Next Article in Journal
Simulation Study of Crystalline Al2O3 Thin Films Prepared at Low Temperatures: Effect of Deposition Temperature and Biasing Voltage
Previous Article in Journal
Reoxidation Behavior of the Direct Reduced Iron and Hot Briquetted Iron during Handling and Their Integration into Electric Arc Furnace Steelmaking: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Annealing Temperature on Microstructure and Magnetocaloric Properties of Gd-Based Metallic Microfibers

School of Materials Science and Engineering, Inner Mongolia University of Technology, Hohhot 010051, China
*
Author to whom correspondence should be addressed.
Metals 2024, 14(8), 874; https://doi.org/10.3390/met14080874 (registering DOI)
Submission received: 5 June 2024 / Revised: 19 July 2024 / Accepted: 23 July 2024 / Published: 29 July 2024

Abstract

:
In this paper, vacuum annealing has been adopted to introduce atomic cluster micro-regions inside Gd-based metallic microfibers to further explore the effect of the structural changes on the magnetocaloric properties and the mechanism which is systematically expressed. The experimental results indicate that the as-prepared Gd-based metallic microfibers have favorable amorphous formation ability and thermal stability. After annealing @ 380 °C, the maximum magnetic entropy change −ΔSmmax, refrigerating capacity (RC), and relative cooling power (RCP) values of the Gd-based metallic microfibers are 7.20 J/kg·K, 459.4 J/kg, and 588.7 J/kg, respectively. Combined with the transmission electron microscopy analysis results, the internal organizational order of the annealed microfibers is significantly altered, and the atomic clusters formed in localized regions, which reduce the magnetocrystalline anisotropy of the microfibers. While under the uni-action of an external magnetic field, the magnetic moment rotation state and magnetic domain structure distribution of the micro-region atoms will be changed obviously, thereby changing the general magnetic properties and magnetocaloric properties of the metallic microfibers. The above research results can promote the engineering application of Gd-based metallic microfibers in the field of magnetic refrigeration.

1. Introduction

Magnetic metallic microfibers have been widely used in the fields of displacement sensors, magnetic sensitive sensors, magnetic refrigeration devices, etc. [1,2,3,4,5]. Magnetic refrigeration technology based on the magnetocaloric effect (MCE) has received a lot of attention due to its unique structure and electromagnetic properties, as well as its advantages in terms of efficiency and environmental protection [6,7,8,9,10,11,12].
The magnetocaloric effect [13,14] refers to the phenomenon that the temperature of a ferromagnet or paramagnet changes with the change in magnetic field intensity in the adiabatic process. MCE can be quantified by a ferromagnetic or paramagnetic spin system near the magnetic ordering temperature. The total entropy of the system is composed of lattice and magnetic entropy. The magnetic moment is arranged and the magnetic entropy is reduced by applying a magnetic field adiabatically. In order to keep the total entropy of the system unchanged, the lattice entropy increases, and an increase in the lattice entropy means an increase in the system temperature. Therefore, the heat is extracted by the heat transfer medium, and then the magnetic field is removed under adiabatic conditions to achieve system cooling.
In amorphous systems, due to the disorder of the atomic arrangement in amorphous materials, there is no macroscopic magnetic anisotropy, so the magnetization process is more uniform. There is no grain boundary in the amorphous alloy, which avoids the decrease in magnetic properties caused by the grain boundary. Therefore, the amorphous systems have high saturation magnetization, Ms, and magnetic permeability, μm, as well as low coercive force, Hc. Moreover, the Curie temperature, TC [15,16], of amorphous magnetic materials is usually higher, which means that they can still maintain good magnetic stability at high temperatures. At the same time, because there are no grains and grain boundaries of crystalline alloys in amorphous materials, the magnetization process is easier to carry out. This makes the amorphous material have a smaller conductivity when magnetized, further reducing the energy loss.
The Gd-based magnetic alloys have been widely studied in terms of the unpaired electrons in the 4f orbital of the element Gd, resulting in a large single-spin magnetic moment [17,18,19,20,21,22,23]. The movement speed of Gd atoms is faster, and it is difficult to form an ordered crystalline structure, but it is more inclined to form a disordered amorphous structure. Pecharsky and Gschneidner [24] found that the TC of Gd5Si2Ge2 alloy is 274 K, which has a large magnetocaloric effect. The maximum magnetic entropy −ΔSmmax is 18.5 J/kg K under the change of a 0~5 T magnetic field.
Compared with thin ribbons and blocks, microfibers prepared by the rotated-dipping process [25,26,27,28] have the advantages of a high degree of amorphization and small diameter. The TC of Gd55Co15Al24Si1Fe5 metallic glass reaches 126 K, which is about 25% higher than that of the original amorphous alloy. In the Gd25RE25Co25Al25RE = Tb, Dy, Ho system, the quaternary high entropy amorphous alloy system shows excellent magnetocaloric effect, and the TC can be adjusted with the change in rare-earth elements, whereas, on account of the high cooling rate during the fabrication process of the metallic microfibers, the large residual stress is generated inside, which has a certain effect on their magnetic properties. Annealing treatment can effectively reduce the residual stress and improve the magnetic domain structure of the metallic microfibers, thereby improving the magnetic properties. In general, the annealing processes commonly used for metallic microfibers include current annealing, magnetic field annealing, laser annealing, vacuum annealing, and so on [29,30]. Gao J.E. et al. [31] induce nanocrystals into the Fe-based amorphous alloys by vacuum annealing, which enhances the soft magnetic properties, with Ms and Hc values of 1.79 T and 11 A/m, respectively.
The research by Zhang R. C. et al. [32] showed that, after annealing at 1373 K for 20 min, the interaction between the nanocrystals and the amorphous phase inside LaFe11.2Si1.8 metallic microfibers formed a magnetic coupling, which increased the range of the secondary magnetic transition temperature; the −ΔSmmax was 6.2 J/kg·K, and the RC was improved. A.F. Manchon-Gord [15] studied the amorphous structure relaxation of Fe70Zr30 amorphous alloy and found that the magnetization derivative curve narrows with annealing treatment. However, the magnetocaloric effect is more manifested in the temperature change that occurs when the magnetic field is magnetized due to the change in the magnetic domain structure. In addition, after heat treatment, the dependence of the domain structure and magnetocaloric effect can be improved by inducing the precipitation of nanocrystals. Luo Q. et al. [33] found that aging treatment will lead to nanocrystallization inside the alloy Gd51Al24Co20Zr4Nb1. After aging at 923 K for 5 h, the internal structure of the alloy changed from amorphous to polycrystalline, and its magnetic properties decreased markedly.
In summary, the purpose of this paper is to analyze the influence of nanocluster structure on magnetic and magnetocaloric behavior in Gd-based metallic microfibers. Therefore, in order to further regulate the behavior of the nanoclusters, the precipitation of nanoclusters was induced by vacuum annealing. By analyzing the structure, morphology, and atomic order of the metallic microfibers, the effect of the metallic microfiber annealing temperature on the magnetic properties and the magnetocaloric behavior of the nanoclusters is obtained, which provides theoretical guidance and technical support for practical engineering applications.

2. Materials and Methods

We adopted Gd, Al, Co, and Ni with a purity of 99.9% as raw materials and determined the proportion according to the nominal composition Gd57Al19Co19Ni5 [34] (expressed as Gd-based metallic microfibers in the article). The non-consumable magnetron tungsten vacuum arc melting furnace was utilized for melting while filled with high-purity Ar gas for protection. After repeated electromagnetic stirring and smelting, copper mold suction casting was carried out to obtain a master alloy preform with a diameter of 8 mm. Gd-based metallic microfibers with a diameter of 35–40 μm and a length of more than 600 mm were prepared with rotated-dipping equipment with a copper roller speed of 1700 r/min and a feed ratio of 15 μm/s. The thermophysical parameters of the metallic microfibers were obtained with differential scanning calorimetry (type: DSC131 Evo, Setaram, Lyon, France) at a heating rate of 20 °C/min, providing a reference for the vacuum annealing process. The microfibers were vacuum-sealed with a vacuum rotary sealing device (type: MRVS-1002, Balab, Wuhan, China), and samples were annealed in a box-type resistance furnace, with the setting of annealing temperatures at 300 °C, 340 °C, and 380 °C for 10 min and air cooled. Figure 1 shows the heat-treatment process for the Gd-based metallic microfibers and the samples’ appearance.
The phase composition of the metallic microfibers was analyzed with a Rigaku D/MAX-2500/PC rotating anode X-ray diffractometer (XRD). The surface morphology of the metallic microfibers was characterized with secondary electron imaging by using an FEI QUANTA 650 FEG scanning electron microscope (SEM) with an accelerating voltage of 20 kV, and the elemental distribution of metallic microfibers was analyzed by energy dispersive spectrometry (EDS). The patterns of the metallic microfibers accordingly proceeded with a JEOL JEM2010 transmission electron microscope (TEM). In addition, the magnetic properties of the microfibers were measured with a SQUID-VSM-type magnetic property measurement system (MPMS).
Generally, magnetic entropy change −ΔSm, RC, and RCP are used to evaluate the magnetocaloric performance of magnetic materials. And −ΔSm can be calculated by bringing the data from the isothermal magnetization curve into Maxwell’s equation [35]:
Δ S m ( T , H ) = S ( T , H ) S ( T , 0 ) = 0 H max ( M T ) H d H
where S—magnetic entropy; Hmax—maximum external magnetic field strength; T—temperature.
The temperature interval corresponding to half of the maximum magnetic entropy change −ΔSmmax value of the metallic microfiber is defined as the working temperature range. The integral of the area of the magnetic entropy change curve in the half maximum width temperature range is defined as the refrigeration capacity, RC. The product of the half peak width δTFWHM and −ΔSm is defined as the relative cooling power, RCP. The refrigerating efficiency of magnetocaloric materials is measured according to the values of RC and RCP [36], which can be expressed as:
R C = T 1 T 2 Δ S m max ( T ) d T
R C P = Δ S m max × δ T F W H M = Δ S m max

3. Results and Discussion

3.1. Structural Analysis of Gd-Based Metallic Microfibers

Figure 2 shows the XRD diffraction patterns of the Gd-based metallic microfibers before and after annealing and the DSC curves. From Figure 2a, the diffuse scattering peak appeared around 2θ = 35°, with no sharp crystal diffraction peaks as a whole, which shows that the microstructure of the metallic microfibers is still in the amorphous state after annealing. A short time of vacuum annealing at the glass transition temperature Tg does not change the surface into a crystal structure. [33] As seen in Figure 2b, the metallic microfibers have two exothermic peaks, there is secondary crystallization during the heating process, and the first crystallization releases more heat than the secondary crystallization. From Figure 2c, it can be seen that an exothermic phenomenon occurs, which is due to the change in the material from the melting state to the solidification state.
As listed in Table 1, the thermophysical parameters of the metallic microfibers are obtained from the DSC curve, where Tg is the glass transition temperature, Tx1 is the initial crystallization temperature, Tx2 is the secondary crystallization temperature, −ΔH is the mixing enthalpy, and ΔT is the glass transition region. It can be seen from the table that the metallic microfibers have high thermal stability. Furthermore, accordingly, Tg, Tx1, and Tx2 are 303 °C, 346 °C, and 384 °C, respectively. The annealing temperatures are determined to be 300 °C, 340 °C, and 380 °C.
Figure 3 indicates the SEM morphology and EDS analysis of the metallic microfibers before and after the annealing process. From Figure 3a,c,e,g, the surface of the metallic microfibers is smooth, uniform, and continuous. The EDS spectra of the metallic microfibers shown in Figure 3b,d,f,h show that the chemical compositions have not changed after annealing. And the positions of the energy spectrum peaks of each element have not deviated, indicating that the annealing process of the appropriate temperature has not altered the chemical composition of the metallic microfibers.
Figure 4 illustrates the TEM analysis of the Gd-based metallic microfibers before and after the annealing process. Figure 4a,b show that the microstructure of the metallic microfibers is uniform in the as-prepared and annealed states at 380 °C, with no nanocrystalline regions. Figure 4c,d show that the SAED patterns of the as-prepared and 380 °C annealed metallic microfibers are both amorphous halos, indicating that the metallic microfibers remain in the amorphous state after annealing.
The autocorrelation function (ACF) is modified to quantitatively characterize the order-degree change of the metallic microfibers both before and after annealing. It is possible to express the domain structure order degree (DSO) of the metallic microfibers as follows:
ψ = ς κ × 100 %
where ς —the number of ordered areas; κ —the total number of divided areas; and κ = 64 .
The HRTEM, FFT, IFFT, and ACF pictures of the metallic microfibers before and after the annealing process are shown in Figure 4e,f. The HRTEM result displays the obvious bright and dark stripes in the metallic microfibers’ local area after annealing, along with the appearance of atomic clusters. On the other hand, the IFFT atomic conformation diagram illustrates the random arrangement of atoms within the metallic microfibers, which is caused by the extreme cooling rate during formation, resulting in the internal stress of the fibers being non-uniformly distributed along the axial direction during condensation. According to the ACF ordered statistics, the order degrees of the as-prepared and 380 °C annealed microfibers are 4.69% and 20.31%, respectively. In conclusion, after annealing at 380 °C, atomic clusters formed due to structural relaxation inside the metallic microfibers, and an orderly micro-domain of atomic arrangement was formed in a certain area, which can affect the magnetic properties, especially in the process of magnetization, hindering the movement of the internal domain wall and the rotation of the magnetic moment of the metallic microfibers, thereby affecting the synergy of the magnetic moment formed under higher external magnetic field conditions. As mentioned above, it provides the structural basis for the change in the magnetic properties in the following research.

3.2. Magnetic Property Analysis of Gd-Based Metallic Microfibers

Figure 5 exhibits the general magnetization curves of the metallic microfibers at 20 K before and after annealing with the applied magnetic field of 0~5 T, from which the Ms of metallic microfibers decreases with the rising of the annealing temperature. At the annealing temperature of 380 °C, the Ms is the smallest, and the value is 155.98 emu/g.
Figure 6 shows the hysteresis loop of the metallic microfibers at 20 K before and after the annealing process and presents that the residual magnetization Mr, coercive force Hc, and magnetic permeability μm of the metallic microfibers increase with the rise in the annealing temperature, and when the annealing temperature is 380 °C, its maximum values are 17.83 emu/g, 125.12 Oe, and 0.58, respectively. At the annealing temperatures of 300 °C and 340 °C, the hysteresis loops are similar to that of the as-prepared state, showing a “Z shape”, and the area of the hysteresis loop indicates small hysteresis loss, while at the annealing temperature of 380 °C, the hysteresis loss increases.
Combining Figure 5 and Figure 6, it can be found that the softly magnetic properties of the microfibers decreased after annealing. This is due to the atomic clusters appearing and forming ordered micro-domains inside the microfibers after the annealing processing, in which there are atomic magnetic moments in different directions from the external magnetic field. Since the spin magnetic moments of the electron are different, the magnetic moments of the atoms are not equal, so the opposite magnetic moments can incompletely offset, resulting in a significant decrease in saturation magnetization.
Figure 7 presents the M-T curves of the metallic microfibers before and after the annealing processing. It can be seen that the TC of the metallic microfibers in the as-prepared state and annealed at 300 °C and 380 °C is 109.3 K, 108.7 K, and 84.4 K, respectively. They all have a rapid transition from ferromagnetism (FM) to paramagnetism (PM) around the TC. In the paramagnetic region, the magnetization of the microfibers disappears and does not vary with the change in the applied magnetic field. In the ferromagnetic region, the heating and cooling M-T curves of the as-prepared metallic microfibers and those annealed at 300 °C overlap, respectively. However, the M-T curves of the metallic microfibers annealed at 380 °C in the temperature range from 20 K to 60 K do not overlap, indicating that magnetothermal hysteresis occurs, which is related to the appearance of a large number of atomic cluster micro-domains inside. The atomic magnetic moment rotation process is affected by the cluster pinning effect, reflecting the magnetocrystalline anisotropy characteristics, so it has strong hysteresis characteristics.
Figure 8 displays the isothermal magnetization curves of the metallic microfibers before and after the annealing process. It can be seen that the changing trend of the metallic microfibers is consistent and has a higher magnetic susceptibility in the temperature range below the TC. The magnetization increases with the rise in the external magnetic field and can quickly reach a saturation state, showing obvious ferromagnetic properties. Around the TC, the metallic microfibers carry out a rapid transition from FM to PM. The magnetic susceptibility of the metallic microfibers is small and completely transitions to the paramagnetic state above TC. Particularly, the metallic microfibers need to be in a relatively high external magnetic field to reach the magnetic saturation state at the annealing temperature of 380 °C, which is caused by the inconsistency in the direction of easy magnetization inside the micro-domains of the atomic clusters formed after annealing and the direction of the magnetic field.
Figure 9 presents the Arrott curves of the Gd-based metallic microfibers before and after annealing. All the curves are approximately aligned in a parallel manner with the positive slopes and without obvious inflection points, showing the typical characteristics of secondary magnetic phase change materials. The results of the Arrott curves [37,38] of the Gd-based metallic microfibers are consistent with the related research results. No change occurs in the type of phase transition in the metallic microfibers, which is still the FM-PM secondary transition, indicating that the metallic microfibers before and after annealing can be used as a magnetic refrigeration working material.
Figure 10 shows the magnetic entropy change curves of the Gd-based metallic microfibers before and after the annealing process, from which the −ΔSm value of the microfibers increases firstly and then decreases with the rise in the external temperature, the value of −ΔSmmax appears near TC, and the curves show symmetry with TC as the axis. This is a typical second-order magnetic phase transition. At the external magnetic field of 5 T, the −ΔSmmax of the as-prepared metallic microfibers is 11.57 J/kg·K. When the annealing temperature increases, the −ΔSmmax change tends to decrease, the value of which is 7.20 J/kg·K at the annealing temperature of 380 °C. This is closely related to the change in the atomic arrangement structure. After the annealing process, the degree of the internal order of the microfibers increases, and the arrangement of the magnetic moments in the micro-regions becomes orderly, which reduces the disorder of the internal magnetic moments. This leads to a decrease in the magnetic entropy of the metallic microfibers.
The RC and RCP of the Gd-based metallic microfibers before and after the annealing process are plotted in Figure 11. When the external magnetic field increases, the RC and RCP of the microfibers increase approximately linearly. At the external magnetic field of Hex = 5 T, the as-prepared metallic microfibers reach a maximum cooling efficiency, and the RC and RCP are 834.14 J/kg and 1138.16 J/kg, respectively. Compared with the RC and RCP values [39] mentioned in previous studies, the cooling capacity has been improved. The RC and RCP of the metallic microfibers after annealing at 300 °C are 667.8 J/kg and 894.6 J/kg, and they are 459.4 J/kg and 588.7 J/kg, respectively, after annealing at 380 °C.
To sum up, the magnetic refrigeration capacity of the Gd-based metallic microfibers before and after the vacuum annealing process is comprehensively evaluated according to magnetocaloric performance indicators such as −ΔSmmax, RC, and RCP, and it can be found that the cooling capacity of the annealed metallic microfibers decreased compared with that of the as-prepared microfibers.
Therefore, the appropriate annealing temperature can improve the magnetic properties of the amorphous alloys. However, the magnetocaloric properties of the amorphous alloys will be reduced due to the change in the internal magnetic domain distribution at high temperatures [31,32], which provides a theoretical basis for the application of Gd-based metallic microfibers in magnetic refrigeration devices.
As listed in Table 2, the −ΔSmmax and RC values of the Gd57Al19Co19Ni5 metallic microfibers are better compared with those alloys. The different transition temperatures of the magnetocaloric materials are applied to different working environment requirements and fields. The TC value of the Gd-based metallic microfibers is about 109 K. The Gd-based metallic microfibers have excellent magnetocaloric properties. The development and performance optimization of Gd-based metallic microfibers can greatly accelerate the engineering application process of magnetic refrigeration technology in various fields. Its application fields include medicine and health care, civil life, precision instruments, and aerospace.

3.3. Mechanism Analysis of Annealing Gd-Based Metallic Microfibers

The mechanism of the effect of the annealing temperature on the magnetic properties of the Gd-based metallic microfibers can be analyzed from the two aspects of magnetic moment arrangement and magnetic domain structure change to further explore the influence of the interaction between the atoms and the evolution of the microstructure on the magnetic properties. For ferromagnetic materials, without an external magnetic field, the vector sum of their internal intrinsic magnetic moments is zero, showing non-magnetism, while under the condition of an external magnetic field, the intrinsic magnetic moment of the material changes and exhibits magnetism. Therefore, the study of the magnetic change in the Gd-based metallic microfibers needs to start with the atomic magnetic moment. The effect of annealing on the magnetic properties of the Gd-based metallic microfibers is explained in Figure 12.
After the vacuum annealing process, the residual internal stress of the metallic microfibers caused by rapid cooling is greatly eliminated, as shown in Figure 12a. Meanwhile, it can be seen from the HRTEM image that, after annealing, there are atomically arranged micro-domains inside where the atomic magnetic moments are aligned. As shown in Figure 12b, the magnetic moment of the microfiber surface is reversed when the magnetic moment rotates from low energy to high energy with the increase in the magnetic field induction intensity. From Figure 12c, it can be seen that vacuum annealing increases the number of ordered atomic micro-regions without crystallization occurring. In addition, the direction of the magnetic moment remains consistent along the direction of the magnetic field. Under the action of an external magnetic field, the domain wall is displaced, as shown in Figure 12d, and the energy consumed is much greater than that at the angle of 90°, so the Ms value of the metallic microfibers decreases. In addition, after the microfiber is magnetized, a circular induced current (eddy current) will be generated, and the eddy current will excite a magnetic field opposite to the external magnetic field, thereby reducing the Ms of the material. After the annealing process of the microfibers, the internal atomic spin interaction is enhanced, the interatomic bonding force is also enhanced, and under the condition of an external magnetic field, the magnetoelastic performance of the microfibers increases, which eventually increases the magnetization work of the metallic microfibers and reduces the Ms. The internal ordering of the annealed microfibers is improved in a certain region, and the ordered arrangement of the magnetic moments reduces the disorder of the internal magnetic moments, thus lowering the −ΔSm value.

4. Conclusions

In summarizing, the structure, general magnetic properties, and magnetocaloric properties of Gd-based metallic microfibers before and after annealing processing were systematically compared and analyzed, and the conclusions are as follows:
  • Gd-based metallic microfibers have a favorable thermal stability and glass-forming ability. The local atoms in the microfibers are arranged in an orderly manner and the ordered micro-regions are formed after annealing @ 380 °C.
  • The Mr, Hc, and μm of the Gd-based metallic microfibers increase with the rising of the annealing temperature, which reached 17.83 emu/g, 125.12 Oe, and 0.58, respectively at the annealing temperature of 380 °C. The −ΔSmmax, RC, and RCP for the microfibers annealed at 380 °C are 7.20 J/kg K, 459.4 J/kg, and 588.7 J/kg, respectively, at the Hex of 5 T.
  • After annealing @ 380 °C, the local atoms in the metallic microfibers are rearranged to generate ordered micro-regions of atomic clusters, which will hinder the movement of the domain wall and the rotation of the magnetic moment during the magnetization process, thereby affecting the synergy of the magnetic moment under the external magnetic field. Simultaneously, the magnetic entropy value of the microfibers decreases with the formation of the locally ordered micro-regions, which ultimately affects the magnetocaloric properties.

Author Contributions

Conceptualization, J.L.; methodology, M.Z. and S.Y.; investigation, S.Y., Z.L. and Y.C.; resources, J.L. and Z.L.; formal analysis, J.L.; writing—original draft preparation, J.L., S.Y., Z.L. and M.Z.; writing—review and editing supervision, J.L., S.Y., Z.L. and Y.C.; project administration, J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (NSFC) under grant no. 52061035, the Key Project of the Natural Science Foundation of Inner Mongolia Autonomous Region (no. 2024ZD07), the Young Leading Talent of “Grassland Talents” Project of Inner Mongolia Autonomous Region (no. QNLJ012010), the Program for Innovative Research Team in Universities of Inner Mongolia Autonomous Region (no. NMGIRT2211), and the Inner Mongolia University of Technology Key Discipline Team Project of Materials Science (no. ZD202012).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Luo, L.; Shen, H.X.; Bao, Y.; Yin, H.; Jiang, S.D.; Huang, Y.J.; Guo, S.; Gao, S.Y.; Xing, D.W.; Li, Z.; et al. Magnetocaloric effect of melt-extracted high-entropy Gd19Tb19Er18Fe19Al25 amorphous microwires. J. Magn. Magn. Mater. 2020, 507, 166856. [Google Scholar] [CrossRef]
  2. Shao, L.L.; Xue, L.; Luo, Q.; Wang, Q.Q.; Shen, B.L. The role of Co/Al ratio in glass-forming GdCoAl magnetocaloric metallic glasses. Materialia 2019, 7, 100419. [Google Scholar] [CrossRef]
  3. Gandara, L.A.B.; Zubiate, L.V.; Flores, D.M.C.; Galindo, J.T.E.; Ornelas, C.; Ramos, M. Tuning magnetic entropy change and relative cooling power in La0.7Ca0.23Sr0.07MnO3 electrospun nanofibers. Nanomaterials 2020, 10, 435. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, F.; Han, K.; Gao, M.; Zhang, Y.; Xu, W.; Huo, J.T.; Zhang, C.J.; Song, L.J.; Wang, J.Q. Magnetocaloric properties of melt-extracted Gd-Co-Al amorphous/crystalline composite fiber. Metals 2022, 12, 1367. [Google Scholar] [CrossRef]
  5. Corte-Leon, P.; Zhukova, V.; Blanco, J.M.; González-Legarreta, L.; Ipatov, M.; Zhukov, A. Stress-induced magnetic anisotropy enabling engineering of magnetic softness of Fe-rich amorphous microwires. J. Magn. Magn. Mater. 2020, 510, 166939. [Google Scholar] [CrossRef]
  6. Gudoshnikov, S.A.; Odintsov, V.I.; Liubimov, B.Y.; Menshov, S.A.; Churukanova, M.N.; Kaloshkin, S.D.; Elmanov, G.N. Method for evaluating the temperature of amorphous ferromagnetic microwires under Joule heating. Measurement 2021, 182, 109783. [Google Scholar] [CrossRef]
  7. Xue, L.; Luo, Q.; Shao, L.L.; Shen, B.L. Magnetocaloric difference between ribbon and bulk shape of Gd-based metallic glasses. J. Magn. Magn. Mater. 2020, 497, 166015. [Google Scholar] [CrossRef]
  8. Yin, H.B.; Huang, Y.J.; Bao, Y.; Jiang, S.D.; Xue, P.; Jiang, S.S.; Wang, H.; Qin, F.X.; Li, Z.; Sun, S.C.; et al. Comparable magnetocaloric properties of melt-extracted Gd36Tb20Co20Al24 metallic glass microwires. J. Alloys Compd. 2020, 815, 150983. [Google Scholar] [CrossRef]
  9. Gudoshnikov, S.A.; Odintsov, V.I.; Liubimov, B.Y.; Menshov, S.A.; Popova, A.V.; Tarasov, V.P. Correlation of electrical and magnetic properties of Co-rich amorphous ferromagnetic microwires after DC Joule heating treatment. J. Alloys Compd. 2020, 845, 156220. [Google Scholar] [CrossRef]
  10. Hou, L.; Xiang, X.Y.; Huang, Y.; Zhang, B.; Jiang, C.; Chen, S.S.; Li, W.H. Influences of oxygen on the magnetocaloric properties of a Fe-based amorphous alloy. Appl. Phys. A 2021, 127, 501. [Google Scholar] [CrossRef]
  11. Tishin, A.M.; Spichkin, Y.I. Recent progress in magnetocaloric effect: Mechanisms and potential applications. Int. J. Refrig. 2014, 37, 223–229. [Google Scholar] [CrossRef]
  12. Yang, Z.Y.; Qin, S.J.; Ye, X.B.; Liu, Z.H.; Guo, Y.J.; Cui, H.Z.; Ge, J.Y.; Long, Y.W.; Zeng, Y.J. Large magnetic entropy change in weberite type oxides Gd3MO7 (M = Nb, Sb, and Ta). Sci. China Phys. Mech. 2022, 65, 247011. [Google Scholar] [CrossRef]
  13. Zhai, Y.Q.; Chen, W.P.; Evangelisti, M.; Fu, Z.; Zheng, Y.Z. Gd-based molecular coolants: Aggregating for better magnetocaloric effect. Aggregate 2024, e520. [Google Scholar] [CrossRef]
  14. Wang, L.; Ouyang, Z.; Li, Z.; Cao, J.; Xia, Z. Large magnetocaloric effect in Gd2Si2O7 and plateau-like magnetic entropy change in Dy2Si2O7. J. Alloys Compd. 2023, 969, 172402. [Google Scholar] [CrossRef]
  15. Manchón-Gordón, A.F.; Blázquez, J.S.; Kowalczyk, M.; Ipus, J.J.; Kulik, T.; Conde, C.F. Effect of thermal treatments below devitrification temperature on the magnetic and magnetocaloric properties in mechanically alloyed Fe70Zr30 powders. J. Non-Cryst. Solids 2023, 609, 122267. [Google Scholar] [CrossRef]
  16. Manchón-Gordón, A.F.; López-Martín, R.; Vidal-Crespo, A.; Ipus, J.J.; Blázquez, J.S.; Conde, C.F.; Conde, A. Distribution of transition temperatures in magnetic transformations: Sources, effects and procedures to extract information from experimental data. Metals 2020, 10, 226. [Google Scholar] [CrossRef]
  17. Zhang, Z.Y.; Tang, Q.; Wang, F.C.; Zhang, H.Y.; Zhou, Y.X.; Xia, A.; Li, H.L.; Chen, S.S.; Li, W.H. Tailorable magnetocaloric effect by Fe substitution in Gd-(Co, Fe) amorphous alloy. Intermetallics 2019, 111, 106500. [Google Scholar] [CrossRef]
  18. Tang, B.Z.; Huang, L.W.; Song, M.N.; Ding, D.; Wang, X.; Xia, L. Compositional dependence of magnetic and magnetocaloric properties of the Gd-Ni binary amorphous alloys. J. Non-Cryst. Solids 2019, 522, 119589. [Google Scholar] [CrossRef]
  19. Wang, X.; Tang, B.Z.; 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]
  20. Anatoly, G.K.; Sergey, P.P.; Roman, D.M.; Alexey, V.L.; Aleksey, S.V.; Vasilii, S.G.; Mari, Y.Y. Large magnetic entropy change in GdRuSi optimal for magnetocaloric liquefaction of nitrogen. Metals 2023, 13, 290. [Google Scholar] [CrossRef]
  21. Duc, N.T.M.; Shen, H.X.; Thiabgoh, O.; Huong, N.T.; Sun, J.F.; Phan, M.H. Melt-extracted Gd73.5Si13B13.5/GdB6 ferromagnetic/antiferromagnetic microwires with excellent magnetocaloric properties. J. Alloys Compd. 2020, 818, 153333. [Google Scholar] [CrossRef]
  22. Wang, X.; Wang, Q.; Tang, B.Z.; Yu, P.; Xia, L.; Ding, D. Large magnetic entropy change and adiabatic temperature rise of a ternary Gd34Ni33Al33 metallic glass. J. Rare Earths 2021, 39, 998. [Google Scholar] [CrossRef]
  23. Uporova, S.; Bykova, V.; Uporova, N. Magnetocaloric effect in Gd60Al25(NiCo)15 bulk metallic glass. J. Non-Cryst. Solids 2019, 521, 119506. [Google Scholar] [CrossRef]
  24. Pecharsky, V.K.; Gschneidner, K.A., Jr. Effect of alloying on the giant magnetocaloric effect of Gd5 (Si2Ge2). J. Magn. Magn. Mater. 1997, 167, L179–L184. [Google Scholar] [CrossRef]
  25. Elham, S.; Baran, S.; Yonghui, Z.; Piotr, B.; Jürgen, E. Fabrication of stainless-steel microfibers with amorphous-nanosized microstructure with enhanced mechanical properties. Sci. Rep. 2022, 12, 10784. [Google Scholar] [CrossRef]
  26. Zhukova, V.; Ipatov, M.; Talaat, A.; Blanco, J.M.; Churyukanova, M.; Zhukov, A. Effect of stress annealing on magnetic properties and GMI effect of Co-rich and Fe-rich microwires. J. Alloys Compd. 2017, 707, 189–194. [Google Scholar] [CrossRef]
  27. Amirabadizadeh, A.; Mardani, R.; Ghanaatshoar, M. The effect of current frequency and magnetic field direction in alternative current-field annealing on the GMI and magnetic properties of Co-based wires. J. Alloys Compd. 2016, 661, 501–507. [Google Scholar] [CrossRef]
  28. Corte-León, P.; Zhukova, V.; Ipatov, M.; Blanco, J.M.; Gonzalez, J.; Zhukov, A. Engineering of magnetic properties of Co-rich microwires by joule heating. Intermetallics 2019, 105, 92–98. [Google Scholar] [CrossRef]
  29. Gonzalez-Legarreta, L.; Corte-Leon, P.; Zhukova, V.; Ipatov, M.; Blanco, J.M.; Churyukanova, M.; Taskaev, S.; Zhukov, A. Route of magnetoimpedance and domain walls dynamics optimization in Co-based microwires. J. Alloys Compd. 2020, 830, 154576. [Google Scholar] [CrossRef]
  30. Pan, L.; Ali, A.; Wang, Y.F.; Zheng, Z.M.; Lv, Y.F. Characterization of effects of heat treated anodized film on the properties of hygrothermally aged AA5083-based fiber-metal laminates. Compos. Struct. 2017, 167, 112–122. [Google Scholar] [CrossRef]
  31. Gao, G.E.; Li, H.X.; Jiao, Z.B.; Wu, Y.; Chen, Y.H.; Yu, T.; Lu, Z.P. Effects of nanocrystal formation on the soft magnetic properties of Fe-based bulk metallic glasses. Appl. Phys. Lett. 2011, 99, 5. [Google Scholar] [CrossRef]
  32. Zhang, R.C.; Qian, M.F.; Waske, A.; Shen, H.X.; Zhang, X.X. Investigating the microstructure and magnetic properties of La-Fe-Si microwires during fabrication and heat treatment process. J. Alloys Compd. 2019, 794, 153–162. [Google Scholar] [CrossRef]
  33. Luo, Q.; Wang, W.H. Magnetocaloric effect in rare earth-based bulk metallic glasses. J. Alloys Compd. 2010, 495, 209–216. [Google Scholar] [CrossRef]
  34. Yao, Y.Y.; Li, Z.; Liu, J.S.; Zhang, M.W.; Zhang, Y.; Wang, F.; Liu, R.; Shen, H.X. Effect of Ni alloying on the microstructure and magnetocaloric properties of Gd-based metallic microfibers. J. Alloys Compd. 2023, 961, 170979. [Google Scholar] [CrossRef]
  35. 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]
  36. Phan, M.H.; Yu, S.C. Review of the magnetocaloric effect in manganite materials. J. Magn. Magn. Mater. 2007, 308, 325–340. [Google Scholar] [CrossRef]
  37. Xue, L.; Shao, L.L.; Luo, Q.; Shen, B.L. Gd25RE25Co25Al25 (RE = Tb, Dy and Ho) high-entropy glassy alloys with distinct spin-glass behavior and good magnetocaloric effect. J. Alloys Compd. 2019, 790, 633–639. [Google Scholar] [CrossRef]
  38. Xue, L.; Li, J.; Yang, W.M.; Yuan, C.C.; Shen, B.L. Effect of Fe substitution on magnetocaloric effects and glass-forming ability in Gd-based metallic glasses. Intermetallics 2018, 93, 67–71. [Google Scholar] [CrossRef]
  39. Liu, J.S.; Wang, Q.X.; Wu, M.J.; Zhang, Y.; Shen, H.X.; Ma, W. Improving the refrigeration capacity of Gd-rich wires through Fe-doping. J. Alloys Compd. 2017, 711, 71–76. [Google Scholar] [CrossRef]
Figure 1. Heat-treatment technique for Gd-based metallic microfibers and samples’ appearance.
Figure 1. Heat-treatment technique for Gd-based metallic microfibers and samples’ appearance.
Metals 14 00874 g001
Figure 2. XRD patterns and DSC curves of Gd-based metallic microfibers: (a) XRD patterns; (b) DSC heating curve; (c) DSC cooling curve.
Figure 2. XRD patterns and DSC curves of Gd-based metallic microfibers: (a) XRD patterns; (b) DSC heating curve; (c) DSC cooling curve.
Metals 14 00874 g002
Figure 3. SEM morphology and EDS analysis of Gd-based metallic microfibers before and after annealing process: (a,b) as-prepared; (c,d) annealing @ 300 °C; (e,f) annealing @ 340 °C; (g,h) annealing @ 380 °C. Among these images, the cross symbol represents the differ EDS spots.
Figure 3. SEM morphology and EDS analysis of Gd-based metallic microfibers before and after annealing process: (a,b) as-prepared; (c,d) annealing @ 300 °C; (e,f) annealing @ 340 °C; (g,h) annealing @ 380 °C. Among these images, the cross symbol represents the differ EDS spots.
Metals 14 00874 g003
Figure 4. TEM analysis of metallic microfibers before and after annealing process: (a) TEM morphology of as-prepared; (b) TEM morphology of annealing @ 380 °C; (c) SEAD of as-prepared; (d) SEAD of annealing @ 380 °C; (e) HRTEM, FFT, IFFT, and ACF images of as-prepared; (f) HRTEM, FFT, IFFT, and ACF images of annealing @ 380 °C. Among these images, the red boxes indicate the sub-images of approximate diffraction spot patterns, and the blue boxes indicate the sub-images of approximate diffraction halo patterns.
Figure 4. TEM analysis of metallic microfibers before and after annealing process: (a) TEM morphology of as-prepared; (b) TEM morphology of annealing @ 380 °C; (c) SEAD of as-prepared; (d) SEAD of annealing @ 380 °C; (e) HRTEM, FFT, IFFT, and ACF images of as-prepared; (f) HRTEM, FFT, IFFT, and ACF images of annealing @ 380 °C. Among these images, the red boxes indicate the sub-images of approximate diffraction spot patterns, and the blue boxes indicate the sub-images of approximate diffraction halo patterns.
Metals 14 00874 g004
Figure 5. General magnetization curves of Gd-based metallic microfibers before and after annealing process (M-H).
Figure 5. General magnetization curves of Gd-based metallic microfibers before and after annealing process (M-H).
Metals 14 00874 g005
Figure 6. Hysteresis loops of Gd-based metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 340 °C; (d) annealing @ 380 °C.
Figure 6. Hysteresis loops of Gd-based metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 340 °C; (d) annealing @ 380 °C.
Metals 14 00874 g006
Figure 7. M-T curves of metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C.
Figure 7. M-T curves of metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C.
Metals 14 00874 g007
Figure 8. Isothermal magnetization curves of Gd-based metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C.
Figure 8. Isothermal magnetization curves of Gd-based metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C.
Metals 14 00874 g008
Figure 9. Arrott curves of metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C.
Figure 9. Arrott curves of metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C.
Metals 14 00874 g009
Figure 10. Magnetic entropy change curves of Gd-based metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C; (d) magnetic entropy change curves of microfibers before and after annealing process when foreign field is 5 T.
Figure 10. Magnetic entropy change curves of Gd-based metallic microfibers before and after annealing process: (a) as-prepared; (b) annealing @ 300 °C; (c) annealing @ 380 °C; (d) magnetic entropy change curves of microfibers before and after annealing process when foreign field is 5 T.
Metals 14 00874 g010
Figure 11. RC and RCP of metallic microfibers before and after annealing process: (a) refrigeration capacity curve; (b) relative cooling capacity curve.
Figure 11. RC and RCP of metallic microfibers before and after annealing process: (a) refrigeration capacity curve; (b) relative cooling capacity curve.
Metals 14 00874 g011
Figure 12. Schematic diagram of the effect of annealing on magnetic properties of Gd-based metallic microfibers: (a) atomic spin after annealing processing; (b) variation in magnetic moment; (c) arrangement of atoms and domain distribution in the microfibers before and after annealing processing; (d) magnetic moment rotation process based on energy minimization principle.
Figure 12. Schematic diagram of the effect of annealing on magnetic properties of Gd-based metallic microfibers: (a) atomic spin after annealing processing; (b) variation in magnetic moment; (c) arrangement of atoms and domain distribution in the microfibers before and after annealing processing; (d) magnetic moment rotation process based on energy minimization principle.
Metals 14 00874 g012
Table 1. Statistics for thermophysical parameters of Gd-based metallic microfibers.
Table 1. Statistics for thermophysical parameters of Gd-based metallic microfibers.
ParameterTg (°C)Tx1 (°C)Tx2 (°C)ΔT (°C)−∆H (J·g−1)
As-Prepared3033464874430.5
Table 2. Statistics for magnetocaloric parameters of Gd-based alloys.
Table 2. Statistics for magnetocaloric parameters of Gd-based alloys.
Compositions−ΔSmmax (J/kg·K)RC (J/kg)RCP (J/kg)Refs.
Gd57Al19Co19Ni511.571138.16834.14This work
Gd70Co10Al206.36538-[4]
Gd36Tb20Co20Al2412.36417-[8]
Gd3SbO73.54--[12]
Gd70Co305.13-277[17]
Gd70Co20Fe102.78-353[17]
Gd71Ni29-804.3-[19]
Gd73.5Si13B13.56.4790885[21]
Gd34Ni33Al3311.06--[22]
Gd60Al25(NiCo)156.31-890[23]
Gd55Co15Al24Si1Fe57.26857-[33]
Gd60Al20Co2010.12698.28936.19[34]
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

Liu, J.; Yu, S.; Zhang, M.; Li, Z.; Cui, Y. Effect of Annealing Temperature on Microstructure and Magnetocaloric Properties of Gd-Based Metallic Microfibers. Metals 2024, 14, 874. https://doi.org/10.3390/met14080874

AMA Style

Liu J, Yu S, Zhang M, Li Z, Cui Y. Effect of Annealing Temperature on Microstructure and Magnetocaloric Properties of Gd-Based Metallic Microfibers. Metals. 2024; 14(8):874. https://doi.org/10.3390/met14080874

Chicago/Turabian Style

Liu, Jingshun, Shiyang Yu, Mingwei Zhang, Ze Li, and Yaqiang Cui. 2024. "Effect of Annealing Temperature on Microstructure and Magnetocaloric Properties of Gd-Based Metallic Microfibers" Metals 14, no. 8: 874. https://doi.org/10.3390/met14080874

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop