Here, we report barocaloric effects in Mn3-xPt1+xalloys saturated at about 60 MPa, the lowest pressure reported in intermetallics to date. A first-order phase transition occurs from the colinear antiferromagnetic phase to the triangular antiferromagnetic phase as temperature decreases. The transition temperature strongly depends on the composition x, ranging from 331 K for x = 0.18 to 384 K for x = 0.04, and is sensitive to pressure, with dTt/dP up to 139 K/GPa. However, the maximum pressure-induced entropy changes are as small as 13.79 J kg-1 K-1, attributed to the mutual cancellation of lattice and magnetic entropy changes. The small driving pressure and total entropy changes are due to the special magnetic geometric frustration.

Keywords

Barocaloric effects, magnetoelastic coupling, magnetic transition, geometrical spin frustration, colinear antiferromagnetic

INTRODUCTION

Refrigeration technology is of great significance for both industry and everyday life. Current refrigeration systems are mostly based on conventional vapor compression technology. Although highly optimized in recent decades, they still have a considerable undesirable impact on the environment[1]. Frequently used refrigerants have thousand-time stronger global warming potentials compared to CO2. To achieve carbon neutrality, solid-state refrigeration technology based on the caloric effects of solids has been proposed as an alternative solution. Various phase transitions caused by some calorimeter materials under external fields are accompanied by huge latent heat, which can be utilized for cooling purposes through designated thermodynamic cycles. Magnetocaloric effects (MCEs) is one of the most studied caloric effects, which is usually linked to magnetic-field-induced first-order transitions. Barocaloric effects (BCEs), as the counterpart and extension of the (MCEs), is defined as the change in the isothermal entropy or adiabatic temperature of the material during the application or withdrawal of the external pressure field. Materials with first-order phase transition are more likely to be the most potential barocaloric effect materials due to the sensitivity of the lattice to pressure.

Initially observed around 2,000 years, BCEs has been found in Pr1-xLaxNiO3[2] and CeSb[3]. Subsequent studies of magneto-elastically coupled materials for MCEs have revealed larger BCEs, such as in magnetic shape memory alloys including NiMnIn[4], La(Fe,Si)13[5,6], Gd5Si2Ge2[7], MnCoGe0.99In0.01[8], FeRh[9,10], and others. These materials exhibit a strong coupling between magnetic and lattice degrees of freedom. Usually, there is a magnetic phase with a larger volume and a magnetic phase with a smaller volume. The application of a sufficiently large hydrostatic pressure induces a change of the system from the large-volume to the small-volume phase, and simultaneously the magnetic phase transition takes place. Typically, the required driving pressures in these systems are as high as several hundred MPa, and a comparable pressure-induced entropy change to that induced by a magnetic field can be obtained.

In recent years, a great variety of materials have been reported with larger BCEs, such as AgI[11], organic-inorganic hybrid chalcogenide [TPrA][Mn(dca)3][12], ferroelectric (NH4)2SO4[13,14], spin-crossover complexes[15-18], and even natural rubber[19,20]. First-principles calculations also predicted sizable BCEs for lithium-ion conductor materials[21], fluorine ion conductor materials[22], and graphene[23]. In plastic crystals, the extensive molecular orientation disorder in plastic crystals leads to huge entropy changes larger than 100 J kg-1 K-1, and the driving pressures have been significantly reduced down to below 100 MPa, for which they are termed as colossal barocaloric effects[24-26].

Antiferromagnetic materials are effective in releasing their entropy change by pressure in addition to the magnetic field[10,27], with remarkably reduced driving pressures, especially in frustrated antiferromagnets. Recent research has found that larger BCEs are observed at phase transitions from frustrated antiferromagnetic (AFM) to paramagnetic states in nitrides (Mn3GaN[28], Mn3NiN[29]) with an anti-perovskite structure. This indicates that even small hydrostatic pressures (as low as 90 MPa in plastic crystals) can effectively act on the AFM system. In this work, we report on the barocaloric properties of Mn3-xPt1+xalloys at first-order phase transitions from low-temperature triangle-lattice frustrated to high-temperature colinear AFM states. The composition-dependent phase transition temperature (Tt) is about 331 K for the Mn2.82Pt1.18. The pressure-dependent calorimetric measurements suggest that entropy changes are saturated at around 60 MPa.

EXPERIMENTS

Polycrystalline samples of Mn3-xPt1+xwith different Mn contents (x = 0.04, 0.08, 0.1, and 0.18) were prepared by arc-melting the high-purity (99.9%) elements under an Ar atmosphere. The true composition was determined using a ThermoFisher iCAP6300 Inductive Coupled Plasma Emission Spectrometer (ICP). The ingots were remelted three times to ensure homogeneity. 2wt.% excessive Mn elements were added to compensate for losses during the melting process. The as-prepared ingots were annealed in encapsulated quartz tubes at 973 K for 120 h and followed by furnace cooling. The nature of the single-phase was checked using a Rigaku MiniFlex 600 X-ray diffractometer. The temperature-dependent X-ray diffraction (XRD) patterns were collected using Bruker D8 Advance X-ray diffractometer. The diffraction patterns were fitted to a cubic unit cell (space group Pm$$ \scriptsize \overline{3}$$m) in Jana2006[30]. The calorimetric data were collected as a function of temperature and pressure using a high-pressure differential scanning calorimeter (µDSC7, Setaram). The samples were enclosed in a high-pressure vessel made of Hastelloy. Constant pressure scans were performed at 0.1, 30, 60, and 90 MPa in the temperature region from 290 to 390 K, respectively. After subtracting the baseline background, the heat flow data can be converted to entropy changes. The magnetic properties are characterized using a Magnetic Properties Measurement System (MPMS-XL, Quantum Design) and a Physical Property Measurement System (PPMS-14T, Quantum Design).

RESULTS AND DISCUSSIONS

Mn3-xPt1+xcrystallizes in the ordered Cu3Au-type structure[31,32], where Mn atoms are located on the face centers of the cubic lattice formed by Pt atoms, as shown in Figure 1A. The compound with the stoichiometric composition magnetically orders into a colinear AFM state at about Tt ~365 K (magnetic transition temperature of Mn3Pt in the literature[32]), where a negligibly small tetragonal distortion is observed. As depicted, magnetic moments carried by Mn atoms are aligned along the c-axis, and the magnetic unit cell is constructed along the c-axis with doubled chemical unit cell. Note that the Mn atoms located on the ab-plane carry no ordered magnetic moment. The four Mn atoms form a square lattice, and the diagonal magnetic moments are parallel while the adjacent ones are anti-parallel. Spaced by the non-magnetic Mn atom, the magnetic moments of the two layers of the square lattices are antiferromagnetically coupled. As the temperature decreases below Tt, such a colinear AFM state transforms into a triangle-lattice AFM state, where magnetic moments are located on the (111) plane and point in the <211> direction, leading to a two-dimensional geometric spin frustration. This arrangement of magnetic moments ensures that the magnetic unit cell is identical to the chemical unit cell. This phase transition is a typical first-order magnetic phase transition, even if the lattice symmetry is maintained[33]. Based on the triangular AFM structure, anomalous Hall effects have been predicted and observed in films as well as bulk single crystals[34,35].

Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in <InlineParagraph>Mn<sub>3-</sub><i><sub>x</sub></i>Pt<sub>1+</sub><i><sub>x</sub></i></InlineParagraph>

Figure 1. (A) Magnetic structures of Mn3Pt alloys with triangular AFM and colinear AFM phases as reported by Tao et al.[27]. (B-D) heat flow vs. temperature curves, the contour plots of the variable temperature XRD patterns and lattice constant vs. temperature curves for sample x = 0.1, respectively. The inset in (C) is an enlarged image of the high Q region.

In the Mn3-xPt1+xsystem, the magnetic properties, in particular, Tt, are strongly dependent on the composition x. Shown in Figure 1B is the heat flow data of the x = 0.1 alloy under ambient pressure, where an endothermic peak is found at 360 K while an exothermic peak at 340 K with a thermal hysteresis of about 20 K, which is a signature of the first-order phase transition. The temperature corresponding to the peak in the heat flow curve is defined as the phase transition temperature. In this paper, we uniformly regard the transition temperature of the cooling process as the phase transition temperature of the sample. The entropy change at this transition is derived from being about 12.31 J kg-1 K-1, which is kind of small compared to other systems that exhibit strong first-order transitions. The reason will be clarified afterward. For example, the entropy change is 22.3 J kg-1 K-1 in Mn3GaN[28] while 43 J kg-1 K-1 in Mn3NiN[29]. The temperature-dependent XRD was used to monitor the lattice distortion during the phase transition. The contour plots of the XRD patterns are shown in Figure 1C as a function of temperature (T) and scattering vector (Q). The patterns can be indexed based on the reported cubic structure. Within our resolution, there is no distinguishable symmetry change in the temperature from 300 to 410 K. Four strong Bragg peaks are observed, and the (210) peak obviously shifts towards the lower Q around 360 K. The determined lattice constant displays a sudden jump at 360 K [Figure 1D], corresponding to relative changes in lattice constant and unit cell volume of 0.75% and 2.26%, respectively. These values are similar to those reported in other literature[36,37]. In addition, the structurally determined Tt is well consistent with the thermally determined one.

After confirming the first-order phase transition as the core of this study typically occurs at x = 0.1, we systematically extend the magnetic characterizations to x = 0.04, 0.08, 0.1, and 0.18. Their field-cooled magnetizations as a function of temperature are plotted in Figure 2A-D. As the temperature decreases, magnetizations abruptly drop at 331 K for x = 0.18, 336 K for x = 0.1, 355 K for x = 0.08, and 384 K for x = 0.04, indicating the first-order phase transitions. As for x = 0.1, the determined Tt at the magnetization is a few kelvins lower than that in the heat flow data. We summarize the Tt values of these four samples, along with their lattice constants at room temperature, in Figure 2E. As x changes from 0.18 to 0.04, Tt monotonically increases while the lattice constant a decrease. Unlike the linear behavior of a, Tt exhibits a saturation feature, remaining nearly constant when x is greater than 0.1. Such a tendency is consistent with the previous report[31,32,36]. This phenomenon is also explained in the literature[31] based on the theoretical phase diagram of Mn3Pt in the molecular field approximation.

Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in <InlineParagraph>Mn<sub>3-</sub><i><sub>x</sub></i>Pt<sub>1+</sub><i><sub>x</sub></i></InlineParagraph>

Figure 2. (A-D) Magnetization curves for samples x = 0.18, 0.1, 0.08, and 0.04, respectively. (E) Curves of phase transition temperature vs. manganese content x (blue squares and red dots indicate measurement by M-T and heat flow, respectively) and lattice constant vs. manganese content x for the samples. (Taking x = 0.08 as an example, the fitting parameters obtained using the lebail method to fit the XRD curve are: GOF = 1.47, Rp = 4.18, Rwp = 5.99). (F) The magnetization loop of x = 0.18. The inset shows the enlarged plot at low fields.

Selecting x = 0.18 as an example, the high-field isothermal magnetization is considered at 300 and 380 K, where the compound is in the triangular AFM and colinear AFM states, respectively. As shown in Figure 2F, the magnetizations essentially obey a linear relation as a function of applied fields up to 14 T, which reflects the dominating AFM interactions. However, the low-field regions (the inset of Figure 2F) show a weak nonlinearity at both temperatures, which might be attributed to the uncancelled moments due to the off-stoichiometry. In particular, there are small remanence and coercivity at 300 K. The exact origin is still unknown. Previous studies have also shown that a ferromagnetic component perpendicular to the (111) plane is allowed on both Mn and Pt atoms, except for the component in the (111) plane pointing in the [112] direction in the magnetic structure of the triangular AFM phase[31]. Furthermore, it has been pointed out that the origin of this ferromagnetism may be because the third site of the colinear AFM phase generates a net magnetic moment due to electron interactions at low temperatures of 100 K or even lower[38]. In contrast to the presence of net moments along the <111> direction, the magnetization curve of the single-crystal material along the <111> direction did not show ferromagnetic behavior[34].

Except for the influence of magnetic fields on the first-order phase transitions, we explore the impact of applied hydrostatic pressures. Heat flow data as a function of temperature are plotted under 0.1, 30, 60, and 90 MPa for x = 0.08 and x = 0.18 in Figure 3A and B, respectively. As the pressure increases, the endothermic and exothermic peaks move toward the high-temperature region simultaneously. The peak intensity has a small increase with pressure, while the peak width has a tendency to narrow, which is especially obvious in the x = 0.18 sample. In addition, the thermal hysteresis at 0.1 MPa for the x = 0.08 sample is the same as that at x = 0.1, which is 19 K. However, the thermal hysteresis at atmospheric pressure for x = 0.18 sample is 9 K and decreases to 5 K at 90 MPa with increasing pressure.

Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in <InlineParagraph>Mn<sub>3-</sub><i><sub>x</sub></i>Pt<sub>1+</sub><i><sub>x</sub></i></InlineParagraph>

Figure 3. (A and B) Heat flow data as a function of temperature for x = 0.08 and x = 0.18 at 0.1, 30, 60, and 90 MPa pressure, respectively. (C and D) Entropy change curves for x = 0.08 and x = 0.18 at different pressures, respectively. (E and F) The temperature-pressure phase diagrams for x = 0.08 and x = 0.18, respectively.

Entropy changes at the phase transition under constant pressure, ΔSP, are determined by integrating the heat flow data. The pressure-induced entropy changes ($$\Delta S_{P_{0}\rightarrow P}$$) when pressure is increased from ambient pressure (P0) to applied pressure (P) is defined as $$\Delta S_{P_{0}\rightarrow P}$$ = ΔSP - $$\Delta S_{P_{0}}$$. Figure 3C and D show $$\Delta S_{P_{0}\rightarrow P}$$ at the final pressure of 30, 60, and 90 MPa for x = 0.08 and 0.18, respectively. It can be seen that more than 80% of the maximum entropy changes are achieved at pressures as low as 30 MPa, and the entropy changes tend to be saturated at about 60 MPa. Note that the maximum entropy of 13.79 J kg-1 K-1 acquired in x = 0.08 at 90 MPa is larger than that in x = 0.18 whereas the so-called reversible region (overlapped region for the cooling and heating curves) is much larger for x = 0.18. The temperature-pressure phase diagrams are constructed based on the heat flow data [Figure 3E and F]. It can be seen that the thermal hysteresis is much smaller in x = 0.18, which is the reason why its reversible region is larger[14]. Furthermore, its thermal hysteresis is obviously reduced at higher pressures.

To understand the uniqueness of this system, Table 1 compares its barocaloric factors to those of other typical compounds, including organic plastic crystals, inorganic salts, magnetocaloric intermetallics, and other frustrated AFM systems. The slope of the phase boundary (dTt/dP) of this system is the largest, and the driving pressure (ΔP) is the smallest among all intermetallics, which are impressively comparable to those of prototype plastic crystal neopentyl glycol with colossal BCEs. However, the larger dTt/dP must lead to smaller entropy changes in terms of the Clausius-Clapeyron relation[39], even though the volume change is not too small. The small entropy change at this first-order phase transition is due to unique magnetic fluctuations in nature. Magnetic fluctuation refers to the fluctuation of magnetic (electron spin) moment in magnetic systems[40]. The interaction between the local moment and the itinerant electron matrix may enhance spin fluctuation. Frustration structures are often accompanied by strong spin fluctuations[41,42]. Neutron diffraction measurements suggest that the ordered moment is 3.3 µB/Mn atom in the colinear AFM state, whereas 2.2 µB/Mn atom in the triangular AFM[37]. The reduction in the latter should be attributed to spin fluctuations due to geometric frustration. As a result, the triangle-lattice AFM state is magnetically less ordered than the colinear AFM one, which leads to an increase of magnetic entropy across Tt. At the same time, the crystal lattice shows a normal contraction, and a reduction of entropy of the lattice subsystem is expected. We infer that the contributions of individuals to the total entropy change partially cancel each other out, and the remaining entropy change represents the overall entropy change of the material. According to the previous theoretic study, the system can be described by a nearest-neighboring exchange interaction J1 and a next-nearest-neighboring exchange interaction J2. J1 is always negative, but J2 can be negative or positive, dependent on the interatomic distance between Mn atoms[31]. At Tt, J2 just changes its sign due to the shrinkage of the Mn-Mn distance. In this sense, such a picture is similar to the exchange striction observed in NiMnIn alloys[43].

Table 1

Performance summary of several typical barocaloric materials

MaterialTt (K)dTt/dP
(K GPa-1)
P (MPa)V/V (%)$$\Delta S_{P_{0}\rightarrow P}$$ (J kg-1K-1)Ref.
NPG31313345-389[24]
C2B10H1227738060-106.2[44]
NH4I24381020-89[45]
NH4SCN364300205128.7[46]
Fe49Rh5131060250112[9,10]
MnNiSi0.61FeCoGe0.3931170260-44[47]
Ni0.95Fe0.05S27475100239.6[48]
Mn3GaN2906590121.6[28]
Mn3NiN26213.52800.435[29]
Mn2.92Pt1.08355139602.26*13.79This work

CONCLUSIONS

In summary, the first-order phase transitions of Mn3-xPt1+x(x = 0.04, 0.08, 0.1, and 0.18) compounds have been studied at varying temperatures, pressures, and magnetic fields. At the phase transitions, both magnetizations and lattice constants showed abrupt drops as the temperature decreased. While the phase transition temperatures decreased at lower Mn content, they increased at higher pressures. This system is highly susceptible to pressure, and the pressure-induced entropy changes are saturated at 60 MPa, which is the lowest among current intermetallics. This may be due to the intense geometric magnetic frustration.

DECLARATIONS

Authors’ contributions

Prepared the samples, collected the data, performed data analysis and contributed to the writing and revisions: Zhao X

Conceived the study, designed the study, and contributed to the writing and revisions: Li B, Zhang K

Collected some of the data and provided technical support: Qi J, Liu P, Zhang Z (Zhang Zhao), Qu L

Revision of articles: Zhang Z (Zhang Zhidong)

Availability of data and materials

The datasets used and analyzed during the current study are available from the corresponding author upon reasonable request.

Financial support and sponsorship

The work was supported by the Ministry of Science and Technology of China (Grant nos. 2021YFB3501201, 2022YFE0109900, and 2020YFA0406002) and the Key Research Program of Frontier Sciences of Chinese Academy of Sciences (Grant no. ZDBS-LY-JSC002).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2023.

REFERENCES

1. Sari O, Balli M. From conventional to magnetic refrigerator technology. Int J Refrig 2014;37:8-15.

2. Müller K, Fauth F, Fischer S, Koch M, Furrer A, Lacorre P. Cooling by adiabatic pressure application in Pr1-xLaxNiO3. Appl Phys Lett 1998;73:1056-8.

3. Strässle T, Furrer A, Lacorre P, Müller K. A novel principle for cooling by adiabatic pressure application in rare-earth compounds. J Alloys Compd 2000;303-304:228-31.

4. Mañosa L, González-Alonso D, Planes A, et al. Giant solid-state barocaloric effect in the Ni-Mn-In magnetic shape-memory alloy. Nat Mater 2010;9:478-81.

5. Mañosa L, González-Alonso D, Planes A, et al. Inverse barocaloric effect in the giant magnetocaloric La-Fe-Si-Co compound. Nat Commun 2011;2:595.

6. Fujieda S, Fujita A, Fukamichi K. Strong pressure effect on the curie temperature of itinerant-electron metamagnetic La(Fe0.88Si0.12)13Hy and La0.7Ce0.3(Fe0.88Si0.12)13Hy. Mater Trans 2009;50:483-6.

7. Yuce S, Barrio M, Emre B, et al. Barocaloric effect in the magnetocaloric prototype Gd5Si2Ge2. Appl Phys Lett 2012;101:071906.

8. Wu RR, Bao LF, Hu FX, et al. Giant barocaloric effect in hexagonal Ni2In-type Mn-Co-Ge-In compounds around room temperature. Sci Rep 2015;5:18027.

9. Stern-taulats E, Gràcia-condal A, Planes A, et al. Reversible adiabatic temperature changes at the magnetocaloric and barocaloric effects in Fe49Rh51. Appl Phys Lett 2015;107:152409.

10. Stern-taulats E, Planes A, Lloveras P, et al. Barocaloric and magnetocaloric effects in Fe49Rh51. Phys Rev B 2014;89:214105.

11. Aznar A, Lloveras P, Romanini M, et al. Giant barocaloric effects over a wide temperature range in superionic conductor AgI. Nat Commun 2017;8:1851.

12. Bermúdez-García JM, Sánchez-Andújar M, Castro-García S, López-Beceiro J, Artiaga R, Señarís-Rodríguez MA. Giant barocaloric effect in the ferroic organic-inorganic hybrid [TPrA][Mn(dca)3] perovskite under easily accessible pressures. Nat Commun 2017;8:15715.

13. Lloveras P, Stern-Taulats E, Barrio M, et al. Giant barocaloric effects at low pressure in ferrielectric ammonium sulphate. Nat Commun 2015;6:8801.

14. Mikhaleva E, Gorev M, Bondarev V, Bogdanov E, Flerov I. Comparative analysis of elastocaloric and barocaloric effects in single-crystal and ceramic ferroelectric (NH4)2SO4. Scripta Mater 2021;191:149-54.

15. Yu C, Huang J, Qi J, et al. Giant barocaloric effects in formamidinium iodide. APL Mater 2022;10:011109.

16. Salgado-beceiro J, Nonato A, Silva RX, et al. Near-room-temperature reversible giant barocaloric effects in [(CH3)4N]Mn[N3]3 hybrid perovskite. Mater Adv 2020;1:3167-70.

17. Ranke P, Alho B, Ribeiro P. First indirect experimental evidence and theoretical discussion of giant refrigeration capacity through the reversible pressure induced spin-crossover phase transition. J Alloys Compd 2018;749:556-60.

18. Szafrański M, Wei W, Wang Z, Li W, Katrusiak A. Research update: tricritical point and large caloric effect in a hybrid organic-inorganic perovskite. APL Mater 2018;6:100701.

19. Bom NM, Imamura W, Usuda EO, Paixão LS, Carvalho AMG. Giant barocaloric effects in natural rubber: a relevant step toward solid-state cooling. ACS Macro Lett 2018;7:31-6.

20. Miliante CM, Christmann AM, Usuda EO, et al. Unveiling the origin of the giant barocaloric effect in natural rubber. Macromolecules 2020;53:2606-15.

21. Sagotra AK, Chu D, Cazorla C. Room-temperature mechanocaloric effects in lithium-based superionic materials. Nat Commun 2018;9:3337.

22. Cazorla C, Errandonea D. Giant mechanocaloric effects in fluorite-structured superionic materials. Nano Lett 2016;16:3124-9.

23. Ma N, Reis MS. Barocaloric effect on graphene. Sci Rep 2017;7:13257.

24. Li B, Kawakita Y, Ohira-Kawamura S, et al. Colossal barocaloric effects in plastic crystals. Nature 2019;567:506-10.

25. Aznar A, Lloveras P, Barrio M, et al. Reversible and irreversible colossal barocaloric effects in plastic crystals. J Mater Chem A 2020;8:639-47.

26. Lloveras P, Tamarit J. Advances and obstacles in pressure-driven solid-state cooling: a review of barocaloric materials. MRS Energy Sustain 2021:8;3-15.

27. Tao K, Song W, Lin J, et al. Giant reversible barocaloric effect with low hysteresis in antiperovskite PdNMn3 compound. Scripta Mater 2021;203:114049.

28. Matsunami D, Fujita A, Takenaka K, Kano M. Giant barocaloric effect enhanced by the frustration of the antiferromagnetic phase in Mn3GaN. Nat Mater 2015;14:73-8.

29. Boldrin D, Mendive-tapia E, Zemen J, et al. Multisite exchange-enhanced barocaloric response in Mn3NiN. Phys Rev X 2018;8:041035.

30. Dusek M, Petricek V. Towards the routine application of computing system Jana2000. Acta Crystallogr A Found Crystallogr 2005;61:c104-5.

31. Krén E, Kádár G, Pál L, Sólyom J, Szabó P, Tarnóczi T. Magnetic structures and exchange interactions in the Mn-Pt system. Phys Rev 1968;171:574-85.

32. Krén E, Kádár G, Pál L, Szabó P. Investigation of the first-order magnetic transformation in Mn3Pt. J Appl Phys 1967;38:1265-6.

33. Tomiyoshi S, Yasui H, Kaneko T, et al. Magnetic excitations in Mn3Pt at high energies by the TOF method. J Magn Magn Mater 1990;90-91:203-4.

34. Zuniga-Cespedes BE, Manna K, Noad HML, et al. Observation of an anomalous hall effect in single-crystal Mn3Pt. Mater Sci 2022;2209:05865.

35. An N, Tang M, Hu S, et al. Structure and strain tunings of topological anomalous hall effect in cubic noncollinear antiferromagnet Mn3Pt epitaxial films. Sci China Phys Mech Astron 2020;63:297511.

36. Yasui H, Kaneko T, Yoshida H, Abe S, Kamigaki K, Mori N. Pressure dependence of magnetic transition temperatures and lattice parameter in an antiferromagnetic ordered alloy Mn3Pt. J Phys Soc Jpn 1987;56:4532-9.

37. Yasui H, Ohashi M, Abe S, et al. Magnetic order-order transformation in Mn3Pt. J Magn Magn Mater 1992;104-107:927-8.

38. Ricodeau JA. Model of the antiferromagnetic-antiferromagnetic transition in Mn3Pt alloys. J Phys F Met Phys 1974;4:1285-303.

39. Boldrin D. Fantastic barocalorics and where to find them. Appl Phys Lett 2021;118:170502.

40. Ehrenreich H, Spaepen F. Solid state physics: advances in research and applications. Amsterdam Boston: Academic Press; 2006.

41. Hemberger J, von Nidda HA, Tsurkan V, Loidl A. Large magnetostriction and negative thermal expansion in the frustrated antiferromagnet ZnCr2Se4. Phys Rev Lett 2007;98:147203.

42. Broholm C, Aeppli G, Espinosa GP, Cooper AS. Antiferromagnetic fluctuations and short-range order in a Kagomé lattice. Phys Rev Lett 1990;65:3173-6.

43. Li B, Ren WJ, Zhang Q, et al. Magnetostructural coupling and magnetocaloric effect in Ni-Mn-In. Appl Phys Lett 2009;95:172506.

44. Zhang K, Song R, Qi J, et al. Colossal barocaloric effect in carboranes as a performance tradeoff. Adv Funct Mater 2022;32:2112622.

45. Ren Q, Qi J, Yu D, et al. Ultrasensitive barocaloric material for room-temperature solid-state refrigeration. Nat Commun 2022;13:2293.

46. Zhang Z, Li K, Lin S, et al. Thermal batteries based on inverse barocaloric effects. Sci Adv 2023;9:eadd0374.

47. Lloveras P, Samanta T, Barrio M, et al. Giant reversible barocaloric response of MnNiSi)1-x(FeCoGe)x (x = 0.39, 0.40, 0.41). APL Mater 2019;7:061106.

48. Greca LG, Lehtonen J, Tardy BL, Guo J, Rojas OJ. Biofabrication of multifunctional nanocellulosic 3D structures: a facile and customizable route. Mater Horiz 2018;5:408-15.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Zhao X, Zhang K, Qi J, Liu P, Zhang Z, Qu L, Zhang Z, Li B. Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in Mn3-xPt1+x. Microstructures 2023;3:2023022. http://dx.doi.org/10.20517/microstructures.2022.46

AMA Style

Zhao X, Zhang K, Qi J, Liu P, Zhang Z, Qu L, Zhang Z, Li B. Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in Mn3-xPt1+x. Microstructures. 2023; 3(3): 2023022. http://dx.doi.org/10.20517/microstructures.2022.46

Chicago/Turabian Style

Zhao, Xueting, Kun Zhang, Ji Qi, Peng Liu, Zhao Zhang, Lin Qu, Zhidong Zhang, Bing Li. 2023. "Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in Mn3-xPt1+x" Microstructures. 3, no.3: 2023022. http://dx.doi.org/10.20517/microstructures.2022.46

ACS Style

Zhao, X.; Zhang K.; Qi J.; Liu P.; Zhang Z.; Qu L.; Zhang Z.; Li B. Low-pressure-driven barocaloric effects at colinear-to-triangular antiferromagnetic transitions in Mn3-xPt1+x. Microstructures. 2023, 3, 2023022. http://dx.doi.org/10.20517/microstructures.2022.46

About This Article

Special Issue

This article belongs to the Special Issue Magnetic Structure and Novel Physical Properties
© The Author(s) 2023. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
545
Downloads
610
Citations
1
Comments
0
3

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 10 clicks
Like This Article 3 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Microstructures
ISSN 2770-2995 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/