Abstract
For magnetic cooling with the goal of hydrogen liquefaction, magnetocaloric materials with outstanding magnetic-entropy and adiabatic temperature changes (ΔST and ΔTad) are required, covering the temperature range from liquid nitrogen to the condensation point of hydrogen at 20 K. Although second-order rare-earth-based intermetallic compounds show large ΔST and ΔTad near 20 K, their performance decreases drastically with increasing temperature. Here, compounds with first-order transition can be beneficial, if the reversibility of the magnetocaloric effect is ensured. In this work, we report that Gd5Si0.25Ge3.75 and Gd5Si0.5Ge3.5 exhibit highly reversible magnetocaloric effects near 55 and 80 K, respectively, in a magnetic field of 5 T despite significant thermal hysteresis. The high reversibility originates from the rapid shift of the transition temperature with magnetic field. Since superconducting coils are widely used to generate magnetic fields up to 5 T in existing magnetic refrigeration prototypes, this work proves that first-order magnetocaloric materials with significant thermal hysteresis can be promising candidates for hydrogen liquefaction in moderate magnetic fields.
Similar content being viewed by others
Introduction
Hydrogen has the highest gravimetric energy density among all chemical fuels (about 120 and 140 MJ kg−1 for lower and higher heating values, respectively). It is environmentally friendly and considered for transportation, use in households, industry, and as an energy storage medium1,2,3. On the other hand, the low density of hydrogen gas leads to an extremely low energy density in relation to volume, which is why high pressures are required to store significant amounts of H24. Alternatively, hydrogen can be stored and transported as a liquid at atmospheric pressure with almost twice the volumetric energy density compared to gaseous storage of 700 bar5. However, the liquefaction is a very energy-intensive process and requires about one-third of a lower heating value6,7. Due to the low efficiency of conventional expansion-based cooling methods at cryogenic temperatures below 100 K, the alternative technology of magnetic cooling has attracted significant interest after a number of discoveries that demonstrated the viability of this idea8. Magnetic refrigeration employs environment-friendly solid-state cooling based on magnetic materials. The device efficiency is directly related to the thermal-magnetic properties of the material, the so-called magnetocaloric effect, which scales with the magnetic field. Despite their cost, superconducting magnets can generate sufficiently strong fields to maximize the useful power of magnetic refrigerators. For the industrial scale, where larger magnetized volumes are required, superconducting coils can even be more cost-efficient than fields produced by permanent magnets, plus they allow for higher field strengths9,10. This is why many groups use superconducting solenoids with up to 5 T for their cryogenic prototypes11,12,13,14,15,16.
The cooling principle is based on the magnetocaloric effect, known for many years17, but the search for suitable materials with the best properties for application began to gain momentum, especially after the discovery of the giant magnetocaloric effect in Gd5Si2Ge218. Dozens of potential material families, such as Heusler alloys19,20,21, Laves phases3,22,23,24,25,26,27,28,29, rare-earth metals30,31, FeRh32,33,34, La-Fe-Si35,36,37, and others38, demonstrating application potential for magnetocaloric cooling both for cryogenics and room temperature. Still, the Gd-Si-Ge family occupies a special place in this list39,40.
The phase diagram of Gd5SixGe1−x as a function of temperature and silicon concentration x was presented in refs. 41,42,43. For Ge-rich compositions below x = 1.2, the high-temperature phase corresponds to the Sm5Ge4-type orthorhombic O(II) structure (space group Pnma). At silicon concentrations below 1, a second-order transition from a paramagnetic (PM) to an antiferromagnetic (AFM) state appears at the Néel temperature TN (from about 128 K for x = 0 to 140 K for x ≈ 1)43. A further first-order phase transition into the Gd5Si4-type orthorhombic O(I) structure (space group Pnma) with ferromagnetic (FM) order occurs below the Curie point, which depends almost linearly on x being TC ≈ 165 K for x = 1.2 and reducing to 20 K when x is close to zero. In the binary compound Gd5Ge4, the AFM-FM magnetostructural transition does not occur. Below 20 K, the material behaves as magnetic glass due to a kinetic arrest (KA)44,45,46. A transition into the FM phase can be induced at about 20 K by applying a magnetic field of about 1 T. Chattopadhyay et al. showed the formation of FM clusters at 1 T, randomly distributed in the AFM matrix, using Hall-probe microscopy, which evidences a heterogeneous magnetic state with glassy behavior47.
Although the O(I) and O(II) crystal structures are very similar, there exist decisive differences42,48,49,50,51,52. For simplifying the description, Smith et al.53 proposed considering the structure of these compounds in terms of so-called slabs, which are flat structural blocks of subnanometer thickness and infinite width in the ac plane (Fig. 1). Each slab consists of five layers of only R (Gd) atoms, only T (T = Ge, Si) atoms, or a combination C of them, lining up as a TRCRT sequence. These slabs are highly robust against deformations, whereas external stimuli promote their relative movements and allow for the only possible structural change54. The distances between the slabs play a crucial role in forming interslab T–T bonds, which influence interlayer ordering through RKKY interactions between the gadolinium magnetic moments50. In the case of the O(I) structure, the T–T bonds are formed and lead to FM order; for the O(II) structure, the bonds are broken, and the magnetic order is AFM. Controlling the interslab distances via silicon substitution allows the transition temperatures to be tuned. Temperature variation, as well as the application of magnetic field and pressure, induce the first-order magnetostructural transition, leading to giant magnetocaloric effects, giant magnetostriction55,56,57, and giant magnetoresistance58,59,60.
Schematic representation of a the O(I) Sm5Ge4-type and b the O(I) Gd5Si4-type orthorhombic structures. The purple boxes show the orthorhombic unit cells with the ab plane lying in the plane of the figure. The dashed gray boxes highlight slabs of highly stable structural blocks that consist of layers formed by rare-earth atoms (R), transition-metal elements (T), and their combinations (C). Relative movements of the slabs (gray arrows) change the interatomic distances between the atoms from the T layers and a strengthening of interslab T–T bonds (orange double lines in (b)). Adapted from Figure 6 in ref. 39.
At the same time, such a transition is accompanied by thermal hysteresis, which leads to irreversibility effects, i.e., an incomplete transition when applying and removing an external stimulus. One of the ways to solve this problem is the application of a multicaloric cycle, when the return to the initial state after application and removal of a magnetic field can be provided by uniaxial stress19,61,62. Another approach is to search for compounds with giant magnetocaloric effects and minimal irreversibility. Guided by the latter, in this paper, we present the results of our study on the adiabatic temperature change ΔTad and the isothermal entropy change ΔST as well as irreversibilities in the Gd5SixGe1−x family for compounds with x = 0, 0.25, and 0.5, in which we observe a large magnetocaloric effect between 20 and 100 K. This is particularly promising when it comes to cryogenic applications working with liquid-nitrogen pre-cooling. These three materials can be used to form a stack that allows magnetocaloric cooling down from 77 K to the boiling point of hydrogen.
Results and discussion
Figure 2a–c shows temperature-dependent magnetization data for Gd5Ge4, Gd5Si0.25Ge3.75, and Gd5Si0.5Ge3.5, respectively. All three samples reveal similar features in the M(T) data. Between 110 and 130 K, there are cusps due to the second-order PM-AFM transition, as indicated by gray dotted lines, which connect the positions of this feature in different fields. The first-order AFM-FM magnetostructural transitions are accompanied by thermal hysteresis and rapid changes in magnetization. Based on these features observed during cooling, we present the phase diagrams in the insets of Fig. 2a–c.
Temperature-dependent magnetization of a Gd5Ge4, b Gd5Si0.25Ge3.75, and c Gd5Si0.5Ge3.5. During cooling and heating, the data were measured under various applied magnetic fields up to 10 T. Arrows around the thermal hysteresis indicate the sweep directions. The gray dotted lines indicate the second-order PM-AFM transitions. The insets in (a–c) show phase diagrams based on the magnetization features extracted from cool-down sweeps. The isothermal entropy changes for d Gd5Ge4, e Gd5Si0.25Ge3.75, and f Gd5Si0.5Ge3.5 are calculated from the magnetization data using Eq. (1). Light thick lines show the envelopes of the reversible areas of ΔST peaks, formed by the overlapping area of up and down sweeps.
For Gd5Ge4, we found additional features (marked by A and KA) that are not present in Gd5Si0.25Ge3.75 and Gd5Si0.5Ge3.5. A splitting of the magnetostructural AFM-FM transition appears in magnetic fields at about 8 T (green area in the inset of Fig. 2a). This additional phase originates from antiferromagnetic short-range correlations in the main ferromagnetic matrix (phase A) as reported in ref. 63. Another feature appears in the heating curve in 1 T. With increasing temperature, the magnetization rises around 15 K, reaches a maximum at around 20 K, and then drops at 26 K. We attribute this behavior to the kinetic-arrest (KA) phase mentioned above. The KA area, reported by S. B. Roy et al. in ref. 44, is depicted in yellow in the inset of Fig. 2a. The shape of the heating curve in 1 T is similar to that reported in ref. 59.
When germanium is substituted by silicon, there are no signs for the presence of the KA and A phases, at least in the measured magnetic-field ranges. There are only the two second-order PM-AFM and first-order AFM-FM transitions (inset in Fig. 2b). The inset in Fig. 2c shows the feature positions obtained from the cooling curves for all three samples. The PM-AFM transition depends only weakly on the Si concentration, with very similar Néel temperatures (TN ≈ 127 K for Gd5Ge4, 128 K for Gd5Si0.25Ge3.75, and 130 K for Gd5Si0.5Ge3.5). The PM-AFM transition shifts to lower temperatures with increasing magnetic field. In contrast, the magnetostructural AFM-FM transition temperatures increase with increasing silicon concentration and increasing magnetic field.
The magnetization at 5 T and 10 K reaches similar values of 185 Am2 kg−1 for all three compounds. Among the samples investigated, Gd5Si0.25Ge3.75 shows the lowest saturation field of less than 1 T. At 0.25 T, the magnetization already reaches about 135 Am2 kg−1 at low temperatures, whereas for Gd5Si0.5Ge3.5 this value is about 70 Am2 kg−1. Gd5Ge4 is in the KA phase up to 1 T and temperatures below 20 K, with a magnetization of not more than 62 Am2 kg−1.
In most materials with a first-order transition, the magnetization changes sharply but continuously, which is why we can use the Maxwell relation to estimate the magnetocaloric effect64,65. From Maxwell relations, we can calculate the isothermal entropy change ΔST for a variation of the magnetic field ΔH = Hf − Hi using
Based on our magnetization measurements, we interpolated the data in 2 K steps and calculated ΔST using Eq. (1) (Fig. 2d–f). Here, we observe a characteristic behavior of first-order transitions with an almost unchanged low-temperature peak edge of ΔST. With increasing magnetic field, this peak grows and expands toward higher temperatures. All three samples show clear hysteresis, evidenced by the shift of the ΔST peaks between the heating and cooling curves. The overlapping area of these peaks for each magnetic field reflects the reversibility region (indicated by light thick lines in Fig. 2d–f). The influence of phase A on ΔST at 10 T for Gd5Ge4 leads to the appearance of a second maximum.
In view of the large magnetization change with temperature and magnetic field of Gd5Si0.25Ge3.75 (Fig. 2b), it is not surprising that this compound exhibits a significant value of ΔST ≈ −23 J kg−1 K−1 already at 1 T. Apparently, the kinetic arrest causes that ΔST for Gd5Ge4 does not even reach −2 J kg−1 K−1 at 1 T. Beyond the KA phase, above 1 T, ΔST gradually increases, reaching about −24 J kg−1 K−1 at 5 T and −30 J kg−1 K−1 at 10 T. For Gd5Si0.25Ge3.75, ΔST reaches −36 J kg−1 K−1 at 5 T and −40 J kg−1 K−1 at 10 T. For Gd5Si0.5Ge3.5, ΔST reaches −37 J kg−1 K−1 at 5 T. Increasing the silicon concentration reduces the width of the AFM phase and, therefore, the full width at half maximum (FWHM) of the ΔST peak. For example, in 5 T the FWHM is about 24, 22, and 18 K for the silicon concentrations x = 0, 0.25, and 0.5, respectively.
We further measured the magnetic-field dependences of ΔTad for selected starting temperatures T0 and maximum fields using pulsed magnetic fields (Fig. 3a–c). The arrows indicate the field-sweep directions. The first-order nature of the transition is reflected by the large hysteresis in most of the ΔTad curves. Depending on T0, the critical fields triggering the magnetocaloric effect shift (see phase diagrams in Fig. 2). For example, for Gd5Ge4 at T0 = 52 K (green curves in Fig. 3a) in fields up to 5 T, ΔTad remains almost zero. For higher field, ΔTad increases strongly until reaching a complete transition to the ferromagnetic phase, where ΔTad saturates. For further analysis, we determined the adiabatic temperature change at maximum field, \(\Delta {T}_{ad}^{max}\). For some T0, due to irreversibility, the final temperature after the field pulse is \(\Delta {T}_{ad}^{fin}\), marked by an open circle for T0 = 8 K in Fig. 3a. Figure 3d–f shows \(\Delta {T}_{ad}^{max}\) and \(\Delta {T}_{ad}^{fin}\) as a function of T0 obtained from our pulse-field data with maximum fields of 5, 10.9, and 22.6 T. For illustration, the hatched areas highlight the irreversible region below \(\Delta {T}_{ad}^{fin}\). For each of the samples, these regions are nearly identical, independent of the maximum field and appear close to the zero-field AFM-FM magnetostructural transition. For Gd5Ge4, we observe an anomalous increase of both \(\Delta {T}_{ad}^{max}\) and \(\Delta {T}_{ad}^{fin}\) below about 20 K. Since ΔST, obtained from the magnetization data, is close to zero (Fig. 2d), there should also be no ΔTad. Indeed, the large values of ΔTad are associated with dissipative losses as previously discussed in ref. 66. Whether the formation of the KA phase is the cause for this behavior is not clear and is not the subject of this paper.
Magnetic-field dependences of adiabatic temperature changes ΔTad measured during various magnetic-field pulses up to 22.6 T for a Gd5Ge4, b Gd5Si0.25Ge3.75, and c Gd5Si0.5Ge3.5. Colors denote different starting temperatures T0. Arrows indicate up and down field sweeps. d–f Extracted ΔTad values at maximum magnetic field (\(\Delta {T}_{ad}^{max}\), filled circles) and immediately after each field pulse (\(\Delta {T}_{ad}^{fin}\), open circles) as a function of initial temperature. Hatched areas below \(\Delta {T}_{ad}^{fin}\) correspond to irreversibilities.
Otherwise, ΔTad for the three compounds is in line with the above-described behavior of ΔST. At 130 K, Gd5Ge4 and Gd5Si0.25Ge3.75 show weak peaks associated with the second-order PM-AFM transition. At 22.6 T, \(\Delta {T}_{ad}^{max}\) is about 5 and 6.5 K for Gd5Ge4 and Gd5Si0.25Ge3.75, respectively. For Gd5Si0.5Ge3.5, the \(\Delta {T}_{ad}^{max}\) peak merges with the magnetostructural peak at 22.6 T. For Gd5Ge4, \(\Delta {T}_{ad}^{max}\) shows a broad peak associated with the AFM-FM transition, reaching about 10.2 K in 22.6 T. At this field, we also could observe a peak splitting due to the appearance of phase A.
For Gd5Si0.25Ge3.75 and Gd5Si0.5Ge3.5, we find much narrower, but larger \(\Delta {T}_{ad}^{max}\) peaks. At 22.6 T, the adiabatic temperature change reaches about 19 K for both compounds, and about 15 K at 10.9 T. Most interestingly, at 5 T, Gd5Si0.5Ge3.5 demonstrates \(\Delta {T}_{ad}^{max}=8.7\,{{{{\rm{K}}}}}\) and \(\Delta {T}_{ad}^{fin}=1.5\,{{{{\rm{K}}}}}\) at T0 = 82 K. At T0 = 55 K, Gd5Si0.25Ge3.75 shows a giant effect of 13 K, while \(\Delta {T}_{ad}^{fin}\) is only 0.9 K, which reflects a minor influence of the irreversibility. However, already at T0 = 50 K, the irreversibility becomes important with \(\Delta {T}_{ad}^{max}=3.9\,{{{{\rm{K}}}}}\) and \(\Delta {T}_{ad}^{fin}=1.4\,{{{{\rm{K}}}}}\). The silicon-doped samples show impressive \(\Delta {S}_{T}^{max}\) and \(\Delta {T}_{ad}^{max}\) values (Fig. 4) in 2 and 5 T, surpassing the results for compounds with similar transition temperatures, such as TbFeSi and DyFeSi (\(\Delta {S}_{T}^{max}=17.5\) and 17.4 J kg−1 K−1, \(\Delta {T}_{ad}^{max}=8.2\) and 7.1 K in 5 T, respectively)67, EuO (17.5 J kg−1 K−1, 6.8 K)68, DyAl2 (16.8 J kg−1 K−1, 6.4 K)69, HoCo2 (23.6 J kg−1 K−1, 6.9 K)23, Eu2In (34 J kg−1 K−1)70, and HoGa (17.1 J kg−1 K−1)71. At the same time, kinetic arrest and irreversibilities lead to a reduced magnetocaloric effect and Gd5Ge4 is inferior to materials with low transition temperatures, such as ErCo2 (\(\Delta {S}_{T}^{max}=35.9\,{{{{\rm{J}}}}}\,{{{{{\rm{kg}}}}}}^{-1}\,{{{{{\rm{K}}}}}}^{-1}\), \(\Delta {T}_{ad}^{max}=7.6\,{{{{\rm{K}}}}}\))23, HoB2 (40.1 J kg−1 K−1, 12 K)72, and ErAl2 (37.8 J kg−1 K−1, 9.8 K)3.
Comparison of \(\Delta {S}_{T}^{\max }\) and \(\Delta {T}_{{{{{\rm{ad}}}}}}^{\max }\) values for the Gd5SixGe1−x family from the present work (stars) with literature data from refs. 3, 23, 67,68,69,70,71,72, 76,77,78,79,80 showing a large magnetocaloric effect at cryogenic temperatures under magnetic-field application of 2 (left) and 5 T (right).
In addition, for a more objective comparison of materials, we evaluated the temperature-averaged entropy change (TEC)73 using ΔST(T) data of our samples and materials from the literature mentioned above and summarized this in Table 1. Most of the materials from that list demonstrate no significant thermal hysteresis. However, this is different for the Gd-Si-Ge system, since it reduces the efficiency of the magnetocaloric thermodynamic cycle in cooling devices. For this reason, we evaluated TEC for our samples on the basis of both sets of data, ΔST(T) during the heating protocol and for the reversible process (overlapping areas of heating and cooling curves). The results are presented in Table 1 for temperature spans of 3 and 10 K in magnetic fields of 1, 2, and 5 T. For the sake of clarity, we sorted the materials according to their transition temperatures and applied a color scheme to reflect the TEC values.
Griffith et al. showed for Gd5Si0.5Ge3.5 values for TEC(3 K) and TEC(10 K) in 1 T of 20 and 11 J kg−1 K−1, respectively, which is higher compared to our estimates from the heating curves (12.4 and 37.8 J kg−1 K−1, respectively)73. However, details on the calculation of the TEC parameter were not provided in that work. In order to avoid overestimates of the material efficiency, we focus on the TEC calculations derived from reversible ΔST curves, which is critical especially at low magnetic-field changes, whereas in 5 T the resulting TEC values differ little between reversible part of ΔST(T) and under heating protocol.
In 1 T, TEC(3 K) and TEC(10 K) for Si-doped samples are inferior to some other compounds. Thus, ErCo2 shows about one and a half times higher values of 19.2 and 9.1 9.1 J kg−1 K−1 compared to Gd5Si0.25Ge3.75 (11.7 and 5.9 J kg−1 K−1). HoCo2 has more than two times higher values than Gd5Si0.5Ge3.5. However, at the same time, these values are comparable and even superior to the TEC values of other materials shown in Table 1. Much higher TEC values are achieved in 5 T, where Gd5SixGe4−x compounds have impressive ΔST peak widths. Especially Gd5Si0.5Ge3.5 and related Gd4ScGe4 stand out in the nitrogen boiling point region. These materials have the advantage of a wide temperature window with a large magnetocaloric effect, making them suitable for use in devices using such moderately high magnetic fields.
In this work, we have studied the magnetocaloric effect in three members of the Gd5SixGe1−x family in high fields. In these compounds, a large magnetocaloric effect arises from the rapid shift of the AFM-FM transition temperatures with magnetic field. Changing the silicon concentration x allows fine-tuning of the Curie point and the choice of a preferred operating temperature range for magnetic cooling. Direct measurements of ΔTad show that Gd5Si0.25Ge3.75 exhibits almost fully reversible ΔTad curves under moderate pulsed magnetic fields of 5 T and higher. Despite the presence of strongly increasing irreversibilities at temperatures below 30 K, Gd5Ge4 still shows reasonable ΔTad and ΔST values in magnetic fields above 2 T. Considering the better cost efficiency of superconducting solenoids on a large scale, moderate fields up to 5 T are common in cryogenic magnetocaloric prototypes. In light of this, the materials discussed here, in particular the silicone-containing compounds, are very attractive for magnetic hydrogen liquefaction.
Methods
We prepared our Gd5Ge4, Gd5Si0.25Ge3.75, and Gd5Si0.5Ge3.5 samples by arc melting high-purity Gd (99.9%), Si (99.9999%), and Ge (99.9999%) from chemPUR. To ensure good homogeneity, we melted the samples five times. The ingots were flipped before each melting step. The sufficient quality of the samples was confirmed by X-ray powder diffraction (Stadi P, Stoe & Cie GmbH) and backscattered electron imaging (Tescan Vega 3 scanning electron microscope). We performed the magnetic measurements using a vibrating sample magnetometer and a Quantum Design physical property measurement system (PPMS). We measured during up and down temperature sweeps at constant magnetic field. For Gd5Ge4 and Gd5Si0.25Ge3.75, we determined the magnetization in fields up to 10 T, for Gd5Si0.25Ge3.75 up to 5 T. We did a smoothing of the M(T) curves and interpolated the data for temperature steps of 2K to reduce the noise in ∂M(T)/∂T and to make ΔST(T) smoother. We measured the adiabatic temperature changes ΔTad at the Dresden High Magnetic Field Laboratory (HLD) using pulsed magnetic fields of different amplitudes up to about 22.6 T. We glued thin chromel-constantan thermocouples between two pieces of the samples using silver paste, allowing the determination of the sample temperature during the short magnetic pulse of about 100 ms74. Due to the irreversibility in the region of the magnetostructural transitions of the studied compounds, we applied a discontinuous measurement protocol of cycling the sample temperatures before each field pulse. We heated the sample to above the transition and then cooled it in zero field to the initial temperature T0 before we applied the magnetic-field pulse. A more detailed description can be found in ref. 75.
Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
References
Edwards, P., Kuznetsov, V., David, W. & Brandon, N. Hydrogen and fuel cells: towards a sustainable energy future. Energy Policy 36, 4356–4362 (2008).
Tzimas, E., Filiou, C., Peteves, S. & Veyret, J.-B. Hydrogen Storage: State-of-the-Art and Future Perspective. EUR 20995 https://publications.jrc.ec.europa.eu/repository/handle/JRC26493 (cat. No. LD-NA-20995-EN-C, 2003).
Liu, W. et al. A study on rare-earth Laves phases for magnetocaloric liquefaction of hydrogen. Appl. Mater. Today 29, 101624 (2022).
Møller, K. T., Jensen, T. R., Akiba, E. & wen Li, H. Hydrogen—a sustainable energy carrier. Prog. Nat. Sci. Mater. Int. 27, 34–40 (2017).
Zhang, T. et al. Hydrogen liquefaction and storage: recent progress and perspectives. Renew. Sustain. Energy Rev. 176, 113204 (2023).
Ohlig, K. & Decker, L. The latest developments and outlook for hydrogen liquefaction technology. AIP Conf. Proc. 1573, 1311–1317 (2014).
Al Ghafri, S. Z. et al. Hydrogen liquefaction: a review of the fundamental physics, engineering practice and future opportunities. Energy Environ. Sci. 15, 2690–2731 (2022).
Zimm, C. et al. Description and performance of a near-room temperature magnetic refrigerator in Advances in Cryogenic Engineering, Vol. 43, 1759–1766 (Springer, 1998).
Bjørk, R., Nielsen, K. K., Bahl, C. R., Smith, A. & Wulff, A. C. Comparing superconducting and permanent magnets for magnetic refrigeration. AIP Adv. 6, 056205 (2016).
López de Toledo, C. H., López, J. M. & Rodriguez, L. G.-T. Exergoeconomic analysis of potential magnet configurations of a magnetic refrigerator operating at 4.2 K. J. Instrum. 17, C10019 (2022).
Kamiya, K. et al. Active magnetic regenerative refrigeration using superconducting solenoid for hydrogen liquefaction. Appl. Phys. Express 15, 053001 (2022).
Natsume, K. et al. Effects of moving magnetic materials in and out of superconducting magnet for active magnetic regenerative refrigeration system. IEEE Trans. Appl. Supercond. 34, 1–5 (2024).
Kim, Y., Park, I. & Jeong, S. Experimental investigation of two-stage active magnetic regenerative refrigerator operating between 77 K and 20 K. Cryogenics 57, 113–121 (2013).
Park, I. & Jeong, S. Development of the active magnetic regenerative refrigerator operating between 77 K and 20 K with the conduction cooled high temperature superconducting magnet. Cryogenics 88, 106–115 (2017).
Holladay, J. et al. Investigation of bypass fluid flow in an active magnetic regenerative liquefier. Cryogenics 93, 34–40 (2018).
Barclay, J. et al. Propane liquefaction with an active magnetic regenerative liquefier. Cryogenics 100, 69–76 (2019).
Smith, A. Who discovered the magnetocaloric effect? Eur. Phys. J. H. 38, 507–517 (2013).
Pecharsky, V. K. & Gschneidner Jr, K. A. Giant magnetocaloric effect in Gd5(Si2Ge2). Phys. Rev. Lett. 78, 4494 (1997).
Liu, J., Gottschall, T., Skokov, K. P., Moore, J. D. & Gutfleisch, O. Giant magnetocaloric effect driven by structural transitions. Nat. Mater. 11, 620–626 (2012).
Kamantsev, A. et al. Giant irreversibility of the inverse magnetocaloric effect in the Ni47Mn40Sn12.5Cu0.5 Heusler alloy. Appl. Phys. Lett. 123, 202405 (2023).
Zhang, F., Bykov, E., Gottschall, T., Van Dijk, N. & Brück, E. Strong magnetoelastic coupling in MnCoSi compounds studied in pulsed magnetic fields. Phys. Rev. B 107, 214431 (2023).
Stein, F. & Leineweber, A. Laves phases: a review of their functional and structural applications and an improved fundamental understanding of stability and properties. J. Mater. Sci. 56, 5321–5427 (2020).
Bykov, E. et al. Magnetocaloric effect in the Laves phases RCo2 (R= Er, Ho, Dy, and Tb) in high magnetic fields. J. Alloy. Compd. 977, 173289 (2024).
Matsumoto, K. et al. Magnetocaloric effect of RCo2 (R: Er, Ho, Dy) compounds for regenerative magnetic refrigeration. AIP Conf. Proc. 850, 1581–1582 (2006).
de Oliveira, N. & von Ranke, P. Magnetocaloric effect in the Laves phase pseudobinaries (Dy1−cRc)Al2 (R= Er and Ho). J. Magn. Magn. Mater. 320, 386–392 (2008).
Ćwik, J. et al. Magnetocaloric performance of the three-component Ho1−xErxNi2 (x=0.25, 0.5, 0.75) Laves phases as composite refrigerants. Sci. Rep. 12, 1–11 (2022).
Tang, X. et al. Magnetic refrigeration material operating at a full temperature range required for hydrogen liquefaction. Nat. Commun. 13, 1817 (2022).
Liu, W. et al. Designing magnetocaloric materials for hydrogen liquefaction with light rare-earth Laves phases. J. Phys. Energy 5, 034001 (2023).
Liu, W. et al. A matter of performance and criticality: a review of rare-earth-based magnetocaloric intermetallic compounds for hydrogen liquefaction. J. Alloys Compd. 995, 174612 (2024).
Gottschall, T. et al. Magnetocaloric effect of gadolinium in high magnetic fields. Phys. Rev. B 99, 134429 (2019).
Terada, N. & Mamiya, H. High-efficiency magnetic refrigeration using holmium. Nat. Commun. 12, 1–6 (2021).
Chirkova, A. et al. Giant adiabatic temperature change in FeRh alloys evidenced by direct measurements under cyclic conditions. Acta Mater. 106, 15–21 (2016).
Nishimura, K., Nakazawa, Y., Li, L. & Mori, K. Magnetocaloric effect of Fe (Rh1−xPdx) alloys. Mater. Trans. 49, 1753–1756 (2008).
Chirkova, A. et al. Magnetocaloric properties and specifics of the hysteresis at the first-order metamagnetic transition in Ni-doped FeRh. Phys. Rev. Mater. 5, 064412 (2021).
Gutfleisch, O., Yan, A. & Müller, K.-H. Large magnetocaloric effect in melt-spun LaFe13−xSix. J. Appl. Phys. 97, 10M305 (2005).
Liu, J. et al. Exploring La(Fe,Si)13-based magnetic refrigerants towards application. Scr. Mater. 67, 584–589 (2012).
Beckmann, B. et al. Multicaloric cryocooling using heavy rare-earth free La(Fe,Si)13-based compounds. ACS Appl. Mater. Interfaces 16, 38208–38220 (2024).
Franco, V. et al. Magnetocaloric effect: from materials research to refrigeration devices. Prog. Mater. Sci. 93, 112–232 (2018).
Mudryk, Y., Pecharsky, V. K. & Gschneidner, K. A. Chapter 262: R5T4 compounds: an extraordinary versatile model system for the solid state science in Including Actinides, Vol. 44 of Handbook on the Physics and Chemistry of Rare Earths (eds Bünzli, J.-C. G. & Pecharsky, V. K.) 283–449 (Elsevier, 2014).
Law, J. Y. & Franco, V. Chapter 332: Modern rare-earth-containing magnetocaloric materials: standing on the shoulders of giant Gd5Si2Ge2 in Including Actinides, Vol. 64 of Handbook on the Physics and Chemistry of Rare Earths (eds Bunzli, J.-C. G. & Kauzlarich, S. M.) 175–246 (Elsevier, 2023).
Pecharsky, V. K. & Gschneidner Jr, K. A. Tunable magnetic regenerator alloys with a giant magnetocaloric effect for magnetic refrigeration from 20 to 290 K. Appl. Phys. Lett. 70, 3299–3301 (1997).
Morellon, L., Blasco, J., Algarabel, P. & Ibarra, M. Nature of the first-order antiferromagnetic-ferromagnetic transition in the Ge-rich magnetocaloric compounds Gd5(SixGe1−x)4. Phys. Rev. B 62, 1022 (2000).
Pecharsky, A., Gschneidner Jr, K., Pecharsky, V. & Schindler, C. The room temperature metastable/stable phase relationships in the pseudo-binary Gd5Si4–Gd5Ge4 system. J. Alloy. Compd. 338, 126–135 (2002).
Roy, S. et al. Evidence of a magnetic glass state in the magnetocaloric material Gd5Ge4. Phys. Rev. B 74, 012403 (2006).
Sharath Chandra, L., Pandya, S., Vishwakarma, P., Jain, D. & Ganesan, V. Frozen magnetostructural order in Gd5Ge4: a calorimetric study. Phys. Rev. B 79, 052402 (2009).
Chattopadhyay, M. K. et al. Metastable magnetic response across the antiferromagnetic to ferromagnetic transition in Gd5Ge4. Phys. Rev. B 70, 214421 (2004).
Chattopadhyay, M. K. et al. Visual evidence of the magnetic glass state and its re-crystallization in Gd5Ge4. Europhys. Lett. 83, 57006 (2008).
Holtzberg, F., Gambino, R. & McGuire, T. New ferromagnetic 5: 4 compounds in the rare earth silicon and germanium systems. J. Phys. Chem. Solids 28, 2283–2289 (1967).
Iglesias, J. E. & Steinfink, H. The crystal structure of Gd5Si4. J. Less-Common Met. 26, 45–52 (1972).
Choe, W. et al. Making and breaking covalent bonds across the magnetic transition in the giant magnetocaloric material Gd5(Si2Ge2). Phys. Rev. Lett. 84, 4617 (2000).
Choe, W., Miller, G. J., Meyers, J., Chumbley, S. & Pecharsky, A. "Nanoscale zippers” in the crystalline solid. Structural variations in the giant magnetocaloric material Gd5Si1.5Ge2.5. Chem. Mater. 15, 1413–1419 (2003).
Choe, W., Pecharsky, A. O., Wörle, M. & Miller, G. J. "Nanoscale zippers” in Gd5(SixGe1−x)4: symmetry and chemical influences on the nanoscale zipping action. Inorg. Chem. 42, 8223–8229 (2003).
Smith, G. S., Johnson, Q. & Tharp, A. Crystal structure of Sm5Ge4. Acta Cryst. 22, 269–272 (1967).
Pecharsky, V. K. & Gschneidner Jr, K. A. Gd5(SixGe1−x) 4: an extremum material. Adv. Mater. 13, 683–686 (2001).
Casanova, F. et al. Scaling of the entropy change at the magnetoelastic transition in \({{{{{\rm{Gd}}}}}}_{5}{({{{{{\rm{Si}}}}}}_{x}{{{{{\rm{Ge}}}}}}_{1-x})}_{4}\). Phys. Rev. B 66, 212402 (2002).
Magen, C., Morellon, L., Algarabel, P., Marquina, C. & Ibarra, M. Magnetoelastic behaviour of Gd5Ge4. J. Phys. Condens. Matter 15, 2389 (2003).
Mudryk, Y., Paudyal, D., Pecharsky, V. K. & Gschneidner, K. A. Magnetostructural transition in Gd5Si0.5Ge3.5: magnetic and x-ray powder diffraction measurements, and theoretical calculations. Phys. Rev. B 77, 024408 (2008).
Levin, E. M., Pecharsky, V. K., Gschneidner, K. A. & Miller, G. J. Electrical resistivity, electronic heat capacity, and electronic structure of Gd5Ge4. Phys. Rev. B 64, 235103 (2001).
Xue, Y. et al. Dependence of magnetization on temperature, magnetoresistance effect and magnetic phase diagram for the layered Gd5Ge4 compound. J. Supercond. Nov. Magn. 22, 637–642 (2009).
Sousa, J. B. et al. Anomalous behavior of the electrical resistivity in the giant magnetocaloric compound \({{{{{\rm{Gd}}}}}}_{5}{({{{{{\rm{Si}}}}}}_{0.1}{{{{{\rm{Ge}}}}}}_{0.9})}_{4}\). Phys. Rev. B 67, 134416 (2003).
Gottschall, T. et al. A multicaloric cooling cycle that exploits thermal hysteresis. Nat. Mater. 17, 929–934 (2018).
Scheibel, F. et al. Multicaloric effect in FeRh, exploiting the thermal hysteresis in a multi-stimuli cycle combining pulsed magnetic field and uniaxial load. J. Appl. Phys. 137, 014901 (2025).
Pérez, N. et al. Griffiths-like phase and magnetic correlations at high fields in Gd5Ge4. Phys. Rev. B 83, 184411 (2011).
Gschneidner Jr, K. & Pecharsky, V. K. Magnetocaloric materials. Annu. Rev. Mater. Res. 30, 387–429 (2000).
Amaral, J. & Amaral, V. On estimating the magnetocaloric effect from magnetization measurements. J. Magn. Magn. Mater. 322, 1552–1557 (2010).
Beckmann, B. et al. Dissipation losses limiting first-order phase transition materials in cryogenic caloric cooling: a case study on all-d-metal Ni (-Co)-Mn-Ti Heusler alloys. Acta Mater. 246, 118695 (2023).
Zhang, H. et al. Large magnetocaloric effects of RFeSi (R= Tb and Dy) compounds for magnetic refrigeration in nitrogen and natural gas liquefaction. Appl. Phys. Lett. 103, 202412 (2013).
Ahn, K., Pecharsky, A., Gschneidner, K. & Pecharsky, V. Preparation, heat capacity, magnetic properties, and the magnetocaloric effect of EuO. J. Appl. Phys. 97, 063901 (2005).
Bykov, E. et al. Magnetocaloric effect in the laves-phase Ho1−xDyxAl2 family in high magnetic fields. Phys. Rev. Mater. 5, 095405 (2021).
Guillou, F. et al. Non-hysteretic first-order phase transition with large latent heat and giant low-field magnetocaloric effect. Nat. Commun. 9, 2925 (2018).
Chen, J., Shen, B., Dong, Q. & Sun, J. Giant magnetocaloric effect in HoGa compound over a large temperature span. Solid State Commun. 150, 157–159 (2010).
de Castro, P. B. et al. Machine-learning-guided discovery of the gigantic magnetocaloric effect in HoB2 near the hydrogen liquefaction temperature. NPG Asia Mater. 12, 1–7 (2020).
Griffith, L. D., Mudryk, Y., Slaughter, J. & Pecharsky, V. K. Material-based figure of merit for caloric materials. J. Appl. Phys. 123, 034902 (2018).
Gottschall, T. et al. Dynamical effects of the martensitic transition in magnetocaloric Heusler alloys from direct ΔTad measurements under different magnetic-field-sweep rates. Phys. Rev. Appl. 5, 024013 (2016).
Mejía, C. S. et al. On the high-field characterization of magnetocaloric materials using pulsed magnetic fields. J. Phys. Energy 5, 034006 (2023).
Zheng, X. Q. et al. Large magnetocaloric effect of NdGa compound due to successive magnetic transitions. AIP Adv. 8, 056425 (2018).
Biswas, A. et al. Correlating crystallography, magnetism, and electronic structure across anhysteretic first-order phase transition in Pr2In. ECS J. Solid State Sci. Technol. 11, 043005 (2022).
Mudryk, Y., Paudyal, D., Liu, J. & Pecharsky, V. K. Enhancing magnetic functionality with scandium: breaking stereotypes in the design of rare earth materials. Chem. Mater. 29, 3962–3970 (2017).
Biswas, A. et al. Unusual first-order magnetic phase transition and large magnetocaloric effect in Nd2In. Phys. Rev. Mater. 6, 114406 (2022).
Liu, W. et al. Large magnetic entropy change in Nd2In near the boiling temperature of natural gas. Appl. Phys. Lett. 119, 022408 (2021).
Acknowledgements
This work was supported by the Helmholtz Association via the Helmholtz-RSF Joint Research Group (Project No. HRSF-0045), and the Clean Hydrogen Partnership and its members within the framework of the project HyLICAL (Grant No. 101101461), the Würzburg-Dresden Cluster of Excellence on Complexity and Topology in Quantum Matter–ct.qmat (EXC 2147, Project No. 390858490), the HLD at HZDR, member of the European Magnetic Field Laboratory (EMFL), and the CRC/TRR 270 (Project-ID 405553726) “HoMMage.”
Funding
Open Access funding enabled and organized by Projekt DEAL.
Author information
Authors and Affiliations
Contributions
E. Bykov: Conceptualization, methodology, validation, formal analysis, investigation, data curation, writing—original draft, writing—review & editing, visualization. W. Liu: Conceptualization, validation, formal analysis, investigation, writing—review & editing. K. Skokov: Methodology, investigation, resources, writing—review & editing, supervision, project administration, funding acquisition. F. Scheibel: Methodology, validation, writing—review & editing. O. Gutfleisch: Resources, supervision, writing—review & editing, project administration, funding acquisition. J. Wosnitza: Resources, supervision, writing—review & editing, project administration, funding acquisition. T. Gottschall: Methodology, supervision, writing—review & editing, project administration, funding acquisition.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Communications Materials thanks the anonymous reviewers for their contribution to the peer review of this work. Primary handling editors: Aldo Isidori. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Bykov, E., Liu, W., Skokov, K. et al. Highly reversible magnetocaloric effect in Gd5Si0.25Ge3.75 and Gd5Si0.5Ge3.5 under moderate magnetic fields for hydrogen liquefaction. Commun Mater 6, 194 (2025). https://doi.org/10.1038/s43246-025-00911-2
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s43246-025-00911-2