Next Article in Journal
Effects of the Volume Fraction of the Secondary Phase after Solution Annealing on Electrochemical Properties of Super Duplex Stainless Steel UNS S32750
Next Article in Special Issue
FeSiCr-Based Soft Magnetic Composites with SiO2 Insulation Coating Prepared Using the Elemental Silicon Powder Hydrolysis Method
Previous Article in Journal
Solidification Kinetics of an Al-Ce Alloy with Additions of Ni and Mn
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Magnetic Field and Hydrostatic Pressure on Metamagnetic Isostructural Phase Transition and Multicaloric Response of Fe49Rh51 Alloy

by
Alexander P. Kamantsev
1,
Abdulkarim A. Amirov
2,*,
Vladislav D. Zaporozhets
3,
Igor F. Gribanov
3,
Aleksay V. Golovchan
3,
Victor I. Valkov
3,
Oksana O. Pavlukhina
4,
Vladimir V. Sokolovskiy
4,5,
Vasiliy D. Buchelnikov
4,
Akhmed M. Aliev
2 and
Victor V. Koledov
1
1
Kotelnikov Institute of Radioengineering and Electronics of Russian Academy of Sciences, 125009 Moscow, Russia
2
Amirkhanov Institute of Physics, Dagestan Scientific Center of Russian Academy of Sciences, 367003 Makhachkala, Russia
3
Galkin Donetsk Institute for Physics and Engineering, 283050 Donetsk, Russia
4
Department of Condensed Matter Physics, Chelyabinsk State University, 454001 Chelyabinsk, Russia
5
Academic Research Center for Energy Efficiency, National University of Science and Technology “MISIS”, 119049 Moscow, Russia
*
Author to whom correspondence should be addressed.
Metals 2023, 13(5), 956; https://doi.org/10.3390/met13050956
Submission received: 17 April 2023 / Revised: 5 May 2023 / Accepted: 8 May 2023 / Published: 15 May 2023
(This article belongs to the Special Issue Feature Papers in Metallic Functional Materials)

Abstract

:
The effect of a high magnetic field up to 12 T and a high hydrostatic pressure up to 12 kbar on the stability of the metamagnetic isostructural phase transition and the multicaloric effect of Fe49Rh51 alloy has been studied. The phase transition temperature shifts under the magnetic field and the hydrostatic pressure on with the rates of dTm0dH = −9.2 K/T and dTm/dP = 3.4 K/kbar, respectively. The magnetocaloric and multicaloric (under two external fields) effects were studied via indirect method using Maxwell relations. The maximum of the entropy change is increasing toward the high temperature region from ∆S~2.5 J/(kg K) at 305 K to ∆S~2.7 J/(kg K) at 344 K under simultaneously applied magnetic field of 0.97 T and hydrostatic pressure of 12 kbar. The obtained results were explained using the first-principle calculations of Gibbs energies and the phonon spectra of the ferromagnetic and the antiferromagnetic phases. Taking into account the low concentration of antisite defects in the calculation cells allows us to reproduce the experimental dTm/dP coefficient.

1. Introduction

The FeRh alloys with the near-equiatomic composition are characterized by a CsCl-type crystal structure with a first-order magnetic phase transition from the antiferromagnetic (AFM) state to the ferromagnetic (FM) at the same average temperature Tm, which depends strongly on the chemical composition [1,2,3]. These alloys exhibit significant changes in the magnetization and lattice parameter without the change in the crystal symmetry in the vicinity of the Tm [4], which can be called the magnetovolume effect at the metamagnetic isostructural phase transition (MIPT). The FeRh alloys are convenient model objects for studying the nature of the MIPT, due to their simple crystal structure and the bright effects observed near Tm. There are several methods to control the MIPT temperature in FeRh alloys such as a magnetic field [5], a hydrostatic pressure [6], an electric field-induced strain [7,8,9], a heat treatment [10,11], a chemical substitution [12], and an ion irradiation [13]. The first studies of the temperature dependences of magnetization, Young’s modulus, and the crystal lattice parameter were carried out via dilatometric method on the equiatomic alloy Fe50Rh50 [2], and the rates of MIPT temperature shift dTm/dP and dTm0dH in the hydrostatic pressure p and the magnetic field μ0H were estimated. The FeRh alloys exhibit high magnetization saturation up to 130 Am2/kg in the FM phase [14], and the MIPT is accompanied by a strong anomaly of the specific heat capacity [15] and an isotropic volume increase of ~1% [2]. The magnetostriction of the Fe50Rh50 alloy was firstly measured in the temperature range 290–400 K in magnetic fields up to 15 T [16]. The published phase diagrams for near-equatomic FeRh alloys as a function of a magnetic field and a pressure present the collected data of the theoretical calculations and the experimental results obtained from the samples with different compositions and fabrication protocols [17,18,19]. For example, the P–T and the μ0H–T diagrams obtained from experimental data for Fe49Rh51 alloy are limited by the hydrostatic pressure of 5 kbar [19] and the magnetic fields of 6 T [20].
The FeRh alloys have a number of unique functional properties in the external magnetic and mechanical fields known as “giant” caloric and multicaloric effects [17,19,20,21,22,23] due to the phenomena near Tm described above. The large magnetocaloric effect (MCE) [21], the elastocaloric effect [22] and the barocaloric effect (BCE) [20] were observed in FeRh alloys near the MIPT temperatures. For the first time, the MCE in the Fe48Rh52 alloy was reported in [24], where the dependence of the critical magnetic field on the MIPT temperature was established and the observed inverse MCE (the decreasing of the sample temperature under increasing of the external magnetic field) was described as a negative factor for constructing the critical magnetic field diagram. The near-equiatomic FeRh alloys were proposed as promising materials for the magnetic cooling at the room temperatures due to record values of MCE, where the values of the adiabatic temperature changes were as follows: ΔTad = −12.9 K at μ0H = 1.95 T in the Fe49Rh51 [21,25]; ΔTad = −20.2 K at μ0H = 8 T in the Fe48Rh52 [26]. In addition, the MCE was studied via the direct method in pulsed fields μ0H = 50 T in the Fe49Rh51 alloy [27]. The BCE in Fe49Rh51 alloys was studied under hydrostatic pressure up to 2.5 kbar [20,28]. The effect of the magnetic field and the hydrostatic pressure on the MIPT trough the electrical resistance measurements were demonstrated in (Fe1−xNix)49Rh51 alloys at the hydrostatic pressure up to 13 kbar [29], as well as in the Fe48Rh52 alloy at the applied magnetic field up to 8 T [30].
One of the promising approaches of the solid-state cooling based on the caloric effects is the multicaloric approach, which is based on the combination of the different external fields (magnetic, mechanical, electric) [31,32]. The multicaloric approach based on the combination of the MCE and the BCE under the external magnetic field up to 5 T and the hydrostatic pressure up to 5 kbar was experimentally used for the Fe49Rh51 alloy [19]. Moreover, the uniaxial mechanical force can be used to improve the efficiency of systems operating on MCEs in magnetic materials with the first-order phase transitions [23,33]. The uniaxial compression in combination with the magnetic field can be used to control hysteresis effects that negatively contribute to the efficiency of magnetic cooling systems [34].
The present work is aimed to investigate the two interrelated phenomena in FeRh alloys within the framework of a single study: (1) the effect of the combination of the magnetic field and the hydrostatic pressure on the temperature of the phase transition; and (2) the measurements of the multicaloric effect in magnetic field combined with the hydrostatical pressure. The FeRh samples with the same Fe49Rh51 composition but with two different heat treatment protocols were used as an object. The ab initio approach, which includes the total energy and the phonons calculations, was employed in order to describe the experimental P–T diagram.

2. Materials and Methods

The ingot of Fe49Rh51 alloy was obtained via induction melting in an argon atmosphere. The samples in shape of the disks with a diameter of 5 mm and thickness of 1 mm were cut out of the ingot. The samples were annealed in the pre-pumped quartz ampoules (~1 Pa) at a temperature of 1273 K. The first series of Fe49Rh51 samples was annealed for 72 h and quenched in ice water following this—the FR72h samples—and the second series was annealed for 48 h and cooled in the furnace—the FR48h samples. The structure and the elemental composition of the samples were confirmed via energy dispersion spectroscopy (EDS) and X-ray diffraction analysis (XRD) and presented in earlier work [35]. The MIPT temperatures of FR72h and FR48h samples were determined using two different methods: (1) the differential scanning calorimetry (DSC) on the calorimeter Netzsch DSC 204F1 Phoenix in zero magnetic field; and (2) the temperature magnetometry on the vibration magnetometer Versa Lab Quantum Design in a low magnetic field of 1 mT. The rate of the temperature change in these experiments was 10 K/min.
The magnetic field dependences of the magnetization of FR72h sample at the different temperatures in vicinity of the MIPT (300–328 K) were measured using the vibration magnetometer Quantum Design PPMS-14T in the magnetic field up to 8 T. The temperature dependences of the electrical resistance of FR72h sample at the different magnetic fields (0–12 T) were measured using the four-point method on Quantum Design PPMS-14T with the insert for the transport properties measurements. The experiments for studies of the multicaloric effects near the MIPT of FR48h sample at simultaneously applied magnetic field (0.155 T or 0.97 T) and hydrostatic pressure (up to 12 kbar) were carried out using the Domenicali-type pendulum magnetometer [36]. The mass of the FR48h sample was approximately 12.5 mg. The non-magnetic high-pressure cells with the sample were mounted at the end of the pendulum. This cell was made of the beryllium bronze, and pure indium (In) was used as the pressure transmitting medium. The pressure in the cell was determined based on the calibration of this system on chromium telluride (CrTe), which has the well-known phase transition temperatures shift on the applied pressure of −6 K/kbar. The MCE and the multicaloric effect were estimated via indirect method with the help of the well-known Maxwell relations.
The theoretical study of the ground-state properties of FeRh alloy was conducted using the density functional theory (DFT) calculations within the projector augmented wave method (PAW) method as implemented in the VASP package [37,38]. To describe the exchange correlation effects, both the generalized-gradient approximation (GGA) in the form of Perdew–Burke–Ernzerhof (PBE) [39] and the recently developed strongly constrained and appropriately normed (SCAN) meta-GGA [40] were used for the exchange correlation functional. The PAW potentials were selected with the following set of valence electrons: 3p63d74s1 for Fe and 4p64d85s1 for Rh. The plane wave cut-off parameter was taken as 700 eV and the Brillouin zone sampling is confined to 8 × 8 × 8 k-mesh generated by the Monkhorst-Pack scheme. The 16-atom cubic supercell with the space group Fm-3m and FM, AFM orders of Fe atoms was taken into account in the geometry optimization calculations. The zero temperature phonon spectrum calculations was performed in the harmonic approximation considering the small displacement method (0.01 Å) as implemented in the Phonopy package [41]. The 2 × 2 × 2 (32 atoms) supercell with four basis atoms in a fcc unit cell for AFM phase and 3 × 3 × 3 (54 atoms) supercell with two basis atoms in a bcc unit cell for FM phase were chosen. In order to investigate the thermodynamic properties of the FeRh under temperature and pressure, the quasi-harmonic Debye model implemented in the GIBBS code [42,43] was applied.

3. Results and Discussion

3.1. Experimental Studies

The DSC thermograms at the heating and cooling protocols for FR72h and FR48h samples are shown in Figure 1 (left axes). The temperatures of the start (AFS and FS) and the finish (AFF and FF) of the MIPT for the FR72h sample were determined from the DSC curves (Figure 1a) as AFS = 317 K, AFF = 308 K at cooling (from FM to AFM state) and FS = 321 K, FF = 330 K at heating (from AFM to FM state). The latent heat of the MIPT for FR72h sample was calculated as the area under the DSC peaks: λC = 4300 J/kg at cooling, and λH = 4400 J/kg at heating (Figure 1a). The corresponding MIPT temperatures for the FR48h sample were determined to be AFS = 316 K, AFF = 295 K at cooling and FS = 306 K, FF = 328 K at heating (Figure 1b). The latent heat of the MIPT for FR48h sample was as follows: λC = 4300 J/kg at cooling, and λH = 3600 J/kg at heating (Figure 1b).
The temperature dependences of the magnetization for FR72h and FR48h samples at heating and cooling in low magnetic field of 1 mT are shown in Figure 1 (right axes). These curves show the temperature hysteresis of the first order MIPT between two phases of Fe49Rh51: AFM at lower temperatures and FM at higher temperatures. The MIPT temperatures obtained from the DSC analysis are consistent with the temperatures from magnetometry data with an accuracy of 1 K (Figure 1a,b). In general, the differences in the heat treatment of the FR72h and FR48h samples greatly affected on AFF and FS temperatures (lower in FR48h).
Figure 2a demonstrates the magnetic field dependences of the magnetization M(μ0H) for FR72h sample at different temperatures from the range 300–328 K in magnetic field of 8 T. The magnetic field-induced MIPT is observed in Figure 2a, similar to that were observed in Fe48Rh52 alloy in [44]. The magnetization hysteresis at the lowest temperature (300 K, the black curve in Figure 2a) is wide because the sample initially is completely in the AFM phase (lower than AFF = 308 K). The magnetization hysteresis at highest temperature (328 K, the yellow curve in Figure 2a) is narrow because, initially, the sample is close to complete transition to the FM phase (FF = 330 K).
Figure 2b demonstrates the temperature dependences of the electrical resistivity for the FR72h sample at heating and cooling at different magnetic fields up to 12 T. These measurements show the shifts in the MIPT temperatures under the high magnetic fields. The temperature hysteresis of the MIPT in Fe49Rh51 alloy according to the measurements results shifts to the lower temperature region with the increasing in magnetic fields, which is explained by the presence of the nuclei of the high-temperature FM phase at the low temperatures. Additionally, as a result, some broadening of the temperature hysteresis is observed with the increasing of the external magnetic field (Figure 3b), which is in good agreement with the results obtained for the Fe48Rh52 alloy [30].
Figure 3 demonstrates the temperature dependences of the magnetization of the FR48h sample at applied hydrostatic pressure from 0 to 12 kbar in magnetic fields of 0.155 T (Figure 3a) and 0.97 T (Figure 3b). The temperature hysteresis of the MIPT in the Fe49Rh51 alloy shifts to the higher temperatures region with the increase in hydrostatic pressure, and there is a certain narrowing of the temperature hysteresis at the same time. This is explained by the strong magneto-elastic coupling between the magnetic and structural subsystems in the solid. The external hydrostatic pressure compresses the crystal lattice and thus blocks the origination of the high-temperature FM phase with the larger unit cell volume. However, studies in such a range of magnetic fields and pressures for FeRh alloys are not known in the literature, and this would certainly expand our knowledge about the nature of the MIPT of these alloys.
If both the magnetization and the entropy are continuous functions of the temperature and the magnetic field, then an infinitesimal change in isobaric-isothermal entropy can be associated with the magnetization M, the magnetic field strength H and the absolute temperature T using one of Maxwell’s relations in integral form [45,46,47]:
Δ S m a g = μ 0 0 H ( M T ) d H
It is possible to calculate the magnetic entropy change ∆S in FR72h sample of Fe49Rh51 alloy connected with the external magnetic field change via Equation (1) using the data of the isothermal magnetization in Figure 2a. The calculation results for μ0H = 1 and 2 T are shown in Figure 4a, while the maximum value was ∆S = 5.6 J/(kg K) at μ0H = 1 T. This result correlates well with the calculated data from the heat capacity measurement in [48]: ∆S = 5.5 J/(kg K) at μ0H = 1 T for Fe49Rh51 alloy.
Additionally, it is possible to calculate the magnetic entropy change ∆S in FR48h sample of Fe49Rh51 alloy connected with the external magnetic field change via Equation (1) using the data of the magnetization under different external pressures in Figure 3b. The calculation results for μ0H = 0.97 T at cooling protocol are shown in Figure 4b, while the maximum obtained value was ∆S = 2.7 J/(kg K) under the external pressure p = 12 kbar. Thus, there is the difference in ∆S more than 2 times for FR72h and FR48h samples at the magnetic field changes of 1 T, which is explained by the difference in the heat treatment of the samples.
The magnetic phase μ0H–T diagram based on the data of the temperature dependence of the electrical resistance for FR72h sample of Fe49Rh51 alloy in high magnetic fields was presented in Figure 5. The MIPT temperatures in different magnetic fields up to 12 T from Figure 2b were plotted on phase μ0H–T diagram. It was established that the magnetic field shifts the MIPT to the region of the lower temperatures with the average rate dTm0dH = −9.2 K/T. The obtained value generally corresponds with known data in the literature for the Fe49Rh51 alloy [20]: dTm0dH = −9.6 K/T in magnetic field up to 6 T.
The phase P–T diagram based on the data of the temperature dependence of the magnetization (at 0.97 T) for FR48h sample of Fe49Rh51 alloy under the high hydrostatic pressure is presented in Figure 5. The MIPT temperatures in different external hydrostatic pressures up to 12 kbar from Figure 3b were plotted in the P–T. diagram. The restores stability of the AFM order and shifts the MIPT to the region of higher temperatures with the average rate dTm/dP = 3.4 K/kbar. The obtained value is lower than the data presented in the literature for the Fe49Rh51 alloy [20]: dTm/dP = 6.4 K/kbar in hydrostatic pressure up to 2.5 kbar—this discrepancy is explained by the difference in the heat treatment of the samples.

3.2. DFT Calculations

Since Fe-based alloys belong to strongly correlated electronic systems, there is an interest in studying the exchange correlation effects on the ground state properties of the AFM and FM phases of FeRh compound. Figure 6a displays the total energy difference as a function of lattice constant with respect to the AFM ground state calculated with PBE and SCAN. To find an optimized lattice parameter, the energy curves are fitted to the Birch-Murnaghan equation of state. As can be seen that both functionals predict the AFM order as energetically favorable. However, for both types of magnetic ordering, the SCAN lattice parameters (2.981 Å and 2.962 Å for FM and AFM) are found to be slightly smaller than those of the PBE one (3.003 Å and 2.988 Å for FM and AFM). It is worth noting that additional electron correlation effects beyond the GGA yield a significantly smaller energy difference ΔE (by approximately an order of magnitude) in comparison to that of PBE. Thus, the calculated SCAN value is ΔE = 2.3 meV/atom. In regard to the magnetic moments, the SCAN slightly overestimates them compared to those for the PBE. This is a known feature of the SCAN functional due to self-interaction corrections, which can be measured using the Coulomb energy U [49,50,51,52,53].
Next, let us discuss the phonon spectra for the AFM phase, which are calculated in the harmonic approximation using the SCAN and PBE functionals. It should be noted that we omit the results for the FM case here, since they show similar behavior without any sign of instability, as reported in earlier DFT calculations (e.g., Refs. [54,55,56]). As evident from Figure 6b, the PBE calculations demonstrate the phonon instability observed at the X point that contradicts the experimental data. On the other side, the meta-GGA calculations lead to the degeneracy of the anomaly of the optical phonon spectrum at the X point, thereby making the AFM phase dynamically stable at 0 K as in a case of FM phase. In general, a similar phonon spectra can be observed for both FM and AFM phases. The calculated Debye temperatures Θ D from the phonon dispersion relations are found to be approximately 370 K and 394 K for FM and AFM phase, respectively.
We now turn to a discussion of the calculated metamagnetic temperature between the AFM and FM phases as a function of applied pressure and compare these with the experimental data shown in Figure 7. To calculate the Gibbs energy for both FM and AFM phases, we consider only a lattice contribution through the quasi-harmonic Debye model [42,43], which allows us to calculate thermodynamic properties as a function of temperature and pressure from the DFT energy-volume E(V) points and Debye temperatures taken from the phonon spectra calculations. It should be noted that the free-energy curves for FM and AFM phases calculated within the PBE E(V) data do not intersect due to a large energy difference between the two phases that is in agreement with earlier theoretical works [54,56]. However, the consideration of SCAN E(V) curves in combination with Θ D as input in the Gibbs energy calculations predict well the transition temperature Tm~310 K under ambient pressure, which is in good agreement with the experiment. This is mainly due to the smallest difference in the total energies of FM and AFM phases. However, an application of hydrostatic pressure up to 12 kbar leads to a sufficient increase in Tm up to 550 K, which contradicts the experimental data. The theoretical rate of change in Tm with pressure is dTm/dP~20 K/kbar, while the experiment gives ~3.4 K/kbar. It should be noted a similar effect of pressure has been observed in earlier theoretical calculations [17,18].
In order to resolve this contradiction, we considered the antisite defects in a high-temperature FM phase, which can occur in an ordered alloy when atoms of different types exchange their positions. For the investigation of point defects, we considered a periodically repeated supercell with 32, 64, 128, and 256 atoms, in which only one antisite pair (Fe–Rh) is embedded. Obviously, the larger the supercell, the lower the concentration of the antisite defect. As an example, the FM composition (Fe15Rh1)–(Rh15Fe1) formed on a 32-atom defect supercell has the antisite defect fraction of 1/16. We denote this defect composition as [email protected]%. In accordance with our calculations, there is no metamagnetic transition for this composition due to a large energy difference between both phases ΔE ≈ 47 meV/atom, whereas it takes place for compositions with a smaller defect concentration. As is evident from the P–T diagram, the transition temperature is only slightly affected by an applied pressure for [email protected]% (the defect supercell with 64 atoms, ΔE ≈ 23 meV/atom). However, the further reduction in defect fraction in the FM phase by half reveals a good agreement between the theory and the experiment for [email protected]% (the defect supercell with 128 atoms, ΔE ≈ 16 meV/atom). We would like to note that the subsequent change in the defect concentration results in an increase in the dTm/dP slope, which becomes close to the case of pure AFM and FM phases, as exemplified by [email protected]% (the defect supercell with 256 atoms, ΔE ≈ 3.7 meV/atom). We are reminded that ΔE for compound with a pure FM and AFM phase is calculated to be 2.3 meV/atom, as shown in Figure 6a. Thus, we can conclude that the theoretical slope of dTm/dP strongly depends on the defect concentration due to a change in ΔE and Θ D between a pure AFM phase and a defect FM one.

4. Conclusions

The effects of the high hydrostatic pressure up to 12 kbar and the high magnetic field up to 12 T on the stability of the MIPT in Fe49Rh51 alloy samples were studied. As expected, the strategy of the combination of the external fields leads to opposite contributions: the magnetic field expands the FM region and shifts the MIPT to the lower temperatures with the rate dTm0dH = −9.2 K/T, and the hydrostatic pressure restores the stability of the AFM order and shifts the MIPT to the higher temperatures with the rate dTm/dP = 3.4 K/kbar. The study of the multicaloric effect demonstrates that the hydrostatic pressure can be used for the tuning of the maximum of the magnetic entropy change, which has prospective applications for the upgrading of the magnetic cooling systems. The obtained experimental PT diagram is explained by using the DFT calculations within the exchange correlation effects beyond the well-known GGA. The experimental dTm/dP slope is reproduced theoretically when the antisite defects were considered in the computational supercell.

Author Contributions

Conceptualization, A.A.A.; Methodology, I.F.G., A.V.G. and V.I.V.; Software, V.V.S. and O.O.P.; Formal analysis, A.A.A. and V.I.V.; Investigation, A.P.K., A.A.A., V.D.Z., A.V.G. and V.V.S.; Writing—original draft, A.P.K. and A.A.A.; Supervision, V.D.B., A.M.A. and V.V.K.; Project administration, V.V.K., V.D.B. and A.M.A. All authors of the manuscript contributed equally to data acquisition and analysis, writing, editing and formatting. All authors have read and agreed to the published version of the manuscript.

Funding

The DSC and the electrical resistivity studies were carried out by A.P.K. and V.V.K. within the framework of the state task of Kotelnikov IRE RAS. V.V.S. acknowledges the financial support from the Priority-2030 Program of NUST “MISiS” (grant No. K2-2022-022) (the total energy calculations). O.O.P. and V.D.B. acknowledge the Ministry of Science and Higher Education of the Russian Federation within the Russian State Assignment, under Contract 075-01493-23-00 (the phonon calculations).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors acknowledge Tino Gottschall from the High Magnetic Field Laboratory (HLD) at the Helmholtz-Zentrum Dresden-Rossendorf (HZDR), a member of the European Magnetic Field Laboratory (EMFL) for the help with the magnetization measurement in high magnetic fields.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hofer, E.M.; Cucka, P. Magnetic properties of Rh-rich FeRh alloy. J. Phys. Chem. Solids 1966, 27, 1552–1555. [Google Scholar] [CrossRef]
  2. Zakharov, A.I.; Kadomtseva, A.M.; Levitin, P.Z.; Ponyatovskii, E.G. Magnetic and magnetoelastic properties of a metamagnetic iron-rhodium alloy. J. Exp. Theor. Phys. 1964, 19, 1348–1353. [Google Scholar]
  3. Belo, J.H.; Pires, A.L.; Araújo, J.P.; Pereira, A.M. Magnetocaloric materials: From micro- to nanoscale. J. Mater. Res. 2019, 34, 134–157. [Google Scholar] [CrossRef]
  4. Zakharov, A. Crystal lattice parameter and structural distortions in Fe-Rh alloy at phase transitions. Fiz. Met. I Metalloved. 1967, 24, 84–90. [Google Scholar]
  5. Lommel, J.M.; Kouvel, J.S. Effects of mechanical and thermal treatment on the structure and magnetic transitions in FeRh. J. Appl. Phys. 1967, 38, 1263–1264. [Google Scholar] [CrossRef]
  6. Vinokurova, L.I.; Vlasov, A.V.; Pardavi-Horváth, M. Pressure effects on magnetic phase transitions in FeRh and FeRhIr alloys. Phys. Status Solidi 1976, 78, 353–357. [Google Scholar] [CrossRef]
  7. Cherifi, R.O.; Ivanovskaya, V.; Phillips, L.C.; Zobelli, A.; Infante, I.C.; Jacquet, E.; Garcia, V.; Fusil, S.; Briddon, P.R.; Guiblin, N.; et al. Electric-field control of magnetic order above room temperature. Nat. Mater. 2014, 13, 345–351. [Google Scholar] [CrossRef]
  8. Amirov, A.A.; Gottschall, T.; Chirkova, A.M.; Aliev, A.M.; Baranov, N.V.; Skokov, K.P.; Gutfleisch, O. Electric-field manipulation of the magnetocaloric effect in a Fe49Rh51/PZT composite. J. Phys. D Appl. Phys. 2021, 54, 505002. [Google Scholar] [CrossRef]
  9. Qiao, K.; Wang, J.; Hu, F.; Li, J.; Zhang, C.; Liu, Y.; Yu, Z.; Gao, Y.; Su, J.; Shen, F.; et al. Regulation of phase transition and magnetocaloric effect by ferroelectric domains in FeRh/PMN-PT heterojunctions. Acta Mater. 2020, 191, 51–59. [Google Scholar] [CrossRef]
  10. Sánchez-Valdés, C.F.; Gimaev, R.R.; López-Cruz, M.; Sánchez Llamazares, J.L.; Zverev, V.I.; Tishin, A.M.; Carvalho, A.M.G.; Aguiar, D.J.M.; Mudryk, Y.; Pecharsky, V.K. The effect of cooling rate on magnetothermal properties of Fe49Rh51. J. Magn. Magn. Mater. 2020, 498, 166130. [Google Scholar] [CrossRef]
  11. Rodionov, V.; Amirov, A.; Annaorazov, M.; Lähderanta, E.; Granovsky, A.; Aliev, A.; Rodionova, V. Thermal hysteresis control in Fe49Rh51 alloy through annealing process. Processes 2021, 9, 772. [Google Scholar] [CrossRef]
  12. Barua, R.; Jiménez-Villacorta, F.; Lewis, L.H. Towards tailoring the magnetocaloric response in FeRh-based ternary compounds. J. Appl. Phys. 2014, 115, 17A903. [Google Scholar] [CrossRef]
  13. Tohki, A.; Aikoh, K.; Iwase, A.; Yoneda, K.; Kosugi, S.; Kume, K.; Batchuluun, T.; Ishigami, R.; Matsui, T. Effect of high temperature annealing on ion-irradiation induced magnetization in FeRh thin films. J. Appl. Phys. 2012, 111, 07A742. [Google Scholar] [CrossRef]
  14. Kouvel, J.S.; Hartelius, C.C. Anomalous Magnetic Moments and Transformations in the Ordered Alloy FeRh. In Proceedings of the Seventh Conference on Magnetism and Magnetic Materials, Phoenix, AZ, USA, 13–16 November 1961; Springer: Boston, MA, USA, 1962; pp. 1343–1344. [Google Scholar] [CrossRef]
  15. Richardson, M.J.; Melville, D.; Ricodeau, J.A. Specific heat measurements on an Fe Rh alloy. Phys. Lett. A 1973, 46, 153–154. [Google Scholar] [CrossRef]
  16. Levitin, R.; Ponomarev, B. Magnetostriction of the Metamagnetic Iron-rhodium Alloy. Sov. J. Exp. Theor. Phys. 1966, 23, 984–985. [Google Scholar]
  17. Mendive-Tapia, E.; Castán, T. Magnetocaloric and barocaloric responses in magnetovolumic systems. Phys. Rev. B—Condens. Matter Mater. Phys. 2015, 91, 224421. [Google Scholar] [CrossRef]
  18. Gruner, M.E.; Entel, P. Simulation of the (p, T) phase diagram of the temperature-driven metamagnet α-FeRh. Phase Transit. 2005, 78, 209–217. [Google Scholar] [CrossRef]
  19. Stern-Taulats, E.; Castán, T.; Planes, A.; Lewis, L.H.; Barua, R.; Pramanick, S.; Majumdar, S.; Mañosa, L. Giant multicaloric response of bulk Fe49Rh51. Phys. Rev. B 2017, 95, 104424. [Google Scholar] [CrossRef]
  20. Stern-Taulats, E.; Planes, A.; Lloveras, P.; Barrio, M.; Tamarit, J.-L.; Pramanick, S.; Majumdar, S.; Frontera, C.; Mañosa, L. Barocaloric and magnetocaloric effects in Fe49Rh51. Phys. Rev. B 2014, 89, 214105. [Google Scholar] [CrossRef]
  21. Nikitin, S.A.; Myalikgulyev, G.; Tishin, A.M.; Annaorazov, M.P.; Asatryan, K.A.; Tyurin, A.L. The magnetocaloric effect in Fe49Rh51 compound. Phys. Lett. A 1990, 148, 363–366. [Google Scholar] [CrossRef]
  22. Nikitin, S.; Myalikgulyev, G.; Annaorazov, M.; 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]
  23. Gràcia-Condal, A.; Stern-Taulats, E.; Planes, A.; Mañosa, L. Caloric response of Fe49Rh51 subjected to uniaxial load and magnetic field. Phys. Rev. Mater. 2018, 2, 84413. [Google Scholar] [CrossRef]
  24. Ponomarev, B.K. Investigation of the antiferro-ferromagnetism transition in an FeRh alloy in a pulsed magnetic field up to 300 kOe. Sov. Phys. JETP 1973, 36, 105–107. [Google Scholar]
  25. Annaorazov, M.P.; Asatryan, K.A.; Myalikgulyev, G.; Nikitin, S.A.; Tishin, A.M.; Tyurin, A.L. Alloys of the FeRh system as a new class of working material for magnetic refrigerators. Cryogenics 1992, 32, 867–872. [Google Scholar] [CrossRef]
  26. Aliev, A.M.; Batdalov, A.B.; Khanov, L.N.; Kamantsev, A.P.; Koledov, V.V.; Mashirov, A.V.; Shavrov, V.G.; Grechishkin, R.M.; Kaul, A.R.; Sampath, V. Reversible magnetocaloric effect in materials with first order phase transitions in cyclic magnetic fields: Fe48Rh52 and Sm0.6Sr0.4MnO3. Appl. Phys. Lett. 2016, 109, 202407. [Google Scholar] [CrossRef]
  27. Kamantsev, A.P.; Amirov, A.A.; Koshkid’ko, Y.S.; Mejía, C.S.; Mashirov, A.V.; Aliev, A.M.; Koledov, V.V.; Shavrov, V.G. Magnetocaloric Effect in Alloy Fe49Rh51 in Pulsed Magnetic Fields up to 50 T. Phys. Solid State 2020, 62, 160–163. [Google Scholar] [CrossRef]
  28. Stern-Taulats, E.; Gràcia-Condal, A.; Planes, A.; Lloveras, P.; Barrio, M.; Tamarit, J.L.; Pramanick, S.; Majumdar, S.; Mañosa, L. Reversible adiabatic temperature changes at the magnetocaloric and barocaloric effects in Fe49Rh51. Appl. Phys. Lett. 2015, 107, 152409. [Google Scholar] [CrossRef]
  29. Kamenev, V.I.; Kamenev, K.V.; Todris, B.M.; Zavadskii, E.A.; Varyukhin, V.N.; Baranov, N.V. Effect of pressure on magnetic phases of (Fe1-xNix)49Rh51. High Press. Res. 2003, 23, 195–199. [Google Scholar] [CrossRef]
  30. Batdalov, A.B.; Aliev, A.M.; Khanov, L.N.; Kamantsev, A.P.; Mashirov, A.V.; Koledov, V.V.; Shavrov, V.G. Specific heat, electrical resistivity, and magnetocaloric study of phase transition in Fe48Rh52 alloy. J. Appl. Phys. 2020, 128, 013902. [Google Scholar] [CrossRef]
  31. Amirov, A.A.; Tishin, A.M.; Pakhomov, O.V. Multicalorics—New materials for energy and straintronics (Review). Phys. Solid State 2022, 64, 383. [Google Scholar] [CrossRef]
  32. Stern-Taulats, E.; Castán, T.; Mañosa, L.; Planes, A.; Mathur, N.D.; Moya, X. Multicaloric materials and effects. MRS Bull. 2018, 43, 295–299. [Google Scholar] [CrossRef]
  33. Gottschall, T.; Bykov, E.; Gràcia-Condal, A.; Beckmann, B.; Taubel, A.; Pfeuffer, L.; Gutfleisch, O.; Mañosa, L.; Planes, A.; Skourski, Y.; et al. Advanced characterization of multicaloric materials in pulsed magnetic fields. J. Appl. Phys. 2020, 127, 185107. [Google Scholar] [CrossRef]
  34. Gottschall, T.; Gràcia-Condal, A.; Fries, M.; Taubel, A.; Pfeuffer, L.; Mañosa, L.; Planes, A.; Skokov, K.P.; Gutfleisch, O. A multicaloric cooling cycle that exploits thermal hysteresis. Nat. Mater. 2018, 17, 929–934. [Google Scholar] [CrossRef] [PubMed]
  35. Amirov, A.A.; Starkov, A.S.; Starkov, I.A.; Kamantsev, A.P.; Rodionov, V.V. Electric field controlled magnetic phase transition in Fe49Rh51 based magnetoelectric composites. Lett. Mater. 2018, 8, 353–357. [Google Scholar] [CrossRef]
  36. Val’kov, V.I.; Varyukhin, D.V.; Golovchan, A.V.; Gribanov, I.F.; Sivachenko, A.P.; Kamenev, V.I.; Todris, B.M. Influence of pressure on the stability of the magnetically ordered states in alloys of the system Mn2-xFexAs0.5P0.5. Low Temp. Phys. 2008, 34, 734–745. [Google Scholar] [CrossRef]
  37. Wien, T.U.; Hauptstrage, W. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal-Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar]
  38. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  39. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  40. Sun, J.; Ruzsinszky, A.; Perdew, J. Strongly Constrained and Appropriately Normed Semilocal Density Functional. Phys. Rev. Lett. 2015, 115, 036402. [Google Scholar] [CrossRef]
  41. Togo, A.; Oba, F.; Tanaka, I. First-principles calculations of the ferroelastic transition between rutile-type and CaCl2-type SiO2 at high pressures. Phys. Rev. B—Condens. Matter Mater. Phys. 2008, 78, 134106. [Google Scholar] [CrossRef]
  42. Otero-De-La-Roza, A.; Luaña, V. Gibbs2: A new version of the quasi-harmonic model code. I. Robust treatment of the static data. Comput. Phys. Commun. 2011, 182, 1708–1720. [Google Scholar] [CrossRef]
  43. Otero-De-La-Roza, A.; Abbasi-Pérez, D.; Luaña, V. Gibbs2: A new version of the quasiharmonic model code. II. Models for solid-state thermodynamics, features and implementation. Comput. Phys. Commun. 2011, 182, 2232–2248. [Google Scholar] [CrossRef]
  44. Kamantsev, A.P.; Koledov, V.V.; Mashirov, A.V.; Dilmieva, E.T.; Shavrov, V.G.; Cwik, J.; Tereshina, I.S.; Lyange, M.V.; Khovaylo, V.V.; Porcari, G.; et al. Properties of metamagnetic alloy Fe48Rh52 in high magnetic fields. Bull. Russ. Acad. Sci. Phys. 2015, 79, 1086–1088. [Google Scholar] [CrossRef]
  45. Tishin, A.M.; Spichkin, Y.I. The Magnetocaloric Effect and Its Applications; Taylor & Francis: Abingdon, UK, 2003; ISBN 0-7503-0922-9. [Google Scholar]
  46. Wang, Y.; Wang, H.; Tan, W.; Huo, D. Magnetization reversal, critical behavior, and magnetocaloric effect in NdMnO3: The role of magnetic ordering of Nd and Mn moments. J. Appl. Phys. 2022, 132, 183907. [Google Scholar] [CrossRef]
  47. Zhao, B.; Hu, X.; Dong, F.; Wang, Y.; Wang, H.; Tan, W.; Huo, D. The Magnetic Properties and Magnetocaloric Effect of Pr0.7Sr0.3MnO3 Thin Film Grown on SrTiO3 Substrate. Materials 2023, 16, 75. [Google Scholar] [CrossRef] [PubMed]
  48. Amirov, A.A.; Cugini, F.; Kamantsev, A.P.; Gottschall, T.; Solzi, M.; Aliev, A.M.; Spichkin, Y.I.; Koledov, V.V.; Shavrov, V.G. Direct measurements of the magnetocaloric effect of Fe49Rh51 using the mirage effect. J. Appl. Phys. 2020, 127, 233905. [Google Scholar] [CrossRef]
  49. Isaacs, E.B.; Wolverton, C. Performance of the strongly constrained and appropriately normed density functional for solid-state materials. Phys. Rev. Mater. 2018, 2, 63801. [Google Scholar] [CrossRef]
  50. Ekholm, M.; Gambino, D.; Jönsson, H.J.M.; Tasnádi, F.; Alling, B.; Abrikosov, I.A. Assessing the SCAN functional for itinerant electron ferromagnets. Phys. Rev. B 2018, 98, 94413. [Google Scholar] [CrossRef]
  51. Buchelnikov, V.D.; Sokolovskiy, V.V.; Miroshkina, O.N.; Zagrebin, M.A.; Nokelainen, J.; Pulkkinen, A.; Barbiellini, B.; Lähderanta, E. Correlation effects on ground-state properties of ternary Heusler alloys: First-principles study. Phys. Rev. B 2019, 99, 14426. [Google Scholar] [CrossRef]
  52. Baigutlin, D.R.; Sokolovskiy, V.V.; Miroshkina, O.N.; Zagrebin, M.A.; Nokelainen, J.; Pulkkinen, A.; Barbiellini, B.; Pussi, K.; Lähderanta, E.; Buchelnikov, V.D.; et al. Electronic structure beyond the generalized gradient approximation for Ni2MnGa. Phys. Rev. B 2020, 102, 45127. [Google Scholar] [CrossRef]
  53. Buchelnikov, V.D.; Sokolovskiy, V.V.; Miroshkina, O.N.; Baigutlin, D.R.; Zagrebin, M.A.; Barbiellini, B.; Lähderanta, E. Prediction of a Heusler alloy with switchable metal-to-half-metal behavior. Phys. Rev. B 2021, 103, 54414. [Google Scholar] [CrossRef]
  54. Wolloch, M.; Gruner, M.E.; Keune, W.; Mohn, P.; Redinger, J.; Hofer, F.; Suess, D.; Podloucky, R.; Landers, J.; Salamon, S.; et al. Impact of lattice dynamics on the phase stability of metamagnetic FeRh: Bulk and thin films. Phys. Rev. B 2016, 94, 174435. [Google Scholar] [CrossRef]
  55. Belov, M.P.; Syzdykova, A.B.; Abrikosov, I.A. Temperature-dependent lattice dynamics of antiferromagnetic and ferromagnetic phases of FeRh. Phys. Rev. B 2020, 101, 134303. [Google Scholar] [CrossRef]
  56. Vieira, R.M.; Eriksson, O.; Bergman, A.; Herper, H.C. High-throughput compatible approach for entropy estimation in magnetocaloric materials: FeRh as a test case. J. Alloys Compd. 2021, 857, 157811. [Google Scholar] [CrossRef]
Figure 1. DSC thermograms in zero magnetic field (left axis) and temperature dependences of the magnetization in low magnetic field of 1 mT (right axis) for (a) FR72h and (b) FR48h samples of Fe49Rh51 alloy.
Figure 1. DSC thermograms in zero magnetic field (left axis) and temperature dependences of the magnetization in low magnetic field of 1 mT (right axis) for (a) FR72h and (b) FR48h samples of Fe49Rh51 alloy.
Metals 13 00956 g001
Figure 2. (a) Magnetization isotherms for FR72h sample of Fe49Rh51 alloy at different temperatures in the 300–328 K range in magnetic field of 8 T. (b) Temperature dependences of the electrical resistivity for FR72h sample of Fe49Rh51 alloy at heating and cooling at different magnetic fields in the 0–12 T range.
Figure 2. (a) Magnetization isotherms for FR72h sample of Fe49Rh51 alloy at different temperatures in the 300–328 K range in magnetic field of 8 T. (b) Temperature dependences of the electrical resistivity for FR72h sample of Fe49Rh51 alloy at heating and cooling at different magnetic fields in the 0–12 T range.
Metals 13 00956 g002
Figure 3. Temperature dependences of the magnetization for FR48h sample of Fe49Rh51 alloy measured under combined hydrostatic pressure in the 0–12 kbar range and magnetic field of (a) 0.155 T and (b) 0.97 T.
Figure 3. Temperature dependences of the magnetization for FR48h sample of Fe49Rh51 alloy measured under combined hydrostatic pressure in the 0–12 kbar range and magnetic field of (a) 0.155 T and (b) 0.97 T.
Metals 13 00956 g003
Figure 4. (a) Temperature dependences of the magnetic entropy change ΔS for FR72h sample at magnetic field changes μ0H = 1 T and 2 T, calculated from Figure 2a using Equation (1). (b) Temperature dependences of ΔS for FR48h sample at μ0H = 0.97 T under applied hydrostatic pressure 0 kbar, 6 kbar, 12 kbar at cooling protocol, calculated from Figure 3b using Equation (1).
Figure 4. (a) Temperature dependences of the magnetic entropy change ΔS for FR72h sample at magnetic field changes μ0H = 1 T and 2 T, calculated from Figure 2a using Equation (1). (b) Temperature dependences of ΔS for FR48h sample at μ0H = 0.97 T under applied hydrostatic pressure 0 kbar, 6 kbar, 12 kbar at cooling protocol, calculated from Figure 3b using Equation (1).
Metals 13 00956 g004
Figure 5. μ0H-T phase diagram for FR72h sample is obtained from Figure 2b and P–T phase diagram for FR48h sample is obtained from Figure 3b.
Figure 5. μ0H-T phase diagram for FR72h sample is obtained from Figure 2b and P–T phase diagram for FR48h sample is obtained from Figure 3b.
Metals 13 00956 g005
Figure 6. (a) Total energy difference versus the lattice parameter for the FM and AFM phases of FeRh alloy calculated within PBE and SCAN. The partial magnetic moments at corresponding optimized lattice parameters are also presented. (b) Phonon dispersion relations for AFM phase of FeRh alloy in the framework of SCAN and PBE calculations.
Figure 6. (a) Total energy difference versus the lattice parameter for the FM and AFM phases of FeRh alloy calculated within PBE and SCAN. The partial magnetic moments at corresponding optimized lattice parameters are also presented. (b) Phonon dispersion relations for AFM phase of FeRh alloy in the framework of SCAN and PBE calculations.
Metals 13 00956 g006
Figure 7. Theoretical and experimental TmP phase diagram for FeRh alloys. Theoretical Tm(P) curves are calculated assuming pure AF and FM phases (label “[email protected]%”) and the FM phase with one antisite pair embedded in the 64- and 128-atom supercell (labels “[email protected]%” and “[email protected]%”, respectively). The experimental curve Tm(P) is obtained by averaging over the FF, FS, AFS, and AFF temperatures shown in Figure 5.
Figure 7. Theoretical and experimental TmP phase diagram for FeRh alloys. Theoretical Tm(P) curves are calculated assuming pure AF and FM phases (label “[email protected]%”) and the FM phase with one antisite pair embedded in the 64- and 128-atom supercell (labels “[email protected]%” and “[email protected]%”, respectively). The experimental curve Tm(P) is obtained by averaging over the FF, FS, AFS, and AFF temperatures shown in Figure 5.
Metals 13 00956 g007
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

Kamantsev, A.P.; Amirov, A.A.; Zaporozhets, V.D.; Gribanov, I.F.; Golovchan, A.V.; Valkov, V.I.; Pavlukhina, O.O.; Sokolovskiy, V.V.; Buchelnikov, V.D.; Aliev, A.M.; et al. Effect of Magnetic Field and Hydrostatic Pressure on Metamagnetic Isostructural Phase Transition and Multicaloric Response of Fe49Rh51 Alloy. Metals 2023, 13, 956. https://doi.org/10.3390/met13050956

AMA Style

Kamantsev AP, Amirov AA, Zaporozhets VD, Gribanov IF, Golovchan AV, Valkov VI, Pavlukhina OO, Sokolovskiy VV, Buchelnikov VD, Aliev AM, et al. Effect of Magnetic Field and Hydrostatic Pressure on Metamagnetic Isostructural Phase Transition and Multicaloric Response of Fe49Rh51 Alloy. Metals. 2023; 13(5):956. https://doi.org/10.3390/met13050956

Chicago/Turabian Style

Kamantsev, Alexander P., Abdulkarim A. Amirov, Vladislav D. Zaporozhets, Igor F. Gribanov, Aleksay V. Golovchan, Victor I. Valkov, Oksana O. Pavlukhina, Vladimir V. Sokolovskiy, Vasiliy D. Buchelnikov, Akhmed M. Aliev, and et al. 2023. "Effect of Magnetic Field and Hydrostatic Pressure on Metamagnetic Isostructural Phase Transition and Multicaloric Response of Fe49Rh51 Alloy" Metals 13, no. 5: 956. https://doi.org/10.3390/met13050956

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