Next Article in Journal
Zeolitic Imidazolate Frameworks Serve as an Interface Layer for Designing Bifunctional Bone Scaffolds with Antibacterial and Osteogenic Performance
Previous Article in Journal
State-of-the-Art 2D and Carbon Nanomaterials in France
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Giant Exchange Bias Effect Due to Enhanced Ferromagnetism Using a Mixed Martensitic Phase in Ni50Mn37Ga13 Spun Ribbons

MOE Key Laboratory for Nonequilibrium Synthesis and Modulation of Condensed Matter, School of Physics, Xi’an Jiaotong University, Xi’an 710049, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(21), 2827; https://doi.org/10.3390/nano13212827
Submission received: 5 October 2023 / Revised: 21 October 2023 / Accepted: 23 October 2023 / Published: 25 October 2023

Abstract

:
The structure of a material is an important factor in determining its physical properties. Here, we adjust the structure of the Ni50Mn37Ga13 spun ribbons by changing the wheel speed to regulate the exchange bias effect of the material. The characterization results of micromorphology and structure show that as the wheel speed increases, the martensite lath decreases from 200 nm to 50 nm, the structure changed from the NM to a NM and 10M mixed martensitic structure containing mainly NM, then changed to NM and 10M where 10M and NM are approaching. Meanwhile, HE first increased and then decreased as the wheel speed increased. The optimum exchange bias effect (HE = 7.2 kOe) occurs when the wheel speed is 25 m∙s−1, mainly attributed to the enhanced ferromagnetism caused by part of 10M in NM martensite, which enhanced the exchange coupling of ferromagnetism and antiferromagnetism. This work reveals the structural dependence of exchange bias and provides a way to tune the magnitude of the exchange bias of Heusler alloys.

1. Introduction

The structure of materials is crucial to studying them and is the main factor affecting their performance. Different structures and microstructures show different properties [1,2]. The most typical examples are the different physical and chemical properties of materials of the carbon family (such as graphite, diamond graphene, etc.) [3,4,5]. The same phenomenon occurs in magnetic materials [6,7,8,9]. For Fe-Co-V soft magnetic alloys, the occurrence of parts of the γ phase in the α phase will increase the coercive force of the alloy. The brittleness will increase when the ordered α’ phase is generated in the material [6,7]. The structure also plays an important role in studying magnetostrictive materials such as Fe-Ga. Excellent magnetostrictive properties will occur when the m-D03 phase appears in the material [8,9]. Therefore, researching and establishing the relationship between material properties and structures is necessary, which is very important for material design.
The exchange bias (EB) effect refers to the phenomenon of a hysteresis loop deviating from the origin’s center after the magnetic material is cooled with an external magnetic field [10,11,12]. The EB effect has many applications in recording device read heads, electronic spin valves, and magnetoresistive sensors [13,14,15]. This effect exists widely in magnetic systems with magnetic interactions, such as antiferromagnetic/ferromagnetic (AFM/FM) thin films [16,17], FM core/AFM shell nanoparticles [18,19], perovskite systems with phase separation [20,21], and Heulser alloys [22,23]. Ni-Mn base Heusler alloys have received extensive attention and research due to their rich structures (monoclinic, tetragonal, orthorhombic, etc.) and magnetic phase transitions in the martensitic phase [24,25,26]. In Ni2Mn1+xZ1−x (In, Ga, Sb, etc.) Heusler alloys, the magnetism mainly depends on the distance of Mn atoms, dMn–Mn < 2.83 Å is AFM, and dMn–Mn ≥ 2.83 Å is FM; different magnetic states will appear in various structures. In general, the Mn-Mn exchange interaction is FM within the regular Mn sublattices, while it is AFM between the Mn atoms occupying the regular Mn sublattice and Z sublattice. In addition, the tetragonal in the martensitic phase would enhance the AFM, resulting in a significant degree of magnetic frustration, and the orthorhombic distortion would enhance the FM [27,28]. In 2007, Mahmud Khan et al. first discovered the EB effect of 284 Oe in the Heusler alloy [22]. Subsequently, the EB effect of the material was improved by adjusting the alloy atomic ratio [29], component substitution [30], heat treatment process [31], and preparation process [32]. However, the relationship between structure and EB effects has not been studied in depth.
In this work, we prepared Ni50Mn37Ga13 spun ribbons by changing the wheel speed through combined arc melting and using a single-roller melt-spinning apparatus. The relationship between the EB effect and the ribbon structure was studied. The results show that the EB effect has a non-monotonic trend of first increasing and then decreasing with the increase in the wheel speed. An optimal EB effect of 7.2 kOe is obtained when the wheel speed is 25 m∙s−1. The structure and microstructure of the material were studied using an X-ray diffractometer (XRD) and transmission electron microscopy (TEM), indicating that the optimal EB effect was caused when the system’s structure was mainly NM in a NM and 10M mixed martensitic structure. This result provides a new method for designing and preparing materials with excellent EB effects.

2. Materials and Methods

The Ni50Mn37Ga13 alloy was formulated according to the stoichiometric ratio of 50:37:13 and prepared via arc melting. The ingots were smelted five times, and the purity of the metal elemental raw materials used was 99.9%. Ribbons were produced in a single-roller melt-spinning apparatus with a copper wheel (radius: 12 cm), and the specific processing parameters are shown in Table 1. The as-cast bulk alloys were remelted in a quartz tube with an opening at the bottom via induction melting. After melting, the molten alloy was ejected on the high-speed rotating copper wheel by pressured argon gas. X-ray diffraction (XRD, Bruker D8 Advance X-ray diffractometer, Germany, Cu Kα, λ = 0.15418 nm) was used to investigate the crystalline nature of Ni50Mn37Ga13 ribbons. A field-emission scanning electron microscope (SEM, JEOL JSM-7000F, Japan) with electron energy disperse spectroscopy (EDS) and transmission electron microscopy (TEM, JEOL JEM-2100, Japan) were used to observe the micromorphology and microstructure of the ribbons. The magnetic properties, including magnetic hysteresis (MH) loops and magnetization temperature (MT) curves, are measured using a superconducting quantum interference device (SQUID) magnetometer (Quantum Design, MPMS-XL-5). A differential scanning calorimetry (DSC) at cooling/heating rates of 10 K min−1 was used to determine the phase transition temperature.

3. Results and Discussion

First, the atom ratio of the ribbons was analyzed using EDS. The result is given in Figure 1a, which illustrates analytical spectra and the corresponding alloy composition. It is seen that the composition of the ribbons with different wheel speeds are all very close to the nominal composition and only deviate from an acceptable level. Figure 1b shows the SEM morphologies imaged on the fractured cross sections of the ribbons. When observing the fractured cross sections, the ribbons were placed with the copper wheel contacting face at the bottom and the free face at the top, as shown in Figure 1b. Obvious grain boundaries can be observed on the cross section of all samples, which implies a typical intergranular fracture feature. For the ribbons with relatively low speeds (5 m∙s−1 and 15 m∙s−1), some equiaxial grains can be observed near the contacting face, while columnar grains appear upwards to the free face. The co-existence of two different types of grains is due to the different cooling rates between the contacting face and the free face due to the relatively thick ribbons. With the increasing wheel speeds, the ribbons become thinner, and the grains also become smaller; only columnar grains can be observed for ribbons with relatively high wheel speeds. The specific thickness of the ribbons is marked in Figure 1b. The thickness of the ribbons when the wheel speed is 5 m∙s−1, 15 m∙s−1, 25 m∙s−1, and 35 m∙s−1 is 140 μm, 44 μm, 21 μm, and 12 μm, respectively.
In order to reveal the structure change of Ni50Mn37Ga13 spun ribbons with wheel speeds of 5~35 m∙s−1, X-ray diffraction was carried out at room temperature, as shown in Figure 2. It can be observed that the crystal structure of the two groups of samples with wheel speeds of 5 and 15 m∙s−1 is similar to the Ni50Mn37Ga13 bulk alloy (Figure 1a,b) and is mainly the tetragonal NM martensitic phase; the characteristic peaks of the XRD diffraction peaks are (112), (200), (202), and (440) [33]. At the same time, some other weak characteristic peaks also appeared in the diffraction peak pattern at the wheel speed of 15 m∙s−1. When the wheel speed increases to 25 m∙s−1, in addition to the characteristic diffraction peak of NM, the characteristic peak of the second phase is further enhanced, showing a martensitic structure of 10M, corresponding to the (200), (0010), and (12 5 - ) [34]. When the wheel speed reaches 35 m∙s−1, the characteristic peak of 10M martensite is further enhanced. The XRD results were further refined using the Rietveld refinement program of Full Prof. The refinements of NM and 10M are based on the structure from the Inorganic Crystal Structure Database (ICSD); the specific ICSD collection code is 153292 and 158168 for NM and 10M, respectively. Small values of the reliability factors of Rietveld refinement (RWP: weighted profile factor and χ2: goodness of fit) indicate a good agreement between experimental and calculated diffraction patterns of samples. The corresponding lattice parameters and volume fractions obtained via the XRD refinements for ribbons with different wheel speeds are summarized in Table 2. The ratios of NM and 10M martensite structures are approximately 98:2, 95:5, 87:13, and 60:40, respectively, when the wheel speed is 5, 15, 25, and 35 m∙s−1. The results show that as the wheel speed increases, the dominance of the NM martensite structure changes to NM and 10M, where 10M and NM are approaching. As the structure changes, the distance between Mn atoms changes, thereby changing the magnetic properties of alloys.
Figure 3a–d shows the micromorphology of Ni50Mn37Ga13 ribbons. Obvious martensite laths appeared in all the ribbons; however, the directions of adjacent martensite laths are inconsistent, indicating that when martensite encounters martensite of another orientation during the formation and growth process, the growth will stop at the junction between the two. At the same time, it can also be observed that the martensite laths inside the grains are not all parallel, and some of the martensite laths inside the grains are in the shape of spearheads, forming a self-collaborative relationship. With the increase in wheel speed, the lath-like martensite went from coarse to fine. This is because during the single-roller spin quenching and stripping process to prepare the ribbons, the different wheel speeds of the copper rollers lead to differences in the cooling speed of the ribbons, resulting in a slightly different latent heat of crystallization and degree of reaction, so martensite with different crystal orientations or different structures appears, like NM and 10M [35,36]. In addition, an EDS mapping analysis was carried out. The bright spots in different colors correspond to Ni, Mn, and Ga, respectively. The results show that all elements were distributed uniformly throughout the area, indicating no component segregation (Figure 3e1–e3).
To further study the micromorphology and microstructure of the spun ribbons, it was characterized and analyzed using TEM, as shown in Figure 4a–d, and the inset shows that all the ribbons have typical martensitic laths, and the shape and thickness of the laths change in detail with the wheel speed. When the wheel speed is 5 m∙s−1, it forms twins in a twin microstructure [37]. The first martensitic twin variant has a width of 500 nm, while the second one has a width of 2 nm. The tetragonal of the corresponding selected area electron diffraction (SAED) pattern was verified as seen in inset Figure 4a; the image is viewed along [021]. The microstructure and diffraction spots of the ribbons with a rotation speed of 15 m∙s−1 are similar to those of 5 m∙s−1, as shown in Figure 4b. It can be observed from Figure 4c,d that as the wheel speed increases, the martensite laths further reduce, and the width decreases to about 100 nm and 50 nm, respectively. The martensite laths with a wheel speed of 25 m∙s−1 show different micromorphology, indicating that the ribbons comprise two types of martensite. As seen in Figure 4c,d, the corresponding SAED represents two sets of overlapping martensitic structures between adjacent martensitic variants. A group of crystal structures is the NM, with nine other diffraction points between the two main diffraction points, which is the 10M crystal structure [38,39]. When the wheel speed is further increased to 35 m∙s−1, 10M are approaching NM. With the increase of the wheel speed, the centrifugal force increases dramatically, thus decreasing the thickness of the ribbons and increasing the cooling rate of molten alloys.
Regarding solidification dynamics, ribbons with a high cooling rate would have a high crystallization rate, which could limit the kinetic grain growth and form a small grain [40]. It is the reason for the decrease in grain size with the increasing wheel speed. Meanwhile, from the perspective of solidification thermodynamics, different wheel speeds can also induce different supercooling degrees and form different martensitic phases, such as NM and 10M [41].
Figure 5 is the DSC analysis spectrum of Ni50Mn37Ga13 ribbons with different wheel speeds. A pair of obvious endothermic and exothermic peaks appear in all DSC curves, corresponding to the martensitic phase transformation and inverse martensitic phase transformation of the alloy. These two peaks are not in the same position, showing a thermal hysteresis for the transformation. A larger thermal hysteresis indicates that the phase transition is the first-order phase transition. The martensitic transition temperature (TM) decreases as the wheel speed increases. The TM for 5 m∙s−1, 15 m∙s−1, 25 m∙s−1, and 35 m∙s−1 ribbons are 610 K, 607 K, 560 K, and 556 K, respectively. The decrease in TM is caused by changes in the material structure.
Figure 6 shows the MT curves of Ni50Mn37Ga13 spun ribbons under zero-field cooling (ZFC) and field cooling (FC) conditions. The ZFC curve cools the sample from 400 K to 10 K without an external field; then, the temperature is raised to 400 K under an external magnetic field (200 Oe). The change in the magnetization of the sample during the heating process was recorded. The FC curve cools the sample from 400 K to 10 K under 200 Oe and then records the change in magnetization intensity during the heating to 400 K under the same external field. It can be seen in Figure 6a–d that all of the ZFC and FC curves have irreversible magnetization behavior, which could be a typical magnetic feature of spin glass. In order to confirm the occurrence of the spin glass further, an AC susceptibility has been tested, as shown in the inset of Figure 6c. The peak position of the AC susceptibility shifts to higher temperatures with increasing frequency, known as the frequency-diffusion phenomenon, a significant feature of the spin glass [23,29]. The ZFC and FC curves overlap when the temperature exceeds the freezing temperature. When the temperature is lower than the freezing temperature, the ZFC curve and the FC curve are separated, and the magnetization of ZFC is significantly lower than FC. The difference between the ZFC and FC curves increases as the temperature decreases. In addition, the freezing behavior of the magnetic moment begins to occur gradually from above the freezing temperature. As the temperature drops below the freezing temperature, the spins in the ZFC are randomly frozen. In the FC process, under the effect of the cooling field, the spins freeze along the magnetic field direction, distinguished from the random direction of the ZFC process. Thus, the energy barrier in the FC process, which must be overcome during the magnetization process, is smaller than in the ZFC process, which causes a split between the ZFC and FC curves [42]. The peak of the ZFC magnetization curve is the temperature of the spin glass phase transition (Tp). Tp reduced with the wheel speed increase and is 130 K, 128 K, 115 K, and 104 K at 5 m∙s−1, 15 m∙s−1, 25 m∙s−1, and 35 m∙s−1, respectively. When the wheel speed increases from 15 m∙s−1 to 25 m∙s−1 and 25 m∙s−1 to 35 m∙s−1, two significant ΔTp mutations occur, which is attributed to changes in the magnetic state caused by structural changes of the NM and 10M martensitic structure. The magnetization intensity of FC increases monotonically with the increase in wheel speed at 10 K, mainly due to the enhanced FM by 10M martensitic structure.
As shown in Figure 7, the relationship between the structure and EB was studied via the MH loops of the ribbons at different wheel speeds. All samples were cooled to 10 K under an external cooling field of 20 kOe to test MH loops, and the measurement range was −50 to 50 kOe. In order to avoid the remanence, the sample was first heated to 400 K before the measurement and then cooled down to the measuring temperature. The HE of the EB effect is calculated using HE = −(HL + HR)/2. Among them, the HL is the left intersection of the loop with the magnetic field axis, and HR is the right intersection with the magnetic field axis. It can be observed from Figure 7a–d that the shape of MH loops with different wheel speeds are similar, and all have obvious shifts along the magnetic field axis. The MH curves show an obvious hysteresis loop under a low magnetic field, and magnetization increases linearly in the region of the high magnetic field. Magnetization loops are attributed to the co-existence of the FM and AFM phases, and the high field linear part is from the contribution of AFM phases [43]. The interaction of the FM and AFM phases causes the EB effect. As shown in Figure 7e, HE shows a non-monotonic trend of first increasing and then decreasing as the wheel speed increases. When the wheel speed is 5 m∙s−1, the HE is 6.3 kOe, and the optimal EB effect with HE = 7.2 kOe appears at the wheel speed 25 m∙s−1. Then, as the wheel speed increases further, the HE is reduced to 3.6 kOe. This was caused by changing the martensitic structure from the NM to the mixed martensitic structure of NM and 10M, then changing to NM and 10M, where 10M and NM are approaching. The optimal EB effect occurs in the ribbon that is mainly NM and contains part of 10M martensite. This is attributed to the enhancement of FM by a small amount of 10M martensitic structure, resulting in an enhanced exchange coupling between FM and AFM. The maximum magnetization (Mmax) is the magnetization measured at the maximum measuring field (50 kOe) and can express the difficulty of the magnetization. Moreover, the Mmax of the spun ribbons increases from 2.3 emu∙g−1 to 2.4 emu∙g−1; when the wheel speed is 5 and 15 m∙s−1, Mmax does not change much. At this time, the martensitic structure is mainly NM. The magnetization significantly varies from 2.4 to 3.0 emu∙g−1 when the wheel speed increases to 25 m∙s−1. This is due to the further formation of more 10M. When the wheel speed is further increased to 35 m∙s−1, the Mmax undergoes a mutation and increases to 5.6 emu∙g−1, caused by a mixed structure of 10M approaching NM.
As shown in Figure 8, the center of the MH loop also shifts toward negative magnetic moments caused by spin glass behavior and the competition between FM and AFM [42]. This shift is strongly dependent on the cooling field, as shown in Figure 8a–d. The centers of the MH curves shift to a negative moment, weaken with the increasing cooling field, and confirm that this shift is a characteristic of the system itself, not caused by the measurement. Furthermore, the EB effect shows a tendency to first increase and then decrease with the increasing cooling field and obtain a peak value under a 20 kOe cooling field (Figure 8e). With the expansion of the cooling field, the FM cluster grows bigger, thus increasing the FM interaction and changing the competition between the FM and AFM [29]; the symmetry along the moment axis and EB effect changes correspondingly. Finally, when the cooling field is increased to 70 kOe, the EB still remains, but the shift of MH loops along the moment axis almost disappears, as shown in (Figure 8d).

4. Conclusions

In summary, we prepared the Ni50Mn37Ga13 spun ribbons by changing the wheel speed (5~35 m∙s−1) through a single-roller melt-spinning apparatus. The martensitic structure changed from NM to NM and 10M, where NM becomes dominant, then changed to NM and 10M, where 10M and NM are approaching. The optimal EB effect with HE = 7.2 kOe appears when the wheel speed is 25 m∙s−1, which is attributed to the structure of the spun ribbon being mainly NM and containing a certain amount of 10M martensite; the emergence of the 10M structure enhances the FM of the system, thereby enhancing the exchange coupling between FM and AFM. Our work concluded the relationship between structure and EB effects, providing an effective way to design functional materials with excellent EB effects.

Author Contributions

F.T.: conceptualization, data collection, data analysis, original draft writing, draft revision; Q.Z.: data collection, data analysis, draft revision; J.G.: data collection, data analysis; S.K.: data analysis; B.L.: data collection; Z.D.: data analysis; M.F.: data analysis; Y.Z.: draft revision; C.Z.: draft revision; K.C.: draft revision; S.Y.: draft revision, project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key Research and Development Program of China (2021YFB3501401), the National Natural Science Foundation of China (Grants No. 51601140 and No. 12304027), the Key Scientific and Technological Innovation Team of Shaanxi Province (2020TD001), and the Innovation Capability Support Program of Shaanxi (No. 2018PT-28 and No. 2017KTPT-04).

Data Availability Statement

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

Acknowledgments

We thank Chenyu Liang at the Instrument Analysis Center of Xi’an Jiaotong University for her assistance with X-ray diffraction analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ding, Q.; Zhang, Y.; Chen, X.; Fu, X.; Chen, D.; Chen, S.; Gu, L.; Wei, F.; Bei, H.; Gao, Y.; et al. Tuning element distribution, structure and properties by composition in high-entropy alloys. Nature 2019, 573, 223–227. [Google Scholar] [CrossRef] [PubMed]
  2. Li, X.; Lu, L.; Li, J.; Zhang, X.; Gao, H. Mechanical properties and deformation mechanisms of gradient nanostructured metals and alloys. Nat. Rev. Mater. 2020, 5, 706–723. [Google Scholar] [CrossRef]
  3. Németh, P.; McColl, K.; Garvie, L.A.J.; Salzmann, C.G.; Murri, M.; McMillan, P.F. Complex nanostructures in diamond. Nat. Mater. 2020, 19, 1126–1131. [Google Scholar] [CrossRef]
  4. Strong, H.M.; Hanneman, R.E. Crystallization of Diamond and Graphite. J. Chem. Phys. 1967, 46, 3668–3676. [Google Scholar] [CrossRef]
  5. Rao, C.N.R.; Biswas, K.; Subrahmanyam, K.S.; Govindaraj, A. Graphene, the new nanocarbon. J. Mater. Chem. 2009, 19, 2457–2469. [Google Scholar] [CrossRef]
  6. Sundar, R.S.; Deevi, S.C. Effect of heat-treatment on the room temperature ductility of an ordered intermetallic Fe–Co–V alloy. Mater. Sci. Eng. A 2004, 369, 164–169. [Google Scholar] [CrossRef]
  7. Rezende, M.F.S.; Lacerda, A.M.; Pfannes, H.-D.; Mansur, R.A.; Ardisson, J.D.; Persiano, A.I.C. Gamma to alpha phase transformation in FeCoV alloys. Hyperfine Interact. 1994, 83, 231–234. [Google Scholar] [CrossRef]
  8. Li, X.; Bao, X.; Yu, X.; Gao, X. Magnetostriction enhancement of Fe73Ga27 alloy by magnetic field annealing. Scripta Mater. 2018, 147, 64–68. [Google Scholar] [CrossRef]
  9. Chen, Y.; Fu, Z.; Wu, Y.; Xu, Y.; Xiao, Y.; Wang, J.; Zhang, R.; Jiang, C. Giant heterogeneous magnetostriction induced by charge accumulation-mediated nanoinclusion formation in dual-phase nanostructured systems. Acta Mater. 2021, 213, 116975. [Google Scholar] [CrossRef]
  10. Meiklejohn, W.H.; Bean, C.P. New magnetic anisotropy. Phys. Rev. 1956, 102, 1413. [Google Scholar] [CrossRef]
  11. Nogués, J.; Schuller, I.K. Exchange bias. J. Magn. Magn. Mater. 1999, 192, 203. [Google Scholar] [CrossRef]
  12. Nayak, A.K.; Nicklas, M.; Chadov, S.; Khuntia, P.; Shekhar, C.; Kalache, A.; Baenitz, M.; Skourski, Y.; Guduru, V.K.; Puri, A.; et al. Design of compensated ferrimagnetic Heusler alloys for giant tunable exchange bias. Nat. Mater. 2015, 14, 679–684. [Google Scholar] [CrossRef] [PubMed]
  13. Skumryev, V.; Stoyanov, S.; Zhang, Y.; Hadjipanayis, G.; Givord, D.; Nogués, J. Beating the superparamagnetic limit with exchange bias. Nature 2003, 423, 850. [Google Scholar] [CrossRef] [PubMed]
  14. Fukami, S.; Zhang, C.; DuttaGupta, S.; Kurenkov, A.; Ohno, H. Magnetization switching by spin–orbit torque in an antiferromagnet–ferromagnet bilayer system. Nat. Mater. 2016, 15, 535. [Google Scholar] [CrossRef]
  15. Barker, J.; Craig, B.; Lamberton, R.; Johnston, A.; Chantrell, R.W.; Heinonen, O. A model of the exchange bias setting process in magnetic read sensors. Appl. Phys. Lett. 2009, 95, 022504. [Google Scholar] [CrossRef]
  16. Feng, J.; Liu, H.F.; Wei, H.X.; Zhang, X.-G.; Ren, Y.; Li, X.; Wang, Y.; Wang, J.P.; Han, X.F. Giant perpendicular exchange bias in a subnanometer inverted (Co/Pt)n/Co/IrMn structure. Phys. Rev. Appl. 2017, 7, 054005. [Google Scholar] [CrossRef]
  17. Migliorini, A.; Kuerbanjiang, B.; Huminiuc, T.; Kepaptsoglou, D.; Muñoz, M.; Cuñado, J.L.F.; Camarero, J.; Aroca, C.; Vallejo-Fernández, G.; Lazarov, V.K.; et al. Spontaneous exchange bias formation driven by a structural phase transition in the antiferromagnetic material. Nat. Mater. 2018, 17, 28. [Google Scholar] [CrossRef]
  18. Binns, C.; Qureshi, M.T.; Peddis, D.; Baker, S.H.; Howes, P.B.; Boatwright, A.; Cavill, S.A.; Dhesi, S.S.; Lari, L.; Kröger, R.; et al. Exchange bias in Fe@Cr core-shell nanoparticles. Nano Lett. 2013, 13, 3334. [Google Scholar] [CrossRef]
  19. Muzzi, B.; Albino, M.; Petrecca, M.; Innocenti, C.; Fernández, C.d.J.; Bertoni, G.; Marquina, C.; Ibarra, M.R.; Sangregorio, C. 3d Metal Doping of Core@Shell Wüstite@ferrite Nanoparticles as a Promising Route toward Room Temperature Exchange Bias Magnets. Small 2022, 18, 2107426. [Google Scholar] [CrossRef]
  20. Sahoo, R.C.; Takeuchi, Y.; Ohtomo, A.; Hossain, Z. Exchange bias and spin glass states driven by antisite disorder in the double perovskite compound LaSrCoFeO6. Phys. Rev. B 2019, 100, 214436. [Google Scholar] [CrossRef]
  21. Xie, L.; Huang, H.L.; Zhang, H.G.; Yu, J.Q.; Zhang, J.Z.; Lu, Y.L. Tunable magnetization reversal and exchange bias in NiCr2O4 ceramics doped with non-magnetic ions. Scripta Mater. 2021, 205, 114210. [Google Scholar] [CrossRef]
  22. Khan, M.; Dubenko, I.; Stadler, S.; Ali, N. Exchange bias behavior in Ni-Mn-Sb Heusler alloys. Appl. Phys. Lett. 2007, 91, 072510. [Google Scholar] [CrossRef]
  23. Wang, B.M.; Liu, Y.; Ren, P.; Xia, B.; Ruan, K.B.; Yi, J.B.; Ding, J.; Li, X.G.; Wang, L. Large exchange bias after zero-field cooling from an unmagnetized state. Phys. Rev. Lett. 2011, 106, 077203. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, H.; Li, Z.; Zhang, Y.; Ni, Z.; Xu, K.; Liu, Y. A large barocaloric effect associated with paramagnetic martensitic transformation in Co50Fe2.5V31.5Ga16 quaternary Heusler alloy. Scripta Mater. 2020, 177, 1–5. [Google Scholar] [CrossRef]
  25. Załeski, K.; Ekholm, M.; Alling, B.; Abrikosov, I.A.; Dubowik, J. Local atomic configuration approach to the nonmonotonic concentration dependence of magnetic properties of Ni2Mn1+xZ1−x (Z = In, Sn, Sb) Heusler alloys. Scripta Mater. 2021, 194, 113646. [Google Scholar] [CrossRef]
  26. Cao, P.; Tian, F.; Li, W.; Vitos, L.; Wang, Y. Ideal superelasticity in Ni-based Heusler alloys. Acta Mater. 2021, 210, 116816. [Google Scholar] [CrossRef]
  27. Entel, P.; Gruner, M.E.; Fähler, S.; Acet, M.; Cahır, A.; Arroyave, R.; Sahoo, S.; Duong, T.C.; Talapatra, A.; Sandratskii, L.; et al. Probing Structural and Magnetic Instabilities and Hysteresis in Heuslers by Density Functional Theory Calculations. Phys. Status Solidi B 2018, 255, 1700296. [Google Scholar] [CrossRef]
  28. Jia, L.; Shen, J.; Li, M.; Wang, X.; Ma, L.; Zhen, C.; Hou, D.; Liu, E.; Wang, W.; Wu, G. Tuning antiferromagnetic exchange interaction for spontaneous exchange bias in MnNiSnSi system. APL Mater. 2017, 5, 126105. [Google Scholar] [CrossRef]
  29. Tian, F.; Ke, X.; Cao, K.; Wang, D.; Zhao, Q.; Li, J.; Dai, Z.; Wang, D.; Zhang, Y.; Zhou, C.; et al. Tailoring exchange bias in reentrant spin glass by ferromagnetic cluster size engineering. APL Mater. 2021, 9, 011104. [Google Scholar] [CrossRef]
  30. Boldrin, M.; Silva, A.G.; Coutrim, L.T.; Jesus, J.R.; Macchiutti, C.; Bittar, E.M.; Bufaical, L. Tuning the spontaneous exchange bias effect with Ba to Sr partial substitution in La1.5(Sr0.5−xBax)CoMnO6. Appl. Phys. Lett. 2020, 117, 212402. [Google Scholar] [CrossRef]
  31. Czaja, P.; Fitta, M.; Przewoznik, J.; Maziarz, W.; Morgiel, J.; Czeppe, T.; Cesari, E. Effect of heat treatment on magnetostructural transformations and exchange bias in Heusler Ni48Mn39.5Sn9.5Al3 ribbons. Acta Mater. 2016, 103, 30–45. [Google Scholar] [CrossRef]
  32. Zhao, D.W.; Li, G.K.; Wang, S.Q.; Ma, L.; Zhen, C.M.; Hou, D.L.; Wang, W.H.; Liu, E.K.; Chen, J.L.; Wu, G.H. Tuning exchange bias by Co doping in Mn50Ni41-xSn9Cox melt-spun ribbons. J. Appl. Phys. 2014, 116, 103910. [Google Scholar] [CrossRef]
  33. Liang, L.; Tong, W.; Xu, J.; Wang, X.; Wang, X.M.; Zhao, Y.; Wang, W.Y.; Wang, H.J.; Tian, J.; Tian, Q.C.; et al. Evidence for austenite to non-modulated martensite transformation crystallography and variant organization in Ni-Mn-Ga-Co ferromagnetic shape memory alloys. Acta Mater. 2023, 254, 119022. [Google Scholar] [CrossRef]
  34. Righi, L.; Albertini, F.; Pareti, L.; Paoluzi, A.; Calestani, G. Commensurate and incommensurate “5M” modulated crystal structures in Ni–Mn–Ga martensitic phases. Acta Mater. 2007, 55, 5237–5245. [Google Scholar] [CrossRef]
  35. Feng, Y.; Chen, H.; Gao, L.; Wang, H.; Bian, X.; Gong, M. Effect of wheel speed and annealing temperature on microstructure and texture evolution of Ni45Mn36.6In13.4Co5 ribbon. Mater. Charact. 2016, 122, 170–176. [Google Scholar] [CrossRef]
  36. Ma, S.C.; Shih, C.W.; Liu, J.; Yuan, J.H.; Lee, S.Y.; Lee, Y.I.; Chang, H.W.; Chang, W.C. Wheel speed-dependent martensitic transformation and magnetocaloric effect in Ni–Co–Mn–Sn ferromagnetic shape memory alloy ribbons. Acta Mater. 2015, 90, 292–302. [Google Scholar] [CrossRef]
  37. Zhou, L.; Schneider, M.M.; Giri, A.; Cho, K.; Sohn, Y. Microstructural and crystallographic characteristics of modulated martensite, non-modulated martensite, and pre-martensitic tweed austenite in Ni-Mn-Ga alloys. Acta Mater. 2017, 134, 93–103. [Google Scholar] [CrossRef]
  38. Ge, Y.; Zarubova, N.; Heczko, O.; Hannula, S.-P. Stress-induced transition from modulated 14M to non-modulated martensite in Ni–Mn–Ga alloy. Acta Mater. 2015, 90, 151–160. [Google Scholar] [CrossRef]
  39. Ge, Y.; Jiang, H.; Sozinova, A.; Soderberga, O.; Lanskaa, N.; Keranenc, J.; Kauppinen, E.I.; Lindroosa, V.K.; Hannulaa, S.-P. Crystal structure and macrotwin interface of five-layered martensite in Ni–Mn–Ga magnetic shape memory alloy. Mater. Sci. Eng. A 2006, 438–440, 961–964. [Google Scholar] [CrossRef]
  40. Quintana-Nedelcos, A.; Sánchez Llamazares, J.L.; Ríos-Jara, D.; Lara-Rodríguez, A.G.; García-Fernández, T. Effect of quenching rate on the average grain size and martensitic transformation temperature in rapidly solidified polycrystalline Ni50Mn37Sn13 alloy ribbons. Phys. Status Solidi A 2013, 210, 2159–2165. [Google Scholar] [CrossRef]
  41. Wua, Y.; Wang, J.; Jianga, C.; Xu, H. Effect of coherent nanoprecipitates on martensitic transformation in Tb doped NiMnGa melt-spun ribbons. Intermetallics 2018, 97, 42–51. [Google Scholar] [CrossRef]
  42. Maniv, E.; Murphy, R.A.; Haley, S.C.; Doyle, S.; John, C.; Maniv, A.; Ramakrishna, S.K.; Tang, Y.-L.; Ercius, P.; Ramesh, R.; et al. Exchange bias due to coupling between coexisting antiferromagnetic and spin-glass orders. Nat. Phys. 2021, 17, 525–530. [Google Scholar] [CrossRef]
  43. Mao, J.; Sui, Y.; Zhang, X.; Su, Y.; Wang, X.; Liu, Z.; Wang, Y.; Zhu, R.; Wang, Y.; Liu, W.; et al. Temperature- and magnetic-field-induced magnetization reversal in perovskite. Appl. Phys. Lett. 2011, 98, 192510. [Google Scholar] [CrossRef]
Figure 1. (a) EDS analyzed the element ratio of the spun ribbons with different wheel speeds; (b) SEM photographs imaged on the fractured cross sections for ribbons with different wheel speeds; the inset is the partially enlarged morphologies of the red box area.
Figure 1. (a) EDS analyzed the element ratio of the spun ribbons with different wheel speeds; (b) SEM photographs imaged on the fractured cross sections for ribbons with different wheel speeds; the inset is the partially enlarged morphologies of the red box area.
Nanomaterials 13 02827 g001
Figure 2. XRD patterns recorded at room temperature for Ni50Mn37Ga13 spun ribbons prepared by different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1.
Figure 2. XRD patterns recorded at room temperature for Ni50Mn37Ga13 spun ribbons prepared by different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1.
Nanomaterials 13 02827 g002
Figure 3. SEM images for Ni50Mn37Ga13 ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1. White dashed line is the grain boundaries; (e1e3) EDS mapping for SEM image (c), the corresponding elemental mapping images of the Ni, Mn, and Ga, respectively.
Figure 3. SEM images for Ni50Mn37Ga13 ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1. White dashed line is the grain boundaries; (e1e3) EDS mapping for SEM image (c), the corresponding elemental mapping images of the Ni, Mn, and Ga, respectively.
Nanomaterials 13 02827 g003
Figure 4. TEM images for Ni50Mn37Ga13 spun ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1; inset figure is the corresponding SAED pattern, the arrows indicate the weak spots between the main spots.
Figure 4. TEM images for Ni50Mn37Ga13 spun ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1; inset figure is the corresponding SAED pattern, the arrows indicate the weak spots between the main spots.
Nanomaterials 13 02827 g004
Figure 5. The DSC curves for Ni50Mn37Ga13 ribbons with different wheel speeds 5 m∙s−1, 15 m∙s−1, 25 m∙s−1, and 35 m∙s−1.
Figure 5. The DSC curves for Ni50Mn37Ga13 ribbons with different wheel speeds 5 m∙s−1, 15 m∙s−1, 25 m∙s−1, and 35 m∙s−1.
Nanomaterials 13 02827 g005
Figure 6. Temperature dependence of magnetization under ZFC/FC processes measured in the presence of a 200 Oe cooling field for Ni50Mn37Ga13 spun ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1, respectively. The inset in (c) is the AC susceptibility for ribbons with a 25 m∙s−1 wheel speed.
Figure 6. Temperature dependence of magnetization under ZFC/FC processes measured in the presence of a 200 Oe cooling field for Ni50Mn37Ga13 spun ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1, respectively. The inset in (c) is the AC susceptibility for ribbons with a 25 m∙s−1 wheel speed.
Nanomaterials 13 02827 g006
Figure 7. MH loops for Ni50Mn37Ga13 spun ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1. (e) Dependency of HE and Mmax with Ni50Mn37Ga13 spun ribbons with different wheel speeds.
Figure 7. MH loops for Ni50Mn37Ga13 spun ribbons with different wheel speeds (a) 5 m∙s−1, (b) 15 m∙s−1, (c) 25 m∙s−1, and (d) 35 m∙s−1. (e) Dependency of HE and Mmax with Ni50Mn37Ga13 spun ribbons with different wheel speeds.
Nanomaterials 13 02827 g007
Figure 8. MH loops for Ni50Mn37Ga13 spun ribbons for 25 m∙s−1 after cooling to 10 K with different cooling fields (a) 10 kOe, (b) 30 kOe, (c) 50 kOe, and (d) 70 kOe. (e) Dependency of HE with Ni50Mn37Ga13 spun ribbons with different cooling fields.
Figure 8. MH loops for Ni50Mn37Ga13 spun ribbons for 25 m∙s−1 after cooling to 10 K with different cooling fields (a) 10 kOe, (b) 30 kOe, (c) 50 kOe, and (d) 70 kOe. (e) Dependency of HE with Ni50Mn37Ga13 spun ribbons with different cooling fields.
Nanomaterials 13 02827 g008
Table 1. The specific processing parameters used in the melt-spinning method.
Table 1. The specific processing parameters used in the melt-spinning method.
Liner Speed (m∙s−1)Rotate Speed (r∙min−1)Induction Current (A)Pressure (Mpa)
15400300.13
2151200300.13
3252000300.13
4352800300.13
Table 2. The corresponding lattice parameters and volume fractions using XRD refinements for ribbons with different wheel speeds.
Table 2. The corresponding lattice parameters and volume fractions using XRD refinements for ribbons with different wheel speeds.
Wheel SpeedsMartensitic Structure a (Å)b (Å)c (Å)α (°)β (°)γ (°)Fraction (%)
5 m∙s−1NM3.828903.828906.7693790909097.85
10M4.317795.7412421.11009094.06709902.15
15 m∙s−1NM3.825493.825496.7903190909094.96
10M4.328605.7304521.219019093.82677905.04
25 m∙s−1NM3.830533.830536.7582790909087.25
10M4.324965.7489721.172779093.831389012.75
35 m∙s−1NM3.843133.843136.73391090909060.19
10M4.334535.7867820.765789093.73939039.81
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

Tian, F.; Zhao, Q.; Guo, J.; Kong, S.; Liu, B.; Dai, Z.; Fang, M.; Zhang, Y.; Zhou, C.; Cao, K.; et al. A Giant Exchange Bias Effect Due to Enhanced Ferromagnetism Using a Mixed Martensitic Phase in Ni50Mn37Ga13 Spun Ribbons. Nanomaterials 2023, 13, 2827. https://doi.org/10.3390/nano13212827

AMA Style

Tian F, Zhao Q, Guo J, Kong S, Liu B, Dai Z, Fang M, Zhang Y, Zhou C, Cao K, et al. A Giant Exchange Bias Effect Due to Enhanced Ferromagnetism Using a Mixed Martensitic Phase in Ni50Mn37Ga13 Spun Ribbons. Nanomaterials. 2023; 13(21):2827. https://doi.org/10.3390/nano13212827

Chicago/Turabian Style

Tian, Fanghua, Qizhong Zhao, Jiale Guo, Sen Kong, Bingjie Liu, Zhiyong Dai, Minxia Fang, Yin Zhang, Chao Zhou, Kaiyan Cao, and et al. 2023. "A Giant Exchange Bias Effect Due to Enhanced Ferromagnetism Using a Mixed Martensitic Phase in Ni50Mn37Ga13 Spun Ribbons" Nanomaterials 13, no. 21: 2827. https://doi.org/10.3390/nano13212827

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