Next Article in Journal
Discrete One-Stage Mechanochemical Synthesis of Titanium-Nitride in a High-Energy Mill
Next Article in Special Issue
Enhanced Plasticity and Corrosion Resistance in Mg-Zn-Ca-Cu Amorphous Alloy Composite via Plasma Electrolytic Oxidation Treatment
Previous Article in Journal
Effects of Water Cooling on the Microstructure of Electron Beam Additive-Manufactured Ti-6Al-4V
Previous Article in Special Issue
Formability and Magnetic Properties of the Binary Nd-Co Amorphous Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Observation of a Broadened Magnetocaloric Effect in Partially Crystallized Gd60Co40 Amorphous Alloy

1
China Energy Guoyuan Electric Power CO., LTD., Beijing 100033, China
2
School of Materials Science and Engineering, Anhui University of Technology, Ma’anshan 243002, China
3
Key Laboratory of Green Fabrication and Surface Technology of Advanced Metal Materials, Anhui University of Technology, Ministry of Education, Ma’anshan 243002, China
*
Author to whom correspondence should be addressed.
Metals 2021, 11(11), 1741; https://doi.org/10.3390/met11111741
Submission received: 26 August 2021 / Revised: 18 October 2021 / Accepted: 26 October 2021 / Published: 30 October 2021
(This article belongs to the Special Issue Forming Ability and Properties of Bulk Metallic Glasses)

Abstract

:
To investigate the effect of crystallization treatment on the structure and magnetocaloric effect of Gd60Co40 amorphous alloy, the melt-spun ribbons were annealed at 513 K isothermally for 20, 40 and 60 min. The results indicate that, with increasing annealing time, the Gd4Co3 (space group P63/m) and Gd12Co7 (space group P21/c) phases precipitated from the amorphous precursor in sequence. In particular, in the samples annealed for 40 and 60 min, three successive magnetic transitions corresponding to the phases of Gd4Co3, Gd12Co7 and remaining amorphous matrix were detected, which induced an overlapped broadened profile of magnetic entropy change (|ΔSM|) versus temperature. Under magnetic field changing from 0 to 5 T, |ΔSM| values of 6.65 ± 0.1 kg−1·K−1 and 6.44 ± 0.1 J kg−1·K−1 in the temperature spans of 180–196 K and 177–196 K were obtained in ribbons annealed for 40 and 60 min, respectively. Compared with the fully amorphous alloy, the enhanced relative cooling power and flattened magnetocaloric effect of partially crystallized composites making them more suitable for the Ericsson thermodynamic cycle.

1. Introduction

As a potential alternate of conventional vapor compression refrigeration, magnetic refrigeration on the basis of magnetocaloric effect (MCE), has attracted a great deal of attention with the advantages of compactness, higher energy efficiency, environmental friendliness and less noise [1,2]. At constant pressure, the total entropy S of a magnetic material is composed of magnetic entropy SM, lattice entropy SL and electronic entropy SE, among which the SM depends on both temperature and applied magnetic field strongly, while the SL and SE usually can be considered temperature dependent only [3].
When a ferromagnetic substance is magnetized isothermally, the alignment of magnetic moment causes enhanced magnetic order and lower SM, and then the system releases heat to the surrounding environment since both SL and SE remain constant. If the magnetizing process is adiabatic, to maintain the total entropy unchanged, the SL and SE increase and the temperature becomes higher [4]. The above mentioned is the principle of MCE, and it is reversible for the demagnetization process. Furthermore, the isothermal magnetic entropy change ΔSM and adiabatic temperature change ΔTad are important parameters to characterize the MCE of magnetic refrigerants [4].
Compared with the reference Carnot cycle, the Ericsson cycle consists of two iso-thermal and two iso-field processes and was proposed to be utilized for temperature ranges higher than 20 K. The influence of SL and SE can be neglected in the two iso-field processes by adding a regenerator to the magnetic refrigeration system [3]. As an ideal material for the Ericsson cycle, its magnetic entropy change |ΔSM| should be a constant value in the refrigeration temperature range, named “table-like” MCE, which is usually achieved by designing multi(bi)-phase magnetocaloric material or monolithic material with multiple successive magnetic phase transitions [1,5].
The typical material for the latter group is heat-treated Gd0.54Er0.46NiAl compounds with hexagonal ZrNiAl-type crystal structure, which exhibit a nearly temperature independent |ΔSM| of 13 J·kg−1·K1 over the temperature region of 15–45 K for a field change of 0–5.32 T [6]. While, for the former group, there are two common strategies to combine multiple phases undergoing neighboring phase transitions, one is manufacturing the composites by artificially mixing, bonding, pressing or sintering [7,8,9,10], and the other is preparing the multi-phase materials by traditional producing techniques, like casting and heat-treatment [11,12,13,14,15].
Multilayer hybrid fabricated by gluing 70 wt% Gd50Co48Fe2 and 30 wt% Gd50Co48Mn2 amorphous ribbons, displays |ΔSM| of ~4.32 J·kg−1·K−1 in the temperature interval of 255 K–275 K under 5 T field [7]. The two |ΔSM| peaks of amorphous Gd65Mn25Si10 ribbons and crystalline Gd are partly overlapping for the Gd65Mn25Si10-Gd/Sn composites synthesized by hop-pressing with the nearly constant value 2.9 J·kg−1·K−1 in a wide temperature span of 220 K–293 K under field change of 0–5 T [8]. Three kinds of REAl2 (RE = Er, Ho, Dy) compounds with appropriate molar ratio were pressed and sintered as a layer structural complex, which presents almost constant of |ΔSM| (~3 J·mol−1·K−1) induced by 5 T magnetic field in the temperature range from ~15 to ~35 K [10].
However, the potential problems of pressing and sintering, including the low density and the resulting intermediate layer formed during solid-state reactions between constituent materials, as well as the differences in the coefficient of thermal expansion, may reduce the efficiency of the thermodynamic cycle and limit the working life of the refrigerant [16]. Compared with the above-mentioned composites, biphasic Gd + GdZn as-melted composite possesses improved thermal contact between phases and shows table-like |ΔSM|-T curves by tuning Gd:Zn ratio (|ΔSM| of ~3.2 J·kg−1·K−1 in temperature range of 266–290 K under 2 T field for the material with nominal composition of Gd75Zn25) [12].
Gd-TM (TM = Co, Fe, Ni, and Mn) amorphous alloys with near-room temperature MCE have attracted more attention in recent years. Generally, the broadened |ΔSM|-T profile can be observed in these materials, owing to the highly disordered structure of amorphous systems, which smears out the magnetic transition [1,11,15,17,18,19,20,21,22,23,24,25,26]. In-situ crystallization treatment of the amorphous ribbons is also utilized to obtain multi-phase magnetic refrigerant, such as crystallized Gd55Co35Ni10 ribbon containing Gd4(Co, Ni)3 and Gd12(Co, Ni)7 phases, and partially crystallized Gd55Co35Mn10 with precipitation of the Gd3Co-type and Gd12Co7-type phases in the amorphous matrix, both of them possess broadened table-like MCE [14,15].
We found that the maximum magnetic entropy change of Gd60Co40 amorphous alloy (8.3 J·kg−1·K−1 [17]) was higher than that of Gd55Co35Ni10 (6.5 J·kg−1·K−1 [15]) and Gd55Co35Mn10 (6.47 J·kg−1·K−1 [14]) amorphous counterparts with analogous Curie temperature (TC). Additionally, according to the Gd-Co binary phase diagram [27], the constituent phases of Gd60Co40 alloy after equilibrium solidification are Gd4Co3 and Gd12Co7 types. The TC of crystalline Gd4Co3, Gd12Co7 and amorphous Gd60Co40 are 220 K [21], 179 K [22] and 193 K [17], respectively, and the intervals between different TC are suitable for achieving the table-like MCE [28,29]. To reveal whether the partially crystallized Gd60Co40 alloy ribbon will exhibit enhanced table-like or broadened MCE, in this study, the influence of crystallization treatment on the structure, magnetic and magnetocaloric properties of Gd60Co40 melt-spun ribbons was investigated.

2. Materials and Methods

Amorphous ribbons with a nominal composition of Gd60Co40 were prepared in two steps. First, the master alloy was fabricated by arc melting the mixtures of high purity metals Gd (99.9 wt%), Co (99.9 wt%), Fe (99.9 wt%) and prealloy B-Fe with a mass ratio of 17.62/81.46 under a Ti gettered argon atmosphere. The alloy ingots were turned over and remelted four times to ensure the homogeneity. Secondly, the ingots were broken into small pieces of 3–4 g, and then the ribbon samples were manufactured by a single-roller melt spinning technique with a surface linear speed of 50 m/s under high-purity argon atmosphere.
The width and thickness of the ribbons were 2–3 mm and 20–50 μm, respectively. In this work, the partially crystallized samples were obtained by annealing the as-spun ribbons at 513 K for 20, 40 and 60 min, and the fully crystallized counterpart was produced through thermal treating at 653 K (the termination temperature of crystallizing exothermic peak on the heat flow curve) for 20 min.
X-ray diffraction (XRD, Bruker D8 Advance, Karlsruhe, Germany) measurements were performed at room temperature using Cu Kα radiation (λ = 0.154178 nm) and a 2θ range of 20–90° with operation conditions of 40 keV and 150 mA. The thermal properties of the ribbons were characterized using differential scanning calorimetry (DSC, Netzsch STA 449F3, Selb, Germany) under the protection of an argon atmosphere with a heating rate of 0.33 K/s. The magnetic properties of the ribbons were detected by a physical property measurement system (PPMS, Quantum Design PPMS-9 T system, San Diego, CA, USA).
In this study, the temperature dependence of the zero-field cooling magnetization (M-T) curve was collected under an applied field of 0.02 T during the heating process from 50 to 350 K. The isothermal initial magnetization (M-H) curves were collected under the applied magnetic field change from 0 to 5 T at selected temperatures from 108 to 248 K. The temperature interval of 4 and 10 K were chosen for the region in vicinity of and far away from TC, respectively, and the scanning speed of the magnetic field was slow enough to ensure accurate recording of the data in the isothermal process. Then, the magnetic entropy change |ΔSM| can be calculated by the M-H curves based on the Maxwell relation [2]:
S M ( T , H ) = S M ( T , H ) S M ( T , 0 ) = 0 H { M ( T , H ) T } H d H
where SM, T, H and M indicate the magnetic entropy, temperature, applied magnetic field and magnetization of the material, respectively. To derive the temperature dependence of |ΔSM|, the numerical approximation of the integral is usually utilized as follows [30,31]:
S M ( T i + 1 + T i 2 ) = i M i + 1 M i T i + 1 T i H i
where Mi and Mi+1 are experimental values of magnetization at temperatures Ti and Ti+1 under external magnetic field Hi.

3. Results and Discussion

3.1. Sturctural and Thermal Characterization

From the XRD pattern of the as-spun Gd60Co40 alloy shown in Figure 1a, there is only one typical diffuse broaden peak without any detectable crystalline peaks, indicating its amorphous structure. This feature was verified by the two exothermal crystallization peaks on DSC curve of the Gd60Co40 as-spun ribbon exhibited in Figure 1b. Additionally, the onset crystallization temperature Tx and the end temperature of crystallization peak with values of 523 and 653 K can be obtained, respectively. Based on these, the annealing temperatures of 513 K (10 K lower than Tx) and 653 K were chosen. As displayed in Figure 1b, the absence of an exothermal peak on the DSC curve of the sample annealed at 653 K for 20 min demonstrates that it is fully devitrified. Then, according to the XRD result (in Figure 1a), two types of crystalline phases Gd4Co3 (space group P63/m) and Gd12Co7 (space group P21/c) were identified.
From Figure 1a, the sharp peaks in XRD patterns of the samples annealed at 513 K for 20, 40 and 60 min indicate the existence of crystals. Combined with the exothermic crystallization process on DSC curves illustrated in Figure 1b, we deduced the microstructure of the alloys as crystals embedded in an amorphous matrix. In the 513 K/20 min annealed ribbon, a broad hump overlapped by one obvious crystalline peak was detected, demonstrating that the precipitation amount is slight, and the crystallites can be identified as Gd4Co3 phase in comparison with the fully crystallized sample. In the 513 K/40 min and 513 K/60 min counterparts, complicated diffraction peaks reveal that the Gd4Co3 and Gd12Co7 phases in-situ crystallized from the amorphous precursor.
Although the annealing temperature is lower than the Tx, the precipitation of Gd4Co3 and Gd12Co7 crystalline phases likely corresponds to the first and second crystallization peaks, respectively, due to heat fluctuation or thermal inertia during crystallization treatment [11]. The DSC curves of the annealed ribbons confirm this assumption, since the primary crystallization peak becomes broader (513 K for 20 min) and then disappears (513 K/40 min and 513 K/60 min) [32,33]. Moreover, with increasing annealing time, the area under the peak gradually decreases, indicating a reduction of the relative content of the amorphous phase in the whole composite material [33].
Utilizing the Scherrer formula [34], grain sizes of ~23 ± 3 nm were estimated for the crystallites in these amorphous-nanocrystalline composites, which were almost unaffected by the annealing time. Furthermore, the similar XRD patterns and DSC curves of the 513 K/40 min and 513 K/60 min annealed ribbons imply their analogous microstructure. In another word, the grain size and transformation volume tend to saturate after certain annealing time, which is attributed to the metastable equilibrium between the remaining amorphous matrix and the crystallites [35].

3.2. Multi-Magnetic Phase Transition

Figure 2a represents the temperature dependence of magnetization (M-T curves) of the annealed Gd60Co40 alloy ribbons under an applied field of 0.02 T. Subsequently, the Curie temperature TC was defined as the temperature corresponding to the minimum of the derivative dM/dT-T curves shown in the Figure 2b. It is evident that all the composite materials experienced two or more magnetic phase transitions during heating.
From the magnified image of the 210–230 K part on dM/dT-T curves (displayed in the inset of Figure 2b), there is a weak peak for 513 K/20 min annealed ribbon, which reveals first magnetic phase transition at 219 K correlated to the magnetic transition of the Gd4Co3 phase (TC = 220 K) [21]. Its second magnetic transition obtained at 196 K is sharp and associated with the amorphous ferrimagnetic phase (TC = 193 K) [17], which is in good agreement with the microstructure consisted by slight amount of crystallites and predominant amorphous matrix, as discussed in Section 3.1.
Owing to the analogous microstructure of the 513 K/40 min and 513 K/60 min annealed samples, their multi-magnetic transition behaviors are similar to each other and three transitions can be observed at temperatures 174 K/194 K/219 K and 176 K/195 K/219 K. Compared with the magnetic transition behavior in the 513 K/20 min annealed sample, an extra transition related to the Gd12Co7 crystalline (TC = 160.8 K) appeared [36]. The deviation of the TC is possibly ascribed to the different microstructure (e.g., crystal size and surrounding phase structure) between the bulk crystalline material and the in-situ precipitated crystallites, in addition, similar phenomena and values of TC were reported in Gd12Co7 melt-spun ribbon (TC = 179 K) [22].
For the fully devitrified ribbon annealed at 653 K for 20 min, the dM/dT-T curve manifested three magnetic phase transitions at 162 K and 218 K associated with Gd12Co7 and Gd4Co3, respectively, which is consistent with the XRD results.

3.3. Magnetocaloric Properties

The isothermal initial magnetization M-H curves under the magnetic field change ΔH = 5 T of partially crystallized Gd60Co40 ribbons with different annealing time are displayed in Figure 3. With raising temperature, the magnetization of all the samples exhibits apparent transition from easy-saturated to linear-field-dependent. The type of the magnetic phase transition was estimated through Arrott plots (M2 vs. H/M) according to Banerjee criteria [37], which is based on the mean-field theory and derived from the M-H isotherms [38]. As shown in Figure 4, the positive slope of all the curves indicates every magnetic transition in the multi-phase alloys is second order magnetic phase transition (SOMT). In comparison with the materials of first order magnetic phase transition (FOMT), such as Gd5(Si2Ge2) and LaFe13−xSix compounds [30,39], the magnetic refrigerants with SOMT possess advantages of negligible thermal and magnetic hysteresis, which make them more suitable for practical application, although their magnetic entropy change is lower [40].
The correlation between magnetic entropy change |ΔSM| and temperature was determined by using Equation (2) to calculate the data of M-H isotherms, and the |ΔSM| vs. T curves of the annealed Gd60Co40 ribbons under the field change from 0 to 5 T are illustrated in Figure 5. The achieved values of maximum magnetic entropy change (|ΔSMpk|) were 7.73, 6.75 and 6.54 J·kg−1·K−1 at temperatures Tpk of 194, 190 and 190 K for samples annealed at 513 K for 20, 40 and 60 min, respectively, which are smaller than those of as-spun Gd60Co40 amorphous alloy (8.3 J·kg−1·K−1) [17].
The Tpk is near to the TC of the amorphous matrix in each partially crystallized alloy, meaning that the amorphous phase makes the predominant contribution to magnetocaloric effect. With increase of the annealing time, the Gd4Co3 and Gd12Co7 phases successively precipitated from the amorphous matrix and resulted in a reduction of the amorphous phase content; therefore, the |ΔSMpk| of the multi-phase alloys decreases.
However, due to the synergistic effects of multi-magnetic phase transition, the |ΔSM|-T curves of the annealed Gd60Co40 alloys in this study were broadened, usually accompanied by wide full temperature width at half maximum (ΔTFWHM) and large relative cooling power (RCP, another parameter to evaluate the MCE as heat transferred between the hot and cold reservoirs in an ideal refrigeration cycle) with expression of RCP = |ΔSMpk| × ΔTFWHM [1]. The RCP of the 513 K/20 min, 513 K/40 min and 513 K/60 min annealed samples were 726.6, 789.8 and 797.9 J·kg−1, respectively. In comparison with that of as-spun Gd60Co40 amorphous alloy (713.8 J·kg−1) [17], the results reveal that the RCP increases with elongation of the annealing time.
As discussed in [23,36], the ΔTFWHM values of the materials in this study are much larger than the temperature span of any real magnetocaloric refrigerator so that the RCP may overestimate their performance in practical applications. In comparison with RCP, the temperature averaged entropy change (TEC) can properly reflect the merit of materials with a broad magnetocaloric response but small magnetic entropy change [41]. This is calculated over a range of temperatures ΔTlift that a material can reasonably support in response to a given field change ΔH, as follows:
T E C ( T l i f t ) = 1 T l i f t max T m i d { T m i d + T l i f t 2 T m i d T l i f t 2 S ( T ) H , T d T }
The value of the temperature at the center of the average, Tmid, is chosen by sweeping over the available ΔS(T)ΔH,T data and selecting the value that maximizes TECTlift) for the given ΔTlift, similar to the evaluation of the maximum energy product of a permanent magnet. In this study, the ΔTlift of 10 K and ΔH of 1 T were chosen, and the TEC(10 K) of the 513 K/20 min, 513 K/40 min and 513 K/60 min annealed samples were 2.34, 1.69 and 1.71 J·kg−1·K−1, respectively, at Tmid of 192, 178 and 188 K. The values are lower than that of Gd and higher than that of La0.813K0.16Mn0.987O3, indicating their magnetocaloric performance is not very good.
Nevertheless, the broadened and flatten MCE with |ΔSM| values of 6.65 ± 0.1 J·kg−1·K−1 and 6.44 ± 0.1 J·kg−1·K−1 within the temperature regions of 180–196 K and 177–196 K observed in samples annealed at 513 K for 40 min and 60 min, enable them to be more suitable for the Ericsson thermodynamic cycle [1]. In comparison with the table-like MCE in other alloys at analogous temperature ranges, such as Gd55Co35Mn10 annealed at 600 K for 30 min (|ΔSM| of 5.46 J·kg−1·K−1 with a temperature range from 137 K to 180 K) [14] and fully crystallized Gd55Co35Ni10 (620 K/30 min) ribbon (|ΔSM| of 5.0 J·kg−1·K−1 with a temperature range from 154 K to 214 K) [15], as listed in Table 1, the |ΔSM| of the materials in this work are larger, but the temperature width of the plateau of |ΔSM|-T is narrower.
On another hand, when compared with single amorphous phase alloys, like Gd75(Fe0.25Co0.75)25 [24] and Gd50(Co69.25Fe4.25Si13B13.5)50 [25], the combined merits of larger or comparative |ΔSM| and broader working temperature range can be observed in these Gd60Co40 annealed samples. The MCE in a temperature range of 160–220 K can be used in the fields of space technology, medicine, biology, life sciences and more [14].
The correlation between |ΔSM| and H follows the power law of | S M | H n for the SOMT materials [42], and n is an exponent depending on both applied field and temperature. Particularly, at temperature T = TC or Tpk, n is field independent, and n(Tpk) can be extracted from the slope of the linear fit of the rescaled ln|∆SMpk| vs. lnH plots [26,43]. Through the fitting results displayed in Figure 6, the values of 0.76 can be obtained for the 513 K/20 min annealed Gd60Co40 alloy, which is close to ~0.75 derived from the experimental data of other soft magnetic amorphous alloys [42], which indicates the prevailing contribution of amorphous matrix and neglectable influence from the slight amount of crystallite in the sample as mentioned above. However, the deviations of 0.90 and 0.91 can be observed in 513 K/40 min and 513 K/60 min annealed counterparts owing to the in-situ precipitated nanocrystalline in the amorphous precursor [44,45].
Assuming that the different magnetic phases are non-interacting, the total magnetic entropy change of the multiphase composites can be computed using a rule-of-mixtures sum of the entropy change in the constituent materials with expression described as [13]:
S M com = α S M 1 + β S M 2 + γ S M 3
where ΔSM1, ΔSM2 and ΔSM3 imply the magnetic entropy change, as well as α, β and γ denote the relative weight fractions of the phases 1, 2 and 3 respectively, with the relation of α + β + γ = 1. According to the experimental data of the amorphous Gd60Co40, crystalline Gd4Co3 and Gd12Co7 [17,21,22], the fitting results of |ΔSM|-T curves under field changing from 0 to 5 T for 513 K/20 min and 513 K/60 min annealed Gd60Co40 alloys were depicted in Figure 7. For comparison, the experimental results are also shown.
It can be seen the calculated results fit the experimental data very well, and the adopted weight fractions of the phases in 513 K/20 min and 513 K/60 min samples are 85 wt% amorphous matrix + 15 wt% Gd4Co3 and 30 wt% amorphous matrix + 36 wt% Gd4Co3 + 34 wt% Gd12Co7, respectively. Although the relative content of each phase is roughly estimated [46], the fraction of phases is significant to construct the broadened MCE in this kind of multiphase materials [29,47]. In this work, with increasing annealing time, the evolution of microstructure in Gd60Co40 amorphous ribbon achieves an appropriate constituent of different phases, resulting in the enhanced magnetocaloric performance.
On basis of the numerical approach provided by A. Smaïli et al. [48], in this study, when the composite is consisted by 63 wt% Gd4Co3 + 26 wt% Gd12Co7 + 11 wt% amorphous Gd60Co40, a nearly flat-shape |ΔSM|-T profile can be observed between 180 K and 220 K. As discussed in Section 3.1, the Gd4Co3 is probably the primary precipitate. Therefore, the possible method to achieve this composite is as following: at first, rapid thermal annealing in temperature range of 523 K–563 K (the onset and end temperatures of the first crystallization peak on DSC curve) to induce large amount of Gd4Co3 crystallites; and then, annealing at 513 K for 40–60 min (similar to the treatment in this work) to remain certain of the content of the amorphous phase. Further research will be carried out in the next step.

4. Conclusions

In summary, amorphous-nanocrystalline Gd60Co40 alloys were synthesized by crystallization treatment of the melt-spun amorphous ribbons. With different annealing times (20, 40 and 60 min) at 513 K, Gd4Co3-type and Gd12Co7-type phases precipitated from the amorphous matrix in sequence; however, the grain size and transformation volume tended to saturate after certain annealing times due to the metastable equilibrium between the crystallites and the remaining amorphous phase. In the samples annealed for 40 min and 60 min, the multi-phase structure consisted of the Gd4Co3, Gd12Co7 and amorphous phases, which resulted in three successive magnetic phase transitions at temperatures of 174 K/194 K/219 K and 176 K/195 K/219 K, respectively.
Owing to the overlap of multi-peaks in the |ΔSM|-T curves, broadened MCE with |ΔSM| values of 6.65 ± 0.1 J·kg−1·K−1 and 6.44 ± 0.1 J·kg−1·K−1 in the temperature spans of 180 K–196 K and 177 K–196 K under a field change of 0–5 T were obtained in ribbons annealed for 40 min and 60 min, respectively, and could be modeled by considering non-interacting phases. The enhanced relative cooling power and flattened magnetocaloric properties of partially crystallized composites enable them to be more suitable for the Ericsson thermodynamic cycle, in comparison with the single amorphous phase counterparts.

Author Contributions

Conceptualization, P.H. and H.Z.; methodology, P.H. and H.Z.; validation, P.H., H.Z. and W.L.; formal analysis, P.H., Z.Z., J.T. and X.Z.; investigation, P.H., Z.Z. and H.Z.; resources, W.L.; data curation, P.H., Z.Z., Y.X. and H.Z.; writing—original draft preparation, P.H. and Z.Z.; writing—review and editing, P.H., Z.Z. and H.Z.; visualization, P.H. and H.Z.; supervision, H.Z. and W.L.; project administration, H.Z.; funding acquisition, H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 51701003.

Data Availability Statement

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

Acknowledgments

Authors would like to express thanks Xuguang Liu from Instruments Center for Physical Science, University of Science and Technology of China in Hefei for help with performing PPMS analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Franco, V.; Blazquez, J.S.; Ipus, J.J.; Law, J.Y.; Moreno-Ramirez, L.M.; Conde, A. Magnetocaloric effect: From materials research to refrigeration devices. Prog. Mater. Sci. 2018, 93, 112–232. [Google Scholar] [CrossRef]
  2. Tishin, A.M.; Spichkin, Y.I. Magnetocaloric Effect and Its Applications; IOP Publishing Ltd.: London, UK, 2003; pp. 14–17. [Google Scholar]
  3. Gomez, J.R.; Garcia, R.F.; Catoira, A.D.; Gomez, M.R. Magnetocaloric effect: A review of the thermodynamic cycles in magnetic refrigeration. Renew. Sust. Energ. Rev. 2013, 17, 74–82. [Google Scholar] [CrossRef]
  4. Pecharsky, V.; Gschneidner, K.; Pecharsky, A.; Tishin, A. Thermodynamics of the magnetocaloric effect. Phys. Rev. B 2001, 64, 144406. [Google Scholar] [CrossRef]
  5. Rashid, T.P.; Nallamuthu, S.; Arun, K.; Curlik, I.; Ilkovic, S.; Dzubinska, A.; Reiffers, M.; Nagalakshmi, R. Magnetocaloric effect over a wide temperature range due to multiple magnetic transitions in GdNi0.8Al1.2 alloy. Eur. Phys. J. Plus 2016, 131, 156. [Google Scholar] [CrossRef]
  6. Takeya, H.; Pecharsky, V.K.; Gschneidner, K.A., Jr.; Moorman, J.O. New type of magnetocaloric effect: Implications on low-temperature magnetic refrigeration using an Ericsson cycle. Appl. Phys. Lett. 1994, 64, 2739–2741. [Google Scholar] [CrossRef]
  7. Ma, L.Y.; Gan, L.H.; Chan, K.C.; Ding, D.; Xia, L. Achieving a table-like magnetic entropy change across the ice point of water with tailorable temperature range in Gd-Co-based amorphous hybrids. J. Alloy. Compd. 2017, 723, 197–200. [Google Scholar] [CrossRef]
  8. Zhong, X.C.; Shen, X.Y.; Mo, H.Y.; Jiao, D.L.; Liu, Z.W.; Qiu, W.Q.; Zhang, H.; Ramanujan, R.V. Table-like magnetocaloric effect and large refrigerant capacity in Gd65Mn25Si10-Gd composite materials for near room temperature refrigeration. Mater. Today Commun. 2018, 14, 22–26. [Google Scholar] [CrossRef]
  9. Tian, H.C.; Zhong, X.C.; Liu, Z.W.; Zheng, Z.G.; Min, J.X. Achieving table-like magnetocaloric effect and large refrigerant capacity around room temperature in Fe78-xCexSi4Nb5B12Cu1 (x = 0–10) composite materials. Mater. Lett. 2015, 138, 64–66. [Google Scholar] [CrossRef]
  10. Hashimoto, T.; Kuzuhara, T.; Sahashi, M.; Inomata, K.; Tomokiyo, A.; Yayama, H. New application of complex magnetic materials to the magnetic refrigerant in an Ericsson magnetic refrigerator. J. Appl. Phys. 1987, 62, 3873. [Google Scholar] [CrossRef]
  11. Shen, X.Y.; Zhong, X.C.; Huang, X.W.; Mo, H.Y.; Feng, X.L.; Liu, Z.W.; Jiao, D.L. Achieving a table-like magnetocaloric effect and large refrigerant capacity in in situ multiphase Gd65Mn25Si10 alloys obtained by crystallization treatment. J. Phys. D Appl. Phys. 2017, 50, 035005. [Google Scholar] [CrossRef]
  12. Law, J.Y.; Moreno-Ramírez, L.M.; Blázquez, J.S.; Franco, V.; Conde, A. Gd+GdZn biphasic magnetic composites synthesized in a single preparation step: Increasing refrigerant capacity without decreasing magnetic entropy change. J. Alloy. Compd. 2016, 675, 244–247. [Google Scholar] [CrossRef]
  13. Fu, H.; Ma, Z.; Zhang, X.J.; Wang, D.H.; Teng, B.H.; Agurgo Balfour, E. Table-like magnetocaloric effect in the Gd-Co-Al alloys with multi-phase structure. Appl. Phys. Lett. 2014, 104, 072401. [Google Scholar]
  14. Mo, H.Y.; Zhong, X.C.; Jiao, D.L.; Liu, Z.W.; Zhang, H.; Qiu, W.Q.; Ramanujan, R.V. Table-like magnetocaloric effect and enhanced refrigerant capacity in crystalline Gd55Co35Mn10 alloy melt spun ribbons. Phys. Lett. A 2018, 382, 1679–1684. [Google Scholar] [CrossRef]
  15. Zhong, X.; Tang, P.; Gao, B.; Min, J.; Liu, Z.; Zheng, Z.; Zeng, D.; Yu, H.; Qiu, W. Magnetic properties and magnetocaloric effects in amorphous and crystalline Gd55Co35Ni10 ribbons. Sci. China Phys. Mech. 2013, 56, 1096–1099. [Google Scholar] [CrossRef]
  16. Balfour, E.A.; Ma, Z.; Fu, H.; Hadimani, R.L.; Jiles, D.C.; Wang, L.; Luo, Y.; Wang, S.F. Table-like magnetocaloric effect in Gd56Ni15Al27Zr2 alloy and its field independence feature. J. Appl. Phys. 2015, 118, 123903. [Google Scholar] [CrossRef] [Green Version]
  17. Zhang, H.Y.; Ouyang, J.T.; Ding, D.; Li, H.L.; Wang, J.G.; Li, W.H. Influence of Fe substitution on thermal stability and magnetocaloric effect of Gd60Co40-xFex amorphous alloy. J. Alloy. Compd. 2018, 769, 186–192. [Google Scholar] [CrossRef]
  18. Zheng, Z.G.; Zhong, X.C.; Yu, H.Y.; Franco, V.; Liu, Z.W.; Zeng, D.C. The magnetocaloric effect and critical behavior in amorphous Gd60Co40-xMnx alloys. J. Appl. Phys. 2012, 111, 07A922. [Google Scholar] [CrossRef]
  19. Foldeaki, M.; Giguere, A.; Gopal, B.R.; Chahine, R.; Bose, T.K.; Liu, X.Y.; Barclay, J.A. Composition dependence of magnetic properties in amorphous rare-earth-metalbased alloys. J. Magn. Magn. Mater. 1997, 174, 295–308. [Google Scholar] [CrossRef]
  20. Liu, G.L.; Zhao, D.Q.; Bai, H.Y.; Wang, W.H.; Pan, M.X. Room temperature table-like magnetocaloric effect in amorphous Gd50Co45Fe5 ribbon. J. Phys. D Appl. Phys. 2016, 49, 055004. [Google Scholar] [CrossRef]
  21. Zhong, X.C.; Gao, B.B.; Liu, Z.W.; Zheng, Z.G.; Zeng, D.C. Amorphous and crystallized (Gd4Co3)100−xBx alloys for magnetic refrigerants working in the vicinity of 200 K. J. Alloy. Compd. 2013, 553, 152–156. [Google Scholar] [CrossRef]
  22. Shishkin, D.A.; Baranov, N.V.; Volegov, A.S.; Gaviko, V.S. Substitution and liquid quenching effects on magnetic and magnetocaloric properties of (Gd1-xTbx)12Co7. Solid State Sci. 2016, 52, 92–96. [Google Scholar] [CrossRef]
  23. Wang, Z.W.; Yu, P.; Cui, Y.T.; Xia, L. Near room temperature magneto-caloric effect of a Gd48Co52 amorphous alloy. J. Alloy. Compd. 2016, 658, 598–602. [Google Scholar] [CrossRef]
  24. Shishkin, D.A.; Gazizov, A.I.; Volegov, A.S.; Gaviko, V.S.; Baranov, N.V. Magnetic properties and magnetocaloric effect of melt-spun Gd75(Co1−xFex)25 alloys. J. Non Cryst. Solids 2017, 478, 12–15. [Google Scholar] [CrossRef]
  25. Bao, Y.; Shen, H.; Xing, D.; Jiang, S.; Sun, J.; Phan, M.-H. Enhanced Curie temperature and cooling efficiency in melt-extracted Gd50(Co69.25Fe4.25Si13B13.5)50 microwires. J. Alloy. Compd. 2017, 708, 678–684. [Google Scholar] [CrossRef] [Green Version]
  26. Duc, N.T.M.; Shen, H.X.; Clements, E.M.; Thiabgoh, O.; Sanchez Llamazares, J.L.; Sanchez-Valdes, C.F.; Huong, N.T.; Sun, J.F.; Srikanth, H.; Phan, M.H. Enhanced refrigerant capacity and Curie temperature of amorphous Gd60Fe20Al20 microwires. J. Alloy. Compd. 2019, 807, 151694. [Google Scholar] [CrossRef]
  27. Baker, H.; Okamoto, H. Alloy Phase Diagrams, 1st ed.; ASM International: Novelty, OH, USA, 1992. [Google Scholar]
  28. Álvarez, P.; Gorria, P.; Sanchez Llamazares, J.L.; Blanco, J.A. Searching the conditions for a table-like shape of the magnetic entropy in magneto-caloric materials. J. Alloy. Compd. 2013, 568, 98–101. [Google Scholar] [CrossRef]
  29. Caballero-Flores, R.; Franco, V.; Conde, A.; Knipling, K.E.; Willard, M.A. Optimization of the refrigerant capacity in multiphase magnetocaloric materials. Appl. Phys. Lett. 2011, 98, 102505. [Google Scholar] [CrossRef] [Green Version]
  30. Shen, B.G.; Sun, J.R.; Hu, F.X.; Zhang, H.W.; Cheng, Z.H. Recent progress in exploring magnetocaloric materials. Adv. Mater. 2009, 21, 4545–4564. [Google Scholar] [CrossRef] [Green Version]
  31. Guo, J.; Xie, L.; Liu, C.; Li, Q.; Huo, J.; Chang, C.; Li, H.; Ma, X. Effect of Co/Ni substituting Fe on magnetocaloric properties of Fe-based bulk metallic glasses. Metals 2021, 11, 950. [Google Scholar] [CrossRef]
  32. Tatchev, D.; Vassilev, T.; Goerigk, G.; Armyanov, S.; Kranold, R. Kinetics of primary crystallization of hypoeutectic amorphous Ni–P alloy studied by in situ ASAXS and DSC. J. Non Cryst. Solids 2010, 356, 351–357. [Google Scholar] [CrossRef]
  33. Kusy, M.; Riello, P.; Battezzati, L. A comparative study of primary Al precipitation in amorphous Al87Ni7La5Zr by means of WAXS, SAXS, TEM and DSC techniques. Acta Mater. 2004, 52, 5031–5041. [Google Scholar] [CrossRef]
  34. Ghosh, J.; Mazumdar, S.; Das, M.; Ghatak, S.; Basu, A.K. Microstructural characterization of amorphous and nanocrystalline boron nitride prepared by high-energy ball milling. Mater. Res. Bull. 2008, 43, 1023–1031. [Google Scholar] [CrossRef]
  35. Jiang, X.Y.; Zhong, Z.C.; Greer, A.L. Particle-size effects in primary crystallization of amorphous Al-Ni-Y alloys. Mater. Sci. Eng. A 1997, 226, 789–793. [Google Scholar] [CrossRef]
  36. Zheng, Z.G.; Zhong, X.C.; Yu, H.Y.; Liu, Z.W.; Zeng, D.C. Magnetic phase transitions and magnetocaloric properties of (Gd12-xTbx)Co7 alloys. J. Appl. Phys. 2011, 109, 07A919. [Google Scholar] [CrossRef]
  37. Banerjee, S.K. On a generalised approach to first and second order magnetic transitions. Phys. Lett. 1964, 12, 16–17. [Google Scholar] [CrossRef]
  38. Shand, P.M.; Bohnet, J.G.; Goertzen, J.; Shield, J.E.; Schmitter, D.; Shelburne, G.; Leslie-Pelecky, D.L. Magnetic behavior of melt-spun gadolinium. Phys. Rev. B 2008, 77, 184415. [Google Scholar] [CrossRef] [Green Version]
  39. Jin, P.; Li, Y.; Dai, Y.; Xu, Z.; Song, C.; Luo, Z.; Zhai, Q.; Han, K.; Zheng, H. Zn and P alloying effect in sub-rapidly solidified LaFe11.6Si1.4 magnetocaloric plates. Metals 2019, 9, 432. [Google Scholar] [CrossRef] [Green Version]
  40. Franco, V.; Blazquez, J.S.; Ingale, B.; Conde, A. The magnetocaloric effect and magnetic refrigeration near room temperature: Materials and models. Annu. Rev. Mater. Res. 2012, 42, 305–342. [Google Scholar] [CrossRef] [Green Version]
  41. Griffith, L.D.; Mudryk, Y.; Slaughter, J.; Pecharsky, V.K. Material-based figure of merit for caloric materials. J. Appl. Phys. 2018, 123, 034902. [Google Scholar] [CrossRef]
  42. Franco, V.; Blázquez, J.S.; Conde, A. Field dependence of the magnetocaloric effect in materials with a second order phase transition: A master curve for the magnetic entropy change. Appl. Phys. Lett. 2006, 89, 222512. [Google Scholar] [CrossRef]
  43. Franco, V.; Conde, A.; Kuz′min, M.D.; Romero-Enrique, J.M. The magnetocaloric effect in materials with a second order phase transition: Are T-C and T-peak necessarily coincident? J. Appl. Phys. 2009, 105, 07A917. [Google Scholar] [CrossRef]
  44. Xia, L.; Tang, M.B.; Chan, K.C.; Dong, Y.D. Large magnetic entropy change and adiabatic temperature rise of a Gd55Al20Co20Ni5 bulk metallic glass. J. Appl. Phys. 2014, 115, 223904. [Google Scholar] [CrossRef]
  45. Franco, V.; Conde, A. Scaling laws for the magnetocaloric effect in second order phase transitions: From physics to applications for the characterization of materials. Int. J. Refrig. 2010, 33, 465–473. [Google Scholar] [CrossRef]
  46. Romero-Muñiz, C.; Franco, V.; Conde, A. Influence of magnetic interactions between phases on the magnetocaloric effect of composites. Appl. Phys. Lett. 2013, 102, 082402. [Google Scholar] [CrossRef]
  47. Alvarez, P.; Llamazares, J.L.S.; Gorria, P.; Blanco, J.A. Enhanced refrigerant capacity and magnetic entropy flattening using a two-amorphous FeZrB(Cu) composite. Appl. Phys. Lett. 2011, 99, 232501. [Google Scholar] [CrossRef] [Green Version]
  48. Smaïli, A.; Chahine, R. Composite materials for Ericsson-like magnetic refrigeration cycle. J. Appl. Phys. 1997, 81, 824–829. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) DSC curves of as-spun and annealed Gd60Co40 ribbons.
Figure 1. (a) XRD patterns and (b) DSC curves of as-spun and annealed Gd60Co40 ribbons.
Metals 11 01741 g001
Figure 2. (a) Temperature dependence of magnetization and (b) dM/dT-T curves of the Gd60Co40 alloys after annealing treatment. The inset of (b) is the magnified image of the part between 210 K and 230 K.
Figure 2. (a) Temperature dependence of magnetization and (b) dM/dT-T curves of the Gd60Co40 alloys after annealing treatment. The inset of (b) is the magnified image of the part between 210 K and 230 K.
Metals 11 01741 g002
Figure 3. M-H isotherms of Gd60Co40 amorphous ribbon after annealing at 513 K for (a) 20, (b) 40 and (c) 60 min under the ΔH of 0–5 T.
Figure 3. M-H isotherms of Gd60Co40 amorphous ribbon after annealing at 513 K for (a) 20, (b) 40 and (c) 60 min under the ΔH of 0–5 T.
Metals 11 01741 g003
Figure 4. Arrott plots of the multi-phase Gd60Co40 alloy annealed at 513 K for (a) 20, (b) 40 and (c) 60 min.
Figure 4. Arrott plots of the multi-phase Gd60Co40 alloy annealed at 513 K for (a) 20, (b) 40 and (c) 60 min.
Metals 11 01741 g004
Figure 5. The relation of |ΔSM| to temperature for the different annealed Gd60Co40 alloy ribbons.
Figure 5. The relation of |ΔSM| to temperature for the different annealed Gd60Co40 alloy ribbons.
Metals 11 01741 g005
Figure 6. Ln–Ln plots of the field dependence of the |ΔSMpk| for calculating n(Tpk) exponent of the annealed Gd60Co40 alloys.
Figure 6. Ln–Ln plots of the field dependence of the |ΔSMpk| for calculating n(Tpk) exponent of the annealed Gd60Co40 alloys.
Metals 11 01741 g006
Figure 7. Temperature dependence of |ΔSM| for the annealed Gd60Co40 alloys: (a) 513 K/20 min and (b) 513 K/60 min, fitted by calculating experimental data of constituent materials 1, 2 and 3, which correspond to amorphous Gd60Co40 [17], crystalline Gd4Co3 [21] and melt-spun Gd12Co7 [22], respectively.
Figure 7. Temperature dependence of |ΔSM| for the annealed Gd60Co40 alloys: (a) 513 K/20 min and (b) 513 K/60 min, fitted by calculating experimental data of constituent materials 1, 2 and 3, which correspond to amorphous Gd60Co40 [17], crystalline Gd4Co3 [21] and melt-spun Gd12Co7 [22], respectively.
Metals 11 01741 g007
Table 1. Magnetocaloric properties of present alloys and some representative materials under applied field change of 0–5 T (A and C stand for amorphous and crystalline, respectively). The ΔTplateau denotes the temperature range of the plateau part of the |ΔSM| vs. T curves.
Table 1. Magnetocaloric properties of present alloys and some representative materials under applied field change of 0–5 T (A and C stand for amorphous and crystalline, respectively). The ΔTplateau denotes the temperature range of the plateau part of the |ΔSM| vs. T curves.
AlloysStructureTC (K)ΔTplateau (K)SMpk| (J·kg−1·K−1)ΔTFWHM (K)RCP (J·kg−1)References
H = 0.02 TΔH = 0–5 T
Gd60Co40 (513 K/20 min)A + C196/219-7.7394726.6This Work
Gd60Co40 (513 K/40 min)A + C174/194/219180–1966.75117789.8This Work
Gd60Co40 (513 K/60 min)A + C176/195/219177–1966.54122797.9This Work
Gd4Co3C220-6.4123787.2[21]
Gd60Co40A193-8.386713.8[17]
Gd12Co7A + C179-7.9--[22]
Gd60Co25Fe15A217178–2284.1200820[17]
Gd55Co35Mn10 (600 K/30 min)A + C123/170137–1805.46123671.6[14]
Gd55Co35Ni10 (620 K/30 min)C158/214154–2145.0--[15]
Gd75(Fe0.25Co0.75)25A194-5.7--[24]
Gd50(Co69.25Fe4.25Si13B13.5)50A170-6.56126826[25]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Han, P.; Zhang, Z.; Tan, J.; Zhang, X.; Xu, Y.; Zhang, H.; Li, W. Observation of a Broadened Magnetocaloric Effect in Partially Crystallized Gd60Co40 Amorphous Alloy. Metals 2021, 11, 1741. https://doi.org/10.3390/met11111741

AMA Style

Han P, Zhang Z, Tan J, Zhang X, Xu Y, Zhang H, Li W. Observation of a Broadened Magnetocaloric Effect in Partially Crystallized Gd60Co40 Amorphous Alloy. Metals. 2021; 11(11):1741. https://doi.org/10.3390/met11111741

Chicago/Turabian Style

Han, Ping, Ziyang Zhang, Jia Tan, Xue Zhang, Yafang Xu, Huiyan Zhang, and Weihuo Li. 2021. "Observation of a Broadened Magnetocaloric Effect in Partially Crystallized Gd60Co40 Amorphous Alloy" Metals 11, no. 11: 1741. https://doi.org/10.3390/met11111741

APA Style

Han, P., Zhang, Z., Tan, J., Zhang, X., Xu, Y., Zhang, H., & Li, W. (2021). Observation of a Broadened Magnetocaloric Effect in Partially Crystallized Gd60Co40 Amorphous Alloy. Metals, 11(11), 1741. https://doi.org/10.3390/met11111741

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