Next Article in Journal
Two-Step Feature Selection for Identifying Developmental Differences in Resting fMRI Intrinsic Connectivity Networks
Previous Article in Journal
Physical and Mechanical Properties of Composite Scaffolds with or without Collagen Impregnation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Properties of NiZn Ferrite Nanofibers Prepared by Electrospinning

1
Department of Advanced Materials Engineering, Gangneung-Wonju National University, Gangneung 25457, Korea
2
Korea Institute of Civil Engineering and Building Technology, Goyang 10223, Korea
3
Research Institute for Dental Engineering, Gangneung-Wonju National University, Gangneung 25457, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(20), 4297; https://doi.org/10.3390/app9204297
Submission received: 3 September 2019 / Revised: 4 October 2019 / Accepted: 10 October 2019 / Published: 12 October 2019
(This article belongs to the Section Materials Science and Engineering)

Abstract

:

Featured Application

Filler for the electromagnetic-interference (EMI) shielding film.

Abstract

When the size of a material is decreased to the nanoscale, the effects of forces that are not influential on a macroscopic scale become increasingly important and the electronic structure is improved. The material then exhibits significantly different physical and chemical properties than in the bulk state. The smaller the size of the material, the more exposure it receives to the nano effects, and the physical properties can be changed via size control. In this study, Ni0.5Zn0.5Fe2O4 ferrite nanofibers were prepared by electrospinning, and the sizes of the prepared samples were controlled to ensure different average diameters by controlling the polymer concentration of the precursor solution. Field emission scanning electron microscope images showed that the samples had average diameters of 224 to 265 nm. The single crystal phase of Ni0.5Zn0.5Fe2O4 and the different crystallite sizes of 13 to 20 nm were confirmed by X-ray diffraction analysis. The magnetization behavior of the samples was measured using a vibrating sample magnetometer and the result confirmed that the samples had different magnetic properties, according to the diameter and crystallite size of the nanofibers. This study suggests that control of magnetic properties and excellent electrical conductivity in a one-dimensional nanostructure can be positively applied to improve the performance of a filler for the electromagnetic-interference shielding film.

Graphical Abstract

1. Introduction

Electronic devices occupy a very important place in modern daily life and the industry. The circuits for these devices have become more portable and functional through increasingly higher levels of integration and complicated structure. Since this trend inevitably makes devices vulnerable to electromagnetic interference, there is a growing demand for simple and efficient electromagnetic wave shielding materials to prevent unnecessary wireless noise. Various studies have reported that, to achieve a thin and light shielding film, a magnetic ceramic filler is more potent than a heavy metal or pyramid-shaped shielding material. Transition metal oxides containing iron have unique electromagnetic, thermal, and mechanical properties, and various studies are underway in areas such as battery anode materials [1,2,3], catalyst materials [4,5], and sensors [6,7]. In particular, soft magnetic spinel ferrite has suitable magnetic properties for a ceramic paint filler to cope with high-frequency environments.
In recent years, there have been many attempts to improve the efficiency of shielding films and some of the methods reported were as follows. Substitute various elements for the bivalent Fe sites in the Fe3O4 spinel ferrite structure [8,9,10,11], control the content of substituted or doped elements [12,13], and optimize the thickness of the shielding film [14,15]. On the other hand, excellent characteristics of a one-dimensional structure like porosity, a large specific area, and improved charge mobility have been attracting more attention because of their potential value in various applications, such as solar cells [16,17], photocatalysts [18,19,20], biosensors, and medical tissues [21,22,23]. Since the improved carrier mobility is considered to have an application value in the field of electromagnetic wave shielding film, several fillers with a one-dimensional nanostructure have been proposed [24,25,26]. To produce a one-dimensional nanostructure, methods such as hydrothermal [27,28], template [29], and anodic oxidation [30,31] are mainly used, and electrospinning have an advantage in that they can fabricate a uniform nanofiber rapidly even with a simple apparatus configuration [32]. At the same time, since the magnetization behavior of the material is closely related to the shielding efficiency and the use of the frequency band, many attempts have been made to control it. Research has focused not only on applying the one-dimensional nanostructure for a geometric advantage but also on optimizing the magnetic behavior by controlling the crystallite size. Some conventional studies have attempted to achieve this goal by increasing or decreasing the crystallite size or nanofiber diameter by controlling the heat treatment temperature [33,34,35]. NiZn ferrites are known to have high saturation magnetization and chemical stability, and attempts have been made to produce NiZn ferrite nanofibers and apply them as microwave absorbers [8,36]. However, there are no reports on how the magnetization behavior in NiZn ferrite nanofibers can be changed by complex microstructural changes, such as diameter and crystallite size resulting from the polymer content in the electrospinning precursor solution.
In this study, we fabricated NiZn ferrite nanofibers by the electrospinning method and controlled the microstructures, the crystallite size, and the nanofiber diameter, by polymer content in the precursor solution. As-spun nanofibers with a mixed composition of transition metal and polymer were calcined at suitable conditions and the calcined oxide nanofibers were investigated to explain the magnetization behavior of the nanofibers with various microstructures. The oxide nanofibers created different average diameters by controlling the polymer content, and the average diameter difference was measured by field emission scanning electron microscope (FE-SEM) image analysis. The presence of the same spinel crystal phase in each sample was confirmed by X-ray diffraction analysis (XRD) and a vibrating sample magnetometer (VSM) measured the magnetization behavior, according to the average diameter change of the nanofibers. Along with the doped and substituted elements, the particle and crystallite size of the structures were the major factors that governed the magnetic properties of the material. Utilizing this point, we found that the frequency band and efficiency of the shielding material could be finely controlled.

2. Materials and Methods

2.1. Fabrication of NiZn Ferrite Nanofibers via Electrospinning

The nanofibers were prepared by a typical electrospinning process as follows. Ni(NO3)2·6H2O (13478-00-7, 98%, EP, Samchun Chemicals, Seoul, Korea), Zn(NO3)2·6H2O (10196-18-6, 98%, EP, Daejung Chemicals, Siheung, Korea), and Fe(NO3)2·9H2O (7782-61-8, 99%, GR, Kanto Chemical, Tokyo, Japan) in a molar ratio of 1:1:4 were mixed in N,N-dimethylmethanamide (DMF, 68-12-2, 99.5%, EP, Daejung chemicals, Siheung, Korea) as a solvent at a 150 mmol/L of molar concentration. The nitrate crystals were mixed using a magnetic stirrer until they were completely dissolved in DMF solvent. Polyvinyl pyrrolidone (PVP, 9003-39-8, 98.3%, M.W. 1,300,000, Alfa Aesar Co., Tewsbury, MA, US) was added to the completely dissolved solution at a weight ratio of 18 wt%, 20 wt%, and 22 wt% and stirred for 48 h to prepare the electrospinning precursor solution. The precursor solutions were labeled A, B, and C, and the compositions of the precursors are shown in Table 1. The prepared precursor was placed in a polypropylene syringe, connected with polyethylene tubing, and mounted on a syringe pump. A stainless-steel nozzle adapter and a 27-gauge plastic nozzle were attached to the end of the tubing, and pressure was applied through the pump to infuse at a constant flow rate of 0.5 mL/h. The nozzle adapter was connected to the power supply, and the aluminum foil collector was grounded. A voltage of 15 kV was applied at the electrode while maintaining a 15-cm electrode-to-electrode distance. The relative humidity was 40% for room temperature. The collected composite nanofiber sheets were dried in a dry oven at 80 °C for 2 h. The dried composite nanofiber sheet was heated to a temperature of 750 °C at a rate of 5 °C/min atmospheric environment, fixed for 4 h, and then cooled to obtain a ceramic nanofiber with a Ni0.5Zn0.5Fe2O4 ferrite single phase. The morphologies and diameters of the nanofiber samples were confirmed by field emission-scanning electron microscopes (FE-SEM) image analysis. The heat treatment temperature of the composite fiber was determined by confirming the pyrolysis behavior using thermogravimetric analysis. The crystal phase was determined by comparing the XRD pattern with the Crystallography Open Database. VSM measured the hysteresis loop of each average diameter sample.

2.2. Characterization

Graphs of pyrolysis behavior of PVP/metal nitrate composite nanofibers were obtained by thermo-gravimetric analysis (TGA, Q500, TA Instruments Co., New Castle, DE, US). The weight change was measured while heating from 20 °C to 750 °C in air at a heating rate of 5 °C/min.
The crystal structure of the NiZn ferrite nanofibers after calcination was analyzed with an XRD diffraction pattern. XRD analysis was carried out by a diffractometer (AXS-D8, Bruker Co., US) and measured at 40 kV – 40 mA current conditions with a Cu Kα radiation source. The measurements were performed at an analysis rate of 0.02°/s from 2θ = 20° to 80°.
The surface morphology and diameter of the calcined NiZn ferrite nanofiber samples were observed by FE-SEM (SU-70, Hitachi Co., Niigata, Japan). Images were collected at an applied voltage of 30 kV and a working distance of 10 mm. In addition, energy dispersive X-ray spectroscopy (EDX, Octane Elite EDS system, EDAX Inc., Mahwah, NJ, USA) and elemental mapping were carried out to confirm the distribution of the elements in the NiZn ferrite nanofibers.
The magnetization behavior of each sample was analyzed at room temperature through VSM. The applied magnetic field was measured from −15 kOe to 15 kOe, the data point was measured once per 500 Oe, the time constant was 0.1, and the analysis time was 1 s/pt.

3. Results

For the heat treatment process, Figure 1 shows the pyrolysis behavior graphs of the PVP/metal nitrate composite nanofiber sheet that were obtained by TGA. In the curve representing the weight to temperature, the weight of about 15 wt% sharply decreased until the temperature of the sample reached about 73 °C. This was due to the loss of moisture absorbed during the process. From this fact, the required temperature of the drying process was calculated at 80 °C immediately after the electrospinning process. The gradual loss curve from 150 °C to 420 °C indicates a range corresponding to the decomposition of PVP and the volatilization of the residual solvent. The mass loss range of pure PVP is 300–600 °C, but the decomposition temperature range of the metal salt and hybrid electro-spun nanofiber sheets shifted to a slightly lower temperature [37]. It is generally considered that the high thermal conductivity of the metal ions between the polymer chains promotes thermal decomposition. At 420 ° C, the mass reduction decreases and then begins to decrease again at 570 °C, which is the point where the NiZn ferrite fiber begins to crystallize into the spinel structure [38]. The reaction ends at 685 °C with a final mass loss of 88 wt%. Regarding these experimental results, the heat treatment temperature of the nanofibers used in the analysis was determined to be 750 °C.
The electro-spun samples from precursors A, B, and C were labeled and analyzed as samples A, B, and C. XRD analysis was used to identify the crystalized structure of the final samples. Figure 2 shows the diffraction patterns of samples A, B, and C calcined at 750 °C for 4 h. Despite the differences in polymer content, the diffraction pattern for the crystalized structure was the same for all three samples. The significant diffraction peaks indexed as (202), (131), (040), (242), (151), (404), and (444), which confirmed that the spinel structure was properly formed. The diffraction pattern was indexed by comparing with the 900-9920 entry of the Crystallography Open Database.
Table 2 shows the results obtained by calculating the XRD diffraction pattern using the Scherrer formula (d = 0.94λ / B cos θ). In this formula, d represents the crystallite size, B is the full width at half maximum (FWHM) of each XRD peak, and λ is the Cu Kα wavelength (0.15478 nm). The table shows the average crystallite size obtained using the peak intensities at (131), (151), and (404) diffraction planes per sample. The crystallite size calculated by the Scherer formula tends to decrease with an increasing PVP content of the precursor solution. According to Xiaolei Song et al., nanofibers with high polymer content in the calcination process increase the size and number of pores and cracks. These defects prevent the growth of a nuclear crystal and make the diffusion in the boundary of primary crystallites difficult [39]. The difference in crystallinity can also be confirmed by the width of the FWHM.
Once all the components except for the metallic element are volatilized during calcination, the fibrous structure may not be maintained even if a suitable crystalized structure is formed. Therefore, it is important to check the morphology through a microscopic observation. Figure 3a, Figure 3b, and Figure 3c are FE-SEM images of the as-spun nanofibers of samples A, B, and C, respectively, and Figure 3d, Figure 3e, and Figure 3f show the calcined Ni0.5Zn0.5Fe2O4 nanofibers. Although as-spun nanofibers were shrunk by calcination, the morphology of isotropic fibers was maintained without breaking the microstructure.
Figure 4 shows the EDX mapping results of the nanofiber samples. Figure 3a, Figure 3b, and Figure 3c show the emission distribution of the characteristic X-ray of Fe, Ni, and Zn, respectively. A comparison of the distribution of each element with the SEM image shown in Figure 4d, confirmed that each component was evenly distributed at the position where the nanofiber existed. The density of each pixel was only related to the degree of the nanofiber overlap. The uniform distribution of each element can support XRD analysis results showing a single crystal structure. The EDX spectrum analysis of the mapping results is shown in Table 3. Due to the characteristics of the electron beam source, high analytical reliability for the magnetic powders cannot be expected, but the chemical composition ratio was very similar to the Ni0.5Zn0.5Fe2O4 phase.
The diameters of Ni0.5Zn0.5Fe2O4 nanofibers were measured with FE-SEM images. To obtain quantification uniformity, 10 specimens for each sample A, B, and C were prepared and FE-SEM analysis was conducted. Four FE-SEM images were taken from each specimen and 5 points in the image with up, down, left, right, and center parts, were selected to measure the diameter. Figure 5 shows a histogram on the distribution of nanofiber diameters obtained by making 200 measurements of samples A, B, and C. The histogram shows that the average diameter of the samples was formed with a constant difference of about 20 nm within the range of 224 to 265 nm. This indicates that the diameter of the final oxide nanofibers could be controlled by adjusting the polymer content. Table 4 shows the average, maximum, minimum, and standard deviation of the diameters of each sample.
Figure 6 shows the magnetization behavior of nanofiber samples with different average diameters measured by VSM. In general, the crystalized structure, chemical composition, oxygen vacancies, particle shape, and the interaction of the domain wall with various defects are considered to be the parameters of saturation magnetization [40,41,42]. Among them, the presence of non-aligned surface spin derived to thermal fluctuation is regarded as the main factor to reduce saturation magnetization, and it is related to the specific surface area [43]. In general, specific surface area is considered the main factor of the magnetization of nanoparticles due to the non-alignment of thermally fluctuating spin. Since one-dimensional nanofibers are a continuum of the smaller crystals, the influence of crystallite size and diameter must be considered. It is known that the saturation magnetization (Ms) increases with the crystallite size of a crystal [44]. The crystallite sizes of our samples were estimated by the Scherrer equation. The crystallite sizes of samples A, B, and C were 20 nm, 16 nm, and 13 nm, respectively. According to this, the Ms of sample A is the largest and should be gradually decreased in the order of samples B and C, but samples B and C were almost at the same levels. The saturation magnetization can be affected by several parameters, such as calcination temperature, crystal size, and crystal structure. Herranz et al. [45] reported that Co-doped LaSrTi oxide films obtained at low oxygen pressures had a higher Ms value than those obtained at high oxygen pressure. Sample C, which had the highest PVP content, must burn more PVP during the calcination process. PVP combustion decreases the oxygen pressure, which causes more oxygen vacancies in the crystalized structure in the formation process. In sample C, the decrease in Ms resulting from the small crystallite size can be compensated by the oxygen vacancies effect. The concentration of the oxygen vacancy in the samples will be measured and discussed in a future study.
To verify the values of the magnetic property of samples A, B, and C, the saturation magnetization, coercivity, and remnant magnetization of each sample were summarized and compared with previous reports on NiZn ferrite nanofibers and nanoparticles. The results are shown in Table 5 [8,46,47,48,49]. Although the range of saturation magnetization and coercivity change varies depending on the chemical composition, such as the content of Ni and Zn ions and the heat treatment temperature, the values of samples A, B, and C were in the general range, which means that samples A, B, and C were properly fabricated by electrospinning and the values of the magnetic property could be controlled. In the previous reports [8,46,47,48,49], the saturation magnetization and coercivity of NiZn ferrite nanofibers tended to be larger than in nanoparticles, which is thought to be due to the shape magnetic anisotropy and the small influence of thermal fluctuations.

4. Conclusions

Ni0.5Zn0.5Fe2O4 ferrite nanofibers were fabricated by electrospinning with a precursor solution of PVP, DMF, metal nitrate, and heat treatment. Since the nanofibers have a smaller specific surface area than the 0-dimensional nanoparticles, we confirmed that the diameter could not be a major factor of Ms value when the difference of the samples was not large. Rather, a large amount of polymer added for the diameter control decreased the crystallite size during the calcination process and caused an unexpected change in the Ms value. On the other hand, we confirmed that, when the diameter difference is not significant, the crystallite size can be a factor leading to the Ms value tendency. The ability to tune the constants related to magnetic forces when making a film for electromagnetic interference shielding with a magnetic ceramic filler means that additional channels are available for optimizing the shielding efficiency and the frequency band. Moreover, the change of the crystallite size, according to the polymer content, provides a clue to controlling the magnetization behavior more economically than the conventional method with the heat treatment temperature control. It is possible to obtain the maximum shielding effect at the same cost by applying a combination of research themes of the size-dependent magnetic characteristics, development of a ferrite composition with excellent magnetic characteristics, and study of thickness to maximize impedance matching.

Author Contributions

Conceptualization, K.-H.N. and W.-Y.C. Methodology, K.-H.N. Validation, K.-H.N. and W.-T.K. Formal analysis, K.-H.N. Investigation, K.-H.N. and W.-T.K. Resources, T.-H.S. and W.-Y.C. Data curation, K.-H.N. Writing—original draft preparation, K.-H.N. Writing—review and editing, W.-Y.C. Visualization, K.-H.N. and W.-T.K. Supervision, W.-Y.C. Project administration, T.-H.S. and W.-Y.C. Funding acquisition, T.-H.S. and W.-Y.C.

Funding

The Korea Agency for Infrastructure Technology Advancement under Construction Technology R&D project (Grant No. 18SCIP-B146646-01) and National Research Foundation of Korea (Grant No. 2019R1I1A3A01057765) supported this work.

Acknowledgments

The authors thank the researchers (Sung-Wook Kim) of the Korea Institute of Civil Engineering and Building Technology, for their time and contributions to the study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chu, Y.-Q.; Fu, Z.-W.; Qin, Q.-Z. Cobalt ferrite thin films as anode material for lithium ion batteries. Electrochim. Acta 2004, 49, 4915–4921. [Google Scholar] [CrossRef]
  2. Ding, Y.; Yang, Y.; Shao, H. High capacity ZnFe2O4 anode material for lithium ion batteries. Electrochim. Acta 2011, 56, 9433–9438. [Google Scholar] [CrossRef]
  3. He, C.; Wu, S.; Zhao, N.; Shi, C.; Liu, E.; Li, J. Carbon-encapsulated Fe3O4 nanoparticles as a high-rate lithium ion battery anode material. ACS Nano 2013, 7, 4459–4469. [Google Scholar] [CrossRef] [PubMed]
  4. Chagas, C.A.; de Souza, E.F.; de Carvalho, M.C.; Martins, R.L.; Schmal, M. Cobalt ferrite nanoparticles for the preferential oxidation of CO. Appl. Catal. A Gen. 2016, 519, 139–145. [Google Scholar] [CrossRef]
  5. Wan, L.S. A Study into Photocatalytic Degradation of Methylene Blue and Glycerol Aqueous Solution over Copper Ferrite Catalyst. Ph.D. Thesis, UMP, Gambang, Malaysia, 2015. [Google Scholar]
  6. Hankiewicz, J.H.; Alghamdi, N.; Hammelev, N.M.; Anderson, N.R.; Camley, R.E.; Stupic, K.; Przybylski, M.; Zukrowski, J.; Celinski, Z.J. Zinc doped copper ferrite particles as temperature sensors for magnetic resonance imaging. Aip Adv. 2017, 7, 056703. [Google Scholar] [CrossRef]
  7. Dumitrescu, A.; Lisa, G.; Iordan, A.; Tudorache, F.; Petrila, I.; Borhan, A.; Palamaru, M.; Mihailescu, C.; Leontie, L.; Munteanu, C. Ni ferrite highly organized as humidity sensors. Mater. Chem. Phys. 2015, 156, 170–179. [Google Scholar] [CrossRef]
  8. Huang, X.; Zhang, J.; Lai, M.; Sang, T. Preparation and microwave absorption mechanisms of the NiZn ferrite nanofibers. J. Alloys Compd. 2015, 627, 367–373. [Google Scholar] [CrossRef]
  9. Huang, X.; Zhang, J.; Xiao, S.; Chen, G. The cobalt zinc spinel ferrite nanofiber: Lightweight and efficient microwave absorber. J. Am. Ceram. Soc. 2014, 97, 1363–1366. [Google Scholar] [CrossRef]
  10. Wu, Y.; Han, M.; Deng, L. Enhancing the microwave absorption properties of Fe–Cu–Nb–Si–B nanocomposite flakes by coating with spinel ferrite NiFe2O4. IEEE Trans. Magn. 2015, 51, 1–4. [Google Scholar] [CrossRef]
  11. Duan, H.-Z.; Zhou, F.-L.; Cheng, X.; Chen, G.-H.; Li, Q.-L. Preparation of hollow microspheres of Ce3+ doped NiCo ferrite with high microwave absorbing performance. J. Magn. Magn. Mater. 2017, 424, 467–471. [Google Scholar] [CrossRef]
  12. Ren, X.; Xu, G. Electromagnetic and microwave absorbing properties of NiCoZn-ferrites doped with La3+. J. Magn. Magn. Mater. 2014, 354, 44–48. [Google Scholar] [CrossRef]
  13. Ghodake, J.; Kambale, R.C.; Shinde, T.; Maskar, P.; Suryavanshi, S. Magnetic and microwave absorbing properties of Co2+ substituted nickel–zinc ferrites with the emphasis on initial permeability studies. J. Magn. Magn. Mater. 2016, 401, 938–942. [Google Scholar] [CrossRef]
  14. Huang, X.; Zhang, J.; Rao, W.; Sang, T.; Song, B.; Wong, C. Tunable electromagnetic properties and enhanced microwave absorption ability of flaky graphite/cobalt zinc ferrite composites. J. Alloys Compd. 2016, 662, 409–414. [Google Scholar] [CrossRef]
  15. Zhao, D.-L.; Lv, Q.; Shen, Z.-M. Fabrication and microwave absorbing properties of Ni–Zn spinel ferrites. J. Alloys Compd. 2009, 480, 634–638. [Google Scholar] [CrossRef]
  16. Yu, K.; Chen, J. Enhancing solar cell efficiencies through 1-D nanostructures. Nanoscale Res. Lett. 2009, 4, 1. [Google Scholar] [CrossRef]
  17. Dharani, S.; Mulmudi, H.K.; Yantara, N.; Trang, P.T.T.; Park, N.G.; Graetzel, M.; Mhaisalkar, S.; Mathews, N.; Boix, P.P. High efficiency electrospun TiO2 nanofiber based hybrid organic–inorganic perovskite solar cell. Nanoscale 2014, 6, 1675–1679. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, Z.; Shao, C.; Li, X.; Wang, C.; Zhang, M.; Liu, Y. Electrospun nanofibers of p-type NiO/n-type ZnO heterojunctions with enhanced photocatalytic activity. ACS Appl. Mater. Interfaces 2010, 2, 2915–2923. [Google Scholar] [CrossRef]
  19. Choi, S.K.; Kim, S.; Lim, S.K.; Park, H. Photocatalytic comparison of TiO2 nanoparticles and electrospun TiO2 nanofibers: Effects of mesoporosity and interparticle charge transfer. J. Phys. Chem. C 2010, 114, 16475–16480. [Google Scholar] [CrossRef]
  20. Chuangchote, S.; Jitputti, J.; Sagawa, T.; Yoshikawa, S. Photocatalytic activity for hydrogen evolution of electrospun TiO2 nanofibers. ACS Appl. Mater. Interfaces 2009, 1, 1140–1143. [Google Scholar] [CrossRef]
  21. Pham, Q.P.; Sharma, U.; Mikos, A.G. Electrospinning of polymeric nanofibers for tissue engineering applications: A review. Tissue Eng. 2006, 12, 1197–1211. [Google Scholar] [CrossRef]
  22. Ma, Z.; Kotaki, M.; Inai, R.; Ramakrishna, S. Potential of nanofiber matrix as tissue-engineering scaffolds. Tissue Eng. 2005, 11, 101–109. [Google Scholar] [CrossRef] [PubMed]
  23. Sill, T.J.; von Recum, H.A. Electrospinning: Applications in drug delivery and tissue engineering. Biomaterials 2008, 29, 1989–2006. [Google Scholar] [CrossRef] [PubMed]
  24. Li, C.-J.; Wang, B.; Wang, J.-N. Magnetic and microwave absorbing properties of electrospun Ba(1−x)LaxFe12O19 nanofibers. J. Magn. Magn. Mater. 2012, 324, 1305–1311. [Google Scholar] [CrossRef]
  25. Zhang, T.; Huang, D.; Yang, Y.; Kang, F.; Gu, J. Fe3O4/carbon composite nanofiber absorber with enhanced microwave absorption performance. Mater. Sci. Eng. B 2013, 178, 1–9. [Google Scholar] [CrossRef]
  26. Han, R.; Li, W.; Pan, W.; Zhu, M.; Zhou, D.; Li, F.-S. 1D magnetic materials of Fe3O4 and Fe with high performance of microwave absorption fabricated by electrospinning method. Sci. Rep. 2014, 4, 7493. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, J.; Yu, H.; Cheng, B.; Zhao, X.; Zhang, Q. Preparation and photocatalytic activity of mesoporous anatase TiO2 nanofibers by a hydrothermal method. J. Photochem. Photobiol. A Chem. 2006, 182, 121–127. [Google Scholar] [CrossRef]
  28. Yu, J.; Kudo, A. Hydrothermal synthesis of nanofibrous bismuth vanadate. Chem. Lett. 2005, 34, 850–851. [Google Scholar] [CrossRef]
  29. Cao, G.; Liu, D. Template-based synthesis of nanorod, nanowire, and nanotube arrays. Adv. Colloid Interface Sci. 2008, 136, 45–64. [Google Scholar] [CrossRef]
  30. Kim, W.-T.; Kim, I.-H.; Choi, W.-Y. Fabrication of TiO2 nanotube arrays and their application to a gas sensor. J. Nanosci. Nanotechnol. 2015, 15, 8161–8165. [Google Scholar] [CrossRef]
  31. Gong, D.; Grimes, C.A.; Varghese, O.K.; Hu, W.; Singh, R.; Chen, Z.; Dickey, E.C. Titanium oxide nanotube arrays prepared by anodic oxidation. J. Mater. Res. 2001, 16, 3331–3334. [Google Scholar] [CrossRef] [Green Version]
  32. Barhoum, A.; Pal, K.; Rahier, H.; Uludag, H.; Kim, I.S.; Bechelany, M. Nanofibers as new-generation materials: From spinning and nano-spinning fabrication techniques to emerging applications. Appl. Mater. Today 2019, 17, 1–35. [Google Scholar] [CrossRef]
  33. Sonsupap, S.; Kidkhunthod, P.; Chanlek, N.; Pinitsoontorn, S.; Maensiri, S. Fabrication, structure, and magnetic properties of electrospun Ce0.96Fe0.04O2 nanofibers. Appl. Surf. Sci. 2016, 380, 16–22. [Google Scholar] [CrossRef]
  34. Saensuk, O.; Phokha, S.; Bootchanont, A.; Maensiri, S.; Swatsitang, E. Fabrication and magnetic properties of NiFe2O4 nanofibers obtained by electrospinning. Ceram. Int. 2015, 41, 8133–8141. [Google Scholar] [CrossRef]
  35. Ponhan, W.; Kadtajun, S.; Khetchompoo, I.; Chinnasa, P. Fabrication and magnetic properties of Co0.05LaTi0.95O3 nanofibers. J. Phys. Conf. Ser. 2018, 1144, 012066. [Google Scholar] [CrossRef]
  36. Wang, C.; Shen, Y.; Wang, X.; Zhang, H.; Xie, A. Synthesis of novel NiZn-ferrite/Polyaniline nanocomposites and their microwave absorption properties. Mater. Sci. Semicond. Process. 2013, 16, 77–82. [Google Scholar] [CrossRef]
  37. Ali, N.N.; Atassi, Y.; Salloum, A.; Charba, A.; Malki, A.; Jafarian, M. Comparative study of microwave absorption characteristics of (Polyaniline/NiZn ferrite) nanocomposites with different ferrite percentages. Mater. Chem. Phys. 2018, 211, 79–87. [Google Scholar] [CrossRef]
  38. Streckova, M.; Mudra, E.; Sebek, M.; Sopcak, T.; Dusza, J.; Kovac, J. Preparation and Investigations of Ni0.2Zn0.8Fe2O4 Ferrite Nanofiber Membranes by Needleless Electrospinning Method. Acta Phys. Pol. A. 2017, 131. [Google Scholar]
  39. Song, X.; Liu, W.; Wang, J.; Xu, S.; Liu, B.; Liu, J.; Ma, Y. Microstructural differences between electrospun alumina borate nanofibers prepared by solutions with different PVP contents. Ceram. Int. 2017, 43, 9831–9837. [Google Scholar] [CrossRef]
  40. Chen, X.; Feng, C.; Wu, Z.L.; Yang, F.; Liu, Y.; Jiang, S.; Li, M.H.; Yu, G.H. Interfacial oxygen migration and its effect on the magnetic anisotropy in Pt/Co/MgO/Pt films. Appl. Phys. Lett. 2014, 104, 052413. [Google Scholar] [CrossRef]
  41. Kumar, L.; Kar, M. Influence of Al3+ ion concentration on the crystal structure and magnetic anisotropy of nanocrystalline spinel cobalt ferrite. J. Magn. Magn. Mater. 2011, 323, 2042–2048. [Google Scholar] [CrossRef]
  42. Jaffari, G.H.; Rumaiz, A.; Woicik, J.; Shah, S.I. Influence of oxygen vacancies on the electronic structure and magnetic properties of NiFe2O4 thin films. J. Appl. Phys. 2012, 111, 093906. [Google Scholar] [CrossRef]
  43. Sun, Y.; Ji, G.; Zheng, M.; Chang, X.; Li, S.; Zhang, Y. Synthesis and magnetic properties of crystalline mesoporous CoFe2O4 with large specific surface area. J. Mater. Chem. 2010, 20, 945–952. [Google Scholar] [CrossRef]
  44. Gong, W.; Li, H.; Zhao, Z.; Chen, J. Ultrafine particles of Fe, Co, and Ni ferromagnetic metals. J. Appl. Phys. 1991, 69, 5119–5121. [Google Scholar] [CrossRef]
  45. Herranz, G.; Ranchal, R.; Bibes, M.; Jaffres, H.; Jacquet, E.; Maurice, J.-L.; Bouzehouane, K.; Wyczisk, F.; Tafra, E.; Basletic, M. Co-doped (La, Sr)TiO3δ: A high Curie temperature diluted magnetic system with large spin polarization. Phys. Rev. Lett. 2006, 96, 027207. [Google Scholar] [CrossRef] [PubMed]
  46. Xiang, J.; Shen, X.Q.; Song, F.Z.; Liu, M.Q. Fabrication and magnetic properties of Ni0.5Zn0.5Fe2O4 nanofibres by electrospinning. Chin. Phys. B 2009, 18, 4960. [Google Scholar]
  47. Fu, J.; Zhang, J.; Peng, Y.; Zhao, C.; He, Y.; Zhang, Z.; Pan, X.; Mellors, N.J.; Xie, E. Wire-in-tube structure fabricated by single capillary electrospinning via nanoscale Kirkendall effect: The case of nickel–zinc ferrite. Nanoscale 2013, 5, 12551–12557. [Google Scholar] [CrossRef] [PubMed]
  48. Morrison, S.A.; Cahill, C.L.; Carpenter, E.E.; Calvin, S.; Swaminathan, R.; McHenry, M.E.; Harris, V.G. Magnetic and structural properties of nickel zinc ferrite nanoparticles synthesized at room temperature. J. Appl. Phys. 2004, 95, 6392–6395. [Google Scholar] [CrossRef] [Green Version]
  49. Mahmoodi, N.M.; Bashiri, M.; Moeen, S.J. Synthesis of nickel–zinc ferrite magnetic nanoparticle and dye degradation using photocatalytic ozonation. Mater. Res. Bull. 2012, 47, 4403–4408. [Google Scholar] [CrossRef]
Figure 1. TGA curve of PVP/metal nitrate composite nanofibers.
Figure 1. TGA curve of PVP/metal nitrate composite nanofibers.
Applsci 09 04297 g001
Figure 2. X-ray diffraction patterns of the PVP/metal nitrate composite nanofibers calcined at 750 °C.
Figure 2. X-ray diffraction patterns of the PVP/metal nitrate composite nanofibers calcined at 750 °C.
Applsci 09 04297 g002
Figure 3. Field emission-scanning electron microscopes (FE-SEM) images of as-spun and calcined Ni0.5Zn0.5Fe2O4 nanofibers obtained from various samples: (a) As-spun nanofibers of sample A, (b) As-spun nanofibers of sample B, (c) As-spun nanofibers of sample C, (d) Calcined Ni0.5Zn0.5Fe2O4 nanofibers of sample A, (e) Calcined Ni0.5Zn0.5Fe2O4 nanofibers of sample B, and (d) Calcined Ni0.5Zn0.5Fe2O4 nanofibers of sample C.
Figure 3. Field emission-scanning electron microscopes (FE-SEM) images of as-spun and calcined Ni0.5Zn0.5Fe2O4 nanofibers obtained from various samples: (a) As-spun nanofibers of sample A, (b) As-spun nanofibers of sample B, (c) As-spun nanofibers of sample C, (d) Calcined Ni0.5Zn0.5Fe2O4 nanofibers of sample A, (e) Calcined Ni0.5Zn0.5Fe2O4 nanofibers of sample B, and (d) Calcined Ni0.5Zn0.5Fe2O4 nanofibers of sample C.
Applsci 09 04297 g003
Figure 4. Energy dispersive X-ray spectroscopy (EDX) elemental mapping of Ni0.5Zn0.5Fe2O4 nanofibers: (a) Fe, (b) Ni, (c) Zn, and (d) FE-SEM image.
Figure 4. Energy dispersive X-ray spectroscopy (EDX) elemental mapping of Ni0.5Zn0.5Fe2O4 nanofibers: (a) Fe, (b) Ni, (c) Zn, and (d) FE-SEM image.
Applsci 09 04297 g004
Figure 5. Histogram of the size distribution of Ni0.5Zn0.5Fe2O4 nanofibers.
Figure 5. Histogram of the size distribution of Ni0.5Zn0.5Fe2O4 nanofibers.
Applsci 09 04297 g005
Figure 6. Magnetization versus magnetic field curves (M-H hysteresis loops) of Ni0.5Zn0.5Fe2O4 nanofibers.
Figure 6. Magnetization versus magnetic field curves (M-H hysteresis loops) of Ni0.5Zn0.5Fe2O4 nanofibers.
Applsci 09 04297 g006
Table 1. The composition and content of the precursor solution. PVP: Polyvinyl pyrrolidone; DMF: N,N-dimethylmethanamide.
Table 1. The composition and content of the precursor solution. PVP: Polyvinyl pyrrolidone; DMF: N,N-dimethylmethanamide.
PrecursorsNi(NO3)2·6H2O
(g)
Zn(NO3)2·6H2O
(g)
Fe(NO3)2·9H2O
(g)
PVP
(g)
DMF
(g)
Precursor A1.451.484.0418100
Precursor B20
Precursor C22
Table 2. Average crystallite size and XRD peak broadening data of each sample. FWHM: full width at half maximum.
Table 2. Average crystallite size and XRD peak broadening data of each sample. FWHM: full width at half maximum.
SamplesPlane (hkl) 2θ (°)FWHM (°)Average Crystallite Size (nm)
Sample A(131)35.250.3320 ± 6
(151)56.720.58
(404)62.470.56
Sample B(131)35.600.4416 ± 3
(151)56.760.69
(404)62.820.68
Sample C(131)35.630.5813 ± 2
(151)56.780.87
(404)62.810.82
Table 3. Energy dispersive X-ray spectroscopy (EDX) spectrum analysis results of Ni0.5Zn0.5Fe2O4 nanofibers.
Table 3. Energy dispersive X-ray spectroscopy (EDX) spectrum analysis results of Ni0.5Zn0.5Fe2O4 nanofibers.
ElementWeight Percent (%) Atomic Percent (%)Net IntensityError (%)
Fe45.8628.07270.801.93
Ni12.597.3355.703.08
Zn14.987.8349.503.08
O26.5756.7700
Table 4. Maximum, minimum, and average diameters, and standard deviations of Ni0.5Zn0.5Fe2O4 nanofibers measured from samples A, B, and C.
Table 4. Maximum, minimum, and average diameters, and standard deviations of Ni0.5Zn0.5Fe2O4 nanofibers measured from samples A, B, and C.
Ni0.5Zn0.5Fe2O4 Nanofibers
SamplesMaximum Diameter (nm)Minimum Diameter (nm)Average Diameter (nm)Standard Deviation (nm)
Sample A355.02112.9224.2645.42
Sample B391.68139.66245.4855.78
Sample C402.52157.54266.4239.87
Table 5. Saturation magnetization, coercivity, and remnant magnetization of Ni0.5Zn0.5Fe2O4 nanofibers measured from samples A, B, and C.
Table 5. Saturation magnetization, coercivity, and remnant magnetization of Ni0.5Zn0.5Fe2O4 nanofibers measured from samples A, B, and C.
SamplesSaturation Magnetization
(emu/g)
Coercivity (Oe)Remnant Magnetization
(emu/g)
Sample A56.65153.129.82
Sample B48.09150.417.37
Sample C48.29169.459.32
NiZn ferrite Nanofibers [8,46,47]37–90.243–175.3-
NiZn ferrite Nanoparticles [48,49]23.95–5616–200-

Share and Cite

MDPI and ACS Style

Na, K.-H.; Kim, W.-T.; Song, T.-H.; Choi, W.-Y. Magnetic Properties of NiZn Ferrite Nanofibers Prepared by Electrospinning. Appl. Sci. 2019, 9, 4297. https://doi.org/10.3390/app9204297

AMA Style

Na K-H, Kim W-T, Song T-H, Choi W-Y. Magnetic Properties of NiZn Ferrite Nanofibers Prepared by Electrospinning. Applied Sciences. 2019; 9(20):4297. https://doi.org/10.3390/app9204297

Chicago/Turabian Style

Na, Kyeong-Han, Wan-Tae Kim, Tae-Hyeob Song, and Won-Youl Choi. 2019. "Magnetic Properties of NiZn Ferrite Nanofibers Prepared by Electrospinning" Applied Sciences 9, no. 20: 4297. https://doi.org/10.3390/app9204297

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