Next Article in Journal
Evolution of Toughening Mechanisms in PH13-8Mo Stainless Steel during Aging Treatment
Next Article in Special Issue
Excellent Magnetocaloric Performance of the Fe87Ce13−xBx (x = 5, 6, 7) Metallic Glasses and Their Composite
Previous Article in Journal
A Phase Field Approach to Two-Dimensional Quasicrystals with Mixed Mode Cracks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Correlation between Magnetocaloric Properties and Magnetic Exchange Interaction in Gd54Fe36B10−xSix Amorphous Alloys

1
Key Laboratory of Green Fabrication and Surface Technology of Advanced Metal Materials, Anhui University of Technology, Ministry of Education, Ma’anshan 243002, China
2
Ma’anshan Shuntai Rare Earth New Materials Co., Ltd., Ma’anshan 243100, China
3
School of Material Science and Engineering, Anhui University of Technology, Ma’anshan 243002, China
4
Wuhu Technology and Innovation Research Institute, Anhui University of Technology, Wuhu 241003, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(10), 3629; https://doi.org/10.3390/ma16103629
Submission received: 3 April 2023 / Revised: 27 April 2023 / Accepted: 5 May 2023 / Published: 10 May 2023
(This article belongs to the Special Issue Advances in Amorphous Alloy)

Abstract

:
Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) amorphous ribbons were fabricated by melt-spinning technique. Based on the molecular field theory, the magnetic exchange interaction was analyzed by constructing the two-sublattice model and deriving the exchange constants JGdGd, JGdFe and JFeFe. It was revealed that appropriate substitution content of Si for B can improve the thermal stability, maximum magnetic entropy change and widened table-like magnetocaloric effect of the alloys, while excessive Si will lead to the split of the crystallization exothermal peak, inflection-like magnetic transition and deterioration of magnetocaloric properties. These phenomena are probably correlated to the stronger atomic interaction of Fe-Si than that of Fe-B, which induced the compositional fluctuation or localized heterogeneity and then caused the different way of electron transfer and nonlinear variation in magnetic exchange constants, magnetic transition behavior and magnetocaloric performance. This work analyzes the effect of exchange interaction on magnetocaloric properties of Gd-TM amorphous alloys in detail.

1. Introduction

Magnetic refrigeration (MR) using solid magnetic material as a refrigerant has the advantages of environmental friendliness, low energy consumption and high efficiency. MR has been regarded as a potential alternative to replace traditional gas compression refrigeration [1]. The basic principle of MR is the intrinsic magnetocaloric effect (MCE) of magnetic materials: when a magnetic material is adiabatically magnetized, its total entropy S (S = SM + SE + SL, where SM, SE and SL denote magnetic entropy, electron entropy and lattice entropy, respectively) retains unchanged; the spin will align parallel to the direction of the applied magnetic field, inducing a decrease in SM as well as an increase in SE and SL, and therefore the enhanced lattice vibration leads to an increase in temperature [2]. The process is reversible during demagnetization. Usually, the isothermal magnetic entropy change ΔSM or adiabatic temperature change ΔTad is utilized to estimate the magnitude of the magnetocaloric effect [3].
Magnetocaloric materials can be classified into first-order magnetic transition (FOMT) and second-order magnetic transition (SOMT) materials, according to the order of ferromagnetic (FM)–paramagnetic (PM) phase transition. FOMT material has a discontinuous magnetic transition process with temperature, which is usually related to the giant magnetocaloric effect (GMCE) [3]; however, narrow operating temperature span, high thermal and magnetic hysteresis, and inferior mechanical stability restrict its practical application. The mainly studied FOMT materials for near-room-temperature MR are Gd5(Si, Ge)4 alloys [4], La(Fe, Si)13 compounds [5], Mn-Fe-P-Si alloys [6], Heusler alloys [7], etc. In contrast, although the MCE of SOMT material is much lower, it usually possesses negligible magnetic hysteresis and nearly zero thermal hysteresis, which is beneficial for reducing energy losses in practical applications. Among the SOMT near-room-temperature refrigerants, Gd is the most representative as a reference material [8]. The high manufacturing cost and low corrosion resistance of Gd limit its practical application in near-room-temperature magnetic refrigeration technology [9].
Amorphous alloys with the unique atomic structure (short-range order and long-range disorder) belong to the SOMT material class in general and show excellent properties, such as neglectable magnetic and thermal hysteresis, tailorable Curie temperature, high mechanical properties and corrosion resistance, high resistivity (associated with low eddy current loss) and simple production process, which makes them good candidates as magnetic refrigerants [10]. Currently, studies on magnetocaloric amorphous alloys for the near-room-temperature region are mainly focused on Gd-based and Fe-based amorphous alloys. Although the Fe-based amorphous alloys have the advantages of suitable Curie temperature (TC) and low cost of raw material, their MCE is generally not high enough; oppositely, many of the Gd-based amorphous alloys exhibit large MCE, but their TC is far from room temperature [11].
It has been revealed that the TC of Gd-Co binary amorphous alloys can be raised from 166 K to 282 K by increasing the amount of Co due to the enhancement of Co-Co exchange interaction [12,13,14]. Additionally, further increment in the transition metals (TM) content through replacing Gd with Fe or Ni can promote the TC to 310 K; however, the clusters or nanocrystals will appear more easily in the alloy, weakening the magnetocaloric properties of the alloys [15,16]. If the concentration of Gd remains constant, minor substitution of Fe or Ni for Co will improve or diminish the TC of Gd-Co amorphous alloys owing to the variation in interaction between the transition metals [17,18]. Meanwhile, the peak value of ΔSM (|ΔSMpk|) of the alloys nearly linearly depends on TC−2/3 [17,19]. However, in our previous work, it was found that excessive Fe replacement of Co results in a decrease in the TC and |ΔSMpk| simultaneously [20], which may be ascribed to the enhancement of antiferromagnetic Gd-Fe exchange interactions and variation in predominant exchange interactions from ferrimagnetism to sperimagnetism [21,22].
On the basis of the molecular field theory (MFT), Gd-TM amorphous alloys can be constructed as a two-sublattice model, and three exchange interaction constants of JGdGd, JTMTM and −JGdTM (the positive and negative signs represent the ferromagnetic and antiferromagnetic exchange, respectively) can be derived by analyzing the temperature dependence of magnetization [23]. Particularly, in the Gd-Fe binary amorphous alloys, the change in Fe content causes the variation in magnetic transition behavior, which can be explained by the different exchange interaction constants [24]. Furthermore, Yano et al. reported that the addition of 10 at.% B in Gd60Fe40 amorphous alloy could manipulate the inversely bent (inflection-like) curve of the magnetization to the normal ferromagnetic curve, which is found to originate from the decrease in Fe magnetic moment and the enhancement of magnetic exchange constant |JGdFe| [25]. It has been demonstrated that co-doping of covalent B and Si modified the magnetic transition behavior and improved the magnetocaloric properties (with a larger |ΔSM| at higher working temperature) of the amorphous (Gd0.6Co0.2Fe0.2)95B2Si3 alloy [26]. In this work, a series of Gd54Fe36B10−xSix (x = 0, 2, 5, 8, and 10) amorphous ribbons were prepared, and the influence of Si substitution for B on magnetic and magnetocaloric properties was investigated. In addition, to interpret the variation in magnetic and magnetocaloric properties with composition, the magnetic exchange interaction was analyzed based on the molecular field theory.

2. Experimental Details

Alloy ingots with nominal compositions of Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) were prepared by arc-melting mixtures of high-purity Gd (99.95 wt%) and Fe (99.95 wt%) metals and pre-alloy BFe and SiFe (mass ratios of B/Fe and Si/Fe were 17.62/81.46 and 22.25/74.53, respectively) under Ti-gettered argon atmosphere. Each ingot was overturned and remelted four times to ensure homogeneity. Then the as-spun ribbons were fabricated by single roller melt-spinning method with a copper wheel linear surface velocity of 50 m/s under a high-purity argon atmosphere. The structure of the as-spun ribbons was determined using an X-ray diffractometer (XRD, Bruker D8 Advance) in the 2θ range of 20°–80° with Cu Kα radiation (λ = 0.154178 nm). Thermal analyses of the samples were carried out using a differential scanning calorimeter (DSC, Netzsch STA499 F3) under the protection of an argon gas flow at a heating rate of 0.33 K/s. A physical property measurement system (PPMS, Quantum Design PPMS Evercool-II) was adopted to measure the temperature dependence of magnetization (M-T) curves under the external magnetic field of 10 Oe and 6 kOe. A superconducting quantum interference device (SQUID, Quantum Design MPMS 3) was utilized to detect the isothermal magnetization (M-H) curves under an applied field change of 0–20 kOe at various selected temperatures in the vicinity of the magnetic transition temperature (Ttr). All the magnetic properties were collected with the direction of the applied field parallel to the surface of the ribbons. To evaluate the magnetocaloric properties, the magnetic entropy change |ΔSM| was calculated from the M-H curves using the Maxwell equation as follows [27]:
Δ S M ( T , H ) = S M ( T , H ) - S M ( T , 0 ) = 0 H M ( T , H ) T H d H
which indicates that the magnetic entropy change ΔSM(T,H) of a specific material is proportional to the derivative of magnetization with respect to temperature under a fixed field and to the magnetic field change. Typically, Equation (1) was numerically approximated as follows [6]:
Δ S M ( T , H ) = i M i ( T n + 1 , H i ) M i ( T n , H i ) T n + 1 T n δ H i
where Mi(Tn+1, Hi) and Mi(Tn, Hi) are experimental values of magnetization at temperatures Tn+1 and Tn under the applied field Hi, respectively.

3. Results and Discussion

3.1. Characterization of Amorphous Structure

Figure 1a shows the XRD patterns of the Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) as-spun ribbons. Only one broad diffraction peak at around 2θ = 33° without obvious peaks corresponding to the crystalline phase was obtained in each sample, which indicates the typical amorphous structure of the as-spun ribbons. The amorphous feature can be confirmed by the crystallization-related exothermic peaks in their DSC curves, as exhibited in Figure 1b, and the onset crystallization temperature (Tx) of Gd54Fe36B10−xSix amorphous alloys is 749, 762, 766, 743 and 686 K for x = 0, 2, 5, 8 and 10 respectively. With increasing content of Si, the Tx increases firstly and then decreases, implying that appropriate co-addition of Si and B effectively enhanced the thermal stability of amorphous Gd54Fe36B10−xSix, while immoderate Si content (x ≥ 8) induced the split of the exothermal peak and even two-step crystallization. There is no obvious glass transition in the DSC curves since the competing transformation to crystalline is predominant under the heating rate of 0.33 K/s [28]. For all the samples, the Tx is high enough to ensure the amorphous structure near room temperature.

3.2. Determination of the Transition Temperature

Figure 2a shows the M-T curves of the Gd54Fe36B10−xSix amorphous ribbons measured under the applied field of 10 Oe. It can be seen that the magnetization reduced with rising temperature, presenting a ferrimagnetic–paramagnetic transition. The magnetic transition temperature Ttr was determined by the inflection-point method, taking the temperature corresponding to the minimum derivative of M-T curve (namely the dM/dT vs. T plot, displayed in the inset of Figure 2a). For the samples with x = 0, 2, 5, 8 and 10, the Ttr is 282, 296, 316, 342 and 364 K, respectively. As illustrated in Figure 2b, the Ttr increased nearly linearly with increasing Si content for the Gd54Fe36B10−xSix amorphous alloys, with the fitting expression of Ttr = 8.1x + 279.7, which is possibly due to the increase in magnetic exchange coupling [19,29,30]. Although only the concentration of non-magnetic B or Si elements changed, it may affect the magnetic moment and exchange interaction in the materials [25]. Similar results have been reported in amorphous alloys Gd65Fe10Co10Al10X5 (X = B, Si) and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) [26,31].

3.3. Magnetocaloric Properties

As indicated by Equation (1), the magnetic entropy change is approximately proportional to the dM/dT, and thereby the M-H isotherms of Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) amorphous alloys at different temperatures near their individual Ttr were measured under a magnetic field changing from 0 to 20 kOe, as exhibited in Figure 3. The sweeping rate of the field was slow enough to ensure that the data were recorded in an isothermal process. The obvious magnetic transition process near the Ttr could be observed in all the samples except Gd54Fe36Si10, which had not achieved a paramagnetic state at 385 K.
The magnetic entropy change as a function of temperature (|ΔSM| vs. T curves) is displayed in Figure 4. The maximum magnetic entropy change (|ΔSMpk|) of Gd54Fe36B10−xSix amorphous alloys under field change of 20 kOe is 1.20, 1.43, 1.25, 1.30 and 1.21 J kg−1K−1 at 260, 250, 245, 240 and 225 K for x = 0, 2, 5, 8 and 10, respectively, which increases firstly and then decreases with increasing content of Si, showing a nonlinear correlation. Gd54Fe36B8Si2 possesses the highest |ΔSMpk|. In our previous study, a similar phenomenon was observed in the (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) series of amorphous alloys [26]. The relative cooling power (RCP) is another parameter for estimating the MCE and energy efficiency of a magnetic refrigerant, and it can be evaluated by the product R C P = | Δ S M pk | × Δ T FWHM , where the ΔTFWHM is the full width at half maximum of the entropy curve [32]. In this work, the values of RCP for the amorphous Gd54Fe36B10−xSix with x = 0, 2, 5, 8 and 10 are 295, 374, 323, 362 and 321 J kg−1 (under the field change of 0–20 kOe), respectively. Owing to the larger value of |ΔSMpk|, the Gd54Fe36B8Si2 shows the highest RCP. Compared with the magnetocaloric properties (including |ΔSMpk| and RCP) of some other materials (with similar working temperature) listed in Table 1, the RCP values of the presently studied materials are higher in spite of their lower |ΔSMpk|, which results from the broadened entropy curve and larger ΔTFWHM. This is the typical characteristic of MCE obtained in Gd-based amorphous alloys with the second-order magnetic transition (SOMT) [18,19,29,30,33,34]. The Arrott plots (M2 vs. H/M) of Gd54Fe36B10−xSix ribbons were derived from the M-H isotherms, as shown in Figure 5. According to Banerjee criteria [35], the slopes of the Arrott plots are positive in the whole temperature range for all the samples, indicating that their magnetic transition is SOMT.
However, the RCP is now recognized to overestimate the actual refrigerating capacity of the materials with a minor magnetic entropy change in an unreasonably broad temperature range [32]. In this regard, the temperature average entropy change (TEC) was introduced as a reliable figure of merit to assess the magnetocaloric efficiency; it is calculated by the following equation [36]:
T E C ( Δ T l i f t ) = 1 Δ T l i f t max T mid + Δ T l i f t 2 T mid Δ T l i f t 2 | Δ S M ( T ) | d T
where ΔTlift is the desired lift temperature of the device and Tmid is the central temperature that maximizes the TECTlift) value for a given ΔTlift. In this research, the ΔTlift was chosen between 10 K and 100 K with an interval of 10 K, and Figure 6a illustrates the variation in TEC values with respect to ΔTlift in the magnetic field change of 20 kOe for the Gd54Fe36B10−xSix amorphous alloys. The correlation between TEC and the content of Si indicates that the partial replacement of B by Si improves the magnetocaloric performances in this series of materials, as reflected by the changing tendency of |ΔSM|. Additionally, the TEC values gradually decrease with increasing ΔTlift for each sample, and similar behavior has been reported previously [6,37,38]. It should be noted that the TEC changes very gently with the different ΔTlift values, which is ascribed to the table-like |ΔSM|(T) curves (the |ΔSM| retains almost constant in a wide temperature range). As revealed in Figure 6b, the TEC(30 K) and |ΔSMpk| of amorphous Gd54Fe36B8Si2 display a similar ΔH dependence, and their values are very close at any magnetic field, demonstrating less loss in the material during the magnetic transition. The obtained TEC(30 K, 20 kOe) and TEC(30 K, 15 kOe) values of Gd54Fe36B8Si2 are 1.42 J kg−1K−1 and 1.07 J kg−1K−1, respectively. Compared with some other materials, the values are lower than those of Gd (TEC(10 K, 10 kOe) = 2.91 J kg−1K−1) [36] but comparable to those of (La0.7Pr0.3)0.8Sr0.2Mn0.9Co0.1O3±δ (TEC(10 K, 20 kOe) = 1.3 J kg−1K−1) [39] and La0.65Nd0.05Ba0.3Mn0.85Cr0.15O3 (TEC(25 K, 20 kOe) = 1.7 J kg−1K−1) [38] and higher than those of Fe63.5Cr10Si13.5B9Nb3Cu1 amorphous alloy (TEC(10 K, 15 kOe) = 0.83 J kg−1K−1) [36,38,39,40]. Although the magnetocaloric performance estimated by TEC is not very good, the table-like MCE with a wide temperature range was observed in all the samples, enabling them to be more suitable for the Ericsson thermodynamic cycle [32].
Table 1. Magnetocaloric properties (the magnetic transition temperature Ttr, maximum magnetic entropy change |ΔSMpk| and its corresponding temperature Tpk and relative cooling power RCP) of amorphized Gd54Fe36B10−xSix ribbons and some other materials in comparison. The A and C denote the amorphous and crystalline states, respectively.
Table 1. Magnetocaloric properties (the magnetic transition temperature Ttr, maximum magnetic entropy change |ΔSMpk| and its corresponding temperature Tpk and relative cooling power RCP) of amorphized Gd54Fe36B10−xSix ribbons and some other materials in comparison. The A and C denote the amorphous and crystalline states, respectively.
AlloysStructureTtr (K)Tpk (K)SMpk| (J kg−1K−1)RCP (J kg−1)Ref.
H = 10 OeΔH = 20 kOe
Gd54Fe36B10A2822601.20295This work
Gd54Fe36B8Si2A2962501.43374This work
Gd54Fe36B5Si5A3162451.25323This work
Gd54Fe36B2Si8A3422401.30362This work
Gd54Fe36Si10A3642251.21321This work
Gd50Co50A267.2~2672.36212.8[33]
Gd55Co35Fe10A268~2681.72337[34]
Gd50Fe45Co5A289.5288.51.85289[18]
GdC292~2955.2226.9[9,41]
Gd5Si2Ge2C276~27618.4195.3[4,9]
LaFe11.05Co0.91Si1.04C282.1~28210.5168.7[5,9]
LaFe11.40Co0.52Si1.09C237.7~23816.8147.4[5,9]

3.4. Magnetic Exchange Interaction

In Figure 2, it can be found that the ferrimagnetic–paramagnetic magnetic transition process becomes broad and gentle with increasing Si content, which is possibly attributed to the variation in Fe magnetic moment and magnetic exchange constant (JGdGd, JGdFe, JFeFe) induced by the replacement of B with Si [25,42]. Additionally, the Ttr obtained in the low magnetic field of 10 Oe is higher than the Tpk achieved under the high magnetic field of 20 kOe, while the Ttr is similar to the Tpk for the other series of alloys, as displayed in Table 1. Especially for the present Si-containing Gd54Fe36B10−xSix amorphous samples, the discrepancy between Ttr and Tpk becomes larger with increasing content of Si, and this may be caused by the varied antiferromagnetic coupling between Fe and Gd sublattices associated with the composition and magnetic field [24,25].
To investigate the origin of these phenomena in detail, an MFT analysis was carried out with the two-sublattice model [43]. First of all, the M-T curves of Gd54Fe36B10−xSix amorphous ribbons were measured under the magnetic field of 6 kOe, which ensures the saturation state of the samples, as exhibited in Figure 7 (open circle). The inflection-like behavior with the characteristic of an inversely bent curve in a relatively wide temperature range can be observed for x = 8 and 10, similar to the transition type revealed in Gd-rich region Gd-Fe amorphous ribbons [25]. In the next step, each sublattice magnetization MGd and MFe and the total magnetization M were calculated by assigning some values to three exchange integration constants, JGdGd, JGdFe and JFeFe, at a certain temperature T, then the M-T curves in the field of 6 kOe was fitted through adopting the nonlinear least square method [44]. The ferrimagnetic model was constructed with the following parameters: The Landé factors of Gd and Fe are ցGd = ցFe = 2. The coordination number Zij (i, j = Gd, Fe) is expressed as ZGdGd = ZFeGd = 7.2 ( = 12 X Gd X Gd + X Fe ) and ZGdFe = ZFeFe = 4.8 ( = 12 X Fe X Gd + X Fe ), where XFe and XGd are atomic content of Fe and Gd respectively. The spin quantum number SGd is 7/2 for the Gd sublattice, while the SFe was derived from Fe magnetic moment μFe (μFe = ցFeSFe), and μFe was evaluated from the magnetization μa at 10 K under 6 kOe by μ a = | X Gd μ Gd - X Fe μ Fe | / 100 (where μGd = ցGdSGd = 7 μB) [45]. As a result, the JGdGd, JGdFe and JFeFe were derived, and the fitting profiles of MGd, MFe and M are depicted in Figure 7 (solid line). It can be seen that all the calculated results are in good accordance with the experimental data.
The content dependence of μFe, JGdGd, −JGdFe, JFeFe, the −JGdFe/JFeFe ratio and |ΔSMpk| for Gd54Fe36B10−xSix amorphous alloys is displayed in Figure 8. The μFe increases from 1.60 μB (for x = 0) to 2.68 μB (for x = 2) first and then decreases to 1.4 μB (for x = 10). Similar results can be found in amorphous alloys Fe56Gd24Si12B8 (μFe ≈ 1.40 μB) and Fe56Gd24B20 (μFe = 1.22 μB) [45,46]. On one side, compared with the B element, Si possesses more covalent electrons (mainly 3p electrons), which possibly intensifies the transfer of electrons to the 3d orbital of Fe, leading to the lower value of μFe [25,46,47]. On the other side, the substitution of Si for B could change the local environment and affect the magnetic moment of Fe atoms [48]. In this study, for the moderate Si content, B may absorb electrons from Fe atoms and promote μFe [49]; for the alloys with high content of Si or B (x = 0 and 10), the p-d hybridization dominates the reduction in μFe [50].
With the replacement of B by Si in amorphous Gd54Fe36B10−xSix alloys, the JGdGd increases slightly, reflecting the decrease in the average distance between Gd atoms [24], which is probably attributed to the stronger atomic interaction between Fe and Si (in comparison with the Fe-B pairs), and more metalloid atoms tend to surround Fe atoms [51]. Furthermore, after the addition of Si, fewer B atoms appear at the nearest neighbor locations around Fe atoms, and as described above, B absorbs electrons from Fe for the lower B content [52].
It can be seen that the introduction of Si has stronger impacts on −JGdFe and JFeFe than on JGdGd. Additionally, both −JGdFe and JFeFe decrease firstly and then increase with increasing Si content; the variation rule is opposite to that of μFe, as displayed in Figure 8. The overlap between 5d electron wave-functions of Gd and 3d electron wave-functions of Fe is considered to be the origin of the antiferromagnetic exchange coupling −JGdFe [53]; therefore, the lost 3d electron of Fe absorbed by the surrounding B (with moderate content of Si, i.e., x = 2, 5 and 8) is in accordance with the weakened 3d–5d interaction [25]. The effect of B addition on JFeFe in the amorphous Gd54Fe36B10 alloys is neglectable [25], which can be interpreted by its statistically random distribution in the structure [24]. However, Si has a stronger atomic interaction with Fe and then preferentially neighbors Fe, resulting in the deviation from the statistical distribution, which is reflected by the great increase in exchange constant JFeFe in the Gd54Fe36Si10 [24]. For the Gd54Fe36B10−xSix samples, partial replacement of B by Si may reduce the distance between Fe atoms, and the exchange interaction decreases according to the Bethe–Slater curve [54].
It has been reported that the inflection-like magnetic transition of a Gd-TM amorphous alloy can be adjusted to normal ferromagnetic–paramagnetic transition by increasing the ratio of −JGdTM/JTMTM when the JGdGd remains constant [55]. As shown in Figure 8, the value of −JGdFe/JFeFe increases first and then decreases with Si content rising from 0 to 10, which can explain the inflection-like M-T curves observed for Gd54Fe36B10−xSix amorphous alloys with x = 8 and 10. Moreover, adding more Si probably makes the atomic structure deviate from the statistical distribution, which leads to compositional fluctuation or localized heterogeneity [24]; then the larger discrepancy between Ttr and Tpk was obtained [56].
In Figure 8, the variation tendency of −JGdFe/JFeFe and |ΔSMpk| with Si content is similar, except for the Gd54Fe36B5Si5; the atomic-scale structure and accurate μFe of this series of alloys need to be clarified further. Nevertheless, it can be revealed that the appropriate substitution of Si for B promotes the value of |ΔSMpk| and table-like MCE of amorphous Gd54Fe36B10−xSix alloys; this can be attributed to the enhancement of the −JGdFe/JFeFe and modified magnetic transition behavior. An excessive amount of Si results in a decline in the −JGdFe/JFeFe, an inflection-like transition and the deterioration of magnetocaloric properties.

4. Conclusions

In summary, the effect of Si substitution for B on thermal stability, magnetic transition behavior and magnetocaloric properties of melt-spun Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) amorphous alloys was researched. With appropriate content of Si, the alloys showed enhanced thermal stability and broadened table-like MCE; with excessive Si, the alloys exhibited poorer thermal stability, inflection-like transition behavior and weakened MCE. Among present alloys, Gd54Fe36B8Si2 possesses the largest values of |ΔSMpk| (1.43 J kg−1K−1), RCP (374 J kg−1) and TEC(30 K) (1.42 J kg−1K−1) under an applied field of 20 kOe, as well as a table-like MCE, which makes it more suitable for the MR with the Ericsson cycle.
The variation in magnetic exchange constants JGdGd, JGdFe and JFeFe was obtained by fitting the temperature dependence of magnetization according to the molecular field theory and two-sublattice model. Substitution of B with Si induces the different ways of electron transfer and different atomic interaction (Fe-Si pairs are stronger than Fe-B), resulting in the nonlinear correlation between μFe, JGdGd, −JGdFe, JFeFe and −JGdFe/JFeFe and Si content. Therefore, the shape of the magnetic transition curve and the magnetocaloric properties changed nonlinearly.

Author Contributions

Conceptualization, H.Z.; Methodology, H.Z.; Validation, H.Z. and J.T.; Formal analysis, H.Z., J.T. and H.L.; Investigation, H.Z., J.T., X.Z. and H.S.; Resources, H.Z.; Data curation, H.Z., J.T. and J.Y.; Writing—original draft, J.T.; Writing—review & editing, H.Z. and H.L.; Visualization, H.Z., J.T., X.Z., J.Y., H.S., Y.Z. and W.C.; Supervision, H.Z.; Project administration, H.Z., W.L. and A.X.; Funding acquisition, W.L. and A.X. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (Grant Nos. 51701003, 52001004 and 52272263) and the Anhui Provincial Key Research and Development Plan (Grant No. 2022a05020016).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Balli, M.; Jandl, S.; Fournier, P.; Kedous-Lebouc, A. Advanced materials for magnetic cooling: Fundamentals and practical aspects. Appl. Phys. Rev. 2017, 4, 021305. [Google Scholar] [CrossRef]
  2. Pecharsky, V.K.; Gschneidner, K.A. Advanced magnetocaloric materials: What does the future hold? Int. J. Refrig. 2006, 29, 1239–1249. [Google Scholar] [CrossRef]
  3. Taguchi, Y.; Sakai, H.; Choudhury, D. Magnetocaloric Materials with Multiple Instabilities. Adv. Mater. 2017, 29, 1606144. [Google Scholar] [CrossRef] [PubMed]
  4. Pecharsky, V.K.; Gschneidner, K.A., Jr. Giant Magnetocaloric Effect in Gd5(Si2Ge2). Phys. Rev. Lett. 1997, 78, 4494–4497. [Google Scholar] [CrossRef]
  5. Skokov, K.P.; Karpenkov, A.Y.; Karpenkov, D.Y.; Gutfleisch, O. The maximal cooling power of magnetic and thermoelectric refrigerators with La(FeCoSi)13 alloys. J. Appl. Phys. 2013, 113, 17A945. [Google Scholar] [CrossRef]
  6. Zheng, Z.G.; Wang, H.Y.; Li, C.F.; Chen, X.L.; Zeng, D.C.; Yuan, S.F. Enhancement of Magnetic Properties and Magnetocaloric Effects for Mn0.975Fe0.975P0.5Si0.5 Alloys by Optimizing Quenching Temperature. Adv. Eng. Mater. 2022, 25, 2200160. [Google Scholar] [CrossRef]
  7. Liu, J.; Gottschall, T.; Skokov, K.P.; Moore, J.D.; Gutfleisch, O. Giant magnetocaloric effect driven by structural transitions. Nat. Mater. 2012, 11, 620–626. [Google Scholar] [CrossRef]
  8. Kitanovski, A. Energy Applications of Magnetocaloric Materials. Adv. Energy. Mater. 2020, 10, 7. [Google Scholar] [CrossRef]
  9. Gottschall, T.; Skokov, K.P.; Fries, M.; Taubel, A.; Radulov, I.; Scheibel, F.; Benke, D.; Riegg, S.; Gutfleisch, O. Making a cool choice: The materials library of magnetic refrigeration. Adv. Eng. Mater. 2019, 9, 1901322. [Google Scholar] [CrossRef]
  10. Luo, Q.; Wang, W.H. Magnetocaloric effect in rare earth-based bulk metallic glasses. J. Alloys Compd. 2010, 495, 209–216. [Google Scholar] [CrossRef]
  11. Tang, B.Z.; Guo, D.Q.; Ding, D.; Xia, L.; Chan, K.C. Large adiabatic temperature rise above the water ice point of a minor Fe substituted Gd50Co50 amorphous alloy. J. Non-Cryst. Solids 2017, 464, 30–33. [Google Scholar] [CrossRef]
  12. Wang, Z.W.; Yu, P.; Cui, Y.T.; Xia, L. Near room temperature magneto-caloric effect of a Gd48Co52 amorphous alloy. J. Alloys Compd. 2016, 658, 598–602. [Google Scholar] [CrossRef]
  13. Wu, C.; Ding, D.; Xia, L.; Chan, K.C. Achieving tailorable magneto-caloric effect in the Gd-Co binary amorphous alloys. AIP Adv. 2016, 6, 035302. [Google Scholar] [CrossRef]
  14. Zhang, C.L.; Wang, D.H.; Han, Z.D.; Xuan, H.C.; Gu, B.X.; Du, Y.W. Large magnetic entropy changes in Gd-Co amorphous ribbons. J. Appl. Phys. 2009, 105, 013912. [Google Scholar] [CrossRef]
  15. Tang, B.; Xie, H.; Li, D.; Xia, L.; Yu, P. Microstructure and its effect on magnetic and magnetocaloric properties of the Co50Gd50-xFex glassy ribbons. J. Non-Cryst. Solids 2020, 533, 119935. [Google Scholar] [CrossRef]
  16. Wang, X.; Tang, B.; Wang, Q.; Yu, P.; Ding, D.; Xia, L. Co50Gd48-xFe2Nix amorphous alloys with high adiabatic temperature rise near the hot end of a domestic magnetic refrigerator. J. Non-Cryst. Solids 2020, 544, 120146. [Google Scholar] [CrossRef]
  17. Song, M.; Huang, L.; Tang, B.; Ding, D.; Zhou, Q.; Xia, L. Enhanced Curie temperature and magnetic entropy change of Gd50Co50−xNix amorphous alloys. Mod. Phys. Lett. B 2020, 34, 2050050. [Google Scholar] [CrossRef]
  18. 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]
  19. Zhang, Z.; Tang, Q.; Wang, F.; Zhang, H.; Zhou, Y.; Xia, A.; Li, H.; Chen, S.; Li, W. Tailorable magnetocaloric effect by Fe substitution in Gd-(Co, Fe) amorphous alloy. Intermetallics 2019, 111, 106500. [Google Scholar] [CrossRef]
  20. 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. Alloys Compd. 2018, 769, 186–192. [Google Scholar] [CrossRef]
  21. Liu, X.Y.; Barclay, J.A.; Gopal, R.B.; Földeàki, M.; Chahine, R.; Bose, T.K.; Schurer, P.J.; LaCombe, J.L. Thermomagnetic properties of amorphous rare-earth alloys with Fe, Ni, or Co. J. Appl. Phys. 1996, 79, 1630–1641. [Google Scholar] [CrossRef]
  22. Elkenany, M.M.; Aly, S.H.; Yehia, S. Magnetothermal properties and magnetocaloric effect in transition Metal-Rich Gd-Co and Gd-Fe amorphous alloys. Cryogenics 2022, 123, 103439. [Google Scholar] [CrossRef]
  23. Hansen, P. Magnetic amorphous alloys. In Handbook of Mangetic Materials; Buschow, K.H.J., Ed.; Elsevier Science Publishers: Berlin, Germany, 1991; Volume 6, pp. 356–369. [Google Scholar]
  24. Yano, K. Molecular field analysis for melt-spun amorphous Fe100-xGdx alloys (18 ≦ X ≦ 60). J. Magn. Magn. Mater. 2000, 208, 207–216. [Google Scholar] [CrossRef]
  25. Yano, K.; Kita, E. Effect of covalent element boron on exchange interaction and magnetic moment in Fe0.4Gd0.6 binary alloy. J. Magn. Magn. Mater. 2004, 272, 1370–1371. [Google Scholar] [CrossRef]
  26. Zhang, H.Y.; Xu, Y.F.; Zhang, Z.Y.; Tan, J.; Zhang, X.; Peng, H.; Xiang, X.J.; Li, H.L.; Xia, A.L.; Li, W.H. Influence of Covalent Element B and Si Addition on Magnetocaloric Properties of Gd-Co-Fe-(B, Si) Amorphous Alloys. Metals 2022, 12, 386. [Google Scholar] [CrossRef]
  27. Sandeman, K.G.; Takei, S. Magnetocaloric Materials and Applications. In Handbook of Magnetism and Magnetic Materials; Coey, M., Parkin, S., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 1489–1526. [Google Scholar]
  28. Ichitsubo, T.; Matsubara, E.; Numakura, H.; Tanaka, K.; Nishiyama, N.; Tarumi, R. Glass-liquid transition in a less-stable metallic glass. Phys. Rev. B 2005, 72, 052201. [Google Scholar] [CrossRef]
  29. Xue, L.; Li, J.; Yang, W.; Yuan, C.; Shen, B. Effect of Fe substitution on magnetocaloric effects and glass-forming ability in Gd-based metallic glasses. Intermetallics 2018, 93, 67–71. [Google Scholar] [CrossRef]
  30. Zheng, Z.G.; Qiu, Z.G.; Zeng, D.C. Effect of Co substitution on thermal stability and magnetocaloric effects of Gd70Al20Fe10−xCox amorphous alloys. Mater. Res. Express 2019, 6, 096109. [Google Scholar] [CrossRef]
  31. Pierunek, N.; Śniadecki, Z.; Marcin, J.; Škorvánek, I.; Idzikowski, B. Magnetocaloric effect of amorphous Gd65Fe10Co10Al10X5 (X = Al, Si, B) alloys. IEEE Trans. Magn. 2014, 50, 2506603. [Google Scholar] [CrossRef]
  32. Franco, V.; Blázquez, 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]
  33. Xia, L.; Wu, C.; Chen, S.H.; Chan, K.C. Magneto-caloric effect of a Gd50Co50 amorphous alloy near the freezing point of water. AIP Adv. 2015, 5, 097122. [Google Scholar] [CrossRef]
  34. Zhong, X.; Huang, X.; Shen, X.; Mo, H.; Liu, Z. Thermal stability, magnetic properties and large refrigerant capacity of ternary Gd55Co35M10 (M = Mn, Fe and Ni) amorphous alloys. J. Alloys Compd. 2016, 682, 476–480. [Google Scholar] [CrossRef]
  35. Banerjee, B.K. On a generalised approach to first and second order magnetic transitions. Phys. Lett. 1964, 12, 16–17. [Google Scholar] [CrossRef]
  36. 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]
  37. Sakka, A.; M’Nassri, R.; Nofal, M.M.; Mahjoub, S.; Cheikhrouhou-Koubaa, W.; Chniba-Boudjada, N.; Oumezzine, M.; Cheikhrouhou, A. Structure, magnetic and field dependence of magnetocaloric properties of Pr0.5 RE0.1Sr0.4MnO3 (RE = Eu and Er). J. Magn. Magn. Mater. 2020, 514, 167158. [Google Scholar] [CrossRef]
  38. Mahjoub, S.; M’nassri, R.; Baazaoui, M.; Hlil, E.K.; Oumezzine, M. Tuning magnetic and magnetocaloric properties around room temperature via chromium substitution in La0.65Nd0.05Ba0.3MnO3 system. J. Magn. Magn. Mater. 2019, 481, 29–38. [Google Scholar] [CrossRef]
  39. Mitrofanov, V.Y.; Estemirova, S.K.; Kozhina, G.A. Effect of oxygen content on structural, magnetic and magnetocaloric properties of (La0.7Pr0.3)0.8Sr0.2Mn0.9Co0.1O3±δ. J. Magn. Magn. Mater. 2019, 476, 199–206. [Google Scholar] [CrossRef]
  40. Shishkin, D.A. Magnetothermal properties of iron-based amorphous alloys type of Fe63.5M10Si13.5B9Nb3Cu1 (M = Cr, Mn, Fe, Co, Ni). Mater. Res. Express 2019, 6, 025201. [Google Scholar] [CrossRef]
  41. Dan’kov, S.Y.; Tishin, A.M.; Pecharsky, V.K.; Gschneidner, K.A., Jr. Magnetic phase transitions and the magnetothermal properties of gadolinium. Phys. Rev. B 1998, 57, 3478–3490. [Google Scholar] [CrossRef]
  42. Wang, Y.F.; Duc, N.T.M.; Feng, T.F.; Wei, H.J.; Qin, F.X.; Phan, M.H. Competing ferromagnetic and antiferromagnetic interactions drive the magnetocaloric tunability in Gd55Co30NixAl15−x microwires. J. Alloys Compd. 2022, 907, 164328. [Google Scholar] [CrossRef]
  43. Yano, K.; Akiyama, Y.; Tokumitsu, K.; Kita, E.; Ino, H. Magnetic moment and Curie temperature for amorphous Fe100−xGdx alloys (18 ≦ x ≦ 60). J. Magn. Magn. Mater. 2000, 214, 217–224. [Google Scholar] [CrossRef]
  44. Madsen, K.; Nielsen, H.B.; Tingleff, O. Methods for Non-Linear Least Squares Problems, 2nd ed.; Informatics and Mathematical Modelling; Technical University of Denmark: Lyngby, Denmark, 2004. [Google Scholar]
  45. Hassini, A.; Lassri, H.; Bouhdada, A.; Ayadi, M.; Krishnan, R.; Mansouri, I.; Chaker, B. Magnetic coupling in amorphous Fe80−xGdxB20 alloys. Phys. B Condens. Matter. 2000, 275, 295–300. [Google Scholar] [CrossRef]
  46. Ishio, S.; Obara, N.; Negami, S.; Miyazaki, T.; Kamimori, T.; Tange, H.; Goto, M. Magnetic properties and exchange interactions in (GdxFe1−x)80Si12B8 amorphous alloys. J. Magn. Magn. Mater. 1993, 119, 271–278. [Google Scholar] [CrossRef]
  47. Lee, J.M.; Jung, J.K.; Lim, S.H. Mean field analysis of exchange coupling in amorphous DyFe2-B alloy ribbons. J. Magn. Magn. Mater. 2001, 234, 133–141. [Google Scholar] [CrossRef]
  48. Lu, S.; Zhang, J.; Duan, H. Effects of B substitution for P on structure and magnetic properties of FePB amorphous alloys by first-principle investigation. Intermetallics 2022, 149, 107674. [Google Scholar] [CrossRef]
  49. Yaocen, W.; Yan, Z.; Makino, A.; Yunye, L.; Kawazoe, Y. Structural and Magnetic Study of Fe76Si9B10P5 Metallic Glass by First Principle Simulation. IEEE Trans. Magn. 2014, 50, 1–4. [Google Scholar] [CrossRef]
  50. Tian, H.; Zhang, C.; Zhao, J.; Dong, C.; Wen, B.; Wang, Q. First-principle study of the structural, electronic, and magnetic properties of amorphous Fe–B alloys. Phys. B Condens. Matter. 2012, 407, 250–257. [Google Scholar] [CrossRef]
  51. De Boer, F.R.; Mattens, W.C.M.; Boom, R.; Miedema, A.R.; Niessen, A.K. Cohesion in Metals: Transition Metal Alloys; Elsevier Science Publishers: Amsterdam, The Netherlands, 1988. [Google Scholar]
  52. Li, H.X.; Lu, Z.C.; Wang, S.L.; Wu, Y.; Lu, Z.P. Fe-based bulk metallic glasses: Glass formation, fabrication, properties and applications. Prog. Mater. Sci. 2019, 103, 235–318. [Google Scholar] [CrossRef]
  53. Campbell, I.A. Indirect exchange for rare earths in metals. J. Phys. F Metal Phys. 1972, 2, L47–L50. [Google Scholar] [CrossRef]
  54. McHenry, M.E.; Willard, M.A.; Laughlin, D.E. Amorphous and nanocrystalline materials for applications as soft magnets. Prog. Mater. Sci. 1999, 44, 291–433. [Google Scholar] [CrossRef]
  55. Hasegawa, R.; Taylor, R.C. Magnetization of amorphous Gd-Co-Ni films. J. Appl. Phys. 1975, 46, 3606–3608. [Google Scholar] [CrossRef]
  56. Schwarz, B.; Mattern, N.; Luo, Q.; Eckert, J. Magnetic properties and magnetocaloric effect of rapidly quenched Gd-Co-Fe-Al alloys. J. Magn. Magn. Mater. 2012, 324, 1581–1587. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) DSC curves of the Gd54Fe36B10−xSix as-spun ribbons.
Figure 1. (a) XRD patterns and (b) DSC curves of the Gd54Fe36B10−xSix as-spun ribbons.
Materials 16 03629 g001
Figure 2. (a) Temperature-dependent magnetization curves of Gd54Fe36B10−xSix amorphous ribbons under an applied field of 10 Oe. The inset presents the corresponding dM/dT-T plots. (b) Correlation between Si content and the TC for Gd54Fe36B10−xSix amorphous alloys.
Figure 2. (a) Temperature-dependent magnetization curves of Gd54Fe36B10−xSix amorphous ribbons under an applied field of 10 Oe. The inset presents the corresponding dM/dT-T plots. (b) Correlation between Si content and the TC for Gd54Fe36B10−xSix amorphous alloys.
Materials 16 03629 g002
Figure 3. Isothermal magnetization curves of Gd54Fe36B10−xSix amorphous ribbons near their TC under the field changing from 0 to 20 kOe.
Figure 3. Isothermal magnetization curves of Gd54Fe36B10−xSix amorphous ribbons near their TC under the field changing from 0 to 20 kOe.
Materials 16 03629 g003
Figure 4. Temperature dependence of the |ΔSM| for Gd54Fe36B10−xSix amorphous alloys under the magnetic field change of 20 kOe.
Figure 4. Temperature dependence of the |ΔSM| for Gd54Fe36B10−xSix amorphous alloys under the magnetic field change of 20 kOe.
Materials 16 03629 g004
Figure 5. Arrott plots of amorphous Gd54Fe36B10−xSix ribbons.
Figure 5. Arrott plots of amorphous Gd54Fe36B10−xSix ribbons.
Materials 16 03629 g005
Figure 6. (a) TEC as a function of ΔTlift for the Gd54Fe36B10−xSix amorphous alloys, under the field change of 0–20 kOe. (b) Magnetic field dependence of the TEC(30 K) and |ΔSMpk| for amorphous Gd54Fe36B8Si2.
Figure 6. (a) TEC as a function of ΔTlift for the Gd54Fe36B10−xSix amorphous alloys, under the field change of 0–20 kOe. (b) Magnetic field dependence of the TEC(30 K) and |ΔSMpk| for amorphous Gd54Fe36B8Si2.
Materials 16 03629 g006
Figure 7. M-T curves for the Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) amorphous alloys under magnetic field of 6 kOe (open circles) and the fitting results (solid line). Calculated data of two sublattice magnetizations (MGd and MFe are denoted as “red line” and “blue line”, respectively) are also shown.
Figure 7. M-T curves for the Gd54Fe36B10−xSix (x = 0, 2, 5, 8, 10) amorphous alloys under magnetic field of 6 kOe (open circles) and the fitting results (solid line). Calculated data of two sublattice magnetizations (MGd and MFe are denoted as “red line” and “blue line”, respectively) are also shown.
Materials 16 03629 g007
Figure 8. Correlation of magnetic moment of Fe, exchange interaction constants JGdGd, −JGdFe and JFeFe, ratio of −JGdFe/JFeFe, and |ΔSMpk| with the content of Si for amorphous Gd54Fe36B10−xSix. The sign of JGdGd and JFeFe is positive but JGdFe is negative.
Figure 8. Correlation of magnetic moment of Fe, exchange interaction constants JGdGd, −JGdFe and JFeFe, ratio of −JGdFe/JFeFe, and |ΔSMpk| with the content of Si for amorphous Gd54Fe36B10−xSix. The sign of JGdGd and JFeFe is positive but JGdFe is negative.
Materials 16 03629 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Tan, J.; Zhang, X.; Yan, J.; Shi, H.; Zhu, Y.; Cheng, W.; Li, H.; Li, W.; Xia, A. Correlation between Magnetocaloric Properties and Magnetic Exchange Interaction in Gd54Fe36B10−xSix Amorphous Alloys. Materials 2023, 16, 3629. https://doi.org/10.3390/ma16103629

AMA Style

Zhang H, Tan J, Zhang X, Yan J, Shi H, Zhu Y, Cheng W, Li H, Li W, Xia A. Correlation between Magnetocaloric Properties and Magnetic Exchange Interaction in Gd54Fe36B10−xSix Amorphous Alloys. Materials. 2023; 16(10):3629. https://doi.org/10.3390/ma16103629

Chicago/Turabian Style

Zhang, Huiyan, Jia Tan, Xue Zhang, Jiazhe Yan, Han Shi, Ye Zhu, Weizhong Cheng, Hailing Li, Weihuo Li, and Ailin Xia. 2023. "Correlation between Magnetocaloric Properties and Magnetic Exchange Interaction in Gd54Fe36B10−xSix Amorphous Alloys" Materials 16, no. 10: 3629. https://doi.org/10.3390/ma16103629

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