Next Article in Journal
Role of Working Temperature and Humidity in Acetone Detection by SnO2 Covered ZnO Nanowire Network Based Sensors
Next Article in Special Issue
Relationship between the Formation of Magnetic Clusters and Hexagonal Phase of Gold Matrix in AuxFe1−x Nanophase Thin Films
Previous Article in Journal
Cellular Uptake of Gold Nanorods in Breast Cancer Cell Lines
Previous Article in Special Issue
Coating of Magnetite Nanoparticles with Fucoidan to Enhance Magnetic Hyperthermia Efficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unidirectional Magnetic Anisotropy in Molybdenum Dioxide–Hematite Mixed-Oxide Nanostructures

1
National Institute of Materials Physics, Atomistilor 405A, 077125 Magurele, Romania
2
Department of Physics, Duquesne University, Pittsburgh, PA 15282, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(6), 938; https://doi.org/10.3390/nano12060938
Submission received: 4 February 2022 / Revised: 7 March 2022 / Accepted: 8 March 2022 / Published: 12 March 2022

Abstract

:
MoO2-Fe2O3 nanoparticle systems were successfully synthesized by mechanochemical activation of MoO2 and α-Fe2O3 equimolar mixtures throughout 0–12 h of ball-milling. The role of the long-range ferromagnetism of MoO2 on a fraction of more defect hematite nanoparticles supporting a defect antiferromagnetic phase down to the lowest temperatures was investigated in this work. The structure and the size evolution of the nanoparticles were investigated by X-ray diffraction, whereas the magnetic properties were investigated by SQUID magnetometry. The local electronic structure and the specific phase evolution in the analyzed system versus the milling time were investigated by temperature-dependent Mössbauer spectroscopy. The substantially shifted magnetic hysteresis loops were interpreted in terms of the unidirectional anisotropy induced by pinning the long-range ferromagnetic order of the local net magnetic moments in the defect antiferromagnetic phase, as mediated by the diluted magnetic oxide phase of MoO2, to those less defect hematite nanoparticles supporting Morin transition. The specific evolutions of the exchange bias and of the coercive field versus temperature in the samples were interpreted in the frame of the specific phase evolution pointed out by Mössbauer spectroscopy. Depending on the milling time, a different fraction of defect hematite nanoparticles is formed. Less nanoparticles supporting the Morin transition are formed for samples exposed to a longer milling time, with a direct influence on the induced unidirectional anisotropy and related effects.

Graphical Abstract

1. Introduction

Molybdenum dioxide, MoO2, as a nanoparticulate system, has various energy-related applications. Other applications are in the fields of catalysis and electrochemistry [1,2,3,4].
Regarding the magnetic properties, nanoparticles of MoO2 could present the diluted magnetism behavior with the diluted ferromagnetic phase persisting above room temperature (RT). It was reported in by Zhang et al. [5] that the functionality of MoO2 can be enlarged by mixing it with other functional nano-systems (e.g., MoS2 or reduced graphene oxide for increasing the capacity of lithium-ion batteries).
The well-known hematite (α-Fe2O3) has various applications and can be used as a semiconductor compound [6], magnetic material [7], catalyst [8], and gas sensor [9]. Incorporation of other active ions into the α-Fe2O3 structure can lead to changes in unit-cell dimensions, particle size, as well as in the magnetic and optical properties. New applications may be developed by combining molybdenum dioxide and hematite in mixed-oxide nanostructures.
The objective of this work is to extend the research on diluted magnetism and to evaluate the possibility to control the RT ferromagnetism of MoO2 via the interfacial magnetic interactions with the antiferromagnetic phase of hematite in the equimolar mixtures of MoO2 and α-Fe2O3 nanoparticles. The exchange interactions between a ferromagnet (FM) and an antiferromagnet (AFM) across an interface, responsible for the unidirectional anisotropy, have attracted much attention because they give rise to both an exchange bias field as well as an enhanced coercive field, if compared with the case of a free (uncoupled) FM nanophase [10,11,12,13,14]. The so-called exchange bias systems have been exploited in spin valve devices such as magnetic read heads [15,16], nonvolatile memories [17,18], and various sensors [19,20]. Nanoparticulate systems involving unidirectional anisotropy are also intensively studied due to the possibility to increase the magnetic stabilization (magnetic blocked regime of fine nanoparticles) or the magnetic hyperthermia effects [21,22].
It is worth mentioning that this paper continues a series of reports on molybdenum dioxide-hematite nano-systems with different molarities [23,24], this time with the focus being on the observation and the explanation of the unidirectional anisotropy phenomena and related effects in the equimolar nanocomposite systems.

2. Materials and Methods

Powders (purchased from Alfa Aesar) of MoO2 (with an average particle size of about 100 nm) and hematite (with an average particle size of about 50 nm) were subjected to milling in a hardened steel vial with 12 stainless-steel balls of the high-energy SPEX 8000 mixer mill. The mass ratio between the 12 balls and the powder-like equimolar mixture was 5:1. Before introducing the powders in the ball-milling chamber, the powders were manually ground in air to obtain a homogeneous mixture (the 0 h sample). MoO2-Fe2O3 nanoparticle systems were successfully synthesized by mechano-chemical activation through ball-milling of MoO2 and α-Fe2O3 equimolar mixtures for 2, 4, 8, and 12 h of ball-milling time [23,24]
The structural characterization and the size evolution of the nanoparticles were investigated by X-ray diffraction (XRD). XRD patterns have been recorded at room temperature with a PANalytic X’Pert Pro MPO powder diffractometer (with Cu-Kα radiation, λ = 1.54187 Å). The lattice parameters a, b, and c were determined via Rietveld refinement of the XRD patterns, in the hypothesis of the Pseudo-Voigt profile of the lines. Additionally, the average crystallite size was calculated using the Scherrer method.
The magnetic properties were investigated by magnetometry using a Superconducting Quantum Interference Device (SQUID), a Quantum Design magnetometer working under the most sensitive reciprocal space option.
The local electronic structure and the specific phase evolution in the analyzed system versus the milling time were investigated by Mössbauer spectroscopy. Temperature-dependent Mössbauer spectroscopy between 5 and 295 K has been performed for a full and comprehensive characterization of the atomistic and magnetic configuration of the Fe-based nanophases. The temperature-dependent spectra were collected by inserting the samples in a closed-cycle He flow cryostat. A 57Co source in the Rhodium matrix was used on a spectrometer working under the constant acceleration mode. The Mössbauer spectra have been fitted by the NORMOS software package (Brand RA (1987) NORMOS program, Internat Rep Angewandte Phys. Universität Duisburg) and the calibration was performed by using an α-Fe metallic foil at RT. Isomer shift (IS) values, reported relative to α-Fe at room temperature (RT), are given in mm/s, similar to the quadrupole splitting (QS) values. In case of sextet-like patterns, the values of the quadrupole corrections, QS*, are provided.

3. Results and Discussions

The lattice parameters, crystallite size, and phase content of XRD patterns for MoO2-α-Fe2O3, subjected to high-energy ball-milling (HEBM) from 0 to 12 h, are presented in Table 1. The Rietveld structural refinement shows that parameter c of hematite presents a slight increase during the milling process due to Mo substituting Fe in the hematite lattice. The crystallite size decreases with the milling time. For example, after a 2 h milling time, the average crystallite size decreased to 39 nm from above 50 nm for α-Fe2O3 and to 72.5 nm from above 100 nm for MoO2. After a milling time of 12 h, the crystallite size decreased down to 17.7 and 50 nm for α-Fe2O3 and MoO2, respectively.
The XRD patterns of the equimolar mixture MoO2 + α-Fe2O3, corresponding to 2 h and respectively 8 h milling times, are presented in Figure 1. Similar to results reported in [23,24], one can notice a decrease of the diffraction peak intensities for both Fe2O3 and MoO2 phases with the increase of the ball-milling time.
This behavior suggests not only a broadening of the diffraction line, but also a substitution of Fe3+ ions with Mo4+ ions in the hematite lattice, and vice versa in the molybdenum dioxide lattice. However, the solubility of MoO2 into Fe2O3 is very limited, as resulting from Table 1. The proportion of Fe2O3 and MoO2 is almost insensitive to the milling time (within the error limits) with 64(1) wt.% corresponding to α-Fe2O3 and 36(1) wt.% to MoO2. Note the slight deviation from the equimolar mixture, with an expected proportion of 59(1) wt.% corresponding to α-Fe2O3 and 41 wt.% to MoO2, with the deviation being lower in case of shorter milling times.
The evolution of magnetization with temperature was measured on heating in a 200 Oe applied field (i.e., 0.02 T magnetic induction), after the samples were firstly cooled in the zero field (ZFC curve). The temperature-dependent magnetization in the same field was subsequently measured during the cooling process (FC curve). The ZFC–FC curves of the considered equimolar mixtures differently milled as well as of the initial components are presented in Figure 2. The divergence between the ZFC and FC curves below 350 K, specific to MoO2 and highlighted in Figure 2a, infer a superparamagnetic behavior assigned to clustered magnetic defects in MoO2 with a large distribution of blocking temperature (even above 350 K). It is important to note that the finite magnetization (although very low) persists up to the highest measuring temperature of 350 K. This supports the long-range magnetic order specific to a diluted magnetic oxide (DMO) and intensively discussed in our works [14,25,26,27]. Such a very weak ferromagnetism is due to the tiny magnetic moments mainly associated to oxygen defects which are interacting at unexpected longer distances by different specific mechanisms (e.g., the bounded magnetic polarons mechanism).
Figure 2b reveals the specific Morin transition (TMorin) of hematite (see the gray squares curve) at about 260 K. The thermal hysteresis associated with first-order thermodynamic Morin transition from a weak ferromagnetic phase at RT to an antiferromagnetic phase below 260 K [28,29] is very well-highlighted, not only for the initial α-Fe2O3 component of the mixture, but also for the 0 h sample (manually ground). Note that the magnetization in the antiferromagnetic state is still finite, as proof for magnetic defects which are always present in this state. A much lower amplitude of the Morin transition can be observed in the sample exposed to a 2 h milling time from Figure 2c, suggesting a reduced number of true hematite nanoparticles which are involved in the transition. In the samples subjected to longer milling times, the ZFC–FC curves show only weak curvatures without visible Morin transitions. A very interesting aspect is related to the behavior of specific magnetization (per mass unit) in the case of the pure α-Fe2O3 component and of the equimolar mixture of the 0 h sample (Figure 2b). Expectedly, the specific magnetization of the mixture should be lower in case of the 0 h mixture due to the relatively lower amount of hematite in this sample. If this expectation is fulfilled below the Morin transition, the behavior is opposite above the transition, where the weak ferromagnetic phase of hematite is formed. The only explanation for an enhanced magnetization above the Morin transition is that the weak ferromagnetic phase increases its magnetic contribution along the field direction due to the presence of the DMO phase, which is characterized by its intrinsic long-range ferromagnetic order. If going on to Figure 2c, the specific behavior of the ZFC–FC curves much better resembles the ones of the long-range magnetic order of the DMO system in Figure 2a than the ones of nanosized hematite shown in Figure 2b, except the two orders of magnitude higher values of magnetization specific to the defect antiferromagnetic phase of hematite. Therefore, as will also be subsequently demonstrated by Mössbauer spectroscopy, we may further assume that the main magnetic contribution of samples from 2 to 12 h comes from long-range ferromagnetic interactions of local net magnetic moments in a defect antiferromagnetic (DAFM) phase of hematite, as mediated by the DMO phase. This high-defect phase of hematite does not support any Morin transition and its specific contribution increases with the milling time.
Hysteresis loops acquired at 5 and 300 K on the considered samples are shown in Figure 3a,b, respectively. It can be observed that in the case of the milled samples, the magnetization at the same high-enough field is increasing with the milling time. At 300 K, the much higher coercive field of pure hematite infers a higher anisotropy of the weak ferromagnetic phase as compared to the antiferromagnetic phase. The presence of a component of high coercivity in the 300 K hysteresis loops of the milled samples where the DAFM phase affords the dominant magnetic contribution supports the fact that a true hematite phase is also present in samples supporting longer mechanical attrition, although the contribution of this component substantially diminishes with the milling time.
An interesting feature, revealed by the zoomed view of the coercive field domain from Figure 3, is the field asymmetry of the hysteresis loops, both at 5 and at 300 K, suggesting enhanced unidirectional anisotropy effects. This led us to measure the hysteresis curves also after cooling the sample in 5 T. Figure 4a shows the hysteresis loops recorded at 5 K, for the sample subjected to milling for 2 h, after cooling without and respectively in the 5 T applied field. Figure 4b shows the hysteresis curves measured after FC of the 2 h milled sample at different temperatures. As the details of the insets show, the asymmetry of the hysteresis loops is significant.
For the sample milled for 2 h, the maximum shift of the field after FC is 385(5) Oe (38.5(5) mT) at 300 K and the maximum coercive field is 550(5) Oe (55.0(5) mT) at the same temperature. In the existing literature [10,11,12,13,14], these asymmetries are related to the interfacial exchange interaction between a ferromagnetic phase and an antiferromagnetic phase. As such, the substantially shifted hysteresis loops were interpreted in terms of the unidirectional anisotropy induced by coupling the DAFM phase specific to disordered hematite to the true hematite-like phase of the nanoparticulate mixture. As an intuitive image, the small size and defect hematite nanoparticles which interact strongly with MoO2 give rise to the DAFM phase, whereas the rest of the larger and better-formed hematite nanoparticles provide the true hematite-like phase, which can still support the Morin transition. There is an interfacial interaction between these two types of nanoparticles which leads to the unidirectional anisotropy.
According to [27], the unidirectional anisotropy involves both an increase of the coercive field and the shift of the hysteresis loop by the so-called exchange bias field (usually negative). The temperature dependence of the exchange bias field and the coercive field obtained in both ZFC and FC procedures is presented in Figure 5.
Within the error limits, the temperature dependences of the exchange bias field as well as of the coercive field obtained after field cooling followed the same trend as after the zero-field cooling. However, after field cooling, these two parameters presented enhanced absolute values, in agreement with the theoretical expectations of the unidirectional pinning mechanisms [30].
Two aspects are clearly pointed out by Figure 5. Firstly, the highest exchange bias and coercive fields, meaning the highest unidirectional anisotropy, were manifested in the 2 h sample, and the average values continuously decreased with the milling time. Secondly, in the mixture milled for 2 h, where the traces for a Morin transition are still evidenced in Figure 2c, the absolute values of the coercive field and exchange bias field increased with the temperature, especially above 200 K, going against the theoretical expectations specific to such dependencies [30,31]. The first aspect can be explained at this point by the assumption that the 2 h sample provided the highest amount of true hematite-like phase, i.e., supporting Morin transition, as illustrated in Figure 2c, and therefore the amount of pinning phase, and hence the pinning effect was maximal in this sample. Moreover, the amount of true hematite like-phase decreased with the milling time, as suggested in Figure 3, and therefore the pinning effects are expected to diminish continuously with the milling time. On the other hand, if looking at the 2 h sample, the anisotropy of the true hematite-like phase was maximal above the Morin transition, and therefore the interfacial pinning effects exerted by this phase are expected to be higher above 200 K.
Direct support for this explanation is based on the formation of two types of magnetic phases having relative amounts controlled by different milling times, as will be shown in the next section dedicated to Mössbauer spectroscopy investigations.
Temperature-dependent Mössbauer spectra have been collected on two representative samples of milled nanocomposite mixtures: the 2 h sample, where traces for the Morin transition were evidenced in Figure 2c, and the 8 h sample, without any visible trace of such transition in Figure 2c. All the collected spectra, which are shown in Figure 6, have been fitted by three spectral components: two external sextets and one central paramagnetic doublet, the last one being assigned to Fe ions entering the MoO2 lattice at the atomic level (counted by the doublet at the lowest temperature) or forming small Fe3+ oxide clusters inside this network (counted by the additional relative contribution of the doublet component at 295 K). According to this supposition, about 2% of Fe enter the MoO2 lattice at the atomic level after a 2 h milling time and 5% after an 8 h milling time, whereas the rest of the other 3% and 10% of Fe ions, respectively, form small clusters of Fe oxide inside the MoO2. However, the exchange bias effects cannot be explained by coupling the diluted FM phase to such antiferromagnetic clusters, while the exchange bias fields decreased with the milling time whereas the amounts of Fe oxide clusters increased with the milling time.
The two sextet components show the following specific behavior: their superposition is very similar below 100 K and is modified at 295 K, therefore inferring a possible magnetic transition between these temperatures. For both samples, at the lowest temperature of 5 K, the best fitting has been obtained via an external sextet with a hyperfine magnetic field, Bhf, of 54.6 T (and QS* of 0.25 mm/s) and an internal sextet with Bhf of 52.8 T (and QS* close to 0.10 mm/s). At 295 K, the two spectral components were characterized by a hyperfine magnetic field of 51.6 T (and QS* close to −0.11 mm/s) and 52.4 T (and QS* close to 0.10 mm/s), with the relative contribution of the 51.6 T sextet being very close to the one of the 54.6 T from the low-temperature spectra. All these spectral characteristics support the assignment of the sextet changing the sign of QS* below 295 K (and with a hyperfine magnetic field of 54.6 T at 5 K and 51.6 at 295 K) to a better-formed hematite phase with Morin transition at above 260 K. The sextet which keeps the value of the quadrupole splitting almost constant versus temperature and presents a very slight decrease of the hyperfine magnetic field with temperature is assigned to an additional defect phase of hematite (as pointed by XRD), which does not support the Morin transition (most probably due to the incorporation of Mo ions into the hematite structure). While the QS* value for this phase is positive and similar to the one of antiferromagnetic phase in pure hematite, a defect antiferromagnetic order involving local net magnetic moments has to be considered for this phase. As a matter of fact, the sextet belonging to well-structured hematite has a relative spectral area of about 57% in the sample milled for 2 h and of only 30% in sample milled for 8 h. The predominance of the pure hematite phase in the sample milled for 2 h also explains the very clearly evidenced Morin transition in the first sample as compared to samples milled for longer times, where the defect hematite phase (without Morin transition) becomes dominant. It is this phase of DAFM hematite which provides the magnetic signal of the samples, of almost similar magnitude at RT and at low temperature. As a direct support for this statement, the magnetization in the low field increased by a factor of about 1.6 in the 8 h sample as compared with the 2 h sample (see Figure 2c), in excellent agreement with the relative spectral contributions of 70% and 43% of the sextets assigned to the DAFM hematite phase in the 8 and 2 h samples, respectively.
Under the same reasoning, the 2 h sample was characterized by 57% of pure hematite pinning phase and 43% of pinned DAFM phase, in contrast to the 8 h sample, where the relative amount of the pinning phase was much lower. Therefore, the unidirectional anisotropy was the highest in the 2 h sample and diminished with the milling time. Moreover, in the same 2 h sample with the highest relative amount of pure hematite phase, there was the strongest contribution of the higher anisotropy specific to the weak ferromagnetic phase above the Morin transition. Therefore, stronger pinning effects and unidirectional anisotropy were observed above 200 K mainly in this sample.

4. Conclusions

Nanocomposites of α-Fe2O3 and MoO2 interacting nanoparticles were obtained from equimolar mixtures via mechanochemical activation. The average size of hematite nanoparticles continuously decreased with the milling time, i.e., from 50 nm before milling down to 17 nm after 12 h of milling. Additionally, parameter c of hematite showed a slight increase during the milling process due to Mo substituting Fe in the hematite lattice. As expected, the structural disorder also increased with the milling time. Magnetic measurements and Mössbauer spectroscopy provide support for the formation of two types of hematite nanoparticles: (i) the less defected ones, denoted as a pure hematite phase which suffers the usual Morin transition, and (ii) the more defected ones, denoted as a DAFM phase which is present over the entire temperature interval and involves long-range ferromagnetic order between local net magnetic moments, as mediated by the diluted magnetic oxide, MoO2 nanophase. The substantially shifted hysteresis loops of the nanocomposite mixture were interpreted in terms of the unidirectional anisotropy induced by pinning the long-range ferromagnetic phase specific to the DAFM nanoparticles to the less defect nanoparticles of pure hematite phase. Both the magnetic order and the anisotropy constant of the pinning phase can be tuned by crossing the specific Morin transition.
Temperature-dependent Mössbauer spectroscopy investigations showed that the relative content of the pure hematite phase and of the DAFM phase, respectively, can be modified by the milling time. In particular, the amount of the pinning phase decreased with the milling time. Consequently, the unidirectional anisotropy and related parameters such as the exchange bias field and the coercive field of the nanocomposite systems decreased with the milling time.

Author Contributions

Conceptualization, V.K., M.S. and F.T.; methodology, V.K.; software, N.I. and M.T.; validation, V.K., M.S., L.D., F.T., M.T. and N.I.; investigation, N.I., F.T., M.T. and L.D.; resources, F.T. and M.S.; data curation, M.S., L.D., N.I. and F.T.; writing—original draft preparation, V.K. and F.T.; writing—review and editing, V.K., F.T. and M.S.; visualization, F.T., M.S., L.D., M.T., N.I. and V.K.; supervision, V.K. and M.S.; project administration, F.T.; funding acquisition, V.K., F.T. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Project PCCDI 47 Contract 597/2018, CPN-III-P2-2.1-PED 493/2020, and PED 486/2020 and Core Program Project PN030101 nr.21N/2019. M.S. acknowledges support from the National Science Foundation, US, from grants DMR-0854794 and DMR-1002627-1.

Data Availability Statement

Raw data supporting the reported results can be found at NIMP.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhu, M.Y.; Xu, X.; Chen, G.; Zhong, Y.; Cai, R.; Li, L.; Shao, Z. Surfactant- Free Self-Assembly of Reduced Graphite Oxide-MoO2 Nanobelt Composites Used as Electrodes for Lithium Ion Batteries. Electrochim. Acta 2016, 211, 972–981. [Google Scholar] [CrossRef]
  2. Zhou, E.; Wang, C.; Zhao, Q.; Li, Z.; Shao, M.; Deng, X.; Liu, X.; Xu, X. Facile Synthesis of MoO2 Nanoparticles as High-Performance Supercapacitor Electrodes and Photocatalysts. Ceram. Int. 2016, 42, 2198–2203. [Google Scholar] [CrossRef]
  3. Zhang, T.; Kong, L.B.; Liu, M.C.; Dai, Y.H.; Yan, K.; Hu, B.; Luo, Y.C.; Kang, L. Design and Preparation of MoO2/MoS2 as Negative Electrode Materials for Supercapacitors. Mater. Des. 2016, 112, 88–96. [Google Scholar] [CrossRef]
  4. Zhang, X.; Hou, Z.; Li, X.; Liang, J.; Zhu, Y.; Qian, Y. MoO2 Nanoparticles as High Capacity Intercalation Anode Material for Long Cycle Lithium Ion Battery. Electrochim. Acta 2016, 213, 416–422. [Google Scholar] [CrossRef]
  5. Zhang, C.; Zhang, P.; Dai, J.; Zhang, H.; Xie, A.; Shen, Y. Facile Synthesis and Electrochemical Properties of MoO2/Reduced Grapheme Oxide Hybrid for Efficient Anode of Lithium Ion Battery. Ceram. Int. 2016, 42, 3618–3624. [Google Scholar] [CrossRef]
  6. Wang, G.X.; Gou, X.L.; Horvat, J.; Park, J. Facile Synthesis and Characterization of Iron Oxide Semiconductor Nanowires for Gas Sensing Application. J. Phys. Chem. C 2008, 112, 15220–15225. [Google Scholar] [CrossRef]
  7. Raffaella, B.; Etienne, S.; Cinzia, G.; Fabia, G.; Mar, G.H.; Miguel, A.G.; Roberto, C.; Pantaleo, D.C. Colloidal Semiconductor/Magnetic Heterostructures Based on Iron-Oxide-Functionalized Brookite TiO2 Nanorods. Phys. Chem. Chem. Phys. 2009, 11, 3680–3691. [Google Scholar] [CrossRef]
  8. Krishnamoorthy, S.; Rivas, J.A.; Amiridis, M.D. Catalytic Oxidation of 1,2-Dichlorobenzene over Supported Transition Metal Oxides. J. Catal. 2000, 193, 264–272. [Google Scholar] [CrossRef]
  9. Sorescu, M.; Diamandescu, L.; Tomescu, A.; Tarabasanu-Mihaila, D.; Teodorescu, V. Structure and Sensing Properties of 0.1SnO2-0.9Fe2O3 System. Mater. Chem. Phys. 2008, 107, 127–131. [Google Scholar] [CrossRef]
  10. Kuncser, V.; Schinteie, G.; Palade, P.; Mustata, I.; Lungu, C.P.; Stefan, N.; Chiriac, H.; Vladoiu, R.; Filoti, G. Spin Configurations and Interfacial Diffusion in Exchange Bias and Spin Valve Systems with Ir–Mn Antiferromagnetic Pinning Layers. Hyp. Int. 2009, 191, 135–141. [Google Scholar] [CrossRef]
  11. Meiklejohn, W.H.; Bean, C.P. New Magnetic Anisotropy. Phys. Rev. 1956, 102, 1413. [Google Scholar] [CrossRef]
  12. Radu, F.; Zabel, H. Exchange Bias Effect of Ferro-Antiferromagnetic Heterostructures. Magnetic Heterostructures, Chapter Springer Tracts. In Modern Physics Springer; Bader, S., Zabel, H., Eds.; Springer: Berlin/Heidelberg, Germany, 2008; Volume 227, pp. 97–184. [Google Scholar]
  13. Nogues, J.; Schuller, I.K. Exchange Bias. J. Magn. Magn. Mater. 1992, 192, 203–232. [Google Scholar] [CrossRef]
  14. Kuncser, V.; Schinteie, G.A.; Kuncser, A.C.; Leca, A.; Scarisoreanu, M.; Morjan, I.; Filoti, G. Physical Mechanisms of Exchange Coupling Effects in Nanoparticulate Diluted Magnetic Oxides Obtained by Laser Pyrolysis. J. Phys. Chem. C 2017, 121, 9063–9069. [Google Scholar] [CrossRef]
  15. Lin, S.; Wang, Z.W.; Zhang, Q.H.; Chen, S.R.; Jin, Q.; Yao, H.B.; Xu, S.; Meng, F.Q.; Yin, X.M.; Wang, C.; et al. Exchange Coupling in Synthetic Anion-Engineered Chromia Heterostructures. Adv. Funct. Mater. 2021, 32, 2109828. [Google Scholar] [CrossRef]
  16. Sharma, S.; Kumar, M.; Srinet, G.; Siqueiros, J.M.; Herrera, O.R. Structural, Raman analysis and exchange bias effects in Mn doped multiferroic Bi0.80La0.10Ca0.10Fe1−xMnxO3 ceramics. Ceram. Int. 2021, 47, 6834–6841. [Google Scholar] [CrossRef]
  17. Wang, L.; Zhang, L.L.; Zhang, X.; Zhong, M.L.; Zhong, Z.C.; Rao, G.H. Tunable reversals of magnetization and exchange bias of perovskite YbCr1−xFexO3 (x = 0–0.15). Ceram. Int. 2019, 45, 6143–6148. [Google Scholar] [CrossRef]
  18. Dong, Y.N.; Zhao, X.N.; Han, X.; Fan, Y.B.; Xie, X.J.; Chen, Y.X.; Bai, L.H.; Dai, Y.Y.; Yan, S.S.; Tian, Y.F. Spin-orbit torque controllable complete spin logic in a single magnetic heterojunction. Appl. Phys. Lett. 2021, 118, 152403. [Google Scholar] [CrossRef]
  19. Guo, Y.; Ouyang, Y.; Sato, N.; Ooi, C.C.; Wang, S.X. Exchange-Biased Anisotropic Magnetoresistive Field Sensor. IEEE Sens. J. 2017, 17, 3309–3315. [Google Scholar] [CrossRef]
  20. Klug, M.J.; Thormahlen, L.; Robisch, V.; Toxvaerd, S.D.; Hoft, M.; Knochel, R.; Quandt, E.; Meyner, D.; McCord, J. Antiparallel exchange biased multilayers for low magnetic noise magnetic field sensors. Appl. Phys. Lett. 2019, 114, 192410. [Google Scholar] [CrossRef]
  21. Kazantseva, N.E.; Smolkova, I.S.; Babayan, V.; Vilcakova, J.; Smolka, P.; Saha, P. Magnetic Nanomaterials for Arterial Embolization and Hyperthermia of Parenchymal Organs Tumors: A Review. Nanomaterials 2021, 11, 3402. [Google Scholar] [CrossRef]
  22. Graca, M.P.; Soreto, S.; Gavinho, S.; Valente, M.; Salgueiro, C.; Nunes, J.; Soares, P.; Lanca, M.; Vieira, T.; Silva, J. Nanomaterials for magnetic hyperthermia. Eur. J. Public Health 2021, 31 (Suppl. 2), ckab120-066. [Google Scholar] [CrossRef]
  23. Trotta, R.; Tolea, F.; Valeanu, M.; Diamandescu, L.; Grabias, A.; Sorescu, M. Magnetic and Hyperfine Properties of Molibdenum Dioxide-Hematite Mixed Oxide Nanostructures. MRS Adv. 2018, 3, 2887–2892. [Google Scholar] [CrossRef]
  24. Knauss, M.; Tolea, F.; Valeanu, M.; Diamandescu, L.; Trotta, R.; Wood, K.; Grabias, A.; Sorescu, M.J. Mechanochemical Synthesis and Characterization of Molibdenum Dioxide-Hematite Nanostructures with Different Molarities. Min. Mat. Charact. Eng. 2018, 6, 587–600. [Google Scholar]
  25. Tolea, F.; Grecu, M.N.; Kuncser, V.; Constantinescu, S.G.; Ghica, D. On the role of Fe ions on magnetic properties of doped TiO2 nanoparticles. Appl. Phys. Lett. 2015, 106, 142404. [Google Scholar] [CrossRef]
  26. Nistor, S.V.; Stefan, M.; Nistor, L.C.; Kuncser, V.; Ghica, D.; Vlaicu, I.D. Aggregates of Mn2+ Ions in Mesoporous Self-Assembled Cubic ZnS:Mn Quantum Dots: Composition, Localization, Structure, and Magnetic Properties. J. Phys. Chem. C 2016, 120, 14454–14466. [Google Scholar] [CrossRef]
  27. Kuncser, V.E.; Stromberg, F.; Acet, M.; Keune, W. Mössbauer Effect Study of Correlation Between Structure and Exchange-Bias Effect In Ferromagnetic Fe∕Antiferromagnetic FeSn2FeSn2 Bilayers. J. Appl. Phys. 2005, 97, 063513. [Google Scholar] [CrossRef]
  28. Gupta, R.K.; Ghosh, K.; Dong, L.; Kahol, P.K. Structural and magnetic properties of nanostructured iron oxid. Phys. e-Low-Dimens. Syst. Nanostruct. 2011, 43, 1095–1098. [Google Scholar] [CrossRef]
  29. Kubaniova, D.; Kubickova, L.; Kmjec, T.; Zaveta, K.; Niznansky, D.; Brazda, P.; Klementova, M.; Kohout, J. Hematite: Morin temperature of nanoparticles with different size. J. Magn. Magn. Mater. 2019, 475, 611–619. [Google Scholar] [CrossRef]
  30. Nogués, J.; Schuller, I.K. Exchange bias. JMMM 1999, 192, 203–232. [Google Scholar] [CrossRef]
  31. Stromberg, F.; Keune, W.; Kuncser, V.; Westerholt, K. Fe Coverage Induced Out-of-plane Spin Component of the Antiferromagnetic Spin Structure in Exchange Biased Fe/FeSn2 Bilayers. Phys. Rev. B 2005, 72, 064440. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the equimolar mixture MoO2-α-Fe2O3, corresponding to milling times of (a) 2 h and (b) 8 h.
Figure 1. XRD patterns of the equimolar mixture MoO2-α-Fe2O3, corresponding to milling times of (a) 2 h and (b) 8 h.
Nanomaterials 12 00938 g001
Figure 2. ZFC–FC thermomagnetic measurements, recorded in 200 Oe applied magnetic field (0.02 T magnetic induction), for (a) MoO2 and (b) Fe2O3 and for the powder mixture manually ground (0 h), (c) for samples obtained after 2, 4, 8, and 12 h milling times.
Figure 2. ZFC–FC thermomagnetic measurements, recorded in 200 Oe applied magnetic field (0.02 T magnetic induction), for (a) MoO2 and (b) Fe2O3 and for the powder mixture manually ground (0 h), (c) for samples obtained after 2, 4, 8, and 12 h milling times.
Nanomaterials 12 00938 g002
Figure 3. Hysteresis curves recorded (a) at 5 and (b) at 300 K. Insets: zoomed view of coercivity.
Figure 3. Hysteresis curves recorded (a) at 5 and (b) at 300 K. Insets: zoomed view of coercivity.
Nanomaterials 12 00938 g003
Figure 4. Hysteresis curves for the sample after milling for 2 h (a) at 5 K measured on ZFC and FC, and (b) measured at different temperatures on FC. Insets: zoomed view of coercivity.
Figure 4. Hysteresis curves for the sample after milling for 2 h (a) at 5 K measured on ZFC and FC, and (b) measured at different temperatures on FC. Insets: zoomed view of coercivity.
Nanomaterials 12 00938 g004
Figure 5. Temperature dependence of the exchange bias field (Hex) and of coercivity (Hc) for samples milled for 2, 4, and 12 h.
Figure 5. Temperature dependence of the exchange bias field (Hex) and of coercivity (Hc) for samples milled for 2, 4, and 12 h.
Nanomaterials 12 00938 g005
Figure 6. Temperature-dependent 57Fe Mössbauer spectra recorded at 5, 100, and 295 K on the equimolar mixture MoO2-Fe2O3 after 2 h (left side) and 8 h (right side) of milling time.
Figure 6. Temperature-dependent 57Fe Mössbauer spectra recorded at 5, 100, and 295 K on the equimolar mixture MoO2-Fe2O3 after 2 h (left side) and 8 h (right side) of milling time.
Nanomaterials 12 00938 g006
Table 1. The lattice parameters, crystallite size, and phase content from XRD patterns by Rietveld refinement and the Scherrer method for mechano-chemically activated MoO2-Fe2O3 nanocomposites at ball-milling times of: 0, 2, 4, 8, and 12 h, respectively.
Table 1. The lattice parameters, crystallite size, and phase content from XRD patterns by Rietveld refinement and the Scherrer method for mechano-chemically activated MoO2-Fe2O3 nanocomposites at ball-milling times of: 0, 2, 4, 8, and 12 h, respectively.
Milling Time
(h)
Lattice Parameters
(Å)
Crystallite Size
(nm)
Phase Content
(wt.%)
abc
05.0343-13.7430>50α-Fe2O3 (63.9)
5.62294.85415.5374>100MoO2 (36.1)
25.0329-13.728638.9Mo: α-Fe2O3 (63.2)
5.61994.85035.537272.5Fe: MoO2 (36.8)
45.0337-13.740329.2Mo: α-Fe2O3 (63.4)
5.62124.84995.540863.3Fe: MoO2 (36.6)
85.0341-13.741323.1Mo: α-Fe2O3 (64.1)
5.61884.84915.544756.5Fe: MoO2 (35.9)
125.0334-13.753717.7Mo: α-Fe2O3 (65.5)
5.61804.84805.545749.9Fe: MoO2 (34.5)
Statistical Errors±0.0007±0.0007±0.0007±1.6±1.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tolea, F.; Sorescu, M.; Diamandescu, L.; Iacob, N.; Tolea, M.; Kuncser, V. Unidirectional Magnetic Anisotropy in Molybdenum Dioxide–Hematite Mixed-Oxide Nanostructures. Nanomaterials 2022, 12, 938. https://doi.org/10.3390/nano12060938

AMA Style

Tolea F, Sorescu M, Diamandescu L, Iacob N, Tolea M, Kuncser V. Unidirectional Magnetic Anisotropy in Molybdenum Dioxide–Hematite Mixed-Oxide Nanostructures. Nanomaterials. 2022; 12(6):938. https://doi.org/10.3390/nano12060938

Chicago/Turabian Style

Tolea, Felicia, Monica Sorescu, Lucian Diamandescu, Nicusor Iacob, Mugurel Tolea, and Victor Kuncser. 2022. "Unidirectional Magnetic Anisotropy in Molybdenum Dioxide–Hematite Mixed-Oxide Nanostructures" Nanomaterials 12, no. 6: 938. https://doi.org/10.3390/nano12060938

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