Magnetic and magnetocaloric properties of Dy1−xErxNi2 solid solutions and their promise for hydrogen liquefaction

Jacek Ćwik *a, Yurii Koshkid’ko a, Kiran Shinde *b, Joonsik Park b, Nilson Antunes de Oliveira c, Michał Babij a and Agata Czernuszewicz d
aInstitute of Low Temperature and Structure Research, PAS, Okólna 2, Wrocław, 50-422, Poland. E-mail: j.cwik@intibs.pl
bDepartment of Materials and Science and Engineering, Hanbat National University, Daejeon 34158, Republic of Korea. E-mail: yourkirans@gmail.com
cInstituto de Física Armando Dias Tavares-Universidade do Estado do Rio de Janeiro, Rua São Francisco Xavier 524, Rio de Janeiro 20550-013, RJ, Brazil
dAmes National Laboratory, U.S. Department of Energy, Iowa State University, Ames, IA 50011, USA

Received 31st May 2024 , Accepted 23rd July 2024

First published on 15th August 2024


Abstract

Magnetic cooling is perceived as an enticing option for the liquefaction of hydrogen. Compared to the currently employed method, it offers significantly enhanced efficiency. However, further research is needed to identify refrigerants suitable for low-temperature applications. Rare-earth nickel, RNi2, intermetallic compounds based on heavy rare earth elements have garnered considerable attention owing to their unique properties linked to highly localized magnetic moments that emanate from the incompletely filled 4f-electron shell of the R atoms. In this work, the impact of simultaneous substitution within the rare earth sublattice on the magnetic and magnetocaloric properties of Dy1−xErxNi2 (x = 0.25, 0.5, 0.75) Laves phase solid solutions has been studied and the critical behavior around ferromagnetic–paramagnetic phase transition was analyzed. The samples were synthesized by arc melting and tested in a wide magnetic field range of up to 14 T. Both direct and indirect methods were used to characterize magnetocaloric properties. Experimental data were compared with theoretical results obtained in the frame of the microscopic model that considers exchange magnetic interaction and crystal electric field anisotropy. At temperatures below 20 K, all samples studied are ferromagnets. As the erbium content increases, Curie temperatures and magnetic entropy changes decrease while temperature changes remain stable. With an increasing magnetic field, the peak value of magnetic entropy change shifts to higher temperatures. At a magnetic field change of 14 T, the largest observed magnetic entropy and temperature changes are equal to 32 J kg−1 K−1 and 13 K, respectively, demonstrating that these solid solutions have high potential for low-temperature applications.


Introduction

The ability to cool on demand has had a profound impact on the modern world. Access to refrigeration and space conditioning has become an integral part of our daily lives. However, cooling has a much broader scope than what we typically think. One of the crucial areas where cooling is essential is gas liquefaction. This process involves reducing the temperature of a substance, often to cryogenic levels, i.e., below 123 K. Liquefied gas has a much higher density, making it easier to store and transport. Consequently, the ability to liquefy gases has become an essential aspect of many industrial and scientific processes.

Hydrogen, the most abundant element on earth, can be used to provide clean and renewable energy. Hydrogen typically does not exist freely in nature. It is generated from a variety of sources, such as fossil fuels, biomass, and water. Hydrogen fuel cells combine oxygen and hydrogen to produce electricity and heat, with water vapor as the sole by-product, making it an environmentally friendly and sustainable energy option. Hydrogen fuel can be used in a variety of applications in industries such as transportation, power generation, and portable devices. It can provide electricity to remote areas without power grids and can also be used in emergencies to power buildings. For practical applications, to facilitate storage and transportation, hydrogen gas needs to be transformed into its liquid state, which increases its energy density from 0.01 MJ L−1 to 8.50 MJ L−1.1 This is achieved by precooling it to 80 K with liquid nitrogen bath and subsequently compressing and expanding the precooled gas until condensation at 20 K. Hydrogen liquefaction is, however, an energy-intensive and expensive process that is typically accomplished by compressing and expanding gas at an efficiency below 25%.2

Magnetic cooling is a promising alternative to traditional liquefaction technology. It can achieve liquefaction efficiencies above 50%,2 and there are no greenhouse gas emissions related to refrigerant leaks. Magnetic cooling is based on the magnetocaloric effect (MCE), a magneto-thermodynamic phenomenon that manifests itself in the absorption or emission of heat by a solid refrigerant when exposed to a changing external magnetic field. The MCE is characterized by the magnetic entropy change, ΔSM, in an isothermal process and the temperature change, ΔTad, in an adiabatic process. Typically, the values of magnetocaloric potentials, ΔSM and ΔTad, are highest near the magnetic ordering temperature and gradually decrease to zero beyond the magnetic phase transition region.

The prospect of high efficiency and low environmental impact resulted in immense focus on the development of magnetic cooling systems.2–4 Extensive research has been conducted to assess the impact of the work cycle,5,6 component design,7,8 and operating parameters9 on system performance. A lot of emphasis has also been placed on developing new magnetocaloric materials that can operate at cryogenic temperatures. To replace hydrogen liquefiers, magnetic coolers need to be able to cool down hydrogen from 80 K to 20 K. This requires the utilization of multistage systems. At each subsequent stage, different magnetocaloric materials with lower Curie temperature and operating temperature range need to be used. The employed magnetic refrigerants should provide continuous magnetocaloric response from 80 K to below 20 K.10 Alternatively, the same results can be achieved by using multilayered regenerators composed of magnetocaloric materials operating in the hydrogen liquefier temperature range.11 Rare-earth nickel, RNi2, intermetallic compounds based on heavy rare earth elements undergoing a second-order ferromagnetic to paramagnetic phase transition are of particular interest. This is because they might show a large reversible MCE in the temperature range of hydrogen liquefaction.12 Their special properties are associated with highly localized magnetic moments originating from the incompletely filled 4f-electron shell of the R atoms. In these compounds, Ni atoms stay in the non-magnetic state.13,14 Because both nickel and rare earth elements are perceived as critical metals with a high supply risk due to their importance in technological applications, it is essential to develop recovery and separation methods for these materials, especially if magnetic cooling technology is commercialized.15

In our previous studies, we found that the crystal lattice of the Dy1−xErxNi2 solid solutions, similar to Tb1−xDyxNi2 (ref. 16) and Tb1−xErxNi2 (ref. 17) systems, can be adequately described by the C15-type (space group F[4 with combining macron]3m) Laves-phase structure with the doubled lattice parameter and that the lattice parameter decreases with increasing Er content from a = 14.295(4) Å for x = 0.25 to a = 14.264(1) Å for x = 0.75.18 Replacing Dy with Er leads to expected magnetic dilution, weakening exchange interactions and decreasing the ordering temperature. The Curie temperatures fall linearly from 16.8 K for Dy0.75Er0.25Ni2 to 10 K for Dy0.25Er0.75Ni2, making these solid solutions useful for a lower temperature range of multistage magnetic hydrogen liquefiers. Based on heat capacity data, Dy0.75Er0.25Ni2 exhibits the maximum magnetic entropy change of 12.3 J kg−1 K−1 around 17 K for a magnetic field change of 2 T. For Dy0.5Er0.5Ni2 and Dy0.25Er0.75Ni2 compositions, the maximum magnetic entropy change is slightly lower, reaching 11.6 and 11.1 J kg−1 K−1, respectively. The maximum values of ΔTad for the 2 T magnetic field change reach 4.1 K for both Dy0.75Er0.25Ni2 and Dy0.5Er0.5Ni2, and 3.9 K for Dy0.25Er0.75Ni2.18

This work continues the investigation of Dy1−xErxNi2 solid solutions and is focused on deepening understanding of their magnetic and magnetocaloric properties in high magnetic fields. Isothermal magnetic entropy changes were calculated based on magnetic measurements, and direct measurements of adiabatic temperature changes were performed in magnetic fields of up to 14 T. The obtained experimental data were compared with theoretical calculations performed employing a model Hamiltonian that includes the crystalline electric field effects and exchange magnetic interaction. Since the MCE is strongly correlated to the phase transition, it is important to explore the magnetic interactions that cause it. The critical behavior around phase transition was investigated through isothermal magnetization measurements. Observed large magnetocaloric quantities and transition temperatures below hydrogen liquefaction temperature showed that Dy1−xErxNi2 solid solutions are strong candidates for low-temperature magnetic refrigerants.

Experimental details

Polycrystalline samples of Dy1−xErxNi2 (x = 0.25, 0.5, and 0.75) solid solutions were prepared as described in ref. 18. The samples used for magnetization measurements had a mass of ∼2 mg and were nearly spherical in shape, allowing for the demagnetizing field to be accounted for. Isothermal magnetization curves, M(μ0H), were measured in magnetic fields up to 14 T at various temperatures below and above TC with a step dT of 2 K. Measurements were performed using a vibration sample magnetometer equipped with a Bitter-type magnet and a step motor.19 For direct magnetocaloric effect measurements (ΔTad), a system described in ref. 20 was used. Experiments were performed on ∼50 mg samples in magnetic fields up to 14 T generated by a Bitter-type magnet. The temperature of the sample was measured using a differential thermocouple, and a Hall sensor placed in the sample holder was used to measure the magnetic field. A thermal screen around the sample helped to minimize heat losses to the environment.

Theoretical formulation

To describe the magnetic properties and the magnetocaloric effect in the doped compound Dy1−xErxNi2, a model Hamiltonian of local interacting magnetic moments with two magnetic sublattices was employed.21 The presence of the two body interactions and the chemical disorder at the rare earth sites complicate the theoretical description of this compound. To simplify the analysis and account for the two-body interactions and doping, we used an extended mean-field approximation.18,22 The Hamiltonian suitable to describe the magnetic properties and the magnetocaloric effect in Dy1−xErxNi2 can be written as:
 
image file: d4tc02247e-t1.tif(1)
where B is the applied magnetic field, image file: d4tc02247e-t2.tif stands for the exchange interaction energy between two Dy ions, image file: d4tc02247e-t3.tif represents the exchange interaction parameters between two Er ions, and image file: d4tc02247e-t4.tif are the exchange interaction parameters between two different rare earth ions. gEr = 1.2 and gDy = 1.33 are the Lande factors as determined by Hund's rule. The term image file: d4tc02247e-t5.tif describes the crystalline electric field.23–26 In cubic symmetry, it can be written as:
 
image file: d4tc02247e-t6.tif(2)
where xc and W are model parameters, Onm are the Stevens’ operators,23 and F4 and F6 are common factors that depend on the angular momentum of rare earth ions.24 For Dy and Er, their values are F4 = 60 and F6 = 13[thin space (1/6-em)]860.

The free energy for the Hamiltonian (1) is FαM = −kBT[thin space (1/6-em)]ln[thin space (1/6-em)]Zα, where image file: d4tc02247e-t7.tif is the partition function and Eαi (α = Dy,Er) are the energy eigenvalues of the Hamiltonian. The magnetic part of the entropy, determined via the usual thermodynamic relation SαM = −(∂FαM/∂T), is:

 
image file: d4tc02247e-t8.tif(3)

In order to calculate the magnetocaloric quantities, besides the magnetic entropy, we also need to include the contributions from the conduction electrons and crystalline lattice. The entropy for the α-sublattice is Sα(T,μ0,H,x) = SαM(T,μ0,H,x) + Slat(T) + Sel(T), where Sel (T) = γT is the contribution from the conduction electrons, γ is the Sommerfeld coefficient and the Slat (T) represents the contribution from the crystalline lattice, which is given in the Debye approximation by:

 
image file: d4tc02247e-t9.tif(4)
where R is the gas constant, and Ni is the number of atoms per unit formula. The total entropy of the doped compound Dy1−xErxNi2 as a function of temperature and impurity concentration is S(T,μ0,H,x) = (1 − x)SDy(T,μ0,H,x) + xSEr(T,μ0,H,x), where SDy(T,μ0,H,x) and SEr(T,μ0,H,x) are the entropy contributions associated with Dy and Er sublattices, respectively. Once the total entropy as a function of temperature for two different values of the applied magnetic field is known, the magnetocaloric quantities, i.e., the isothermal entropy change ΔS(T,μ0ΔH,x) = S(T,μ0H2,x) − S(T,μ0H1,x) where μ0ΔH = μ0H2μ0H1 and the temperature change ΔTad(T,μ0ΔH,x) = T2T1 under the adiabatic condition S(T,μ0H2,x) = S(T,μ0H1,x) can be determined.

Results and discussion

Magnetic measurements

All investigated Dy1−xErxNi2 (x = 0.25, 0.5, and 0.75) solid solutions are characterized by a ferromagnetic order and relatively low Curie temperatures below the temperature of liquid hydrogen. Fig. 1 shows the magnetization versus temperature dependence at an applied field of μ0H = 0.5 T. The M(T) curves exhibit a single magnetic transition from a paramagnetic (PM) to a ferromagnetic (FM) state as the temperature decreases. The transition temperatures TC, determined from the critical point in the derivative dM/dT curves, are found to be 16.2, 12.7, and 10.2 K for Dy0.75Er0.25Ni2, Dy0.5Er0.5Ni2, and Dy0.25Er0.75Ni2, respectively (Fig. 1 inset). These results are comparable to those previously estimated from heat capacity and magnetic measurements.18
image file: d4tc02247e-f1.tif
Fig. 1 Magnetization of Dy1−xErxNi2 (x = 0.25, 0.5, and 0.75) solid solutions as a function of temperature at a magnetic field of 0.5 T. The inset shows the Curie temperatures determined from dM/dT vs. temperature.

The magnetization curves measured at 4.2 K are presented in Fig. 2. The magnetization shows a rapid increase in low fields (<2 T) for all samples. However, at higher fields (>6 T), it varies linearly with the field, and saturation is not achieved even at the maximum field of 14 T. The saturation magnetization, MS, values were determined from the M vs. 1/μ0H curves extrapolated to 1/μ0H = 0 (see Fig. 3 inset). The obtained MS values decrease with the increase of Er content from 8.2μB per f.u. for Dy0.75Er0.25Ni2 to 7.8μB per f.u. for Dy0.25Er0.75Ni2. The results are ∼1.5μB below the expected theoretical values calculated assuming that nickel is non-magnetic and the magnetic moments originate from the rare-earth sublattice.


image file: d4tc02247e-f2.tif
Fig. 2 Magnetization of Dy1−xErxNi2 (x = 0.25, 0.5 and 0.75) solid solutions and binary DyNi2, ErNi2 measured at 4.2 K. The inset shows the high-field region of M vs. μ0H plots.

image file: d4tc02247e-f3.tif
Fig. 3 Magnetization of the Dy0.5Er0.5Ni2 solid solution near the ordering temperature obtained with a temperature step of 2 K at various magnetic fields between 0 and 14 T. The inset shows M vs. (μ0H)−1 dependences measured at T = 4.2 K.

According to our previous studies on parent compounds,27,28 the reduced MS moment results from partial quenching of the orbital moment by the crystalline field, and this effect increases with a decrease in magnetic ordering temperatures. In the case of DyNi2 with a 4f10 electron configuration, the measured MS value is 0.2μB lower than the theoretical calculations.27 In contrast, the ErNi2 sample with a 4f12 electron configuration and the lower magnetic ordering temperature TC of 6.5 K, at 4.2 K and 14 T shows an MS of 7.7μB per f.u. (ref. 28). This value is ∼12% higher than the value reported by Farrell et al.29 and close to the one reported for the ribbon sample by Llamazares et al.30 It can be assumed that the low MS values in Dy1−xErxNi2 solid solutions are due to the Er sublattice. To confirm our assumption, neutron diffraction is necessary. The spontaneous magnetic moments, MSP, also decrease with the increasing Er content as a consequence of the lower total angular momentum of Er atoms than Dy atoms. The calculated values are 7.3, 6.3, and 6.2μB per f.u. for Dy0.75Er0.25Ni2, Dy0.5Er0.5Ni2, and Dy0.25Er0.75Ni2, respectively. The magnetic and structural parameters obtained for the investigated Dy1−xErxNi2 solid solutions, together with parent DyNi2 and ErNi2 compounds, are listed in Table 1.

Table 1 Space groups, transition temperatures (TC), theoretical and experimental saturation magnetic moments (MS), and spontaneous magnetic moments (MSP) for Dy1−xErxNi2 (x = 0.25, 0.5, and 0.75) solid solutions and parent DyNi2 and ErNi2 compounds
Compound Space group T C (K) M S (μB per f.u.) M SP (μB per f.u.) Ref.
Theor. Expt.
DyNi2 Fd[3 with combining macron]m 21.8 10.0 9.8 8.9 27
Dy0.75Er0.25Ni2 F[4 with combining macron]3m 16.2 9.75 8.2 7.3 Present
Dy0.5Er0.5Ni2 F[4 with combining macron]3m 12.7 9.5 8.0 6.3 Present
Dy0.25Er0.75Ni2 F[4 with combining macron]3m 10.2 9.25 7.8 6.2 Present
ErNi2 Fd[3 with combining macron]m 6.2 9.0 7.7 6.1 28


Magnetization isotherms were measured at selected temperatures near ordering temperatures at applied magnetic fields up to 14 T. The measurements were performed for all investigated solid solutions. Fig. 3 shows, as an example, the magnetization isotherms for Dy0.5Er0.5Ni2 from 4.2 to 30 K with a 2 K step. We noticed that at low temperatures, well below TC, the magnetization increases abruptly in the low-field range and then gradually rises linearly with increasing magnetic field. No saturation is observed even at the maximum external magnetic field of 14 T. Above TC, M(μ0H) shows a downward curvature and increases slowly, with no saturation-like behavior observed.

High-field magnetocaloric properties

To identify magnetocaloric properties, we used indirect and direct methods. The magnetic entropy change (ΔSM) was calculated from magnetization data, and the adiabatic temperature change (ΔTad) was measured directly. The experimental results were compared with theoretical calculations.

Using the magnetization isotherms obtained close to the Curie temperature, the change in magnetic entropy was calculated for each Dy1−xErxNi2 solid solution by applying the integrated Maxwell's relation:

 
image file: d4tc02247e-t10.tif(5)

In order to calculate the magnetocaloric properties using the previously presented model, a set of model parameters had to be selected. The exchange parameters image file: d4tc02247e-t11.tif were chosen to fit the experimental data for the magnetic ordering temperature. The values of these parameters were image file: d4tc02247e-t12.tif. The crystalline electric field parameters were xc = −0.1, W = 0.019 meV for the intermetallic compound DyNi2, and xc = −0.54, W = 0.05 meV for ErNi2. These values were obtained by fitting experimental data of specific heat capacity.18,31,32 The Sommerfeld coefficient was chosen as γ = 20 mJ mol−1 K−2 for all investigated concentrations. These values fall in the range commonly used in the literature.21,32

Fig. 4 compares experimental data and results of theoretical calculations of the isothermal entropy change resulting from the variation of the magnetic field, ranging from 1 to 14 T, for all solid solutions studied. The theoretical calculations are in reasonable agreement with the experimental data. For all the samples, −ΔSM reaches relatively large quantities near the temperature of magnetic phase transitions. Increasing Er content leads to a decrease in the Curie temperature, resulting in a significant shift of the peak of −ΔSM towards lower temperatures. At the magnetic field change of 14 T, the maximum experimental −ΔSM values are 33.5 J kg−1 K−1 at 20 K, 28 J kg−1 K−1 at 16.9 K, and 28.9 J kg−1 K−1 at 13.0 K for Dy0.75Er0.25Ni2, Dy0.5Er0.5Ni2, and Dy0.25Er0.75Ni2, respectively. A similar trend has been observed for a field change of 10 T, where −ΔSM for Dy0.75Er0.25Ni2 is 30.2 J kg−1 K−1 at 16.9 K, and for Dy0.5Er0.5Ni2 and Dy0.25Er0.75Ni2, −ΔSM values are close to 26 J kg−1 K−1 at 15 and 13 K, respectively. It should also be noted that the values obtained at 10 T are lower than the ones obtained for the initial DyNi2 and ErNi2 compounds, where −ΔSM reaches 34 (ref. 27) and 29.1 J kg−1 K−1,30 respectively (see Table 2). The reason behind reduced magnetocaloric properties of Dy1−xErxNi2 solid solutions is likely due to the formation of a superstructure during crystallization. This effect was also observed in other solid solutions, such as Tb1−xErxNi2 (ref. 17) or Er1−xHoxNi2.33 The characteristic quantity, −ΔSM/μ0ΔH, was observed to decrease with the increase in the external magnetic field change. For magnetic field change rising from 2 to 14 T, −ΔSM/μ0ΔH decreases from 6.3 J kg−1 K−1 to 2.4 J kg−1 K−1 for Dy0.75Er0.25Ni2, from 5.2 J kg−1 K−1 to 2.0 J kg−1 K−1 for Dy0.5Er0.5Ni2, and from 6.1 J kg−1 K−1 to 2.1 J kg−1 K−1 for Dy0.25Er0.75Ni2.


image file: d4tc02247e-f4.tif
Fig. 4 Isothermal magnetic entropy change, −ΔSM, vs. temperature of Dy0.75Er0.25Ni2 (a), Dy0.5Er0.5Ni2 (b), and Dy0.25Er0.75Ni2 (c) for the magnetic field change between 1 and 14 T. The symbols represent experimental data obtained from magnetization measurements, and the solid lines show the results of theoretical calculations.
Table 2 Experimental results of the adiabatic temperature change (ΔTad) and magnetic entropy change (−ΔSM) caused by the magnetic field change μ0ΔH for Dy1−xErxNi2 (x = 0.25, 0.5 and 0.75) solid solutions and parent DyNi2 and ErNi2 compounds. Tmax is the temperature at which ΔTad(T) reaches its maximum and image file: d4tc02247e-t22.tif is the temperature at which −ΔSM reaches its maximum
Compound μ 0ΔH (T) ΔTad −ΔSM Ref.
T max (K) ΔTad (K) ΔTad/μ0ΔH (K T−1)

image file: d4tc02247e-t23.tif

−ΔSM (J kg−1 K−1) −ΔSM/μ0ΔH (J kg−1 K−1 T−1)
DyNi2 2 21.7 3.6 1.8 19.3 11.1 5.6 27
5 20.8 7.1 1.4 19.1 23.0 4.6 27
10 21.3 10.6 1.1 19.1 34.0 3.4 27
Dy0.75Er0.25Ni2 2 16.3 3.8 1.9 16.9 12.6 6.3 Present
5 16.3 7.3 1.5 16.9 22.0 4.4 Present
10 16.3 10.8 1.1 16.9 30.2 3.0 Present
14 16.3 12.6 0.9 20.0 33.5 2.4 Present
Dy0.5Er0.5Ni2 2 12.7 3.8 1.9 12.9 10.4 5.2 Present
5 12.7 7.3 1.5 12.9 19.4 3.9 Present
10 12.7 10.8 1.1 15.0 26.0 2.6 Present
14 16.9 28.0 2.0 Present
Dy0.25Er0.75Ni2 2 9.1 3.0 1.5 9.2 12.1 6.1 Present
5 10.1 6.4 1.3 9.2 19.5 3.9 Present
10 10.2 10.5 1.1 13.0 25.5 2.6 Present
14 10.3 12.2 0.9 13.0 28.9 2.1 Present
ErNi2 2 6.5 3.4 1.7 6.2 15.1 7.6 30
5 7.8 6.2 1.2 6.2 24.3 4.9 30
10 10.0 9.8 1.0 6.2 29.1 2.9 30


For all studied samples, as the external magnetic field increases, the peak corresponding to the maximum entropy shifts toward a higher temperature range. This behavior is associated with the electronic entropy and electronic heat capacity, which exhibit strong nonlinear dependencies on both temperature and magnetic field in the low-temperature range, where the electronic heat capacity coefficient drastically changes under the influence of external magnetic field or due to the coexistence of magnetic phase transitions, as is the case with the solid solutions being investigated. Other Laves-phase solid solutions that are magnetically ordered at low temperatures also showed similar behavior.17,33

Fig. 5 shows theoretical and directly measured ΔTad results. The ΔTad values measured for field applications are plotted against the initial sample temperature for magnetic fields changing from 0 up to 14 T for Dy0.75Er0.25Ni2 and Dy0.25Er0.75Ni2, and from 0 up to 10 T for Dy0.5Er0.5Ni2. The experimentally observed ΔTad data are in good agreement with the theoretical calculations. As expected, the field application leads to an increase in the sample temperature. One can notice that the maximum adiabatic temperature changes for Dy0.75Er0.25Ni2 and Dy0.5Er0.5Ni2 are similar. For instance, with a magnetic field change of 10 T, ΔTad for both Dy0.75Er0.25Ni2 and Dy0.5Er0.5Ni2 reaches 10.8 K at temperatures of 16.7 and 13.5 K, respectively. The values for Dy0.25Er0.75Ni2 are slightly lower, with a 10.5 K temperature change at 10 K. The values obtained at 1 and 2 T are similar to the ones previously calculated from heat capacity measurements and published in ref. 18.


image file: d4tc02247e-f5.tif
Fig. 5 Adiabatic temperature change, ΔTad, vs. temperature of Dy0.75Er0.25Ni2 (a), Dy0.5Er0.5Ni2 (b), and Dy0.25Er0.75Ni2 (c) at various magnetic field changes, μ0ΔH, up to 14 T. The symbols represent experimental data obtained by direct measurements, while the solid lines correspond to data calculated using the model Hamiltonian.

Additionally, it was observed that for Dy0.25Er0.75Ni2, the temperature at which ΔTad reaches its maximum value shifts toward higher temperatures as the magnetic field increases. The behavior is similar to that seen earlier for the parent ErNi2 compound.28 It is caused by a combination of factors such as low magnetic ordering temperature, the magnitude of the magnetic field, and the shape of the heat capacity curve above the Curie temperature. In the low-temperature region, the lattice heat capacity increases in proportion to T3, while at higher temperatures, this growth slows down. Such displacements were reported previously for a high magnetic field change mentioned in ref. 34 and 35. This behavior was not observed for the other two studied compounds. The characteristic quantity, ΔTad/μ0ΔH, was observed to decrease with the increase in the external magnetic field change, similar to −ΔSM/μ0ΔH. For the magnetic field change rising from 2 to 14 T, ΔTad/μ0ΔH decreases from 1.9 K T−1 to 0.9 K T−1 for Dy0.75Er0.25Ni2, from 1.9 K T−1 to 1.1 K T−1 for Dy0.5Er0.5Ni2, and from 1.5 K T−1 to 0.9 K T−1 for Dy0.25Er0.75Ni2. All the experimental −ΔSM and ΔTad results are presented in Table 2.

Critical behavior

The critical behavior of the FM–PM, second-order magnetic transition (SOMT), can be analyzed with a set of three critical exponents, β, γ, and δ,36,37 associated with the magnetocaloric properties of materials. To gain a comprehensive understanding of the intrinsic magnetic interactions within Dy1−xErxNi2 (x = 0.25, 0.5, 0.75), different techniques were employed to study the critical behavior near the Curie temperature. According to the Arrott–Noakes equation,38 the magnetization of a ferromagnetic material undergoing the second-order phase transition is expressed as,
 
image file: d4tc02247e-t13.tif(6)
where A and B are temperature-dependent parameters, β and γ are the critical exponents, and ε is the reduced temperature calculated as,
 
image file: d4tc02247e-t14.tif(7)

The critical exponents β, γ, and δ are related to the spontaneous magnetization (MSP) below TC, the inverse magnetic susceptibility (χ0−1) above TC, and the critical magnetization isotherm M at TC, respectively:39,40

 
MSP(T) = M0(−ε)β, ε < 0, T < TC,(8)
 
image file: d4tc02247e-t15.tif(9)
 
M = RH1/δ,  at ε = 0, T = TC,(10)
where the critical amplitudes are represented as constants M0, h0/M0, and R.

Based on the Arrott–Noakes equation, if the values of the critical exponents β and γ are approached, the modified Arrott plots (MAP) of M1/βversus (μ0H/M)1/γ will form a set of parallel lines at temperatures near TC with a line at T = TC passing through the coordinate origin. However, as evidenced in Fig. 6, the Arrott plots of M2versus μ0H/M (β = 0.5 and γ = 1.0) for all investigated solid solutions do not exhibit the expected characteristics.41 Therefore, we can conclude that the mean-field model is not suitable for our materials. The reason behind such behavior is the existence of the short-range FM order in the studied compounds.


image file: d4tc02247e-f6.tif
Fig. 6 Arrott plots for Dy0.75Er0.25Ni2 (a), Dy0.5Er0.5Ni2 (b), and Dy0.25Er0.75Ni2 (c) obtained near ferromagnetic–paramagnetic phase transition temperature with a temperature step of 2 K.

Fig. 7 shows the M1/βversus (μ0H/M)1/γ plots for Dy0.75Er0.25Ni2 at different temperatures, using three different theoretical models: (a) 3D-Heisenberg model (β = 0.365, γ = 1.386), (b) tricritical mean-field model (β = 0.265, γ = 1), and (c) 3D-ising model (β = 0.325, γ = 1.24).42 Different results for each approach as shown in Fig. 7 prove that identifying the most appropriate theoretical model for the studied solutions is a challenging task.


image file: d4tc02247e-f7.tif
Fig. 7 Modified Arrott plots of M1/βvs. (μ0H/M)1/γ for M(μ0H) data of the Dy0.75Er0.25Ni2 solid solution with exponents expected for (a) 3D Heisenberg (β = 0.365 and γ = 1.386), (b) tricritical MF (β = 0.265 and γ = 1.0), and (c) 3D ising (β = 0.325 and γ = 1.241) models.

The relative slope (RS) method is frequently utilized to determine which model best describes the magnetic isotherms. This involves normalizing the slope of magnetic isotherms by the slope at TC, as indicated by the following relation,

 
image file: d4tc02247e-t16.tif(11)
where S(T) and S(TC) are the slopes of isotherms around and at TC, respectively. Fig. 8 compares the RS versus T plots for the Dy0.75Er0.25Ni2 sample obtained using four different models. The most suitable model is expected to have the RS value closest to the unity.43,44 The tricritical mean-field model meets this requirement, and therefore, we will exclusively examine the critical behavior of Dy0.75Er0.25Ni2 under this model.


image file: d4tc02247e-f8.tif
Fig. 8 Relative slope (RS) defined as RS = S(T)/S(TC) of the Dy0.75Er0.25Ni2 solid solution as a function of temperature.

The spontaneous magnetization MSP(T) and the inverse magnetic susceptibility χ0−1(T) were extracted from the modified Arrott plots for the tricritical mean-field model. The MSP(T) and χ0−1(T) values as the intercepts with the M1/β and (μ0H/M)1/γ axes, respectively, were calculated by linear extrapolation in a higher magnetic field region. The obtained values of MS(T) and χ0−1(T) are plotted against temperature (T) as shown for the Dy0.75Er0.25Ni2 sample in Fig. 9(a). By fitting eqn (8) and (9), the corresponding critical exponents are determined. The results indicate β = 0.21 with TC = 16.3 K and γ = 0.76 with TC = 16.1 K.


image file: d4tc02247e-f9.tif
Fig. 9 (a) Temperature dependence of (left) spontaneous magnetization MSP and (right) inverse initial magnetic susceptibility χ0−1. (b) The Kouvel–Fisher plots of temperature dependence of (left) spontaneous magnetization MSP and (right) inverse initial magnetic susceptibility χ0−1 at different magnetic-field ranges up to 3 T. The results provided are for Dy0.75Er0.25Ni2 with fitting curves shown in red.

The accuracy and reliability of the β, γ, and TC values can also be verified independently using the Kouvel–Fisher (K–F) method.45,46 The Kouvel–Fisher plot does not require any prior knowledge of TC. Instead, it is derived from the intercept of the fitted straight lines on the temperature axis, and the critical exponents can be directly determined from the following Kouvel–Fisher relations,

 
image file: d4tc02247e-t17.tif(12)
 
image file: d4tc02247e-t18.tif(13)

Based on these equations, a straight line with slopes 1/β and 1/γ, respectively, should be obtained by plotting MSP(T)[dMSP(T)/dT]−1 and χ−10(T)[dχ0−1(T)/dT]−1 against T. Fig. 9(b) shows the Kouvel–Fischer plot of the Dy0.75Er0.25Ni2 sample. The linear fitting provides the estimated exponents and TC as β = 0.21, TC = 16.4 K and γ = 0.93, TC = 16.5 K.

In the critical region, the critical exponents should follow the standard critical state and scaling equation, which is given as:47

 
image file: d4tc02247e-t19.tif(14)
where f+ and f are magnetic state functions for the M(μ0H) data at the temperatures above and below TC, respectively. Moreover, the re-normalized magnetization and field can be expressed as m = M|ε|β and h = μ0H|ε|−(β+γ), respectively.48,49 The m vs. h curves near the TC should yield two independent branches, one below TC and the other above TC, as depicted in Fig. 10(a). Fig. 10(b) shows the m2versus h/m curves for T < TC and T > TC. It is evident that all curves fall on two independent branches: T < TC and T > TC. The critical exponent δ can be estimated by employing the Widom scaling relation,50
 
image file: d4tc02247e-t20.tif(15)


image file: d4tc02247e-f10.tif
Fig. 10 (a) Scaling plots of renormalized magnetization m = M|ε|−βvs. renormalized field h = μ0H|ε|−(β+γ) below and above TC for Dy0.75Er0.25Ni2. The inset shows the same plots on a log–log scale. (b) m2vs. h/m below and above TC.

Utilizing the critical exponents β and γ obtained through the MAP and K–F methods in eqn (15), the values of critical exponent δ were determined as 4.61 and 5.42, respectively. Table 3 compares the critical exponents for Dy0.75Er0.25Ni2 obtained using various theoretical models. The discrepancies between experimentally determined critical exponents and theoretical values can be attributed to several factors, such as the finite size effect, sample purity and quality, and the critical region. The finite size effects in real experimental systems, which are not present in theoretical models assuming infinite size, can cause deviations in critical exponents. The sample preparation method produces impurities and defects in the materials that introduce additional interactions not considered in theoretical models. The critical region, where scaling laws and critical exponents are valid, can be very narrow. Experiments may not always probe sufficiently close to the critical point due to practical limitations, leading to observed deviations from the theoretical critical behavior. The characteristic exponents of Dy0.75Er0.25Ni2 coincide well with the tricritical mean-field model. The findings suggest the presence of short-range FM interactions in this compound.51–55 It should be noted that a recent study conducted by Fan et al.54 analyzed the critical behavior of ErNi2 compounds, and their findings indicate that the critical behaviors of these compounds also align most closely with the tricritical mean-field model with final critical exponents determined as β = 0.164(2) and γ = 0.821(9).

Table 3 Comparison of β, γ, and δ values of the Dy0.75Er0.25Ni2 solid solution obtained using available standard models
Sample/model Method T C (K) β γ δ
Dy0.75Er0.25Ni2 MAP 16.3 0.21 0.76 4.61
KF 16.4 0.21 0.93 5.42
WSR 5.7
Mean-field model Theory 0.5 1.0 3.0
3D-Heisenberg model Theory 0.365 0.386 4.8
3D-Ising model Theory 0.325 1.24 4.82
Tricritical mean-field model Theory 0.265 1.0 5.0


In the case of SOMT materials, there is an additional critical exponent, n, which can be specified by the magnetic field dependence of −ΔSmaxM, which follows a scaling law,56

 
ΔSmaxMa(μ0H)n,(16)
where n is a scaling exponent for the field dependence of the peak in the isothermal magnetic entropy change as shown in Fig. 11. The value of n was extracted by fitting with eqn (16), resulting in n = 0.62 for Dy0.75Er0.25Ni2. Alternatively, n can be calculated using the following equation because it is field- and temperature-dependent, as shown in the inset of Fig. 11:
 
image file: d4tc02247e-t21.tif(17)


image file: d4tc02247e-f11.tif
Fig. 11 Field dependence (μ0H ≤ 3 T) of the maximum magnetic entropy change around the transition temperature for Dy0.75Er0.25Ni2. The inset shows the temperature dependence on local exponent n measured at different fields.

According to mean field theory (MFT), the exponent n(T,μ0H) for the long-range ferromagnet tends to achieve the lowest value of 0.67 at Curie temperature. However, at TTC and TTC, it tends to be 1 and 2, respectively.57,58 For Dy0.75Er0.25Ni2, n initially slightly decreases as temperature increases, reaching a minimum around TC. Further temperature increase causes a rapid rise of n. The observed value of n near TC is 0.62 at μ0ΔH = 3 T, which is lower than the value for the mean-field model (0.67) and indicates the occurrence of short-range magnetic ordering. In the Fig. 11 inset, there is no overshoot of n = 2 near the transition temperature for Dy0.75Er0.25Ni2, confirming that this alloy follows SOMT behavior.

Conclusions

Solid solutions of Dy1−xErxNi2 with x = 0.25, 0.50, and 0.75 were prepared using the arc-melting method and their magnetic and magnetocaloric properties in a wide magnetic field range, up to 14 T, have been studied theoretically and experimentally. The magnetic measurements confirmed that the substitution of Dy atoms for Er atoms noticeably modifies the magnetic behavior of the Dy1−xErxNi2 solutions. The presence of Er atoms in the rare-earth sublattice results in common magnetic dilution, which weakens exchange interactions and reduces the ordering temperature. The TC decreases from 16.2 K for Dy0.75Er0.25Ni2 to 10.2 K for Dy0.25Er0.75Ni2. The saturation magnetic moments, estimated from the magnetization curves at a temperature of 4.2 K, exhibit a decreasing trend with an increase in Er content and are approximately 1.5μB lower than theoretically anticipated values. The reduced moment is attributed to partial quenching of the orbital moment by the crystalline field, and this effect increases as the magnetic ordering temperatures decrease.

The critical behavior of the Dy0.75Er0.25Ni2 solid solution under magnetic field variations of up to 3 T was studied using various techniques. The study was validated by using the scaling theory and Widom scaling relation. The results show that the critical exponents are close to the theoretical prediction of the tricritical mean-field model. These findings suggest that the Dy0.75Er0.25Ni2 solid solution exhibits short-range ferromagnetic interactions.

The experimentally obtained magnetocaloric properties, both the isothermal magnetic entropy change and the adiabatic temperature change, demonstrate good agreement with the theoretical results calculated using the model Hamiltonian. The highest entropy change was observed for Dy0.75Er0.25Ni2. At the maximum applied magnetic field change of 14 T, the maximum entropy change was of 33.5 J kg−1 K−1 at 20 K. Dy0.5Er0.5Ni2 and Dy0.25Er0.75Ni2 −ΔSM reached slightly lower values of 28 J kg−1 K−1 at 16.9 and 28.9 at 13.0 K, respectively. At the 2 T field change, the maximum −ΔSM values were between 10.4 and 12.6 J kg−1 K−1, depending on the composition. The maximum observed adiabatic temperature changes upon a 10 T magnetic field change were 10.8 K, 10.8 K, and 10.5 K for Dy0.75Er0.25Ni2, Dy0.5Er0.5Ni2, and Dy0.25Er0.75Ni2, respectively, measured at 16.3, 12.7, and 10.2 K. At a field of 2 T, the maximum ΔTad values were between 3 and 3.8 K, depending on the composition.

The observed magnetocaloric properties are high and reach their maximum within a narrow temperature range below 20 K. This leads us to believe that Dy1−xErxNi2 (x = 0.25, 0.5, 0.75) solid solutions may be a viable option for refrigerants in magnetic cryocoolers suitable for hydrogen liquefaction.

Data availability

The data are available from the corresponding author upon reasonable request.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The work was supported by the National Science Center, Poland through the OPUS Program under Grant No. 2019/33/B/ST5/01853. N. A. de Oliveira acknowledges the financial support from CNPq and FAPERJ.

References

  1. G. Valenti, Hydrogen liquefaction and liquid hydrogen storage, Compendium of Hydrogen Energy, Woodhead Publishing, 2016, pp. 27–51 Search PubMed .
  2. K. Kamiya, K. Matsumoto, T. Numazawa, S. Masuyama, H. Takeya, A. T. Saito, N. Kumazawa, K. Futatsuka, K. Matsunaga, T. Shirai, S. Takada and T. Iida, Active magnetic regenerative refrigeration using superconducting solenoid for hydrogen liquefaction, Appl. Phys. Express, 2022, 15, 53001 CrossRef CAS .
  3. A. Kitanovski, Energy Applications of Magnetocaloric Materials, Adv. Energy Mater., 2020, 10, 1903741 CrossRef CAS .
  4. C. Archipley, J. Barclay, K. Meinhardt, G. Whyatt, E. Thomsen, J. Holladay, J. Cui, I. Anderson and S. Wolf, Methane liquefaction with an active magnetic regenerative refrigerator, Cryogenics, 2022, 128, 103588 CrossRef CAS .
  5. J. A. Barclay and W. A. Steyert, Active magnetic regenerator, US4332135, 1982 Search PubMed .
  6. J. R. Gómez, R. F. Garcia, A. D. M. Catoira and M. R. Gómez, Magnetocaloric effect: a review of the thermodynamic cycles in magnetic refrigeration, Renewable Sustainable Energy Rev., 2013, 17, 74–82 CrossRef .
  7. A. Czernuszewicz, J. Kaleta, D. Kołosowski and D. Lewandowski, Experimental study of the effect of regenerator bed length on the performance of a magnetic cooling system, Int. J. Refrig., 2019, 97, 49–55 CrossRef CAS .
  8. R. Bjørk, C. R. H. Bahl, A. Smith and N. Pryds, Review and comparison of magnet designs for magnetic refrigeration, Int. J. Refrig., 2010, 33, 437–448 CrossRef .
  9. A. Czernuszewicz, L. Griffith, J. Slaughter and V. Pecharsky, Balancing performance of active magnetic regenerators: a comprehensive experimental study of aspect ratio, particle size, and operating conditions, J. Phys. Energy, 2023, 5, 024008 CrossRef .
  10. P. J. Shirron, Applications of the magnetocaloric effect in single-stage, multistage and continuous adiabatic demagnetization refrigerators, Cryogenics, 2014, 62, 130–139 CrossRef CAS .
  11. D. P. Quijano, C. I. Ferreira and E. Brück, Layering strategies for active magnetocaloric regenerators using MnFePSi for heat pump applications, Therm. Eng., 2023, 232, 120962 CrossRef .
  12. W. Liu, T. Gottschall, F. Scheibel, E. Bykov, N. Fortunato, A. Aubert, H. Zhang, K. Skokov and O. Gutfleisch, Designing magnetocaloric materials for hydrogen liquefaction with light rare-earth Laves phases, J. Phys. Energy, 2023, 5, 034001 CrossRef .
  13. H. R. Kirchmayr and E. B. Landolt-Börnstein, New Series III/19d2, ed. Wijn, H. P. J., Springer, Berlin, Germany, 1990 Search PubMed .
  14. K. H. J. Buschow, Ferromagnetic Materials, 1, ed. Wohlfarth, E. P., Amsterdam, The Netherlands, North-Holland, 1980 Search PubMed .
  15. C. Tunsu and M. Petranikova, Perspectives for the recovery of critical elements from future energy-efficient refrigeration materials, J. Cleaner Prod., 2018, 197, 232–241 CrossRef CAS .
  16. J. Ćwik, Y. Koshkid’ko, N. A. de Oliveira, A. Mikhailova and K. Nenkov, Effect of composition changes on the structural, magnetic and thermodynamic properties in Tb1−xDyxNi2 intermetallic compounds, J. Alloys Compd., 2018, 769, 588 CrossRef .
  17. J. Ćwik, Y. Koshkid’ko, K. Nenkov, E. Tereshina-Chitrova and N. Kolchugina, Correlation between the structure and thermomagnetic properties of pseudo-binary (Tb,Er)Ni2 solid solutions, J. Alloys Compd., 2021, 859, 157870 CrossRef .
  18. J. Ćwik, Y. Koshkid’ko, K. Nenkov, A. Mikhailova, M. Małecka, T. Romanova, N. Kolchugina and N. A. de Oliveira, Experimental and theoretical analysis of magnetocaloric behavior of Dy1−xErxNi2 intermetallics (x = 0.25, 0.5, 0.75) and their composites for low-temperature refrigerators performing an Ericsson cycle, Phys. Rev. B, 2021, 103, 214429 CrossRef .
  19. V. I. Nizhankovskii and L. B. Lugansky, Vibrating sample magnetometer with a step motor, Meas. Sci. Technol., 2007, 18, 1533 CrossRef CAS .
  20. Y. Koshkid'ko, J. Ćwik, T. I. Ivanova, S. A. Nikitin, M. Miller and K. Rogacki, Magnetocaloric properties of Gd in fields up to 14 T, J. Magn. Magn. Mater., 2017, 433, 234 CrossRef .
  21. N. A. de Oliveira and P. J. von Ranke, Theoretical aspects of the magnetocaloric effect, Phys. Rep., 2010, 489, 89 CrossRef CAS .
  22. N. A. de Oliveira, Magnetocaloric effect in rare earth doped compounds, J. Alloys Compd., 2008, 455, 81 CrossRef CAS .
  23. K. W. H. Stevens, Matrix elements and operator equivalents connected with the magnetic properties of rare earth ions, Proc. Phys. Soc., London, Sect. A, 1952, 65, 209 CrossRef .
  24. M. T. Hutchings, Point charge calculations of energy levels of magnetic ions in crystalline electric fields, Solid State Phys., 1964, 16, 227 CAS .
  25. H. G. Purwins and A. Lesson, Magnetic properties of (rare Earth)Al2 intermetallic compounds, Adv. Phys., 1990, 39, 309 CrossRef CAS .
  26. K. R. Lea, M. J. M. Leask and W. P. Wolf, The raising of angular momentum degeneracy of f-electron terms by cubic crystal fields, J. Phys. Chem. Solids, 1962, 26, 1381 CrossRef .
  27. J. Ćwik, Y. Koshkid’ko, N. A. de Oliveira, K. Nenkov, A. Hackemer, E. Dilmieva, N. Kolchugina, S. Nikitin and K. Rogacki, Magnetocaloric effect in Laves-phase rare-earth compounds with the second-order magnetic phase transition: Estimation of the high-field properties, Acta Mater., 2017, 133, 230–239 CrossRef .
  28. J. Ćwik, Y. Koshkid’ko, K. Nenkov, E. A. Tereshina and K. Rogacki, Structural, magnetic and magnetocaloric properties of HoNi2 and ErNi2 compounds ordered at low temperatures, J. Alloys Compd., 2018, 735, 1088–1095 CrossRef .
  29. J. Farrell and W. E. Wallace, Magnetic Properties of Intermetallic between the Lanthanides and Nickel Compounds or Cobalt, Inorg. Chem., 1966, 5, 105–109 CrossRef CAS .
  30. J. L. S. Llamazares, P. Ibarra-Gaytán, C. F. Sánchez-Valdés, D. Ríos-Jara and P. Álvarez-Alonso, Magnetocaloric effect in ErNi2 melt-spun ribbons, J. Rare Earths, 2020, 38, 612–616 CrossRef .
  31. E. A. Goremychkin, I. Natkaniec, E. M. Muhle and D. Chistyakov, Crystal field splitting observed in RNi2 compounds (R = Pr, Nd, Tb, Er, Ho, Tm) by means of inelastic neutron scattering, J. Magn. Magn. Mater., 1989, 81, 63–71 CrossRef CAS .
  32. P. J. von Ranke, V. K. Pecharsky and K. A. Gschneidner Jr., Influence of the crystalline electrical field on the magnetocaloric effect of DyAl2, ErAl2, and DyNi2, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 58, 12110 CrossRef CAS .
  33. J. Ćwik, Y. Koshkid’ko, B. Weise and A. Czernuszewicz, High-field magnetic and magnetocaloric properties of pseudo-binary Er1−xHoxNi2 (x = 0.25–0.75) solid solutions, J. Alloys Compd., 2023, 968, 172297 CrossRef .
  34. V. K. Pecharsky, K. A. Gschneidner Jr., A. O. Pecharsky and A. M. Tishin, Thermodynamics of the magnetocaloric effect, Phys. Rev. B, 2001, 64, 144406 CrossRef .
  35. T. Kihara, Y. Kohama, Y. Hashimoto, S. Katsumoto and M. Tokunaga, Adiabatic measurements of magneto-caloric effects in pulsed high magnetic fields up to 55 T, Rev. Sci. Instrum., 2013, 84, 074901 CrossRef CAS PubMed .
  36. X. Yang, J. Pan, S. Liu, X. He and M. Yang, Three-dimensional critical behavior and magnetocaloric effect in Yb14MnSb11, J. Magn. Magn. Mater., 2023, 582, 170982 CrossRef CAS .
  37. R. Skini, H. Baaziz, A. Tozri, M. Abdel-Hafiez and A. Hassan, Magnetocaloric effect and critical behavior in La0.8K0.2MnO3 nanoparticle, Results Phys., 2021, 30, 104861 CrossRef .
  38. A. Arrott and J. E. Noakes, Approximate equation of state for nickel near its critical temperature, Phys. Rev. Lett., 1967, 19, 786–789 CrossRef CAS .
  39. H. E. Stanley, Introduction to Phase Transitions and Critical Phenomena, Oxford University Press, London, 1971 Search PubMed .
  40. S. Pinninti, P. Sarita and G. J. Naga Raju, Magnetocaloric effect near room temperature and critical behaviour of Fe doped MnCo0.7Fe0.3Ge, Solid State Commun., 2021, 327, 114211 CrossRef CAS .
  41. X. Yang, J. Pan, S. Liu, M. Yang, L. Cao, D. Chu and K. Sun, Critical behavior and anisotropic magnetocaloric effect of the quasi-one-dimensional hexagonal ferromagnet PrCrGe3, Phys. Rev. B, 2021, 103, 104405 CrossRef CAS .
  42. T. V. Manh, K. P. Shinde, D. Nanto, H. Lin, Y. Pham, D. S. Razaq, D. R. Munazat, B. Kurniawan, S. C. Yu, K. C. Chung and D. H. Kim, Critical behavior and magnetocaloric effect in La0.7Ba0.25Nd0.05Mn1−xCuxO3, AIP Adv., 2019, 9, 035345 CrossRef .
  43. M. S. Anwar, I. Hussain, S. N. Khan, P. Fournier and B. Heun Koo, Magnetic critical behavior and room temperature magnetocaloric effect in Ba1.9Pr0.1FeMoO6 double perovskite compound, Mater. Res. Bull., 2023, 161, 112151 CrossRef CAS .
  44. Y. Ounza, S. El Ouahbi, R. Moubah, M. Lassri, M. Sajieddine, M. Oubla, I. Abdelrhafor, E. K. Hlil and H. Lassri, Critical behavior of LaBiCaMn2O7 and its relation with magnetocaloric effect, Appl. Phys. A: Mater. Sci. Process., 2023, 129, 418 CrossRef CAS .
  45. J. S. Kouvel and M. E. Fisher, Detailed magnetic behavior of nickel near its Curie point, Phys. Rev., 1964, 136, A1626 CrossRef .
  46. J. J. Mboukam, M. B. Tchoula Tchokonté, A. K. H. Bashir, B. M. Sondezi, B. N. Sahu, A. M. Strydom and D. Kaczorowski, Critical behavior in Nd2Pt2In studied using the magnetocaloric effect: Comparison with the conventional method, Mater. Res. Bull., 2020, 122, 110604 CrossRef CAS .
  47. H. Liu, J. Fan, D. Hu, J. L. Sanchez Llamazares, P. Gorria, C. Ma, Y. Zhu and H. Yang, Critical behavior of the cubic ErNi2 Laves compound nearby the Ferro-paramagnetic phase transition, Mater. Res. Bull., 2023, 164, 112239 CrossRef CAS .
  48. K.-Y. Hou, Q.-Y. Dong, L. Su, X.-Q. Zhang and Z.-H. Cheng, Three-dimensional Heisenberg critical behavior in amorphous Gd65Fe20Al15 and Gd71Fe3Al26 alloys, J. Alloys Compd., 2019, 788, 155–162 CrossRef CAS .
  49. S. N. Kaul, Static critical phenomena in ferromagnets with quenched disorder, J. Magn. Magn. Mater., 1985, 53, 5 CrossRef CAS .
  50. H. E. Stanley, Scaling, universality, and renormalization: three pillars of modern critical phenomena, Rev. Mod. Phys., 1999, 71, S358 CrossRef CAS .
  51. L. Zhang, B. S. Wang, Y. P. Sun, P. Tong, J. Y. Fan and C. J. Zhang, L. Pi, Y.H. Zhang, Critical behavior in the anti-perovskite ferromagnet AlCMn3, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 104419 CrossRef .
  52. L. Zhang, D. Menze, C. M. Jin, H. F. Du, M. Ge, C. J. Zhang, L. Pi, M. L. Tian and Y. H. Zhang, Critical behavior of the single-crystal helimagnet MnSi, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 91, 024403 CrossRef .
  53. M. E. Fisher, The theory of equilibrium critical phenomena, Rep. Prog. Phys., 1967, 30, 615 CrossRef CAS .
  54. H. Liu, J. Fan, D. Hu, J. L. Sánchez Llamazares, P. Gorria, C. Ma, Y. Zhu and H. Yang, Critical behavior of the cubic ErNi2 Laves compound nearby the Ferro-paramagnetic phase transition, Mater. Res. Bull., 2023, 164, 112239 CrossRef CAS .
  55. J. Yang and Y. P. Lee, Critical behavior in Ti-doped manganites LaMn1−xTixO3 (0.05 ≤ x ≤ 0.2), Appl. Phys. Lett., 2007, 91, 142512 CrossRef .
  56. T. L. Phan, T. V. Manh, H. R. Park, B. W. Lee, S. C. Yu, H. Yang, C. Li, H.-G. Piao, Y. D. Zhang, D. H. Manh and N. T. Dang, Magnetocaloric effect in Ba-doped LaCoO3 cobaltites showing second-order phase transitions, J. Magn. Magn. Mater., 2021, 539, 168378 CrossRef CAS .
  57. V. Franco, J. S. Blázquez, J. J. Ipus, J. Y. Law, L. M. Moreno-Ramírez and A. Conde, Magnetocaloric effect: from materials research to refrigeration devices, Prog. Mater. Sci., 2018, 93, 112–232 CrossRef .
  58. J. Y. Law, V. Franco and L. M. Moreno-Ramirez, et al., A quantitative criterion for determining the order of magnetic phase transitions using the magnetocaloric effect, Nat. Commun., 2018, 9, 2680 CrossRef PubMed .

This journal is © The Royal Society of Chemistry 2024