Next Article in Journal
Effect of Microstructure on High Cycle Fatigue Behavior of 211Z.X-T6 Aluminum Alloy
Previous Article in Journal
Relation between the Fatigue and Fracture Ductile-Brittle Transition in S500 Welded Steel Joints
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Covalent Element B and Si Addition on Magnetocaloric Properties of Gd-Co-Fe-(B,Si) Amorphous Alloys

1
Key Laboratory of Green Fabrication and Surface Technology of Advanced Metal Materials, Anhui University of Technology, Ma’anshan 243002, China
2
School of Materials Science and Engineering, Anhui University of Technology, Ma’anshan 243002, China
*
Author to whom correspondence should be addressed.
Metals 2022, 12(3), 386; https://doi.org/10.3390/met12030386
Submission received: 5 January 2022 / Revised: 17 February 2022 / Accepted: 21 February 2022 / Published: 23 February 2022

Abstract

:
The effect of covalent element addition on the magnetic and magnetocaloric properties of ferrimagnetic Gd60Co20Fe20 amorphous alloy was studied. Particularly, the co-doping of B and Si promoted the magnetocaloric performance (with a larger magnetic entropy change |ΔSM| at higher working temperature) of the initial Gd60Co20Fe20 amorphous alloy. This is possibly ascribed to the modified magnetic transition behavior. Additionally, the broadened magnetocaloric effect with |ΔSM| of 2.25 J kg−1 K−1 at 222.5 K under a magnetic field change of 0–2 T and high relative cooling power of 396.0 J kg−1 were obtained for (Gd0.6Co0.2Fe0.2)95B2Si3. These properties reveal that the material could be a good candidate as a magnetic refrigerant suitable for active magnetic refrigerators.

1. Introduction

Compared with traditional gas compression refrigeration, magnetic refrigeration (MR) exhibits advantages of environmental friendliness, compactness, less noise, and higher efficiency and has attracted worldwide attention [1]. The working principle of MR is the intrinsic magnetocaloric effect (MCE) of magnetic materials, meaning when an external magnetic field is applied isothermally, the magnetic moment is parallelly aligned and the magnetic entropy decreases. Under the adiabatic condition, the total entropy of the magnetic materials remains unchanged, so the lattice and electron entropy increase to balance the variation in magnetic entropy, and then the system temperature becomes higher. The reverse process occurs when the magnetic field is removed. Therefore, the isothermal magnetic entropy change |ΔSM| and the adiabatic temperature change ΔTad are important parameters used to characterize the MCE [2].
Most of the near-room-temperature magnetic refrigerants are ferromagnetic, and the MCE reaches its maximum close to the Curie temperature (TC) [3]. Based on the order of magnetic transition, magnetocaloric materials can be classified into two types, namely first-order magnetic transition (FOMT) materials and second-order magnetic transition (SOMT) materials. FOMT materials undergo a discontinuous change in magnetization with varying temperature, so generally, they show a large |ΔSM|, e.g., Gd5Si2Ge2 [4], Mn-Fe-P-Si alloy [5], Ni-Mn-Si-based alloy [6], and LaFe11.83Mn0.32Si1.30Hx [7]. However, their drawbacks are the sharp drop in |ΔSM| when the temperature deviates from TC, and the existence of magnetic and thermal hysteresis, which would reduce the refrigeration efficiency. In the case of SOMT materials, the magnetization changes with temperature continuously. Although their |ΔSM| is lower than that of FOMT materials, they possess many merits, such as low thermal and magnetic hysteresis and a wider working temperature range, ensuring good magnetocaloric performance. The representative material is pure metal Gd, known as “prototype refrigerant” [3].
In general, magnetic amorphous alloys belong to SOMT materials. Except for the above-mentioned advantages, the unique atomic structure (long-range disordered and short-range ordered) provides amorphous alloys with properties of compositional tailoring, low eddy current losses, high corrosion resistance, etc. [8]. In the past 2 decades, Gd- and Fe-based amorphous alloys have attracted more attention as candidates for near-room-temperature magnetic refrigeration [1]. Among Gd-based amorphous alloys, the ferrimagnetic Gd-Co-Fe alloy system shows higher magnetocaloric performance than many Fe-based samples near room temperature [9]. Under an applied field changing from 0 to 2 T, the peak value of the magnetic entropy change (|∆SMpk|) of Gd48Co50Fe2 and Gd50Co45Fe5 amorphous alloys is 1.98 and 1.85 J kg−1 K−1 at temperatures of 297 and 289 K, respectively [10,11].
However, the magnetic and magnetocaloric properties of Gd-Co-Fe amorphous alloys are sensitive to the Fe content: (1) When the element Co is substituted by a small amount of Fe, TC increases to be closer to room temperature and |ΔSM| decreases moderately [11,12]; (2) further replacement of Co with Fe (e.g., Gd60Co20Fe20) may cause the change in the magnetic exchange constant of antiferromagnetic Gd–Fe interaction (JGd-Fe) and then the deviation of the working temperature from TC, which means the magnetic entropy change and working temperature drop simultaneously [13,14,15]. It has been reported that the magnetic transition behavior can be modified by adding covalent element B to Gd-Fe amorphous alloys, owing to the increase in JGd-Fe and decrease in the Fe magnetic moment [16]. In this study, the influence of adding covalent elements B and Si on the magnetic and magnetocaloric properties of Gd60Co20Fe20 amorphous alloy was studied, with the expectation to improve its magnetocaloric performance by modifying the internal magnetic properties.

2. Materials and Methods

The raw materials used to prepare the master alloys with nominal compositions of Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) were pure metals Gd (99.9 wt%), Co (99.9 wt%), and Fe (99.9 wt%) and the prealloys of BFe and SiFe (the mass ratios of B to Fe and Si to Fe in BFe and SiFe were 17.62:81.46 and 22.25:74.53, respectively). The master alloys were produced using a vacuum arc-melting furnace under the protection of a high-purity argon atmosphere, and the ingots were remelted four times to ensure their compositional homogeneity. The small pieces of the ingots were put into quartz tubes and remelted through high-frequency induction in a high-purity argon atmosphere, and then ribbons with a width of 1.5–2 mm and a thickness of 25–30 μm were manufactured by the single roller melt-spinning method at a wheel surface speed of 50 m/s.
Phase structure characterization of the as-spun ribbons was carried out by X-ray diffraction (XRD; D8 Advance, Bruker, Karlsruhe, Germany) using Cu Kα radiation (λ = 0.15418 nm) and a 2θ range of 20°–80°. Thermal properties were measured by differential scanning calorimetry (DSC; STA 449F3, NETZSCH, Selb, Germany) at a heating rate of 0.33 K/s with high-purity argon gas purged. In-plane magnetic hysteresis loops at different temperatures were detected by a magnetic property measurement system (MPMS; MPMS 3, Quantum Design, San Diego, CA, USA). To investigate the magnetic transition behavior of the samples, the temperature dependence of the magnetization (MT) curve was recorded during the heating process from 100 K to 380 K under a magnetic field of 0.001 T after zero-field cooling (ZFC). The magnetocaloric performance of the ribbons was determined by the temperature dependence of the magnetic entropy change |ΔSM| obtained through indirect measurement using the physical property measurement system (PPMS; PPMS-9T, Quantum Design, San Diego, CA, USA). The magnetization isotherms (MH curves) at selected temperatures were observed under the applied magnetic field changing from 0 to 2 T. Subsequently, according to the MH curves, the |ΔSM| was calculated by the Maxwell relation [17]:
Δ S M T , H = S M T , H S M T , 0 = 0 H M T , H T H d H
Usually, the numerical approximation of the integral of the MH curves with small, discrete temperature intervals was used as the following equation [18]:
Δ S M T i + 1 + T i 2 = i M i + 1 M i T i + 1 T i Δ H i ,
where Mi and Mi+1 are experimental values of magnetization at temperatures Ti and Ti+1 under an applied magnetic field Hi, respectively. All the measurements of magnetic properties were performed with the magnetic field transverse to the alloy ribbons.

3. Results and Discussion

3.1. Structural and Thermal Characterization

Figure 1a exhibits the XRD patterns of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) as-spun ribbons. The typical broadened diffuse peaks indicate the amorphous structure of the samples. However, a few minor sharp peaks can be found. The ones marked with triangles and circles are possibly associated with Gd4Co3 and Gd12Co7 phases, respectively [19]. The precipitation of different crystalline phase resulted in the compositional variation of the amorphous matrix and a slight shift of the diffuse peak for different alloys.
The exothermic peak (corresponding to crystallization) on the DSC curves shown in Figure 1b confirms the amorphous structure of the as-spun ribbons. The initial crystallization temperature (Tx) was determined by the intersection of the extrapolated base line and the steepest tangent to the first exothermic peak. In this research, Tx of (Gd0.6Co0.2Fe0.2)95Si5, (Gd0.6Co0.2Fe0.2)95B2Si3, and (Gd0.6Co0.2Fe0.2)95B5 was 555, 596, and 597 K, respectively, higher than that of Gd60Co20Fe20 (544 K). Therefore, the addition of covalent element B or Si enhances the thermal stability of (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous alloys, and (Gd0.6Co0.2Fe0.2)95B2Si3 possesses the best thermal stability [20]. In addition, the Tx of each sample was high enough to maintain its amorphous structure at room temperature. At higher temperature, more exothermic peaks were obtained, which is ascribed to the formation of different phases at different stages of heating [21]. With increasing B content, the exotherms became sharper and moved closer to each other.

3.2. Magnetic Properties

Figure 2a displays the ZFC MT curves of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloys under a magnetic field of 0.001 T, respectively. With increasing temperature, the magnetic transition from ferrimagnetic to paramagnetic could be observed in all the alloys near the Curie temperature (TC). In this study, we used the inflection point method to determine TC, i.e., the temperature corresponding to the minimum of the dM/dTT curves, as illustrated in Figure 2b. For x = 0, 2, and 5, the values of TC were 256, 230, and 212 K, lower than that of Gd60Co20Fe20 (TC = 338 K). Moreover, Gd60Co20Fe20 showed a broad magnetic transition with respect to temperature, which became narrower after adding B and Si. The magnetic transition behavior is closely related to the internal magnetic exchange interaction [11,15,22,23], and the phenomenon can be interpreted as follows: The introduction of B weakens the magnetic moment of Fe and enhances the antiferromagnetic exchange interaction between Gd and Fe, and then results in a reduction in TC and a sharpened magnetic transition process [16]. The covalent element Si may have an analogous effect.

3.3. Magnetocaloric Properties

The isothermal magnetization (MH) curves of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous alloys at different selected temperatures under an applied magnetic field change of 0–2 T are shown in Figure 3a–d. In this study, a temperature interval of 15 K and 2 K (or 5 K) was chosen for the temperature region far away from and in the vicinity of TC, respectively. It can be seen that with increasing temperature, the magnetization decreased and the magnetic state gradually changed from ferrimagnetism to paramagnetism. ΔSM can be calculated from the data of the MH curves on basis of the Equation (2), and the temperature dependence of |ΔSM| is shown in Figure 4. Under the magnetic field changing from 0 to 2 T, the |∆SMpk| of (Gd0.6Co0.2Fe0.2)95Si5, (Gd0.6Co0.2Fe0.2)95B2Si3, and (Gd0.6Co0.2Fe0.2)95B5 amorphous alloys was 1.92, 2.25, and 2.27 J kg−1 K−1, respectively, higher than that of Gd60Co20Fe20 (1.71 J kg−1 K−1). The temperature corresponding to |∆SMpk| under 2 T was denoted as Tpk, and the values of Tpk for x = 0, 2, and 5 and Gd60Co20Fe20 were 205, 222.5, 202.5, and 205 K, respectively. Therefore, the co-addition of Si and B not only improves |∆SMpk| but also raises the working temperature. The large inconsistency between Tpk and TC is worth noting, which is possibly attributed to the local inhomogeneity in the amorphous matrix, such as nanocrystalline (observed in Figure 1a), clusters, or free volume [15,24].
As listed in Table 1, the |∆SMpk| of the (Gd0.6Co0.2Fe0.2)95B2Si3 amorphous ribbon was larger than that of Gd60Co30Fe10, Gd75(Fe0.75Co0.25)25, and some Fe-based amorphous counterparts with similar Tpk but lower than that of Gd-Co-based amorphous alloys with a small amount of Fe or without Fe, such as Gd60Co35Fe5 and Gd56Co44. Another parameter used to evaluate the MCE is the relative cooling power (RCP), which represents the heat transfer between the hot and cold reservoirs in an ideal refrigeration cycle, and can be described as:
R C P = Δ S M p k × Δ T F W H M ,
where ΔTFWHM is defined as the full width at half maximum of the |ΔSM|–T curve [1]. A large RCP of 396.0 J kg−1 was obtained in the (Gd0.6Co0.2Fe0.2)95B2Si3 sample, and the value of |∆SM| was nearly constant (2.20 ± 0.05 J kg−1 K−1) among a wide temperature range of 198–233 K, i.e., the so-called table-like MCE, which makes it more suitable for Ericsson-cycle refrigeration [25]. It is probable that the exchange interaction of local magnetic clusters with different ordering temperature results in a broad magnetic transition, leading to the broadened |ΔSM|–T curve and significant RCP enhancement [11]. In this study, the (Gd0.6Co0.2Fe0.2)95B2Si3 amorphous alloy exhibited superior performance compared to the partially crystallized Gd55Ni10Co35 (|∆SM| maintained the value of ~2.03 J kg−1 K−1 within 179–205 K) and dual-phase Gd48Co50Nb2 ribbons (|∆SM| maintained its value of ~1.7 J kg−1 K−1 within a temperature range of 220–265 K) [26,27].
To determine the magnetic transition behavior in the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous alloys, Arrott plots (M2 vs. H/M) converted from the isotherms are illustrated in Figure 5. According to the Banerjee criterion, the second-order transition can be identified when the slope of all the plots is positive, while the negative slope corresponds to the first-order transition [33]. In this case, all the slopes of Arrott plots were positive, which demonstrates the magnetic phase transition is of the second order. Figure 6 shows the hysteresis loops detected at 10, 200, 300, and 400 K for Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x ribbons. The samples were almost soft magnetic with negligible hysteresis at 10 and 200 K and were paramagnetic at 400 K.
For the second-order magnetic phase transition, a universal relation of the field dependence of the magnetic entropy change was proposed by Oesterreicher and Parker, as the following equation [34]:
Δ S M H n ,
where n is an exponent correlated with both H and T and can be locally calculated as:
n H , T = d l n Δ S M d l n H
In the particular case of T = TC or Tpk, the n exponent becomes field independent [35]. Figure 7 exhibits the fitting results of the ln|∆SMpk| vs. lnH plots, and the values of n(Tpk) for the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloys were 0.93, 0.95, 0.87, and 0.87. It can be found all the n exponents deviate from the value of 2/3 corresponding to mean field theory and the value of 0.72 ± 0.05 predicted in some amorphous alloys [36,37,38] but are close to the values observed in some Fe-containing Gd-based metallic glasses, e.g., Gd50Co45Fe5 (n = 0.837) [10] and Gd60Fe20Al20 (n = 0.869) [14]. The deviation is probably ascribed to the local inhomogeneity of the structure, since the critical exponents for the mean-field model are highly sensitive to the microstructure of the amorphous alloys [10,14,39].

4. Conclusions

To summarize, Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloy ribbons were fabricated, and their magnetic and magnetocaloric properties were investigated. The addition of covalent B or Si to Gd60Co20Fe20 causes a reduction in TC and sharpens the magnetic transition process. Especially, (Gd0.6Co0.2Fe0.2)95B2Si3 co-doped with B and Si exhibits enhanced magnetocaloric performance with a larger |ΔSMpk| of 2.25 J kg−1 K−1 at a higher working temperature of 222.5 K under an applied field change of 0–2 T. Due to the existence of a small amount of Gd4Co3 or Gd12Co7 phase precipitated from the amorphous matrix, the local structural inhomogeneity leads to a broadened |ΔSM|–T curve and a high RCP of 396.0 J kg−1. The results indicate that the co-addition of B and Si to Gd60Co20Fe20 modifies the magnetic transition behavior and improves the magnetocaloric properties of the material, making the (Gd0.6Co0.2Fe0.2)95B2Si3 amorphous alloy a better candidate for magnetic refrigeration applications.

Author Contributions

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

Funding

The work was supported by the Natural Science Foundation of China (grant no. 51701003) and the Innovation and Entrepreneurship Training Program for College Students of the Anhui University of Technology (grant nos. S201910360193 and 202110360023).

Data Availability Statement

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

Acknowledgments

Authors would like to express gratitude to Xuguang Liu (from the Instruments Center for Physical Science, University of Science and Technology of China) and Wensen Wei (from the Hefei Institutes of Physical Science, Chinese Academy of Science) for help with performing PPMS and MPMS analyses, respectively.

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. Kitanovski, A. Energy applications of magnetocaloric materials. Adv. Energy Mater. 2020, 10, 1903741. [Google Scholar] [CrossRef]
  3. 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. Energy Mater. 2019, 9, 1901322. [Google Scholar] [CrossRef] [Green Version]
  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. Hu, S.Y.; Miao, X.F.; Liu, J.; Ou, Z.Q.; Cong, M.Q.; Haschuluu, O.; Gong, Y.Y.; Qian, F.J.; You, Y.R.; Zhang, Y.J.; et al. Small hysteresis and giant magnetocaloric effect in Nb-substituted (Mn, Fe)2(P,Si) alloys. Intermetallics 2019, 114, 106602. [Google Scholar] [CrossRef]
  6. Moreno-Ramirez, L.M.; Diaz-Garcia, A.; Law, J.Y.; Giri, A.K.; Franco, V. Hysteresis, latent heat and cycling effects on the magnetocaloric response of (NiMnSi)0.66(Fe2Ge)0.34 alloy. Intermetallics 2021, 131, 107083. [Google Scholar] [CrossRef]
  7. Morrison, K.; Sandeman, K.G.; Cohen, L.F.; Sasso, C.P.; Basso, V.; Barcza, A.; Katter, M.; Moore, J.D.; Skokov, K.P.; Gutfleisch, O. Evaluation of the reliability of the measurement of key magnetocaloric properties: A round robin study of La(Fe,Si,Mn)Hδ conducted by the SSEEC consortium of European laboratories. Int. J. Refrig. 2012, 35, 1528–1536. [Google Scholar] [CrossRef] [Green Version]
  8. 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]
  9. Huang, L.W.; Tang, B.Z.; Ding, D.; Wang, X.; Xia, L. Achieving high adiabatic temperature change at room temperature in a Gd48Co50Fe2 amorphous alloy. J. Alloys Compd. 2019, 811, 152003. [Google Scholar] [CrossRef]
  10. Tang, B.Z.; Xie, H.X.; Li, D.M.; 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]
  11. 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]
  12. 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]
  13. 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]
  14. 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. Alloys Compd. 2019, 807, 151694. [Google Scholar] [CrossRef]
  15. Yuan, F.; Li, Q.; Shen, B.L. The effect of Fe/Al ratio on the thermal stability and magnetocaloric effect of Gd55FexAl45−x (x = 15–35) glassy ribbons. J. Appl. Phys. 2012, 111, 07A937. [Google Scholar] [CrossRef]
  16. 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]
  17. Krenke, T.; Duman, E.; Acet, M.; Wassermann, E.F.; Moya, X.; Manosa, L.; Planes, A. Inverse magnetocaloric effect in ferromagnetic Ni-Mn-Sn alloys. Nat. Mater. 2005, 4, 450–454. [Google Scholar] [CrossRef] [Green Version]
  18. McMichael, R.D.; Ritter, J.J.; Shull, R.D. Enhanced magnetocaloric effect in Gd3Ga5−xFexO12. J. Appl. Phys. 1993, 73, 6946–6948. [Google Scholar] [CrossRef]
  19. 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. [Google Scholar] [CrossRef]
  20. Son, H.; Yoo, G.; Mustaghfiroh, Q.; Kim, D.H.; Choi-Yim, H. Effect of Substituting Hf for Zr on Fe-Co-M-Nb-B (M = Zr, Hf) Amorphous Alloys with High Saturation Magnetization. Metals 2022, 12, 12. [Google Scholar] [CrossRef]
  21. Suryanarayana, C.; Inoue, A. Bulk Metallic Glasses, 1st ed.; CRC Press; Taylor & Francis Group: Boca Raton, FL, USA, 2011; pp. 188–214. [Google Scholar]
  22. 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]
  23. 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]
  24. Schwarz, B.; Podmilsak, B.; Mattern, N.; Eckert, J. Magnetocaloric effect in Gd-based Gd60FexCo30−xAl10 metallic glasses. J. Magn. Magn. Mater. 2010, 322, 2298–2303. [Google Scholar] [CrossRef]
  25. Du, Y.S.; Zhang, C.H.; Lu, Y.M.; Li, L.; Li, J.Q.; Ma, L.; Rao, G.H. Table-like magnetocaloric effect and large refrigerant capacity in Nd6Fe13Pd1−xAgx compounds. Intermetallics 2021, 130, 107062. [Google Scholar] [CrossRef]
  26. Yu, P.; Chen, L.S.; Xia, L. Phase separation and its effect on the magnetic entropy change profile in an amorphous Gd48Co50Nb2 alloy. J. Non-Cryst. Solids 2018, 493, 82–85. [Google Scholar] [CrossRef]
  27. Wang, Y.F.; Qin, F.X.; Luo, Y.; Wang, H.; Peng, H.X. Tuning of magnetocaloric effect and optimization of scaling factor for Gd55Ni10Co35 amorphous microwires. J. Alloys Compd. 2018, 761, 1–7. [Google Scholar] [CrossRef]
  28. Fang, Y.; Yu, Z.; Peng, G.; Feng, T. Near room-temperature magnetocaloric effect in amorphous Fe-Sc alloys: The effect of minor Co additions. J. Non-Cryst. Solids 2019, 505, 211–214. [Google Scholar] [CrossRef]
  29. Alouhmy, M.; Moubah, R.; Alouhmy, G.; Abid, M.; Lassri, H. Effects of hydrogen implantation on the magnetocaloric properties of amorphous FeZr films. Vacuum 2021, 186, 110063. [Google Scholar] [CrossRef]
  30. 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]
  31. 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]
  32. 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] [Green Version]
  33. Qin, F.X.; Bingham, N.S.; Wang, H.; Peng, H.X.; Sun, J.F.; Franco, V.; Yu, S.C.; Srikanth, H.; Phan, M.H. Mechanical and magnetocaloric properties of Gd-based amorphous microwires fabricated by melt-extraction. Acta Mater. 2013, 61, 1284–1293. [Google Scholar] [CrossRef]
  34. Oesterreicher, H.; Parker, F.T. Magnetic cooling near Curie temperatures above 300 K. J. Appl. Phys. 1984, 55, 4334–4338. [Google Scholar] [CrossRef]
  35. Franco, V.; Conde, A.; Kuz’min, M.D.; Romero-Enrique, J.M. The magnetocaloric effect in materials with a second order phase transition: Are TC and Tpeak necessarily coincident? J. Appl. Phys. 2009, 105, 07A917. [Google Scholar] [CrossRef]
  36. 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]
  37. Duc, N.T.M.; Shen, H.X.; Clements, E.; Thiabgoh, O.; Llamazares, J.L.S.; Sanchez-Valdes, C.F.; Huong, N.T.; Sun, J.F.; Srikanth, H.; Phan, M.H. Critical magnetic and magnetocaloric behavior of amorphous melt-extracted Gd50(Co69.25Fe4.25Si13B13.5)50 microwires. Intermetallics 2019, 110, 106479. [Google Scholar] [CrossRef]
  38. Zhang, H.; Li, R.; Xu, T.; Liu, F.; Zhang, T. Near room-temperature magnetocaloric effect in FeMnPBC metallic glasses with tunable Curie temperature. J. Magn. Magn. Mater. 2013, 347, 131–135. [Google Scholar] [CrossRef]
  39. Zheng, Z.G.; Zhong, X.C.; Liu, Z.W.; Zeng, D.C.; Franco, V.; Zhang, J.L. Magnetocaloric effect and critical behavior of amorphous (Gd4Co3)1−xSix alloys. J. Magn. Magn. Mater. 2013, 343, 184–188. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) DSC curves of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) as-spun ribbons.
Figure 1. (a) XRD patterns and (b) DSC curves of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) as-spun ribbons.
Metals 12 00386 g001
Figure 2. MT curves of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloys under an applied magnetic field of (a) 0.001 T and (b) the correlating curves of dM/dT vs. T.
Figure 2. MT curves of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloys under an applied magnetic field of (a) 0.001 T and (b) the correlating curves of dM/dT vs. T.
Metals 12 00386 g002
Figure 3. The MH curves of the (a) Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous ribbons with x = (b) 0, (c) 2, and (d) 5 at selected temperatures.
Figure 3. The MH curves of the (a) Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous ribbons with x = (b) 0, (c) 2, and (d) 5 at selected temperatures.
Metals 12 00386 g003
Figure 4. The correlation between magnetic entropy change and temperature of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous alloys.
Figure 4. The correlation between magnetic entropy change and temperature of the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous alloys.
Metals 12 00386 g004
Figure 5. Arrott plots of amorphous alloys, (a) Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x with x = (b) 0, (c) 2, and (d) 5, derived from the MH isotherms.
Figure 5. Arrott plots of amorphous alloys, (a) Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x with x = (b) 0, (c) 2, and (d) 5, derived from the MH isotherms.
Metals 12 00386 g005
Figure 6. Magnetic hysteresis loops of (a) Gd60Co20Fe20 and (bd) (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloys at different temperatures.
Figure 6. Magnetic hysteresis loops of (a) Gd60Co20Fe20 and (bd) (Gd0.6Co0.2Fe0.2)95BxSi5−x (x = 0, 2, 5) amorphous alloys at different temperatures.
Metals 12 00386 g006
Figure 7. Ln–Ln plots of the field dependence of |ΔSMpk| with the calculated n(Tpk) exponent for the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous ribbons.
Figure 7. Ln–Ln plots of the field dependence of |ΔSMpk| with the calculated n(Tpk) exponent for the Gd60Co20Fe20 and (Gd0.6Co0.2Fe0.2)95BxSi5−x amorphous ribbons.
Metals 12 00386 g007
Table 1. Magnetic and magnetocaloric properties of various materials with a similar working temperature range (A and C indicate amorphous and crystalline structures, respectively.).
Table 1. Magnetic and magnetocaloric properties of various materials with a similar working temperature range (A and C indicate amorphous and crystalline structures, respectively.).
AlloysStructureTC
(K)
Tpk
(K)
SMpk|
(J kg−1 K−1)
ΔTFWHM
(K)
RCP
(J kg−1)
Reference
μ0H = 0.001 Tμ0H = 2 T
(Gd0.6Co0.2Fe0.2)95B5A212202.52.27162367.7This work
(Gd0.6Co0.2Fe0.2)95B2Si3A230222.52.25176396.0This work
(Gd0.6Co0.2Fe0.2)95Si5A2562051.92190364.8This work
Gd60Co20Fe20A3382051.71178304.4This work
Gd48Co50Nb2A220/264~2201.7145~246.5[26]
Gd55Ni10Co35
(533 K annealing)
A + C2122122.03148~300[27]
Fe89Co1Sc10A1962101.24182225.7[28]
(Fe93Zr7)89H11A200~2151.008145.6146.75[29]
Gd75(Fe0.75Co0.25)25A + C223~2202.0--[30]
Gd60Co30Fe10A239222.52.08154~320[31]
Gd60Co35Fe5A222222.52.91104~303[31]
Gd56Co44A216~2153.3485~284[32]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, H.; Xu, Y.; Zhang, Z.; Tan, J.; Zhang, X.; Peng, H.; Xiang, X.; Li, H.; Xia, A.; Li, W. Influence of Covalent Element B and Si Addition on Magnetocaloric Properties of Gd-Co-Fe-(B,Si) Amorphous Alloys. Metals 2022, 12, 386. https://doi.org/10.3390/met12030386

AMA Style

Zhang H, Xu Y, Zhang Z, Tan J, Zhang X, Peng H, Xiang X, Li H, Xia A, Li W. Influence of Covalent Element B and Si Addition on Magnetocaloric Properties of Gd-Co-Fe-(B,Si) Amorphous Alloys. Metals. 2022; 12(3):386. https://doi.org/10.3390/met12030386

Chicago/Turabian Style

Zhang, Huiyan, Yafang Xu, Ziyang Zhang, Jia Tan, Xue Zhang, Hui Peng, Xinji Xiang, Hailing Li, Ailin Xia, and Weihuo Li. 2022. "Influence of Covalent Element B and Si Addition on Magnetocaloric Properties of Gd-Co-Fe-(B,Si) Amorphous Alloys" Metals 12, no. 3: 386. https://doi.org/10.3390/met12030386

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