Next Article in Journal
MgO–ZrO2 Ceramic Composites for Silicomanganese Production
Previous Article in Journal
Formation of Composite Coatings on Galvanized Steel from Organosilane Solutions Using Electrophoresis and Sol–Gel Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Cation Distributions of Zn-doped Normal Spinel MgFe2O4 Ferrite and Its Magnetic Properties

1
School of Mathematics and Physics, Lanzhou Jiaotong University, Lanzhou 730070, China
2
Guangdong Provincial Key Laboratory of Optical Information Material and Technology & Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, China
3
Electron Microscopy Centre of Lanzhou University and Key Laboratory of Magnetism and Magnetic Materials of the Ministry of Education, School of Materials and Energy, Lanzhou University, Lanzhou 730000, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(7), 2422; https://doi.org/10.3390/ma15072422
Submission received: 11 February 2022 / Revised: 16 March 2022 / Accepted: 21 March 2022 / Published: 25 March 2022
(This article belongs to the Topic Synthesis and Applications of Nanowires)

Abstract

:
Determining the exact occupation sites of the doping ions in spinel ferrites is vital for tailoring and improving their magnetic properties. In this study, the distribution and occupation sites of cations in MgFe2O4 and Zn-doped MgFe2O4 ferrite are imaged by Cs-STEM. The experimental STEM images along [001], [011] and [111] orientations suggest that the divalent Mg2+ cations occupy all A sites, and the trivalent Fe3+ cations occupy all B sites in MgFe2O4 ferrite prepared by electrospinning, which is consistent with the normal spinel structure. We further clarify that the preferred sites of dopant Zn2+ ions are Fe3+ crystallographic sites in the Zn-doped MgFe2O4 ferrite nanofibers. Magnetic measurements show that Zn doping affects the spin states of the Fe3+, and the Fe3+-O2−-Fe3+ super-exchange interaction leads to enhancements in the magnetization and reduction in the Curie temperature. Our work should contribute a significant step toward eventually realizing the practical application of doped spinel ferrites.

1. Introduction

The direct imaging of atomic structures, especially in the site preference of substituted ions, is vital for magnetic materials to correctly explain their magnetic performance and provide guidance for potential commercial applications, including microwave devices, magnetic storage media, ferrofluids and biomedical devices [1,2,3,4,5,6,7]. It is known that the structural, electrical and magnetic properties of materials are highly sensitive to the conditions of their preparation, compositions and magnetic interactions, which strongly depend on the distribution of cations [8,9,10,11,12]. Magnetic spinel ferrites are of great interest in addressing the fundamental relationship between magnetic properties and their crystal structures, and they have received significant attention for their unique magnetic properties, such as a high Curie temperature, large magnetocrystalline anisotropy, a large Kerr effect and low magnetic losses [13,14,15,16]. Spinel ferrites are represented as [M1−xFex]tet[MxFe(2−x)]octO4 with two different co-ordination polyhedra for cations, namely tetrahedral and octahedral sites, where x represents the degree of inversion if (x = 0) is normal, (x = 1) is inverse and (0 < x < 1) is partially inversed. Twenty-four metallic ions occupied two interstitial crystallographic sites, including tetrahedral and octahedral sites. They showed a distribution of parallel and antiparallel magnetic moments in two sub-lattices, which resulted in a net magnetic moment per formula unit and eventually determined the unique magnetic properties of the spinel ferrites. Their magnetic properties could be systematically varied by changing the identity of the divalent M2+ cation or by a partial substitution while maintaining the basic crystal structure.
Biocompatible magnesium ferrite (MgFe2O4), an important functional soft magnetic (Ms ~ 27.8 emu/g; Hc ~ 9 Oe; Tc ~ 713 K) [17,18] and an n-type semiconductor material with a direct band gap of 1.9 eV [19], is an important member of the spinel family. It is widely used in transformer cores, humidity sensors, catalysts, coil cores, heterogeneous catalysis and sensors [20,21]. Recently, it has been reported in the literature that this ferrite could be applied to thermal coagulation treatment using an alternating magnet [22]. Many attempts have been made to reduce their cost and further improve their magnetic performance. Among these, ion doping has proven to be an effective method to tailor the magnetic performance of spinel ferrites [23,24]. The magnetic and dielectric properties of substituted ferrites directly depend on the electronic configuration of the dopant cations and on their preference for occupying the different Fe3+ sub-lattices of the spinel structure, such as ZnxMg1−xFe2O4 [25,26,27,28]. Non-magnetic Zn2+ ions are selected as doping ions to further improve the magnetic performance of the MgFe2O4 ferrite. However, the preferred site of the dopant Zn atoms is indistinct, so a satisfactory explanation of the magnetic properties of Zn-doped MgFe2O4 is still a matter of debate. Therefore, it is necessary to image their specific distributions in a sub-ångstrom resolution. Among the techniques for detecting the distributions and the preferred site of the doping ions [29,30,31,32,33,34,35], aberration-corrected transmission electron microscopy (Cs-corrected TEM) and scanning transmission electron microscopy (Cs-corrected STEM) equipped with energy-dispersive X-ray analysis (EDX) have recently made significant progress in space and time resolution. These techniques achieve a surprisingly high spatial resolution below 0.5 Å and readily realize a direct imaging and electronic-state detection of individual atoms and even chemical bonds. In this study, we employed Cs-STEM to systematically study the microstructure of MgFe2O4 nanofibers and the influence of the substitution and occupancy of doped Zn2+ on their magnetic properties.
In this study, the occupation sites and the occupancies of trace dopants in Zn-doped MgFe2O4 nanofibers were directly and precisely observed by Cs-STEM, which demonstrated that dopant Zn2+ cations preferred to randomly occupy Fe3+ crystallographic sites rather than Mg2+ sub-lattices to modify magnetic properties. The findings indicated that the Zn2+ substitution results in an increase in magnetization and a decrease in the Curie temperature. These results offer insights into interactions between dopant atoms and the magnetic properties in spinel structures and provide guidance for designing magnetoelectric multiferroics applications.

2. Materials and Methods

Synthesis of MFO and MZFO nanofibers. Single-particle-chain normal spinel MgFe2O4 (MFO) and MgZn0.2Fe1.8O4 (MZFO) nanofibers were synthesized by electrospinning. A typical precursor solution was used (A.R., Alfa-Aesar Inc., Ward Hill, NJ, USA), which contained 0.1 mmol of magnesium nitrate hexahydrate, 0.2 mmol of iron nitrite nonahydrate (0.1 mmol magnesium nitrate hexahydrate, 0.18 mmol iron nitrite nonahydrate and 0.02 mmol zinc nitrate hexahydrate); 5 mL of N, N-dimethyl formamide (DMF, A.R., Tianjin Chemical Corp., Tianjin, China) was dissolved into a glass container marked as A solution. Next, 1 mL of A solution, 1 mL of ethanol alcohol and 0.1786 g of poly vinylpyrrolidone (PVP, Mw ≈ 1,300,000, Alfa-Aesar Inc., Ward Hill, NJ, USA) were dissolved into a 5 mL vessel combined with sufficient stirring for 4~5 h. A typical precursor solution was successfully obtained. The electrospinning process was performed with 18 kV DC voltage with a 15 cm gap between the needle tip (Jinan Qinlu Pharmaceutical Technology Co., Ltd., Jinan, China) and the collector; the feed rate was 0.4 mL⋅h−1. The electrospun polymer composite nanofibers were first heated at 200 °C for 2 h and then calcined at 900 °C for 3 h with a heating rate of 1 °C⋅min−1 in air. The samples were finally cooled to room temperature with the same rate of 1 °C⋅min−1.
Characterization of MFO and MZFO nanofibers. The morphology and atomic-level crystal structures of the individual single-particle-chain MFO and MZFO nanofibers were systemically characterized with an aberration-corrected scanning transmission electron microscope (FEI Titan Cubed Themis G2 300, FEI, Hillsboro, OR, USA), which operated at 300 kV and was equipped with a monochromator, Gatan image filter (GIF Quantum ER/965, Gatan, CA, USA), EDX (Bruke EDX, Bruke, VA, USA). And an X-ray diffraction instrument (XRD, Philips X’pert ProMPD, Almelo, The Netherlands). Magnetic properties were investigated using a superconducting quantum interference magnetometer (SQUID, MPMS XL-7, Quantum Design, San Diego, CA, USA) and a magnetic property measurement system (MPMS, SQUID-VSM, Quantum Design, San Diego, CA, USA).

3. Results and Discussions

Single-particle-chain MgFe2O4 nanofibers (MFO) and Zn-doped MFO nanofibers (MZFO) were prepared using the electrospinning method. All nanofibers had a continuous structure and uniform chemical composition (ESI Figure S1). The average diameter of individual particles on the MFO and MZFO nanofibers was approximately 100 nm, with a single-crystal structure ranging from 90 to 105 nm (ESI Figures S2 and S3). MFO nanofibers were found to have a spinel structure, and Zn elements were successfully doped into the MFO nanofibers to form a single-phase MZFO with no impurity phase (ESI Figure S4). Figure 1a,b show the typical unit cell structure model of the ideal normal spinel MFO ferrite. It has a cubic unit cell structure containing 56 ions in total, in which the 32 O anions are closely packed, and 24 metal cations can be distributed in two different crystallographic sites, including a tetrahedral (A) site and an octahedral (B) site. As shown in Figure 1a, in the MFO unit cell ‘ball-and-stick’ model of the normal spinel structure, all the divalent Mg2+ cations occupy tetrahedral A sites, and all the trivalent Fe3+ cations occupy the octahedral B sites, in which the orange balls represent Mg cations (the atomic radius is 1.6 Å) and the bright blue spheres represent Fe cations (the atomic radius is 1.27 Å). Figure 1b shows the polyhedral model, which provides the corresponding polyhedral sites where different atoms are located.
The atomic-scale cation occupations and distributions of individual MFO single-particle-chain nanofibers were then captured using Cs-STEM. As shown in Figure 2a–c, atomic-scale images of an individual MFO nanoparticle, along with [001], [011] and [111] orientations, were taken by HAADF-STEM (Cubed Titan G2 60-300, FEI, Hillsboro, OR, USA).
It is noted that the contrast of the HAADF-STEM image is sensitive to the effective atomic number Z (ZMg < ZMg–Fe < ZFe) and the number of atoms (the length of the column). The atomic number of iron (Z = 26) was significantly higher than magnesium (Z = 12) when the number of atoms on the atomic column was the same or similar, so the atomic columns of pure Fe and mixed Mg-Fe were much brighter than the pure Mg atomic columns. However, when the number of atoms on the atomic column greatly differed, the contrast of images of different atomic columns was sensitive to the number of atoms rather than the atomic number. Figure 1a shows that the Mg atomic column located in the tetrahedral sites is clearly darker than the surrounding Fe atomic column, and the Mg atoms (1.6 Å) are easily distinguished because they are larger in size than the Fe atoms (1.27 Å). The atomic images in the three crystal axes are all consistent with the cation occupancy and distribution of the ideal normal spinel structure. The corresponding unit cell models are shown in Figure 2g–i, which clearly show that all tetrahedral sites are occupied by Mg cations, and all octahedral sites are occupied by Fe cations. Figure 2d–f further show the line intensity profiles of ‘1’, ‘2’ and ‘3’ marked in the light blue color, in which ‘Mg’ represents the pure Mg atomic column at the tetrahedral A sites, ‘Fe’ represents the pure Fe atomic column in the octahedral B sites, and ‘Mg & Fe’ represents mixed Mg-Fe columns. The intensity distribution of the atomic columns further indicates that the Fe atomic columns at the octahedral B sites are brighter than the pure Mg atomic columns at the tetrahedral A sites, which is consistent with the theoretical atomic occupation of the ideal normal spinel MFO structure.
In parallel with the observed approach, the specific preferred sites of the dopant Zn atoms in the MFO nanofibers were further imaged. Figure 3a–c show typical HADDF-STEM atomic images of the MZFO nanofibers along the [001], [011] and [111] orientations. It is clear that MZFO nanofibers retain a spinel crystal structure in contrast to the pure MFO nanofibers. Figure 3a shows the atomic STEM image of the MZFO nanofibers projected along the [001] orientation, revealing that the overall atomic image contrast of MZFO nanofibers is essentially the same as that of the MFO nanofibers. It should be noted that the number of atoms on each column is the same in this direction, which presents an ideal situation to identify the preferred sites of dopant atoms. It can be seen that the brightness of the tetrahedral Mg atomic columns is nonetheless darker, but the octahedral Fe atomic column is brighter than that of the MFO. The line intensity distribution was obtained from the blue rectangular box ‘1’. As Figure 3d shows, the difference between the intensities of the two peaks is obviously larger than that of the MFO nanofibers (Figure 2d). Due to the atomic number Z (30) of the Zn atom being greater than the atomic number Z (26) of the Fe atom, the brightness of the octahedral atomic column increases. If the Zn atoms replace the Mg atoms at the tetrahedral sites, the brightness of the atomic column at the tetrahedral sites inevitably increases, thereby reducing the intensity difference between the two peaks. Furthermore, the atomic radius of a Zn ion (1.39 Å) and that of an Fe ion (1.27 Å) are relatively similar, making it easier for atomic substitution to occur. Therefore, it is believed that the dopant Zn atoms randomly replace the Fe atoms at the octahedral sites. The corresponding atomic unit cell model is shown in Figure 3g, in which the blue atoms represent the doping Zn atoms. Figure 3b shows the atomic image along the [011] crystal orientation, showing that the octahedral sites of the atomic column, marked by the yellow dashed circle in the image, are brighter than other octahedral atoms, and its corresponding sites are marked with the numbers ‘2’ and ‘4’ in the atomic unit cell model (Figure 3h). Moreover, the atomic tailing found around the Fe atoms at the octahedral sites, indicated by the yellow dashed rectangular box, may be caused by the irregular occupancy of the doping Zn atoms, which prevented the Zn-Fe atomic columns from being arranged in a straight line. The corresponding positions are indicated by the number ‘5’ in the atomic unit cell model (Figure 3h). Figure 3e is the line intensity distribution diagram shown in the blue rectangle ‘2’ in Figure 3b, and it indicates the specific distribution of the atomic column intensity at the tetrahedral and octahedral sites. Tilting the nanofiber to the [111] crystal axis, shown in Figure 3c, clearly shows that there are six octahedral Fe atomic columns around the Mg-Fe mixed atomic column, forming a regular hexagon, represented by the numbers ‘1–6’ in the atomic unit cell model (Figure 3i). These atomic columns should show the same brightness because each atomic column contains the same number of atoms. However, the octahedral atomic column marked by the yellow dashed circle in Figure 3c is obviously brighter than the surrounding columns. Its atomic unit cell model is indicated by the numbers ‘3’ and ‘6’ in Figure 3i. In addition, the line intensity analysis of the Fe atoms at the octahedral site marked by rectangle ‘3’ in Figure 3c further indicates the specific substitution position of dopant Zn atoms, which match well with the peak intensity distribution shown in Figure 3f and the unit cell model numbered as ‘1’ and ‘6’ atomic columns in Figure 3i. The Fe atom columns with the same number of atoms have different contrasts due to the substitution of Zn atoms, which makes pure Fe atom columns become Zn-Fe mixed-atom columns, further contributing to the increase in the brightness of the atomic column. The above results further prove that the MZFO nanofibers retain a normal spinel crystal structure, and the Zn substitutions occur in the Fe atomic columns at the octahedral sites.
In order to further verify the correlation between crystal structure and magnetic properties, SQUID and MPMS technologies were used to characterize the magnetic properties of the normal spinel MFO nanofibers and Zn-doped MFO nanofibers. The external magnetic field (3 T) dependence on the magnetization of the MZFO and MFO nanofibers at 300 K and 5 K is shown in Figure 4a,b, respectively. A quantitative analysis indicated that the saturation magnetization (Ms) and coercivity (Hc) of MZFO nanofibers at 300 K were 54.7 emu/g (2.30 μB/fu) and 82 Oe, respectively, and the saturation magnetization (Ms) and coercivity (Hc) at 5 K were 92.2 emu/g (3.87 μB/fu) and 145 Oe, respectively, compared to the pure MFO nanofibers with 42.3 emu/g (1.78 μB/fu) 102 Oe at 300 K and 60.8 emu/g (2.55 μB/fu) 167 Oe at 5 K. It was found that the saturation magnetization (μB/f.u.) of MZFO nanofibers increased by 0.52 μB/f.u. and 1.32 μB/f.u. at 300 K and 5 K, respectively, and the coercivity had a slight tendency to increase.
The increase in saturation magnetization Ms and coercivity Hc should relate to the substitution of Zn2+ cations for Fe3+ cations at the octahedral sites, including the occupied sites; canted spin structure; and the super-exchange interaction between the cations and the O2- anions [27], of which the exchange interactions should be the main cause. According to Neel’s theory of the sub-lattice [36], the cations located at the tetrahedral and octahedral sites indicate two sub-lattice positions, A and B, respectively, and the magnetic moments of cations at the tetrahedral A sites and the octahedral B sites are aligned in antiparallel in the spinel structure. The net magnetization (Mnet) is calculated by the following:
Mnet = ∑MB − ∑MA
where ∑MB and ∑MA represent the moment sum of cations located at octahedral sites and tetrahedral sites, respectively. The magnetization of the nanofibers depends on the distribution of the magnetic Fe3+ ions among the two sites because the Mg2+ and Zn2+ ions are non-magnetic.
The exchange interactions between two sub-lattices have different intensities; the antiferromagnetic A-B interaction is stronger than the ferromagnetic B-B interaction and A-A interaction, and the ferromagnetic A-A interaction is the weakest. In the normal spinel MgFe2O4 ferrites, all non-magnetic Mg2+ ions occupy the tetrahedral A sites, and all magnetic Fe3+ ions occupy the octahedral B sites. In the present study, Zn-doped MgFe2O4 ferrites, non-magnetic Zn2+ ions preferred occupying the octahedral Fe3+ site. A small amount of Zn2+ ions substituting Fe3+ ions increased the valence of the surrounding trivalent Fe3+ ions to Fe3+x ions, which may have contributed to the increase in the saturation magnetization. The coercivity was related to magnetocrystalline anisotropy (K1 ~ −2.46 × 104 erg/cm3 for MgFe2O4) and crystallite size in spinel magnetic ferrites. Therefore, the slight increase in coercivity could be attributed to the enhancement in magnetocrystalline anisotropy of ferrites due to the substitutions of Zn2+ ions in MgFe2O4 ferrite. The minor change in coercivity means that Zn doping does not change the soft magnetic properties of MgFe2O4 ferrite.
The Curie temperature of MFO and MZFO nanofibers was further measured in the temperature range of 300–950 K at the applied external magnetic field of 3 T, as shown in Figure 4d. The first order derivative was taken to find out the Tc. It was evident that the Tc of MFO nanofibers was 725 K, which was higher than the value reported for MFO (713 K) in the literature [18]. When the Zn2+ ions were doped, the Tc of MZFO ferrite dropped to 690 K. The substitution of Zn ions for magnetic Fe ions reduced the Fe3+-O2--Fe3+ super-exchange interaction, eventually lowering the Curie temperature. The paramagnetic Curie point Tp of the MFO and MZFO nanofibers were 904 K and 852 K, respectively, as shown in Figure 4d. When T > Tp, the nanofibers were in a completely pure paramagnetic state.

4. Conclusions

In summary, we proved that the MFO nanofibers synthesized by electrospinning have a normal spinel structure by directly observing the occupancy and distribution of cations using Cs-STEM. The atomic HAADF-STEM images, combined with the theoretical atomic model and line intensity distribution, revealed that the dopant Zn2+ cations only randomly substitute for the Fe3+ sites rather than non-magnetic Mg2+ sub-lattices sites. Furthermore, the magnetic characterization showed that the Zn doping affects the spin states of the Fe3+, and the Fe3+-O2−-Fe3+ super-exchange interaction leads to an enhancement in magnetization and a reduction in the Curie temperature. We demonstrated that the doping of cations in spinel ferrites is an effective method to improve magnetic properties through tailoring the spin states of the cations and the occupation sites of the trace dopants, which provides guidance to modify the magnetic properties for the development of the next generation of multi-functional magneto-electric devices.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma15072422/s1.

Author Contributions

X.Z.: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Software, Visualization, Writing—original draft, Writing—review and editing. Z.H.: Formal analysis, Visualization. J.J.: Investigation, Methodology. L.G.: Methodology, Software. J.Z.: Supervision, Formal analysis, Funding acquisition. Y.P.: Supervision, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (52101224, 51771085, 51571104 and 51801087), the Youth Science and Technology Fund Project of Gansu (20JR10RA257), the Innovation Ability Promotion Project of Higher Education Institutions in Gansu (2021B-097), the Youth Science Fund of Lanzhou Jiaotong University (1200060926), the Experimental Teaching Reform Project of Lanzhou Jiaotong University (20210125002), Science and Technology Project of Gansu Province, China (21JR7RA494, 21JR7RA491), Fundamental Research Funds for the Central Universities (lzujbky-2020-58).

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ferre, R.; Ounadjela, K.; George, J.M.; Piraux, L.; Dubois, S. Magnetization processes in nickel and cobalt electrodeposited nanowires. Phys. Rev. B Condens. Matter 1997, 56, 14066–14075. [Google Scholar] [CrossRef]
  2. Satyanarayana, L.; Reddy, K.M.; Manorama, S.V. Nanosized spinel NiFe2O4: A novel material for the detection of liquefied petroleum gas in air. Mater. Chem. Phys. 2003, 82, 21. [Google Scholar] [CrossRef]
  3. Tartaj, P.; Morales, M.D.; Sabino, V.V.; Carreño, T.G.; Serna, C.J. The preparation of magnetic nanoparticles for applications in biomedicine. J. Phys. D Appl. Phys. 2003, 36, 182. [Google Scholar] [CrossRef]
  4. Peng, J.; Hojamberdiev, M.; Xu, Y.; Cao, B.; Wang, J.; Wu, H. Hydrothermal synthesis and magnetic properties of gadolinium-doped CoFe2O4 nanoparticles. J. Magn. Magn. Mater. 2011, 323, 133. [Google Scholar] [CrossRef]
  5. Rana, S.; Gallo, A.; Srivastava, R.S.; Misra, R.D.K. On the suitability of nanocrystalline ferrites as a magnetic carrier for drug delivery: Functionalization, conjugation and drug release kinetics. Acta Biomater. 2007, 3, 233. [Google Scholar] [CrossRef]
  6. Kotnala, R.K.; Shah, J.; Mathpal, M.C.; Gupta, D.; Purohit, L.P.; Kishan, H. Role of modified active surface sites of magnesium ferrite for humidity sensing. J. Optoelect. Adv. Mater. 2009, 11, 296–301. [Google Scholar]
  7. Anantharamaiah, P.N.; Joy, P.A. Tuning of the magnetostrictive properties of cobalt ferrite by forced distribution of substituted divalent metal ions at different crystallographic sites. J. Appl. Phys. 2017, 121, 093904. [Google Scholar] [CrossRef]
  8. Hemeda, O.M.; Mostafa, N.Y.; Abd Elkader, O.H.; Ahmed, M.A. Solubility limits in Mn–Mg ferrites system under hydrothermal conditions. J. Magn. Magn. Mater. 2014, 364, 39–46. [Google Scholar] [CrossRef]
  9. Patil, J.Y.; Mulla, I.S.; Suryavanshi, S.S. Gas response properties of citrate gel synthesized nanocrystalline MgFe2O4: Effect of sintering temperature. Mater. Res. Bull. 2013, 48, 778–784. [Google Scholar] [CrossRef]
  10. Kotnala, R.K.; Shah, J.; Gupta, R. Colossal humidoresistance in ceria added magnesium ferrite thin film by pulsed laser deposition. Sens. Actuat. B 2013, 181, 402–409. [Google Scholar] [CrossRef]
  11. Khan, M.A.; Islam, M.U.; Ishaque, M.; Rahman, I.Z. Magnetic and dielectric behavior of terbium substituted Mg1−xTbxFe2O4 ferrites. J. Alloys Compd. 2012, 519, 156–160. [Google Scholar] [CrossRef]
  12. Iqbal, M.J.; Ahmada, Z.; Melikhov, Y.; Nlebedim, I.C. Effect of Cu–Cr co-substitution on magnetic properties of nanocrystalline magnesium ferrite. J. Magn. Magn. Mater. 2012, 324, 1088–1094. [Google Scholar] [CrossRef]
  13. Wang, Y.C.; Ding, J.; Yi, J.B.; Liu, B.H.; Yu, T.; Shen, Z.X. High-coercivity Co-ferrite thin films on (100)-SiO2 substrate. Appl. Phys. Lett. 2004, 84, 2596–2598. [Google Scholar] [CrossRef] [Green Version]
  14. Sivakumar, N.; Narayanasamy, A.; Shinoda, K.; Chinnasamy, C.N.; Jeyadevan, B.; Greneche, J.M. Electrical and magnetic properties of chemically derived nanocrystalline cobalt ferrite. J. Appl. Phys. 2007, 102, 013916. [Google Scholar] [CrossRef]
  15. Reddy, R.A.; Rao, K.R.; Rajesh Babu, B.; Kumar, G.K.; Rajesh, C.; Chatterjee, A.; Jyothi, N.K. Structural, electrical and magnetic properties of cobalt ferrite with Nd3+ doping. Rare Met. 2019, 41, 240–245. [Google Scholar] [CrossRef]
  16. Xiang, J.; Shen, X.; Zhu, Y. Effects of Ce3+ doping on the structure and magnetic properties of Mn-Zn ferrite fibers. Rare Met. 2009, 28, 151–155. [Google Scholar] [CrossRef]
  17. Reyes-Rodrígues, P.Y.; Cortés-Hernándeza, D.A.; Escobedo-Bocardoa, J.C.; Almanza-Roblesa, J.M.; Sánchez-Fuentesa, H.J.; Jasso-Terána, A.; León-Pradoa, L.E.D.; Méndez-Nonellb, J.; Hurtado-López, G.F. Structural and magnetic properties of MgZn ferrites (Mg1−xZnxFe2O4) prepared by sol-gel method. J. Magn. Magn. Mater. 2017, 427, 268–271. [Google Scholar] [CrossRef]
  18. Ghodake, U.R.; Chaudhari, N.D.; Kambale, R.C.; Patil, J.Y.; Suryavanshi, S.S. Effect of Mn2+ substitution on structural, magnetic, electric and dielectric properties of Mg-Zn ferrites. J. Magn. Magn. Mater. 2016, 407, 60–68. [Google Scholar] [CrossRef]
  19. Fan, W.Q.; Li, M.; Bai, H.Y.; Xu, D.B.; Chen, C.; Li, C.F.; Ge, Y.L.; Shi, W.D. Fabrication of MgFe2O4/MoS2 heterostructure nanowires for photoelectrochemical catalysis. Langmuir 2016, 32, 1629–1636. [Google Scholar] [CrossRef]
  20. West, A.R. Basic Solid State Chemistry, 2nd ed.; John Wiley and Sons Ltd.: New York, NY, USA, 1988. [Google Scholar]
  21. Busca, G.; Finocchio, E.; Lorenzelli, V.; Trombetta, M.; Rossini, S.A. IR study of alkene allylic activation on magnesium ferrite and alumina catalysts. J. Chem. Soc. Faraday Trans. 1996, 92, 4687–4693. [Google Scholar] [CrossRef]
  22. Konishi, K.; Maehara, T.; Kamimori, T.; Aono, H.; Naohara, T.; Kikkawa, H.; Watanabe, Y.; Kawachi, K. Heating ferrite powder with AC magnetic field for thermal coagulation therapy. J. Magn. Magn. Mater. 2004, 2428, 272–276. [Google Scholar] [CrossRef]
  23. Fatemeh, S.M.; Abdol, M.D.; Abbas, R.; Hamid, B.; Orhan, Z.; Barrett, S.D. Structural, magnetic, and in vitro inhibitory characteristics of Ce-substituted MnFe2O4 nanoparticles. Appl. Phys. A 2021, 127, 600. [Google Scholar]
  24. Davarpanah, A.M.; Rahdar, A.; Azizi Dastnae, M.; Zeybek, O.; Beyzaei, H. (1−x)BaFe12O19/xCoFe2O4 hard/soft magnetic nanocomposites: Synthesis, physical characterization, and antibacterial activities study. J. Mol. Struct. 2019, 1175, 445–449. [Google Scholar] [CrossRef]
  25. Gao, J.M.; Cheng, F.Q. Effect of preparation conditions on the structure and magnetic properties of metal-doped magnesium ferrites synthesized from laterite leaching solutions. J. Supercond. Nov. Magn. 2018, 31, 1475–1482. [Google Scholar] [CrossRef]
  26. Rahman, S.; Samanta, S.; Errandonea, D.; Yan, S.; Yang, K.; Lu, J.L.; Wang, L. Pressure-induced structural evaluation and insulator-metal transition in the mixed spinel ferrite. Phys. Rev. B 2017, 95, 024107. [Google Scholar] [CrossRef]
  27. Dwivedi, G.D.; Tseng, F.; Chan, L.; Chatterjee, S.L.; Ghosh, A.K.; Yang, H.D.; Chatterjee, S. Signature of ferroelectricity in magnetically ordered Mo-doped CoFe2O4. Phys. Rev. B 2010, 82, 134428. [Google Scholar] [CrossRef]
  28. Manohar, A.; Krishnamoorthi, C. Photocatalytic study and superparamagnetic nature of Zn-doped MgFe2O4 colloidal size nanocrystals prepared by solvothermal reflux method. J. Photochem. Photobiol. B Biol. 2017, 173, 456–465. [Google Scholar] [CrossRef]
  29. Sawatzky, G.A.; Van Der Woude, F.; Morrish, A.H. Mössbauer study of several ferrimagnetic syinels. Phys. Rev. 1969, 187, 747–757. [Google Scholar] [CrossRef]
  30. Kedem, D.; Rothem, T. Internal fields in nickel ferrite. Phys. Rev. lett. 1967, 18, 165–166. [Google Scholar] [CrossRef]
  31. Wechsler, B.A.; Von Dreele, R.B. Structure refinements of Mg2TiO4, MgTiO3 and MgTi2O5 by time-of-flight neutron powder diffraction. Acta Cryst. 1989, B45, 542–549. [Google Scholar] [CrossRef]
  32. Moyer, J.A.; Vaz, C.A.F.; Negusse, E.; Arena, D.A.; Henrich, V.E. Controlling the electronic structure of Co1−xFe2+xO4 thin films through iron doping. Phys. Rev. B 2011, 83, 035121. [Google Scholar] [CrossRef] [Green Version]
  33. Carta, D.; Casula, M.F.; Falqui, A.; Loche, D.; Mountjoy, G.; Sangregorio, C.; Corrias, A. A structural and magnetic investigation of the inversion degree in ferrite nanocrystals MFe2O4 (M = Mn, Co, Ni). J. Phys. Chem. C 2009, 113, 8606–8615. [Google Scholar] [CrossRef]
  34. Subías, G.; Cuartero, V.; García, J.; Blasco, J.; Lafuerza, S.; Pascarelli, S.; Mathon, O.; Strohm, C.; Nagai, K.; Mito, M.; et al. Investigation of pressure-induced magnetic transitions in CoxFe3−xO4 spinels. Phys. Rev. B 2013, 87, 094408. [Google Scholar] [CrossRef] [Green Version]
  35. Loukya, B.; Negi, D.S.; Dileep, K.; Pachauri, N.; Gupta, A.; Datta, R. Effect of Bloch wave electron propagation and momentum-resolved signal detection on the quantitative and site-specific electron magnetic chiral dichroism of magnetic spinel oxide thin films. Phys. Rev. B 2015, 91, 134412. [Google Scholar] [CrossRef] [Green Version]
  36. Neel, L. Magnetic properties of ferrites: Ferrimagnetism and antiferromagnetism. Ann. Phys. Paris 1948, 3, 137–198. [Google Scholar] [CrossRef]
Figure 1. Crystal structure of normal spinel MFO ferrite. (a) Ball-and-stick model: the cyan and orange spheres represent Fe cations at B sites and Mg cations at A sites, respectively, and O anions are in red. (b) Polyhedral model depicting an FCC cubic network of O anions with Mg2+ and Fe3+ cations. The tetrahedral A sites and octahedral B sites are marked in orange and cyan, respectively.
Figure 1. Crystal structure of normal spinel MFO ferrite. (a) Ball-and-stick model: the cyan and orange spheres represent Fe cations at B sites and Mg cations at A sites, respectively, and O anions are in red. (b) Polyhedral model depicting an FCC cubic network of O anions with Mg2+ and Fe3+ cations. The tetrahedral A sites and octahedral B sites are marked in orange and cyan, respectively.
Materials 15 02422 g001
Figure 2. HAADF-STEM images and perspective views of normal spinel MFO structure. (ac) The HAADF-STEM images of the MFO nanofibers projected along [001], [011] and [111] orientations, respectively. (df) Line intensity profiles for the atomic columns at tetrahedral A sites and octahedral B sites along blue line ‘1’, ‘2’ and ‘3’ in (ac), respectively. (gi) Perspective view of unit cells along [001], [011] and [111] orientations, respectively.
Figure 2. HAADF-STEM images and perspective views of normal spinel MFO structure. (ac) The HAADF-STEM images of the MFO nanofibers projected along [001], [011] and [111] orientations, respectively. (df) Line intensity profiles for the atomic columns at tetrahedral A sites and octahedral B sites along blue line ‘1’, ‘2’ and ‘3’ in (ac), respectively. (gi) Perspective view of unit cells along [001], [011] and [111] orientations, respectively.
Materials 15 02422 g002
Figure 3. HAADF-STEM image and perspective views of Zn-doped normal spinel MFO. (ac) Experimental HAADF-STEM images of MZFO nanofibers observed along the [001], [011] and [111] crystal ribbon axis directions. (df) The linear intensity distribution of the atomic column at tetrahedral A sites and octahedral B sites along the blue lines ‘1’, ‘2’ and ‘3’ in (ac), respectively. (gi) Perspective view of unit cells along [001], [011] and [111] orientations, respectively. The blue balls represent the Zn-doped atoms.
Figure 3. HAADF-STEM image and perspective views of Zn-doped normal spinel MFO. (ac) Experimental HAADF-STEM images of MZFO nanofibers observed along the [001], [011] and [111] crystal ribbon axis directions. (df) The linear intensity distribution of the atomic column at tetrahedral A sites and octahedral B sites along the blue lines ‘1’, ‘2’ and ‘3’ in (ac), respectively. (gi) Perspective view of unit cells along [001], [011] and [111] orientations, respectively. The blue balls represent the Zn-doped atoms.
Materials 15 02422 g003
Figure 4. Magnetic properties of the spinel structure MFO and MZFO single-particle-chain nanofibers. (a) The hysteresis loops of MFO and MZFO single-particle-chain nanofibers measured under an external magnetic field of 3 T and a temperature of 300 K. (b) The hysteresis loops of MFO and MZFO single-particle-chain nanofibers measured under an external magnetic field of 3 T and a temperature of 5 K. (c) The plot of saturation magnetization (μB/f.u.) vs. temperature of MFO and MZFO single-particle-chain nanofibers. (d) The saturation magnetization as a function of temperature measured at the 3 T external magnetic field and the temperature range of 300–950 K to obtain the Curie temperature.
Figure 4. Magnetic properties of the spinel structure MFO and MZFO single-particle-chain nanofibers. (a) The hysteresis loops of MFO and MZFO single-particle-chain nanofibers measured under an external magnetic field of 3 T and a temperature of 300 K. (b) The hysteresis loops of MFO and MZFO single-particle-chain nanofibers measured under an external magnetic field of 3 T and a temperature of 5 K. (c) The plot of saturation magnetization (μB/f.u.) vs. temperature of MFO and MZFO single-particle-chain nanofibers. (d) The saturation magnetization as a function of temperature measured at the 3 T external magnetic field and the temperature range of 300–950 K to obtain the Curie temperature.
Materials 15 02422 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zeng, X.; Hou, Z.; Ju, J.; Gao, L.; Zhang, J.; Peng, Y. The Cation Distributions of Zn-doped Normal Spinel MgFe2O4 Ferrite and Its Magnetic Properties. Materials 2022, 15, 2422. https://doi.org/10.3390/ma15072422

AMA Style

Zeng X, Hou Z, Ju J, Gao L, Zhang J, Peng Y. The Cation Distributions of Zn-doped Normal Spinel MgFe2O4 Ferrite and Its Magnetic Properties. Materials. 2022; 15(7):2422. https://doi.org/10.3390/ma15072422

Chicago/Turabian Style

Zeng, Xue, Zhipeng Hou, Jiaqi Ju, Lei Gao, Junwei Zhang, and Yong Peng. 2022. "The Cation Distributions of Zn-doped Normal Spinel MgFe2O4 Ferrite and Its Magnetic Properties" Materials 15, no. 7: 2422. https://doi.org/10.3390/ma15072422

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