Next Article in Journal
Hysteretic Photochromic Switching (HPS) in Doubly Doped GaN(Mg):Eu—A Summary of Recent Results
Next Article in Special Issue
In Situ Generation of Fluorescent Copper Nanoclusters Embedded in Monolithic Eggshell Membrane: Properties and Applications
Previous Article in Journal
3D Printing of Resin Material for Denture Artificial Teeth: Chipping and Indirect Tensile Fracture Resistance
Previous Article in Special Issue
Enhanced Photoelectrochemical Properties of Ti3+ Self-Doped Branched TiO2 Nanorod Arrays with Visible Light Absorption
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic and Mössbauer Spectroscopy Studies of Zinc-Substituted Cobalt Ferrites Prepared by the Sol-Gel Method

1
College of Medical Informatics, Hainan Medical University, Haikou 571199, China
2
Guangxi Key Laboratory of Nuclear Physics and Nuclear Technology, Guangxi Normal University, Guilin 541004, China
3
College of Physics and Technology, Guangxi Normal University, Guilin 541004, China
4
Sate Key Laboratory for Chemistry and Molecular Engineering of Medicinal Resources, Guangxi Normal University, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2018, 11(10), 1799; https://doi.org/10.3390/ma11101799
Submission received: 18 August 2018 / Revised: 16 September 2018 / Accepted: 19 September 2018 / Published: 21 September 2018
(This article belongs to the Special Issue Advanced Functional Nanomaterials and Their Applications)

Abstract

:
Zinc ion-substituted cobalt ferrite powders Co1−xZnxFe2O4 (x = 0–0.7) were prepared by the sol-gel auto-combustion process. The structural properties and magnetic of the samples were investigated with X-ray diffraction (XRD), superconducting quantum interference device, and a Mössbauer spectrometer. The results of XRD showed that the powder of a single cubic phase of ferrites calcined when kept at 800 °C for 3 h. The lattice constant increases with increase in Zn concentration, but average crystallite size does not decrease constantly by increasing the zinc content, which is related to pH value. It was confirmed that the transition from ferrimagnetic to superparamagnetic behaviour depends on increasing zinc concentration by Mössbauer spectra at room temperature. Magnetization at room temperature increases for x ≤ 0.3, but decreases for increasing Zn2+ ions. The magnetization of Co0.7Zn0.3Fe2O4 reached maximum value (83.51 emu/g). The coercivity decreased with Zn2+ ions, which were doped on account of the decrease of the anisotropy constant.

1. Introduction

Cobalt ferrite is important magnetic material. It is a hard ferromagnetic material, which has high coercivity of 5000 Oe, a high Curie temperature (TC) of 520 °C, moderate saturation magnetization of approximately 80 emu/g, a high anisotropy constant of 2.65 × 106–5.1 × 106 J/m3, and a high magneto-strictive of −225 ppm [1,2]. Moreover, cobalt ferrite exhibits high electromagnetic performance, a large magneto-optic effect, excellent mechanical hardness, and chemical stability [3]. Cobalt ferrite has been widely used as a high-density recording medium, because it is a well-known hard magnetic material [4]. Furthermore, cobalt ferrite is a promising candidate for medical applications, such as magnetic drug delivery, magnetic resonance imaging (MRI), radio-frequency hyperthermia and medical diagnostics [2], microwave and magneto-optic devices, and high-frequency catalysis and applications [1,3].
The saturation magnetization, electrical resistivity, coercivity, permittivity, and permeability of cobalt ferrite can be modified with partial replacement of non-magnetic zinc cations. The Zn2+ ion of non-magnetic substituted cobalt ferrite leads to a decrease in its saturation magnetization, Curie temperature, and coercivity [2]. Direct current (DC) electrical resistivity increases and the dielectric constant of the ferrite Co1−xZnxFe2O4 decreases with an increase in Zinc content, but the DC electrical resistivity decreases by increasing the calcination temperature, which ensures the semi-conductor performance of the sample [5]. Literature [6] has reported the permittivity and permeability of Co0.5Zn0.5Fe2O4 between 10 MHz to 1.0 GHz, which shows that the material is a potential wave absorber of electromagnetic interference (EMI). Veverka et al. [7] studied ferrite Co1−xZnxFe2O4+γ; when x = 0.6, they observed a transition temperature of 310–334 K for the paramagnetic state, which suggests that the magnetic fluid hyperthermia can be applied in a self-controlled regime. The maximum magnetoresistance (MR) has been observed for the Zn0.8Co0.2Fe2O4 polycrystalline sample [8].
In this paper, ferrite powders Co1−xZnxFe2O4 (x = 0–0.7) were synthesized with the sol-gel auto-combustion method. The aim was to study variation in structural and magnetic performance of cobalt ferrite powders with partial substitution of non-magnetic zinc cations.

2. Experimental

2.1. Sample Synthesis

Zinc ion-substituted cobalt ferrite powders Co1−xZnxFe2O4 (x = 0–0.7) were synthesized with the chemical method of sol-gel auto-combustion. The raw materials were analytical grade Co(NO3)2·6H2O, Zn(NO3)2·6H2O, Fe(NO3)3·9H2O, citric acid (C6H8O7·H2O), and ammonia (NH3·H2O). The molar ratio of citric acid to metal nitrates was 1:1. The citric acid and metal nitrates were respectively added to deionized water. Ammonia was added to adjust the pH value (about 7) of the metal nitrates solution. The solution was put into a thermostat water bath and maintained at 80 °C under continuous stirring to form the dried gel. In the process of stirring, citric acid was dripped continually. The gel was dried in a dry-oven of 120 °C for two hours, being burnt from the self-propagating combustion to become loose powder. The loose powder was ground and calcined at 200 °C, 400 °C, 600 °C, and 800 °C for 3 h.

2.2. Characterization

The structure and crystallite sizes were characterized by XRD (D/max-2500V/PC, Rigaku Corporation, Tokyo, Japan) in the 2θ range 20–70° using Cu-Kα radiations (λ = 0.15405 nm). The crystallite sizes were calculated with Scherrer’s formula: D = kλ/hcosθ where D, k, h, and θ are the average diameter, the shape factor, the half intensity width of the relevant diffraction peak, and diffraction angle, respectively. The micrographs were obtained by scanning electron microscopy (Nova NanoSEM 430, FEI Corporation, Hillsboro, OR, USA). Saturation magnetization was measured by Quantum Design MPMS series XL-7(Quantum Design Corporation, San Diego, CA, USA). The Mössbauer spectrum was performed by Mössbauer spectroscop (Fast Tec PC-mossII, FAST Corporation, Oberhaching, Bavaria, Germany, in constant acceleration mode. The γ-rays were provided by a 57Co source in a rhodium matrix. The hyperfine parameters, magnetic hyperfine field (Hhf), isomer shift (I.S.), quadrupole shift (Q.S.), relative area (A0), and line width (Г) were obtained by the fitted spectra through the Mösswinn 3.0 program, and the calibration was relative to 25 μm thick high purity alpha iron.

3. Results and Discussion

3.1. X-ray Diffraction Analysis

The X-ray diffraction (XRD) patterns of Co1−xZnxFe2O4 (x = 0–0.7) ferrites calcined at 800 °C for 3 h are shown in Figure 1. The XRD of the samples exhibited single-phase spinel structure. No impurity peak was detected in these samples. Table 1 shows that the lattice parameter increased by increasing the zinc content. The increase of the lattice parameter is probably due to the radius of the Zn2+ ions (0.74 Å) being larger than that of the Co2+ ions (0.72 Å) [8,9]. Zn2+ ions prefer to enter the tetrahedral A site, while Co2+ ions prefer to enter the octahedral B site in the Co1−xZnxFe2O4 ferrite [10,11]. On the basis of the earlier studies [12], the spinel structure can be assigned in the synthesized materials, which can be expressed as:
( Zn x 2 + Fe 1 x 3 + ) [ Co 1 x 2 + Fe 1 + x 3 + ] O 4 2 ,
where the parentheses indicate cations in the tetrahedral A sites, and the cations in the square bracket are in the octahedral B sites. The theoretical lattice parameter (ath) was estimated using the relation related to the radii of the tetrahedral and octahedral sites (rA, rB) [9]:
a t h = 8 3 3 [ ( r A + R 0 ) + 3 ( r B + R 0 ) ] ,
where R0 is the oxygen radius (R0 = 1.32 Å) [10,11,12], rA and rB are the tetrahedral radii and octahedral radii, respectively. On the basis of the ion distribution model (1), rA and rB are calculated:
r A = x   r Zn 2 + + ( 1 x ) r Fe 3 + r B = 1 2 [ ( 1 x )   r Co 2 + + ( 1 + x ) r Fe 3 + ]
We estimated the value of ath using Equations (2) and (3). There are deviations between the theoretical and the experimental lattice parameter, because cobalt ions (Co2+) not only occupy the octahedral site (B), but a few cobalt ions (Co2+) occupy the tetrahedral site (A) [10]. However, the experimental lattice parameter aexp and the theoretical lattice parameter ath all increase with the increasing of Zn2+ ions, as shown in the Figure 2. Average crystallite size tends to decrease with the increasing of Zn concentration, for the following reasons [13]: when zinc is introduced into the system, it will liberate more heat and decrease the crystal surface molecular concentration, thereby have a large impact on the grain growth. The preferences of the cations are not fully satisfied, which may be obstructed by grain growth. The higher bond energy of Co2+-O2−, as compared with Zn2+-O2−, lead to the particle sizes of the samples decrease with the increasing of zinc content. However, average crystallite size not decrease continuously with an increase in Zn ions, and is related to the pH value. When the pH value is too high, there is a lot of gas created by the ammonia, leading to an increase in porosity and a decrease in average crystallite size [14].
The X-ray density was estimated by the following relation [15]:
ρ x = 8 M N a 3 ,
where M is relative molecular mass, a is the lattice parameter, and N is the Avogadro’s number. Table 1 shows the X-ray density decreases with a Zn2+ concentration for x ≤ 0.4. The atomic weight of zinc is greater than the cobalt, so the increase of the lattice parameter leads to the decrease in X-ray density.
The XRD of Co0.9Zn0.1Fe2O4 sintered at different temperatures are shown in Figure 3. All the samples are of the single-phase cubic spinel structure. Table 2 shows the X-ray density of Co0.9Zn0.1Fe2O4 sintered at different temperature. The additional phase was not detected. For all the samples, the lattice parameter showed no significant changes. The average crystallite size of Co0.9Zn0.1Fe2O4 increases with an increase in the calcining temperature. When the burning temperature is below 600 °C the diffraction peaks of Co0.9Zn0.1Fe2O4 in our result were sharper and narrower relative to the diffraction peaks of Co0.5Zn0.5Fe2O4 in [11]. It indicates that when we prepared zinc-substituted cobalt ferrite powders with the experimental method of sol-gel auto-combustion, the sample without calcining had a good crystallinity [16].

3.2. Structures and Grain Sizes

The micrographs of CoFe2O4 annealed at 800 °C from scanning electron microscopy (SEM) are shown in Figure 4. It can be observed that the distribution of grains were almost uniform in size and well crystallized. Figure 5 shows the histogram of the grain size distribution of CoFe2O4 ferrites. The average grain size of CoFe2O4 estimated by a statistical method is approximately 96.26, which shows that CoFe2O4 ferrite powers are nanoparticles.

3.3. Mössbauer Spectroscopy

The Figure 6 shows the Mössbauer spectra at room temperature for Co1−xZnxFe2O4. The hyperfine parameters, isomer shift (IS), magnetic hyperfine field (Hhf), quadrupole shift (QS), relative area (A0), and line width (Г), were obtained by fitted spectra using the Mösswin 3.0 software program [17], and calibration was relative to a 25 μm thick sample of high purity alpha iron. For the Co1−xZnxFe2O4 with 0 ≤ x ≤ 0.3, the Mössbauer spectra are two sextets of normal Zeeman splits, which is attributable to Fe3+ ions at the tetrahedral A site and octahedral B site, indicating the ferrimagnetic properties of the samples. The sextet are assigned to the Fe3+ ions of the tetrahedral A site and the octahedral B site, and the octahedral B site isomer shift is larger than that of tetrahedral A site. Maybe it is due to the fact that the bond separation of A site Fe3+ ions is smaller than that of the B site Fe3+ ions, which have the smaller orbits overlapping Fe3+ ions and O2+ ions at the octahedral B site, resulting in smaller covalency for Fe2+ ions in octahedral B site [16,17,18]. Other studies have shown that the value of Fe3+ (S = 1/2, 3/2, 5/2) ions is 0.1–0.5 mm/s, while for Fe2+ (S = 2) ions the value is 0.6–1.7 mm/s [19]. As seen in Table 3, the value of iron is Fe3+ state in our study.
The magnetic hyperfine field (H) of A and B sites decreases with an increase in non-magnetic zinc substitution, It was also observed that the HA decreases with a larger rate than HB [20,21], because non-magnetic ion Zn2+-substituted cobalt ferrite goes to the A site [11]. In all of the samples, the Mössbauer spectra quadrupole shift is close to zero, which indicates that the ferrite is close to cubic symmetry. The A Mössbauer absorption area decreases and the B Mössbauer absorption area increases with increasing zinc concentration, since Zn2+ substitutes for cobalt ferrite and occupies the A site, leading to Fe3+ from the A site transferring to the B site. When 0.4 ≤ x ≤ 0.6, the spectra of Co1−xZnxFe2O4 is only the B magnetic sextet—the magnetic sextet of A site vanishes, which indicates the Fe3+ ions only occupy the octahedral B site [22]. When x = 0.6, the Mössbauer spectrum showed the relaxation effects features and was fitted one single sextet. When x = 0.7, the Mössbauer spectrum of the sample consisted only of a central doublet, and it exhibited superparamagnetic character. The central doublet can be due to the nonmagnetic nearest neighbors of Fe3+ ions, as the Fe3+ ions of magnetic isolation do not take part in long-range magnetic ordering [22,23,24].
Figure 7 shows room-temperature Mössbauer spectrums of Co0.9Zn0.1Fe2O4 powders calcined at different temperatures. Spectra of all samples sintered at different temperatures were fitted with two sextet sub-patterns. The research results of others [25] show that Mössbauer spectrums of Co0.9Zn0.1Fe2O4 calcined at different temperatures display a transition from paramagnetic doublet to ferrimagnetic sextet. Table 4 shows that while the Mössbauer parameters have no significant change for Co0.9Zn0.1Fe2O4 calcined at different temperatures, the magnetic hyperfine field increases slightly with increasing annealing temperatures. The X-ray patterns show that Co0.9Zn0.1Fe2O4 calcined at different temperatures has a good crystallinity, and that average crystallite size increase with increasing the calcining temperature. Therefore, the magnetic hyperfine field that has changed for the ferrite powders can be attributed to the variation of average crystallite size as a function of sintering temperature [25,26].

3.4. Magnetic Analysis

The room temperature hysteresis loops of Co1−xZnxFe2O4 are shown in Figure 8. It is observed from Table 5 that magnetization at 1000 Oe increases initially (up to x = 0.3) and then decreases as Zn content x increases.
When x = 0.3, the magnetization maximum value is 83.51 emu/g, the results are almost equal to the literature [11]. The magnetization could be expressed by the following relation [27]:
σ s = 5585 × n B M ,
where nB is the magnetic moment and M is relative molecular mass. The relative molecular mass of Co1−xZnxFe2O4 decreases as Zn content x increases. The variation of magnetic moment nB can be explained with Néel’s theory. The magnetic moment of Fe3+, Co2+, and Zn2+ ions are 5 μB, 3 μB, and 0 μB [9,11], respectively. According to two sub-lattice models of Néel’s theory, using the cation distribution of ( Zn x 2 + Fe 1 x 3 + ) [ Co 1 x 2 + Fe 1 + x 3 + ] O 4 2 (1), the magnetic moment nB is expressed as [9,11,28]:
n B = M B M A = ( 1 + x ) × 5 + ( 1 x ) × 3 ( 1 x ) × 5 = 3 + 7 x ,
where MA and MB are the A and B sub-lattice magnetic moments. Figure 9 shows the change in the experimental and theoretical magnetic moment with Zn content x. From the Equation (6), the magnetic moment nB increases with the increasing Zn content, and according to the relation (5) the theoretical magnetization monotonously increases as Zn content x increases. The change of the experimental and theoretical magnetization are in a good agreement with each other for x ≤ 0.3. When x ≥ 0.4, the experimental magnetization decreases as Zn content x increases, which can be explained with the Yafet–Kittel (YK) three sub-lattice model [9,29]. It is a reasonable that the sample appears to spin around the arrangement of the magnetic moment on B-sites when the content of nonmagnetic Zn2+ substitute cobalt ferrite is too high (x ≥ 0.4), leading to the B–B interaction increases; consequently, the A–B interaction decreases and subsequent decreases in magnetization.
It is observed from Table 5 that the coercivity of CoFe2O4 is 1005.32 Oe, and the coercivity is less than 100 Oe with Zn content x = 0.5 or 0.6. When x = 0.7, the coercivity is nearly zero, indicating the sample has turned from hard to soft magnetic materials. The magnetic state is from ferrimagnetic to superparamagnetic, which is in a good agreement with the result of the Mössbauer spectra. The relationship of the coercivity HC, the magnetic anisotropy K1, and the magnetization MS is [13]:
H C = 2 K 1 M S
From Table 5, we know that regardless of whether the magnetization decreases or increases, the coercivity decreases by increasing the Zn concentration. As a result, the decrease of coercivity is attributed to the reduction of the magnetic anisotropy. The anisotropy contribution comes from Co2+ ions of the octahedral B site, due to the electron configuration of Co2+ being 3d7 [13], as well as the ion’s spin and incompletely frozen orbital angular momentum coupling [30,31]. The Zn2+ of the 3d10 electron configuration has a zero angular momentum (l = 0), and it does not contribute to magneto-crystalline anisotropy. From our earlier research [32], the average particle size of CoFe2O4 (x = 0) is between 85.3 and 102.4 nm, and with increasing Zn content the particle size increases [23,33,34]. The sample’s magneto-crystalline anisotropy decreases with increasing Zn2+ ions substituted for Co2+ ions, with subsequent decreases in the coercivity [35,36]. Kamali et al. synthesized NiFe2-xAlxO4 ferrites and observed the complicated cation distributions in this ferrite system as a function of x. The relationship between the electronic ground state, magnetism, and cation distributions is explained in terms of a model [37,38,39].

4. Conclusions

The XRD indicates that the ferrite Co1−xZnxFe2O4 calcined at 800 °C is a single-phase cubic spinel structure. The increase of the lattice parameter is attributed to replacement of smaller Co2+ ions by larger Zn2+ ions. The XRD patterns of ferrite Co0.9Zn0.1Fe2O4 sintered at different temperature indicate that the ferrite prepared with the experimental method of sol-gel auto-combustion have good crystallinity. Room temperature Mössbauer spectra reveal a transition from ferrimagnetic behavior to superparamagnetic behavior by increasing the zinc concentration for Co1−xZnxFe2O4 calcined at 800 °C. The spectra of Co0.9Zn0.1Fe2O4 calcined at different temperatures are fitted with two sextet sub-patterns. The variation of the magnetic hyperfine field can be attributed to the average crystallite size change with annealing temperature. The magnetization increases initially (up to x = 0.3) and then decreases with increasing Zn content. When x = 0.3, the magnetization maximum value is 83.51 emu/g. The change of the magnetization can be explained with Néel’s theory and the Yafet–Kittel model. As Zn content x increases, the coercivity and remanence is close to zero, indicating that the sample shows superparamagnetic character.

Author Contributions

Q.L. and J.X. contributed equally to this work. Q.L. and Y.H. participated in experimental design. Q.L. and J.L. performed the experiments. J.L., H.Y. and F.Y. contributed reagents/ materials/analysis tools. J.X., H.Y., and F.Y. collected data. J.L. and Y.H. are co-corresponding authors and contributed equally to this study. All authors discussed the results and commented on the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 11364004, No. 11164002), the Hainan Provincial Natural Science Foundation of China (No. 818MS065), the Science research project of Hainan higher education institutions (Grant No. Hnjg2017-41). The project was funded by the Guangxi Key Laboratory of Nuclear Physics and Nuclear Technology.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mohamed, R.M.; Rashada, M.M.; Haraz, F.A.; Sigmund, W. Structure and magnetic properties of nanocrystalline cobalt ferrite powders synthesized using organic acid precursor method. J. Magn. Magn. Mater. 2010, 322, 2058–2064. [Google Scholar] [CrossRef]
  2. Amiri, S.; Shokrollah, H. The role of cobalt ferrite magnetic nanoparticles in medical science. Mater. Sci. Eng. 2013, 33, 1–8. [Google Scholar] [CrossRef]
  3. Sanpo, N.; Berndt, C.C.; Wen, C.; Wang, J. Transition metal-substituted cobalt ferrite nanoparticles for biomedical applications. Acta Biomater. 2013, 9, 5830–5837. [Google Scholar] [CrossRef] [PubMed]
  4. Chia, C.H.; Zakaria, S.; Yusoffa, M.; Goh, S.C.; Haw, C.Y.; Ahmadi, Sh.; Huang, N.M.; Lim, H.N. Size and crystallinity-dependent magnetic properties of CoFe2O4 nanocrystals. Ceram. Int. 2010, 36, 605–609. [Google Scholar] [CrossRef]
  5. Gul, H.; Abbasi, A.Z.; Amin, F.; Anis-ur-Rehman, M.; Maqsood, A. Structural, magnetic and electrical properties of Co1−xZnxFe2O4 synthesized by co-precipitation method. J. Magn. Magn. Mater. 2007, 311, 494–499. [Google Scholar] [CrossRef]
  6. Waje, S.B.; Hashim, M.; Yusoff, W.D.W.; Abbas, Z. Sintering temperature dependence of room temperature magnetic and dielectric properties of Co0.5Zn0.5Fe2O4 prepared using mechanically alloyed nanoparticles. J. Magn. Magn. Mater. 2010, 322, 686–691. [Google Scholar] [CrossRef]
  7. Veverka, M.; Veverka, P.; Jirák, Z.; Kaman, O.; Knížek, K.; Maryško, M.; Závěta, K. Synthesis and magnetic properties of Co1−xZnxFe2O4+γ nanoparticles as materials for magnetic fluid hyperthermia. J. Magn. Magn. Mater. 2010, 322, 2386–2389. [Google Scholar] [CrossRef]
  8. Hossain, A.A.; Tabata, H.; Kawai, T. Magnetoresistive properties of Zn1−xCoxFe2O4 ferrites. J. Magn. Magn. Mater. 2008, 320, 1157–1162. [Google Scholar] [CrossRef]
  9. Jadhav, S.S.; Shirsath, S.E.; Patange, S.M.; Jadhav, K.M. Effect of Zn substitution on magnetic properties of nanocrystalline cobalt ferrite. J. Appl. Phys. 2010, 108, 093920. [Google Scholar] [CrossRef]
  10. Hemeda, O.M.; El-Saadawy, M. Effect of gamma irradiation on the structural properties and diffusion coefficient in Co-Zn ferrite. J. Magn. Magn. Mater. 2003, 256, 63–68. [Google Scholar] [CrossRef]
  11. Liu, Y.; Zhu, X.G.; Zhang, L.; Min, F.F.; Zhang, M.X. Microstructure and magnetic properties of nanocrystalline Co1−xZnxFe2O4 ferrites. Mater. Res. Bull. 2012, 47, 4174–4180. [Google Scholar] [CrossRef]
  12. Deraz, N.M.; Alarifi, A. Structural, morphological and magnetic properties of nano-crystalline zinc substituted cobalt ferrite system. J. Anal. Appl. Pyrolysis 2012, 94, 41–47. [Google Scholar] [CrossRef]
  13. Sharifi, I.; Shokrollahi, H. Nanostructural, magnetic and Mössbauer studies of nanosized Co1−xZnxFe2O4 synthesized by co-precipitation. J. Magn. Magn. Mater. 2012, 324, 2397–2403. [Google Scholar] [CrossRef]
  14. Zhang, Y.; Lan, Z.; Yu, Z. Synthesis of NiZn ferrite nanoparticles by sol-gel method. Mater. Rev. 2006, 20, 49–51. [Google Scholar]
  15. Gul, H.; Maqsood, A. Structural, magnetic and electrical properties of cobalt ferrites prepared by the sol-gel route. J. Alloys Compd. 2008, 465, 227–231. [Google Scholar] [CrossRef]
  16. Barpanda, P.; Yamashita, Y.; Yamada, Y.; Yamada, A. High-throughput solution combustion synthesis of high-capacity LiFeBO3 cathode. J. Electrochem. Soc. 2013, 160, 3095–3099. [Google Scholar] [CrossRef]
  17. Zhang, Y.; Lin, J.; Wen, D. Structure, Infrared Radiation Properties and Mössbauer Spectroscopic Investigations of Co0.6Zn0.4NixFe2-xO4 Ceramics. J. Mater. Sci. Technol. 2010, 26, 687–692. [Google Scholar] [CrossRef]
  18. Siddique, M.; Butt, N.M. Effect of particle size on degree of inversion in ferrites investigated by Mössbauer spectroscopy. Phys. Rev. B Condens. Matter 2010, 405, 4211–4215. [Google Scholar] [CrossRef]
  19. Kumar, S.; Farea, A.M.M.; Batoo, K.M.; Lee, C.G.; Koo, B.H.; Yousef, A. Alimuddin. Mössbauer studies of Co0.5CdxFe2.5-xO4 (0.0–0.5) ferrite. Phys. Rev. B Condens. Matter 2008, 403, 3604–3607. [Google Scholar] [CrossRef]
  20. He, Y.; Lei, C.; Lin, Q.; Dong, J.; Yu, Y.; Wang, L. Mössbauer and Structural properties of La-substituted Ni0.4Cu0.2Zn0.4Fe2O4 nanocrystalline ferrite. Sci. Adv. Mater. 2015, 7, 1809–1815. [Google Scholar] [CrossRef]
  21. Singhal, S.; Barthwal, S.K.; Chandra, K. XRD, magnetic and Mössbauer spectral studies of nano size aluminum substituted cobalt ferrites (CoAlxFe2-xO4). J. Magn. Magn. Mater. 2006, 306, 233–240. [Google Scholar] [CrossRef]
  22. Bayoumi, W. Structural and electrical properties of zinc-substituted cobalt ferrite. J. Mater. Sci. 2007, 42, 8254–8261. [Google Scholar] [CrossRef]
  23. Lee, S.W.; Ryu, Y.G.; Yang, K.J.; Jung, K.D.; An, S.Y.; Kim, C.S. Magnetic properties of Zn2+ substituted ultrafine Co-ferrite grown by a sol-gel method. J. Appl. Phys. 2002, 91, 7610. [Google Scholar] [CrossRef]
  24. Guivar, J.A.R.; Sanches, E.A.; Bruns, F.; Sadrollahi, E.; Morales, M.A.; López, E.O.; Litterst, F.J. Vacancy ordered γ-Fe2O3, nanoparticles functionalized with nanohydroxyapatite: XRD, FTIR, TEM, XPS and Mössbauer studies. Appl. Surf. Sci. 2016, 389, 721–734. [Google Scholar] [CrossRef]
  25. Kamali, S.; Bringas, E.; Hah, H.Y.; Bates, B.; Johnson, J.A.; Johnson, C.E.; Stroeve, P. Magnetism and Mössbauer study of formation of multi-core γ-Fe2O3 nanoparticles. J. Magn. Magn. Mater. 2018, 451, 131–136. [Google Scholar] [CrossRef]
  26. Chae, K.P.; Kim, W.K.; Lee, S.H.; Lee, Y.B. Crystallographic and magnetic properties of TixCo1+xFe2-2xO4 ferrite powders. J. Magn. Magn. Mater. 2001, 232, 133–138. [Google Scholar] [CrossRef]
  27. Birajdar, D.S.; Mane, D.R.; More, S.S.; Kawade, V.B.; Jadhav, K.M. Structural and magnetic properties of ZnxCu1.4-xMn0.1Fe1.2O4 ferrites. Mater. Lett. 2005, 59, 2981–2985. [Google Scholar] [CrossRef]
  28. Verma, S.; Chand, J.; Batoo, K.M.; Singh, M. Structural, magnetic and Mössbauer spectral studies of aluminum substituted Mg-Mn-Ni ferrites (Mg0.2Mn0.5Ni0.3AlyFe2-yO4). J. Alloys Compd. 2013, 551, 715–721. [Google Scholar] [CrossRef]
  29. Hossain, A.A.; Amin, M.R.; Tanaka, H. Increase in initial permeability due to substitution of high spin cations in nanocrystalline Ni-Mg ferrites. J. Magn. Magn. Mater. 2013, 334, 124–129. [Google Scholar] [CrossRef]
  30. He, Y.; Yang, X.; Lin, J.; Lin, Q.; Dong, J. Mössbauer spectroscopy, Structural and magnetic studies of Zn2+ substituted magnesium ferrite nanomaterials prepared by sol-gel method. J. Nanomater. 2015, 5, 854840. [Google Scholar] [CrossRef]
  31. Gözüak, F.; Köseoğlu, Y.; Baykal, A.; Kavas, H. Synthesis and characterization of Co1−xZnxFe2O4magnetic nanoparticles via a PEG-assisted route. J. Magn. Magn. Mater. 2009, 321, 2170–2177. [Google Scholar] [CrossRef]
  32. Lin, Q.; Lin, J.; He, Y.; Wang, R.; Dong, J. The structural and magnetic properties of gadolinium doped CoFe2O4 nanoferrites. J. Nanomater. 2015, 8, 294239. [Google Scholar]
  33. Naseri, M.G.; Saion, E.B.; Ahangar, H.A.; Shaari, A.H.; Hashim, M. Simple synthesis and characterization of cobalt ferrite nanoparticles by a thermal treatment method. J. Nanomater. 2010, 75, 907686. [Google Scholar]
  34. Soibam, I.; Phanjoubam, S.; Prakash, C. Mössbauer and magnetic studies of cobalt substituted lithium zinc ferrites prepared by citrate precursor method. J. Alloys Compd. 2009, 475, 328–331. [Google Scholar] [CrossRef]
  35. Lin, J.; He, Y.; Lin, Q.; Wang, R.; Chen, H. Microstructural and Mössbauer spectroscopy Studies of Mg1−xZnxFe2O4 (x = 0.5, 0.7) nanoparticles. J. Spectrosc. 2014, 2014, 540319. [Google Scholar] [CrossRef]
  36. Motavallian, P.; Abasht, B.; Abdollah-Pour, H. Zr doping dependence of structural and magnetic properties of cobalt ferrite synthesized by sol-gel based Pechini method. J. Magn. Magn. Mater. 2018, 451, 577–586. [Google Scholar] [CrossRef]
  37. Kamali, S.; Shih, K.; Barbiellini, B.; Wang, Y.J.; Kaprzyk, S.; Itou, M.; Sakurai, Y. Extracting cation distributions in NiFe2-xAlxO4 Solid Solutions using magnetic Compton scattering. J. Phys. Condens. Matter 2015, 27, 456003. [Google Scholar] [CrossRef] [PubMed]
  38. Kamali, S. Spin structure, magnetism, and cation distributions of NiFe2-xAlxO4 solid solutions. J. Magn. Magn. Mater. 2017, 433, 155–161. [Google Scholar] [CrossRef]
  39. Hu, Z.P.; Xu, Y.B.; Tan, X.D.; Peng, F.; Gu, X.L.; Zou, Y.; Yu, S.C. Effect of rapid heating on a cold-rolled Mn–Al transformation-induced plasticity steel with coarse delta-ferrite. Sci. Adv. Mater. 2017, 9, 1953–1959. [Google Scholar]
Figure 1. Room-temperature XRD patterns of Co1−xZnxFe2O4 calcined at 800 °C.
Figure 1. Room-temperature XRD patterns of Co1−xZnxFe2O4 calcined at 800 °C.
Materials 11 01799 g001
Figure 2. The variation of the theoretical and experimental lattice parameter for Co1−xZnxFe2O4 (x = 0–0.7).
Figure 2. The variation of the theoretical and experimental lattice parameter for Co1−xZnxFe2O4 (x = 0–0.7).
Materials 11 01799 g002
Figure 3. The XRD of Co0.9Zn0.1Fe2O4 sintered at different temperatures.
Figure 3. The XRD of Co0.9Zn0.1Fe2O4 sintered at different temperatures.
Materials 11 01799 g003
Figure 4. Scanning electron microscopy (SEM) micrographs of CoFe2O4 annealed at 800 °C.
Figure 4. Scanning electron microscopy (SEM) micrographs of CoFe2O4 annealed at 800 °C.
Materials 11 01799 g004
Figure 5. Histogram of grain size distribution of CoFe2O4 annealed at 800 °C.
Figure 5. Histogram of grain size distribution of CoFe2O4 annealed at 800 °C.
Materials 11 01799 g005
Figure 6. The Mössbauer spectra of Co1−xZnxFe2O4 sintered at 800 °C.
Figure 6. The Mössbauer spectra of Co1−xZnxFe2O4 sintered at 800 °C.
Materials 11 01799 g006
Figure 7. The Mössbauer spectrum of Co0.9Zn0.1Fe2O4 calcined at different temperatures.
Figure 7. The Mössbauer spectrum of Co0.9Zn0.1Fe2O4 calcined at different temperatures.
Materials 11 01799 g007
Figure 8. Room temperature hysteresis loops of Co1−xZnxFe2O4 calcined at 800 °C.
Figure 8. Room temperature hysteresis loops of Co1−xZnxFe2O4 calcined at 800 °C.
Materials 11 01799 g008
Figure 9. Variation in experimental and theoretical magnetic moment with Zinc content x.
Figure 9. Variation in experimental and theoretical magnetic moment with Zinc content x.
Materials 11 01799 g009
Table 1. XRD date of Co1−xZnxFe2O4 calcined at 800 °C.
Table 1. XRD date of Co1−xZnxFe2O4 calcined at 800 °C.
Content (x)Lattice Parameter (Å)Average Crystallite Size (Å)Density (g/cm3)
08.354975565.3468
0.18.394355085.2693
0.28.408844985.2421
0.38.410824555.2384
0.48.412564895.2351
0.58.444214085.3187
0.68.427664825.3501
0.78.451583605.3048
Table 2. XRD date of Co0.9Zn0.1Fe2O4 sintered at different temperatures.
Table 2. XRD date of Co0.9Zn0.1Fe2O4 sintered at different temperatures.
Temperature (°C)Lattice Parameter (Å)Average Crystallite Size (Å)Density (g/cm3)
unsintered8.382242705.2921
2008.405693125.2480
4008.386443145.2842
6008.384643325.2876
8008.394355085.2693
Table 3. Mössbauer parameters of Co1−xZnxFe2O4 sintered at 800 °C.
Table 3. Mössbauer parameters of Co1−xZnxFe2O4 sintered at 800 °C.
Content (x)ComponentI.S. (mm/s)Q.S. (mm/s)H (T)Γ (mm/s)A0 (mm/s)
0Sextet (A)0.237−0.00448.9460.36033
Sextet (B)0.375−0.02445.6950.32267
0.1Sextet (A)0.2470.01948.4590.34426
Sextet (B)0.343−0.02745.2270.37874
0.2Sextet (A)0.2480.03847.7350.38621
Sextet (B)0.340−0.01642.5320.35279
0.3Sextet (A)0.2350.05547.5080.42911
Sextet (B)0.306−0.05038.9460.33889
0.4Sextet (B)0.288−0.02234.9110.375100
0.5Sextet (B)0.3090.000228.60.375100
0.6Sextet (B)0.346−0.00518.80.291100
0.7Double0.3470.430500.357100
Table 4. Mössbauer parameters of Co0.9Zn0.1Fe2O4 calcined at different temperatures.
Table 4. Mössbauer parameters of Co0.9Zn0.1Fe2O4 calcined at different temperatures.
Temperature (°C)ComponentI.S. (mm/s)Q.S. (mm/s)H (T)Γ (mm/s)A0 (mm/s)
unsinteredSextet (A)0.2250.05648.3560.39422.5
Sextet (B)0.328−0.05043.9050.33677.5
200Sextet (A)0.2490.00748.4490.39122.5
Sextet (B)0.348−0.01143.7120.35177.5
400Sextet (A)0.2370.01648.4580.40618.9
Sextet (B)0.324−0.03044.1540.38781.1
600Sextet (A)0.2340.00848.5820.39623.1
Sextet (B)0.3460.00144.5630.36676.9
Table 5. Magnetic data for Co1−xZnxFe2O4 calcined at 800 °C.
Table 5. Magnetic data for Co1−xZnxFe2O4 calcined at 800 °C.
Content (x)MS (emu/g)HC (Oe)Mr1 (emu/g)nB
072.581005.3334.713.05
0.180.46703.9133.143.39
0.282.61402.2426.463.49
0.383.51301.7525.003.54
0.476.97100.7323.643.27
0.563.6375.6215.082.71
0.641.2325.394.061.76
0.720.090.240.130.86
1 Mr is remanent magnetization.

Share and Cite

MDPI and ACS Style

Lin, Q.; Xu, J.; Yang, F.; Lin, J.; Yang, H.; He, Y. Magnetic and Mössbauer Spectroscopy Studies of Zinc-Substituted Cobalt Ferrites Prepared by the Sol-Gel Method. Materials 2018, 11, 1799. https://doi.org/10.3390/ma11101799

AMA Style

Lin Q, Xu J, Yang F, Lin J, Yang H, He Y. Magnetic and Mössbauer Spectroscopy Studies of Zinc-Substituted Cobalt Ferrites Prepared by the Sol-Gel Method. Materials. 2018; 11(10):1799. https://doi.org/10.3390/ma11101799

Chicago/Turabian Style

Lin, Qing, Jianmei Xu, Fang Yang, Jinpei Lin, Hu Yang, and Yun He. 2018. "Magnetic and Mössbauer Spectroscopy Studies of Zinc-Substituted Cobalt Ferrites Prepared by the Sol-Gel Method" Materials 11, no. 10: 1799. https://doi.org/10.3390/ma11101799

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