Next Article in Journal
Dissimilar Laser Welding of a NiTi Shape Memory Alloy to Ti2AlNb
Next Article in Special Issue
Effect of Strain Rate and Extrinsic SIZE Effect on Micro-Mechanical Properties of Zr-Based Bulk Metallic Glass
Previous Article in Journal
Anodized Biomedical Stainless-Steel Mini-Implant for Rapid Recovery in a Rabbit Model
Previous Article in Special Issue
Influence of Cold Rolling Process and Chemical Composition on the Mechanical Properties and Corrosion Behavior of Zr-Based Metallic Glasses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Mn Substitution Fe on the Formability and Magnetic Properties of Amorphous Fe88Zr8B4 Alloy

1
College of Engineering, Shanghai Polytechnic University, Shanghai 201209, China
2
Institute of Materials, Shanghai University, Shanghai 200072, China
3
Center for Advanced Microanalysis, Shanghai University, Shanghai 200444, China
*
Authors to whom correspondence should be addressed.
Metals 2021, 11(10), 1577; https://doi.org/10.3390/met11101577
Submission received: 27 August 2021 / Revised: 22 September 2021 / Accepted: 28 September 2021 / Published: 3 October 2021
(This article belongs to the Special Issue Forming Ability and Properties of Bulk Metallic Glasses)

Abstract

:
Elemental substitution is commonly used to improve the formability of metallic glasses and the properties of amorphous alloys over a wide compositional range. Therefore, it is essential to investigate the influence of element content change on the formability as well as magnetic and other properties. The purpose is to achieve tailorable properties in these alloys with enhanced glass forming ability. In this work, the glass-forming ability (GFA) and magnetic properties of the minor Mn-substituted Fe88Zr8B4 amorphous alloy were investigated. The addition of Mn improving the amorphous forming ability of the alloy. With the addition of Mn, the magnetic transition temperature, saturation magnetization and the magnetic entropy changes (−ΔSm) peaks decreased simultaneously, which is possibly caused by the antiferromagnetic coupling between Fe and Mn atoms. The dependence of −ΔSmpeak on Tc displays a positive correlation compared to the −ΔSmpeak- Tc−2/3 relationship proposed by Belo et al.

1. Introduction

With the deterioration of the global energy crisis and environmental pollution, it has become more and more urgent to seek clean energy, renewable energy and energy-saving technologies in recent years [1,2]. The traditional refrigeration system using gas compression/expansion technology is a large energy consumer, low efficiency and not environmentally friendly due to ozone-depleting chlorofluorocarbons. The magnetic refrigerator using a solid refrigerant is more energy efficient (up to 30% reduction of energy loss), more environmentally friendly (free of freon) and has a larger energy density (conducive to miniaturization) when compared to the freon compression machine [2,3]. Therefore, it is essential to find materials with excellent magnetocaloric properties because magnetic refrigeration technology is developed based on the magnetocaloric effect (MCE) of these magnetic refrigerants [4].
The MCE characteristics of the prepared materials so far are usually evaluated by two factors: maximum magnetic entropy change (−ΔSmpeak) and refrigeration capacity (RC). Intermetallic compounds, such as Gd5(SixGe1−x)4, La(Fe,M)13 (M = Si,Co,Al), MnFe(P1−xAsx), NiMnGa and LaMnO3 [5,6,7,8,9,10], exhibit a giant −ΔSmpeak but relatively low RC due to their sharp and narrow magnetic entropy change (−ΔSm) peaks. Additionally, the MCE of the intermetallic compounds is closely related to phase structure, which means the Curie temperature (Tc) of these intermetallic refrigerants can hardly be tuned without deteriorating their −ΔSmpeak, just like the tunable mechanical properties in the Zr-Cu system [11]. On the other hand, metallic glasses, such as Gd-based and Fe-based amorphous alloys [12,13,14,15,16,17,18,19,20,21,22,23,24,25], usually exhibit huge RC but a relatively low −ΔSmpeak due to their broad −ΔSm humps. In addition, the Tc and −ΔSmpeak of these amorphous alloys are tailorable by randomly selecting the composition of the alloys within a certain composition range, such as the compositional-induced structural change in the ZrNi system [26].
It has been found that the constant −ΔSm peak in a long temperature range above and below the Curie temperature, namely, a table-like −ΔSm profile, is the most ideal form for magnetic refrigerants working in an Ericsson cycle [16,17,18,19,20]. Obviously, an approximately flat −ΔSm profile is difficult to achieve in intermetallic compounds not only because of their narrow −ΔSm peak but also due to the difficulty of tuning their Tc without dramatically decreasing their −ΔSmpeak. In contrast, a table-like −ΔSm profile can be easily achieved in amorphous composites with an appropriate Curie temperature and −ΔSm peak due to their tunable Tc and −ΔSmpeak over a wide compositional range.
Microalloying and elemental substitution are the usual methods used to improve the formability of metallic glasses and to change the properties of amorphous alloys over a wide compositional range [14,15,19,20,21], and therefore, they are very important for achieving tailorable properties in these based alloys, accompanied by better glass-forming ability. In this study, we prepared Fe88Zr8B4Mnx as-spun ribbons by adding Mn to lower the Fe content in based Fe88Zr8B4 samples. The glass formability and magnetic properties of the Fe88Zr8B4Mnx metallic glasses were studied in comparison with those of the Fe88Zr8B4 glassy alloy. Based on these results, the effect of Mn alloying on the formability and MCE of the original base alloys, accompanied by the interaction mechanism between the internal elements, were investigated.

2. Materials and Methods

A series of button-shaped metal samples with element compositions of Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) were prepared by heating the high-purity raw materials until fully integrated at least four times in a non-consumable electrode high vacuum arc melting furnace that filled with high purity argon after the oxygen absorption protection operation of the titanium ingot. The Fe88−xZr8B4Mnx ribbons (typically 2–3 mm wide and ~40 μm thick) were prepared in a protective atmosphere of argon by ejecting the melts from the quartz tube on a rotating copper wheel under a pure Ar atmosphere. The structural characteristics of the obtained alloys were ascertained by means of X-ray diffraction (XRD) on a Rigaku diffractometer (model D/max-2550, Tokyo, Japan) using Kα radiation of copper, and the thermodynamic parameters were tested on the differential scanning calorimetry (DSC) on a NETZSCH calorimeter (model 404C, Selb, Germany) with the heating curves obtained at 0.333 K/s. Magnetic measurements were performed on the vibrating sample magnetometer (VSM) of a Physical Properties Measurement System (PPMS model 6000, Quantum Design, San Diego, CA, USA), and the magnetization vs. temperature and the isothermal magnetization curves were obtained.

3. Results and Discussion

3.1. Descriptive of the Experimental Results

The preliminary structure verification information of the prepared Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) ribbons characterized by the broadened diffraction hump at about 43°, as shown in the curves in the XRD patterns in Figure 1, roughly explains the amorphous trait of the samples. From another point of view, the glassy feature, shown in Figure 2a, was further ascertained on the DSC curves that displayed a clear endothermic glass transition point that appeared before the crystallization peak. Figure 2b illustrates the melting behaviors of the Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) alloys. The glass transition temperature (Tg), crystallization temperature (Tx) and liquid temperature (Tl) of the glassy ribbons obtained from their DSC traces are listed in Table 1. For comparison, in Table 1 the thermodynamic properties of the Fe88Zr8B4 glassy ribbon have been recalculated. As such, the reduced glass transition temperature (Trg = Tg/Tl) and the parameter γ (= Tx/(Tg + Tl)), both of which are usually applied as the major reference for the glass formability (GFA) of the alloys [27,28], can be obtained accordingly to be 0.503 and 0.353 for Fe86Zr8B4Mn2, 0.506 and 0.355 for Fe83Zr8B4Mn5, 0.517 and 0.353 for Fe80Zr8B4Mn8 and 0.512 and 0.356 for Fe78Zr8B4Mn10. The GFA of the Fe88Zr8B4 glassy alloy was generally enhanced by Mn substitution for Fe, which agrees well with the multicomponent rule for glass forming [4]. The glass formation enthalpy (ΔHamor) of the Fe88−xZr8B4Mnx (x = 0, 2, 5, 8, 10) alloys were calculated using Miedema′s model, which can be used as a reference to investigate the mechanism of alloy’s GFA in more detail by comparing the internal bonding changes after the addition of new elements [28]. The calculated ΔHamor shows a decreasing trend with a value of about −2.85 kJ/mol for x = 0, −3.26 kJ/mol for x = 2, −3.84 kJ/mol for x = 5, −4.68 kJ/mol for x = 8 and −5.13 kJ/mol for x = 10. As the formability of metallic glass is closely related to ΔHamor, the enhanced GFA by Mn addition is most likely attributed to a decrease in ΔHamor as the Mn content increases. The composition of alloys was ascertained by chemical analysis. In order to ascertain the homogeneity of the samples, we selected an amorphous ribbon with lower glass formability, that is, the Fe78Zr8B4Mn10 ribbon, for HRTEM examination. Figure 3 shows the HRTEM image of the Fe78Zr8B4Mn10 ribbon. The ribbon is homogeneous with a disordered atomic configuration.
The relationship between magnetization (M) and temperature (T) of the amorphous Fe88−xZr8B4Mnx samples was measured, respectively, from 200 to 380K after the operation by a zero-field cooling method. By performing a derivative process on the curves in Figure 4, the Curie temperature of a series of Fe88Zr8B4 samples with Mn addition was found to be 283 K for x = 2, 263 K for x = 5, 238 K for x = 8 and 225 K for x = 10. The Tc of the metallic glasses with Mn added, as seen in the embedded image in the upper right corner of Figure 4, had a decreasing trend compared with the Fe88Zr8B4 samples. It is known that the magnetic properties of Fe-based amorphous alloys exhibit a close correlation with the direct 3d interactions of Fe atoms [19,20,21,22,23]. As with the situation shown in the FeZrB series amorphous alloys, although the reduction of Fe content reduces the number of interactions between 3d atoms, the introduction of other atoms can enhance the 3d interaction between Fe to some extent and thereby enhance the magnetic properties of the FeZrB amorphous alloys. A completely different trend of the Mn addition on the Fe-Zr-B amorphous alloys, however, induced antiferromagnetic coupling between Fe and Mn atoms [29], which deteriorated the Fe-Fe interactions and therefore led to a decrease in the Curie temperature of the Fe88Zr8B4 metallic glass.
Figure 5 shows the hysteresis loops of the amorphous Fe88−xZr8B4Mnx (x = 0, 2, 5, 8 and 10) samples measured at 200 K under 5 Tesla. The ribbons are all soft magnetic and exhibit negligible coercivity, indicating that the new series of FeZrB alloys is similar to the based alloys that are potential materials for practical application because they can easily be magnetized and demagnetized. The saturation magnetization (Ms) of the Fe88−xZr8B4Mnx glassy ribbons is about 109 Am2/kg for x = 0, 94.6 Am2/kg for x = 2, 80.3 Am2/kg for x = 5, 65.4 Am2/kg for x = 8 and 58.33 Am2/kg for x = 10. To facilitate observation, Figure 5 plots the variation curve of Ms and Mn content inset in the lower right corner. The Ms of the samples decreased dramatically as the Mn content increased, which indicates reduced magnetic moments of the glassy samples induced by antiferromagnetic coupling between Fe and Mn atoms.
Considering that the saturation magnetization is closely related to the magnetic moment of the material, the reduced magnetic moments by Mn addition may also result in the deteriorated magnetocaloric properties of the glassy Fe88−xZr8B4Mnx samples. By measuring the isothermal M-H curves (with H = 5 T in the present work) of different samples, a series of different Arrott plots (converted from the M-H curves) under different temperatures was obtained to investigate the magnetic phase change character. The Arrott plots of the Fe88−xZr8B4Mnx (x = 2, 5, 8 and 10) glassy samples, as shown in Figure 6, display a positive slope with a nearly C-shape, which demonstrates a second-order magnetic phase transition (MPT) during the ferromagnetic–paramagnetic transition in all the samples [30].
According to Maxwell’s relations, we can calculate the magnetic entropy change (−ΔSm) of these amorphous alloys.
Δ S m ( T , H ) = S m ( T , H ) S m ( T , 0 ) = 0 H ( M H ) H d H
The −ΔSm-T curves of the amorphous Fe88−xZr8B4Mnx samples are plotted in Figure 7: (a) for Fe86Zr8B4Mn2, (b) for Fe83Zr8B4Mn5, (c) for Fe80Zr8B4Mn8 and (d) for Fe78Zr8B4Mn10 to further study the MCE of the alloys. The samples exhibited a typical lambda shape with a maximum in the vicinity of Tc. The wide −ΔSm peak of the series of Fe88−xZr8B4Mnx glassy ribbons is in accordance with the characteristics of the second-order MPT [12,13,14,19,20,21]. The −ΔSmpeak of the Fe88−xZr8B4Mnx glassy ribbons that increased from 1 to 5T is displayed in Figure 7. As shown in Table 2, the MCE behavior of the Fe88−xZr8B4Mnx samples can be described by the −ΔSmHn relationship [24,25]. Fitting the relationship curve of the −ΔSm and H after logarithmic processing, the n exponent of the Fe88−xZr8B4Mnx alloys under different temperatures was ascribed from the fitted curves, as the value plotted in the n-T curves in Figure 7e. All the amorphous samples displayed the typical magnetocaloric behaviors of soft magnetic metallic glasses: n is nearly 1 at low temperatures when the sample is ferromagnetic, then gradually reduced to a minimum value near Tc and finally increased dramatically to a value up to 2 at the paramagnetic range. The n in the vicinity of Tc, as listed in Table 2, is about 0.769 for Fe88Zr8B4, 0.745 for Fe86Zr8B4Mn2, 0.736 for Fe83Zr8B4Mn5, 0.740 for Fe80Zr8B4Mn8 and 0.736 for Fe78Zr8B4Mn10, all of which are roughly in accordance with the ones of other fully amorphous alloys [13,14,20,21,24,25,31,32].
As predicted above, adding the antiferromagnetic element Mn created a new antiferromagnetic coupling and reduced the overall magnetic moments, and thus, deterioration the −ΔSmpeak of the Fe88−xZr8B4Mnx metallic glasses is mostly due to the reduced magnetic moments caused by adding the antiferromagnetic element Mn. Figure 8 shows the −ΔSmpeak vs Tc−2/3 plots at each magnetic field for the Fe88-xZr8B4Mnx amorphous samples. Combining the results obtained in other Fe-based metallic glasses [3,18,19,20,21,22,23,24,25], the dependence of −ΔSmpeak on Tc displays a positive correlation in the Fe-Zr-B-based metallic glasses, which is contrary to the −ΔSmpeak-Tc−2/3 relationship proposed by Belo et al. [12,13,14,15,16,33]. This is probably because some magnetic parameters including Tc and magnetization in Fe-based metallic glasses are dominated by the direct coupling between the Fe-based elements. This factor, which enhances the interaction between Fe atoms, simultaneously improves the Curie temperature, Ms and −ΔSmpeak, and vice versa. In contrast, the factors in RE-based or containing amorphous alloys are more complicated in view of the complex interaction of 4f layer electrons of RE with other TM electrons. Overall, the combination of the direct interaction between TM atoms and the indirect interactions between RE-RE and RE-TM atoms results in the complex magnetic performance of the RE-TM-based metallic glasses.

3.2. Discussion

As predicted above, the Fe88−xZr8B4Mnx ribbon is homogeneous with a disordered atomic configuration. Adding the antiferromagnetic element Mn creates a new antiferromagnetic coupling and reduces the overall magnetic moments, and thus, the deterioration of the −ΔSmpeak of the Fe88-xZr8B4Mnx metallic glasses is mostly due to the reduced magnetic moments caused by adding the antiferromagnetic element Mn. Figure 7 shows the −ΔSmpeak vs Tc−2/3 plots at each magnetic field for the Fe88−xZr8B4Mnx amorphous samples. Combining the results obtained in other Fe-based metallic glasses [3,18,19,20,21,22,23,24,25], the dependence of −ΔSmpeak on Tc displays a positive correlation in the Fe-Zr-B-based metallic glasses, which is contrary to the −ΔSmpeak-Tc−2/3 relationship proposed by Belo et al. [12,13,14,15,16,33]. This is probably because some magnetic parameters including Tc and magnetization in Fe-based metallic glasses are dominated by the direct coupling between the Fe-based elements. This factor, which enhances the interaction between Fe atoms, simultaneously improves the Curie temperature, Ms and −ΔSmpeak, and vice versa. In contrast, the factors in RE-based or containing amorphous alloys are more complicated in view of the complex interaction of 4f layer electrons of RE with other TM electrons. Overall, the combination of the direct interaction between TM atoms and the indirect interactions between RE-RE and RE-TM atoms results in the complex magnetic performance of the RE-TM-based metallic glasses.

3.3. Discussion

As predicted above, the adding of antiferromagnetic element Mn creates a new antiferromagnetic coupling and reduces the overall magnetic moments, and thus the deteriorates of the −ΔSmpeak of the Fe88−xZr8B4Mnx metallic glasses is mostly due to the reduced magnetic moments by the adding of antiferromagnetic element Mn. Figure 7 shows the −ΔSmpeak vs. Tc−2/3 plots at each magnetic field for the Fe88−xZr8B4Mnx amorphous samples. Combining the results obtained in other Fe-bases metallic glasses [3,18,19,20,21,22,23,24,25], the dependence of −ΔSmpeak on Tc displays a positive correlation in the Fe-Zr-B-based metallic glasses which is contrary to the −ΔSmpeak-Tc−2/3 relationship proposed by Belo et al. [12,13,14,15,16,33]. This is probably because that some magnetic parameters including Tc and magnetization in Fe-based metallic glasses are dominated by the direct coupling between the based elements Fe. The factor, which enhances the interaction between Fe atoms, will simultaneously improve the Curie temperature, Ms and −ΔSmpeak, and vice versa. In contrast, the factors in RE based or containing amorphous alloys are more complicated in view of the complex interaction of 4f layer electrons of RE with other TM electrons. all above, the combination of the direct interaction between TM atoms, indirect interactions between RE-RE and RE-TM atoms makes the complex magnetic performance of the RE-TM based metallic glasses.

4. Conclusions

In summary, we obtained the amorphous Fe88−xZr8B4Mnx (x = 2, 5, 8,10) ribbons using the single roller melt spinning method with an average thickness of about 30~40 μm. The metal glass-forming performance of the Fe88Zr8B4 alloy was enhanced by substituting Mn for Fe, which is supposed to be related to the alloys, and it was enhanced by the decrease in ΔHamor with increasing Mn content. To investigate the influence mechanism of Mn substitution on the MCE of the Fe88Zr8B4 amorphous alloy, the magnetocaloric properties of the Fe88−xZr8B4Mnx samples and Fe88Zr8B4 amorphous alloy were systematically analyzed. The Curie temperature, Ms and −ΔSmpeak were found to decrease simultaneously with the addition of Mn. It is supposed that substituting Mn for Fe, which induced the antiferromagnetic coupling between Fe and Mn atoms, deteriorated the Fe-Fe interactions and therefore lead to a decrease in the Tc, Ms and −ΔSmpeak of the Fe88Zr8B4 amorphous alloys. The exponent n (−ΔSmHn) in the vicinity of Tc, or the other magnetic state of the Fe88-xZr8B4Mnx samples, is similar to the other fully amorphous alloys, which indicates the second-order magnetic phase transition and the amorphous structure of the series amorphous alloys. The dependence of −ΔSmpeak on Tc displayed a positive correlation in the Fe-based amorphous alloys due to the direct coupling between the Fe-based elements, which is different from the relationship proposed by Belo et al. in rare-earth-containing alloys, in view of the complex interaction of 4f layer electrons of RE with other TM electrons.

Author Contributions

Conceptualization, L.X.; methodology, Q.W.; investigation, X.W.; writing—original draft preparation, B.T.; writing—review and editing, D.D.; measurement, L.C.; funding acquisition, L.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Nature Science Foundation of China (Grant Nos. 51271103 and 51671119), the Shanghai Polytechnic University 2021 Youth Academic Backbone Training Project (EGD21QD17) and the Shanghai Educational Science Research Project (C2021062).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This research was provided technical support by the Center for Advanced Microanalysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, L.; Chen, Z.-G.; Dargusch, M.; Zou, J. High Performance Thermoelectric Materials: Progress and Their Applications. Adv. Energy Mater. 2018, 8, 1701797. [Google Scholar] [CrossRef]
  2. Franco, V.; Blázquez, J.S.; Ipus, J.J.; Law, J.Y.; Moreno-Ramírez, L.M.; Conde, A. Magnetocaloric effect: From materials re-search to refrigeration devices. Prog. Mater. Sci. 2018, 93, 112–232. [Google Scholar] [CrossRef]
  3. De Oliveira, N.A.; von Ranke, P.J. Theoretical aspects of the magnetocaloric effect. Phys. Rep. 2010, 489, 89–159. [Google Scholar] [CrossRef]
  4. Tishin, A.M.; Spichkin, Y.I. The Magnetocaloric Effect and Its Applications; CRC Press, Taylor and Francis Group, LLC: Boca Raton, FL, USA, 2003. [Google Scholar]
  5. Pecharsky, V.K.; Gschneidner, K.A. Giant Magnetocaloric Effect in Gd5(Si2Ge2). Phys. Rev. Lett. 1997, 78, 4494–4497. [Google Scholar] [CrossRef]
  6. Shen, B.G.; Sun, J.R.; Hu, F.X.; Zhang, H.W.; Cheng, Z.H. Recent Progress in Exploring Magnetocaloric Materials. Adv. Mater. 2009, 21, 4545–4564. [Google Scholar] [CrossRef] [Green Version]
  7. Balli, M.; Fruchart, D.; Gignoux, D. Optimization of La(Fe,Co)13−xSix based compounds for magnetic refrigeration. J. Phys. Condens. Matter 2007, 19, 236230. [Google Scholar] [CrossRef]
  8. Tegus, O.; Bao, L.-H.; Song, L. Phase transitions and magnetocaloric effects in intermetallic compounds MnFeX (X = P, As, Si, Ge). Chin. Phys. B 2013, 22, 037506. [Google Scholar] [CrossRef]
  9. Planes, A.; Mañosa, L.; Acet, M. Magnetocaloric effect and its relation to shape-memory properties in ferromagnetic Heusler alloys. J. Phys. Condens. Matter 2009, 21, 233201. [Google Scholar] [CrossRef] [Green Version]
  10. 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]
  11. Ghidelli, M.; Orekhov, A.; Bassi, A.L.; Terraneo, G.; Djemia, P.; Abadias, G.; Nord, M.; Béché, A.; Gauquelin, N.; Verbeeck, J.; et al. Novel class of nanostructured metallic glass films with superior and tunable mechanical properties. Acta Mater. 2021, 213, 116955. [Google Scholar] [CrossRef]
  12. Luo, Q.; Wang, W.H. Magnetocaloric effect in rare earth-based bulk metallic glasses. J. Alloys Compd. 2010, 495, 209–216. [Google Scholar] [CrossRef]
  13. 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]
  14. 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]
  15. Tang, B.; Liu, X.; Li, D.; Yu, P.; Xia, L. Effect of Ni substitution on the formability and magnetic properties of a Gd50Co50 amorphous alloy. Chin. Phys. B 2020, 29, 056401. [Google Scholar] [CrossRef]
  16. Ma, L.Y.; Gan, L.H.; Chan, K.C.; Ding, D.; Xia, L. Achieving a table-like magnetic entropy change across the ice point of wa-ter with tailorable temperature range in Gd-Co-based amorphous hybrids. J. Alloys Compd. 2017, 723, 197–200. [Google Scholar] [CrossRef]
  17. 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]
  18. Lai, J.; Zheng, Z.; Zhong, X.; Franco, V.; Montemayor, R.; Liu, Z.; Zeng, D. Table-like magnetocaloric effect of Fe88−xNdxCr8B4 composite materials. J. Magn. Magn. Mater. 2015, 390, 87–90. [Google Scholar] [CrossRef]
  19. Chen, L.S.; Zhang, J.Z.; Wen, L.; Yu, P.; Xia, L. Outstanding magnetocaloric effect of Fe88−xZr8B4Smx (x = 0, 1, 2, 3) amorphous alloys. Sci. China Phys. Mech. Astro. 2018, 61, 056121. [Google Scholar] [CrossRef]
  20. Gan, L.H.; Ma, L.Y.; Tang, B.Z.; Ding, D.; Xia, L. Effect of Co substitution on the glass forming ability and magnetocaloric effect of Fe88Zr8B4 amorphous alloys. Sci. China Phys. Mech. Astron. 2017, 60, 076121. [Google Scholar] [CrossRef]
  21. Yu, P.; Zhang, J.Z.; Xia, L. Effect of boron on the magneto-caloric effect in Fe91−xZr9Bx (x = 3, 4, 5) amorphous alloys. J. Mater. Sci. 2017, 52, 13948–13955. [Google Scholar] [CrossRef]
  22. Wang, Y.; Bi, X. The role of Zr and B in room temperature magnetic entropy change of FeZrB amorphous alloys. Appl. Phys. Lett. 2009, 95, 262501. [Google Scholar] [CrossRef]
  23. Mishra, D.; Gurram, M.; Reddy, A.; Perumal, A.; Saravanan, P.; Srinivasan, A. Enhanced soft magnetic properties and mag-netocaloric effect in B substituted amorphous Fe–Zr alloy ribbons. Mater. Sci. Eng. B. 2010, 175, 253–260. [Google Scholar] [CrossRef]
  24. Franco, V.; Borrego, J.M.; Conde, C.F.; Conde, A.; Stoica, M.; Roth, S. Refrigerant capacity of FeCrMoCuGaPCB amor-phous alloys. J. Appl. Phys. 2006, 100, 083903. [Google Scholar] [CrossRef] [Green Version]
  25. Franco, V.; Blázquez, J.S.; Conde, A. The influence of Co addition on the magnetocaloric effect of Nanoperm-type amor-phous alloys. J. Appl. Phys. 2006, 100, 064307. [Google Scholar] [CrossRef]
  26. Ghidelli, M.; Gravier, S.; Blandin, J.-J.; Pardoen, T.; Raskin, J.-P.; Mompiou, F. Compositional-induced structural change in ZrxNi100−x thin film metallic glasses. J. Alloys Compd. 2014, 615, S348–S351. [Google Scholar] [CrossRef]
  27. Turnbull, D. Under what conditions can a glass be formed? Contemp. Phys. 1969, 10, 473–488. [Google Scholar] [CrossRef]
  28. Lu, Z.P.; Liu, C.T. Glass Formation Criterion for Various Glass-Forming Systems. Phys. Rev. Lett. 2003, 91, 115505. [Google Scholar] [CrossRef]
  29. Takeuchi, A.; Inoue, A. Calculations of Amorphous-Forming Composition Range for Ternary Alloy Systems and Analyses of Stabilization of Amorphous Phase and Amorphous-Forming Ability. Mater. Trans. 2001, 42, 1435–1444. [Google Scholar] [CrossRef] [Green Version]
  30. Trhlík, M.; De Moor, P.; Erzinkyan, A.L.; Gurevich, G.M.; Parfenova, V.P.; Schuurmans, P.; Severijns, N.; Vanderpoorten, W.; Vanneste, L.; Wouters, J. Antiferromagnetic coupling between the Mn and Fe nearest neighbours in ferromagnetic Mn:Pt-10 at.% Fe. J. Phys. Condens. Matter 1992, 4, 9181–9190. [Google Scholar] [CrossRef]
  31. Banerjee, S.K. On a generalized approach to first and second order magnetic transitions. Phys. Lett. 1964, 12, 16–17. [Google Scholar] [CrossRef]
  32. Law, J.Y.; Franco, V.; Moreno-Ramírez, L.M.; Conde, A.; Karpenkov, D.Y.; Radulov, I.A.; Skokov, K.; Gutfleisch, O. A quantitative criterion for determining the order of magnetic phase transitions using the magnetocaloric effect. Nat. Commun. 2018, 9, 1–9. [Google Scholar] [CrossRef] [PubMed]
  33. Belo, J.H.; Amaral, J.S.; Pereira, A.M.; Amaral, V.S.; Araújo, J.P. On the Curie temperature dependency of the magneto-caloric effect. Appl. Phys. Lett. 2012, 100, 242407. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) as-spun ribbons.
Figure 1. XRD patterns of the Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) as-spun ribbons.
Metals 11 01577 g001
Figure 2. DSC traces of the Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) alloys with the heating curves obtained at 0.333 K/s. (a) shows the glass transition temperature (Tg), crystallization temperature (Tx) and (b) is the liquid temperature (Tl).
Figure 2. DSC traces of the Fe88−xZr8B4Mnx (x = 2, 5, 8, 10) alloys with the heating curves obtained at 0.333 K/s. (a) shows the glass transition temperature (Tg), crystallization temperature (Tx) and (b) is the liquid temperature (Tl).
Metals 11 01577 g002
Figure 3. The HRTEM image of the Fe78Zr8B4Mn10 amorphous alloy.
Figure 3. The HRTEM image of the Fe78Zr8B4Mn10 amorphous alloy.
Metals 11 01577 g003
Figure 4. The temperature dependence of the magnetization (M-T) curves of the Fe88−xZr8B4Mnx samples, inset the compositional dependence of Tc in Fe88-xZr8B4TMx (TM=Co, Mn) metallic glasses.
Figure 4. The temperature dependence of the magnetization (M-T) curves of the Fe88−xZr8B4Mnx samples, inset the compositional dependence of Tc in Fe88-xZr8B4TMx (TM=Co, Mn) metallic glasses.
Metals 11 01577 g004
Figure 5. The hysteresis loops of the amorphous Fe88−xZr8B4Mnx (x = 0, 2, 5, 8 and 10) samples measured at 200 K under 5 Tesla. The inset shows the relationship between the saturation magnetization (Ms) of the Fe88−xZr8B4Mnx amorphous ribbons and the Mn content.
Figure 5. The hysteresis loops of the amorphous Fe88−xZr8B4Mnx (x = 0, 2, 5, 8 and 10) samples measured at 200 K under 5 Tesla. The inset shows the relationship between the saturation magnetization (Ms) of the Fe88−xZr8B4Mnx amorphous ribbons and the Mn content.
Metals 11 01577 g005
Figure 6. The Arrott plots of the Fe88−xZr8B4Mnx (x = 2, 5, 8 and 10) glassy samples. (a) for Fe86Zr8B4Mn2, (b) for Fe83Zr8B4Mn5, (c) for Fe80Zr8B4Mn8 and (d) for Fe78Zr8B4Mn10.
Figure 6. The Arrott plots of the Fe88−xZr8B4Mnx (x = 2, 5, 8 and 10) glassy samples. (a) for Fe86Zr8B4Mn2, (b) for Fe83Zr8B4Mn5, (c) for Fe80Zr8B4Mn8 and (d) for Fe78Zr8B4Mn10.
Metals 11 01577 g006
Figure 7. The −ΔSm-T curves of the amorphous Fe88−xZr8B4Mnx samples from 1 to 5T: (a) for Fe86Zr8B4Mn2, (b) for Fe83Zr8B4Mn5, (c) for Fe80Zr8B4Mn8 and (d) for Fe78Zr8B4Mn10 (e) the n-T curves of the Fe88−xZr8B4Mnx samples obtained by the −ΔSmHn relationship.
Figure 7. The −ΔSm-T curves of the amorphous Fe88−xZr8B4Mnx samples from 1 to 5T: (a) for Fe86Zr8B4Mn2, (b) for Fe83Zr8B4Mn5, (c) for Fe80Zr8B4Mn8 and (d) for Fe78Zr8B4Mn10 (e) the n-T curves of the Fe88−xZr8B4Mnx samples obtained by the −ΔSmHn relationship.
Metals 11 01577 g007
Figure 8. The −ΔSmpeak vs. Tc−2/3 plots under a field of 1T, 2T, 3T, 4T, 5T for the Fe88−xZr8B4Mnx amorphous samples.
Figure 8. The −ΔSmpeak vs. Tc−2/3 plots under a field of 1T, 2T, 3T, 4T, 5T for the Fe88−xZr8B4Mnx amorphous samples.
Metals 11 01577 g008
Table 1. Thermal properties of Fe88−xZr8B4Mnx amorphous samples.
Table 1. Thermal properties of Fe88−xZr8B4Mnx amorphous samples.
Fe88−xZr8B4MnxTg (K)Tx (K)Tl (K)TrgγRef.
x = 078784016110.4890.350[19]
x = 279483715780.5030.353Present work
x = 579784215750.5060.355
x = 881184315750.5150.353
x = 1080684715730.5120.356
Table 2. Magnetic and magnetocaloric properties of Fe88−xZr8B4Mnx amorphous samples.
Table 2. Magnetic and magnetocaloric properties of Fe88−xZr8B4Mnx amorphous samples.
Fe88−xZr8B4MnxTc (K)−ΔSmpeak (J·kg−1·K−1)nRef.
1 T2 T3 T4 T5 T
x = 02910.871.52.062.573.040.769[19]
x = 22830.8671.4711.9992.4802.9310.745Present work
x = 52630.8181.3801.8652.3062.7150.736
x = 82380.6941.1721.5851.9602.3090.740
x = 102250.6891.1521.5511.9192.2610.736
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, X.; Wang, Q.; Tang, B.; Ding, D.; Cui, L.; Xia, L. Effect of Mn Substitution Fe on the Formability and Magnetic Properties of Amorphous Fe88Zr8B4 Alloy. Metals 2021, 11, 1577. https://doi.org/10.3390/met11101577

AMA Style

Wang X, Wang Q, Tang B, Ding D, Cui L, Xia L. Effect of Mn Substitution Fe on the Formability and Magnetic Properties of Amorphous Fe88Zr8B4 Alloy. Metals. 2021; 11(10):1577. https://doi.org/10.3390/met11101577

Chicago/Turabian Style

Wang, Xin, Qiang Wang, Benzhen Tang, Ding Ding, Li Cui, and Lei Xia. 2021. "Effect of Mn Substitution Fe on the Formability and Magnetic Properties of Amorphous Fe88Zr8B4 Alloy" Metals 11, no. 10: 1577. https://doi.org/10.3390/met11101577

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