Next Article in Journal
Hepatotoxic Evaluation of N-(2-Hydroxyphenyl)-2-Propylpentanamide: A Novel Derivative of Valproic Acid for the Treatment of Cancer
Next Article in Special Issue
Photocatalytic Degradation Studies of Organic Dyes over Novel Cu/Ni Loaded Reduced Graphene Oxide Hybrid Nanocomposite: Adsorption, Kinetics and Thermodynamic Studies
Previous Article in Journal
Walnut Kernel Oil and Defatted Extracts Enhance Mesenchymal Stem Cell Stemness and Delay Senescence
Previous Article in Special Issue
Structural and Magnetic Properties of Perovskite Functional Nanomaterials La1−xRxFeO3 (R = Co, Al, Nd, Sm) Obtained Using Sol-Gel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Nonmagnetic Zn2+ Ion and RE Ion Substitution on the Magnetic Properties of Functional Nanomaterials Co1−yZnyRExFe2−xO4 (RE = La, Sm, Gd) by Sol–Gel

1
College of Biomedical Information and Engineering, Hainan Medical University, Haikou 571199, China
2
College of Physics and Technology, Guangxi Normal University, Guilin 541004, China
3
Department of Civil Engineering, Jiangxi Water Resources Institute, Nanchang 330013, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(17), 6280; https://doi.org/10.3390/molecules28176280
Submission received: 5 July 2023 / Revised: 3 August 2023 / Accepted: 23 August 2023 / Published: 28 August 2023
(This article belongs to the Special Issue Functional Sol-Gel Composites: Preparation and Applications)

Abstract

:
Magnetic Functional Nanomaterials Co1−yZnyRExFe2−xO4 (RE (rare-earth) = La,Sm,Gd) were prepared using the sol–gel combustion method. XRD characterization confirms that the ferrite samples we synthesized are single-phase cubic structures. The variation in the average crystalline size and lattice parameter is related to RE ion doping. The Mössbauer spectra of CoRExFe2−xO4 are two sets of magnetic six-wire peaks that indicate the ferrimagnetic behavior of the sample. The calcination temperature greatly influences the absorption area of Mössbauer for CoFe2O4, indicating that the calcination temperature affects the iron ion content at the octahedral B and tetrahedral A sites. Additionally, scanning electron microscopy measurements of the substituted specimens reveal that the ferrite powders are nanoparticles. With an increase in RE ions, the coercivity increases, and the saturation magnetization changes obviously. The XRD characterization of Co0.7Zn0.3LaxFe2−xO4 shows that the main crystalline phase of the sample is the cubic spinel structure phase, and there are fewer secondary crystalline phases. The lattice parameter tends to decrease with the substitution of La3+ ions. The average grain size decreased significantly with the increase in La content. From ferrimagnetic state transition to relaxation behavior, the hyperfine magnetic field decreases in La concentration by room temperature Mössbauer spectra. With the substitution of La3+ ions, both the saturation magnetization and coercivity of the samples were reduced, and the coercivity of all samples was lower.

1. Introduction

CoFe2O4 is an important magnetostriction and magnetic material [1,2,3,4]. CoFe2O4 has been widely used as magnetic recording material [5,6,7,8]. Mössbauer effect spectra are a suitable technique to detect the local environment of iron nuclei around various divalent cations consisting of a close-packed oxygen arrangement [9,10,11,12,13]. The rare-earth ions exhibit interesting magnetic and magnetostrictive properties [14,15,16,17,18], playing an essential role in the magnetocrystalline anisotropy of ferrite. The RE ions of 4f elements substitute Fe3+ ions of 3d elements in ferrites, exhibiting the strong 3d–4f spin–orbital coupling [19,20,21,22,23]. Kumar et al. [10] studied the magnet anisotropy on the La3+ substitution effect of CoFe2−xLaxO4 ferrites. Mariano et al. [11] synthesized the magnetic nanoparticles Fe2−xCoSmxO4 and analyzed the variation of physical and magnetic properties with Sm content. Simultaneously, Rana et al. [12] investigated the dielectric properties, revealing a strong dependence on Gd3+ substitution in nano cobalt ferrite. Mixed ferrite has vast applications in a wide range, from frequencies to radio microwaves [24,25,26,27]. They play an irreplaceable role in magnetic recording, computer memory, and microwave devices due to their low eddy current loss and high resistivity [28,29,30,31]. In nanomaterials containing transition metals with 3D band structures, magnetic carriers are 3D shell electrons that migrate from one atom to another [32,33,34,35]. The magnetic moments are localized in individual atoms [36,37]. The dielectric constants and permeability of Co0.5Zn0.5Fe2O4 in the range of 10 MHz to 1.0 GHz show the potential of the material as an electromagnetic interference absorber [38]. For Co1−xZnxFe2O4+γ of the composition x = 0.6, its paramagnetic transition temperature is 310–334 K, indicating that magnetic fluid hyperthermia is suitable for self-control states [39]. For synthesizing nanometer ferrite powder, the sol–gel method is a simple and economical process.
To sum up, the substitution of nonmagnetic Zn2+ ions and rare-earth ions will have a greater impact on magnetic properties. The above literature mainly studies the influence of Zn ion or rare-earth ion single-component doping on sample properties, while some of the literature has focused on the influence of co-doped cobalt ferrite with nonmagnetic Zn ions and rare-earth ions on sample magnetism. Therefore, the co-doping of nonmagnetic Zn ions and rare-earth ions will be studied in this paper. It has many advantages, for example, inexpensive raw material, low external energy consumption, homogenous highly reactive powder, and good crystallinity [40]. In this paper, Co1−yZnyRExFe2−xO4 (RE = La, Sm, Gd) functional nanomaterials were prepared by sol–gel auto combustion method, and the effects of nonmagnetic zinc ion and rare-earth ion substitution on their structural magnetic properties were studied.

2. Results and Discussion

2.1. The XRD Analysis

Figure 1 shows the XRD diffraction pattern of CoFe2O4 sintered at different temperatures, from which no impurity peak is observed, confirming that CoFe2O4 possesses a single spinel-structure ferrite.The average crystallite size increased, and the lattice parameter showed changes for CoFe2O4 sintered at different temperatures, as indicated in Table 1. Previous studies [4,20,21] have shown that the diffraction peaks of XRD are not sharp for CoFe2O4 calcined at low temperatures.
Ferrite doped with nonmagnetic rare-earth ions exhibits strong spin–orbit coupling (3d–4f) since the rare-earth ions plan an important role in determining the magnetocrystalline anisotropy [8,9,10,11]. In our results, however, the diffraction peaks of XRD are sharp for CoFe2O4 without burning. The samples without calcination still display excellent crystallinity.
Figure 2 depicts the X-ray diffraction (XRD) analysis of CoRExFe2−xO4 ferrites (x = 0, 0.02; RE = La, Sm, Gd). CoRExFe2−xO4 ferrites possess a single spinel structure (JCPDS card numbers 22-1086). No other impurity peaks were observed in the XRD pattern of the sample. The lattice constant of CoRExFe2−xO4 ferrites is larger than that of CoFe2O4 ferrites due to the larger ionic radius of RE3+ ions (rLa3+ = 1.03 Å, rSm3+ = 0.96 Å, rGd3+ = 0.938 Å) than that of Fe3+ ions (0.645 Å) [10,12,13,14]. Table 2 presents the corresponding data. The average crystallite size of the investigated samples, CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd), estimated by the Debye–Scherrer formula [7,13], ranges from 37.4 nm to 55.6 nm. When RE ions are employed as substitutes, the average crystallite size decreases, which aligns with other reports of the literature [14,15,16].
The XRD density is calculated using the formula [17,18,19]:
ρ x = 8 M N a 3
where M is the relative molecular mass, N is Avogadro’s number, and a is the lattice parameter. The XRD density decreases with RE3+ substitution, as shown in Table 2. Doping RE ions leads to an increase in relative molecular mass, and according to Equation (1), the lattice parameter will also increase. The decrease in XRD density can be attributed to the increasing lattice parameter.
Figure 3 displays the XRD patterns of Co0.7Zn0.3LaxFe2−xO4 (x = 0~0.20) ferrites calcined at 800 °C for 3 h. For all the samples with 0 ≤ x ≤ 0.05, the XRD patterns shown are single phase with spinel structure. When 0.07 ≤ x ≤ 0.20, XRD results confirm that the main phase is the cubic spinel phase, and there is a small amount of LaFeO3 in the impurity phase. The XRD intensity of LaFeO3 increases by increasing the La content, which is due to the La3+ ions having bigger ionic radii and very low solubility in spinel lattice [41]. Table 3 indicate XRD data for Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Figure 4 indicate that the lattice parameter tends to decrease with the substitution of La3+ ions. It may be that there are secondary phases forming in the grain boundary. Other articles have reported similar results for rare-earth substituted ferrites [10].Average crystallite size tends to decrease with the substitution of La3+ ions; similar results are introduced in the other literature [15]. Due to the lower bond energy of Fe3+-O2 compared to La3+-O2−, the entry of La3+ ions into the lattice to form La3+-O2 bonds requires more energy. Therefore, for ferrite replaced by La3+ ions, more energy is required to complete crystallization and grain growth. Table 3 shows the trend that density decreases with La3+ concentration for all samples.The relative atomic weight of La is greater than that of Fe, so the relative molecular weight of the sample increases with the increase in La doping amount. The increase in X-ray density is attributed to an increase in relative molecular weight and a decrease in lattice parameters [42]. The increase in X-ray density also indicates improved densification [43].
The X-ray patterns of Co0.7Zn0.3La0.01Fe1.99O4 calcined at different temperatures are displayed in Figure 5 and Table 4. The samples are all single-phase structures of spinel ferrite, and no other impurities were found. The lattice parameters of all samples showed no significant changes, but the average crystallite size showed an increasing trend with the increase in calcination temperature. The diffraction peaks of Co0.7Zn0.3La0.01Fe1.99O4 without burning are very sharp. The samples without calcination still display excellent crystallinity.

2.2. Structures and Grain Sizes

Figure 6 depicts the SEM images of Functional Nanomaterials CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C. The sample is well crystallized and displays almost uniform grain sizes. With the substitution of RE3+ ions, some cobalt ferrite particles become agglomerated, which may be due to the magnetic interactions between CoRExFe2−xO4 particles [11]. Figure 7 shows a histogram of the grain size distribution of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd). Statistical methods were used to analyze the average grain size of CoFe2O4, CoLa0.02Fe1.98O4, CoSm0.02Fe1.98O4, and CoGd0.02Fe1.98O4, which were estimated as approximately 96.26, 54.18, 57.36, and 46.15 nm, respectively. From this, it can be seen that the ferrite samples we prepared are nanoparticles, as their average grain size is less than 100 nm. They are slightly larger than the average crystallite size of X-ray, so each particle is composed of several crystallites [22].
The SEM micrographs and grain size distribution diagram of CoFe2O4 sintered at 1000 °C are depicted in Figure 8. The average grain size of CoFe2O4 ferrite sintered at 1000 °C is estimated to be approximately 137.5 nm, which is larger than the average grain size (96.26 nm) of the cobalt ferrite annealed at 800 °C. This indicates that the average grain size of the CoFe2O4 sample increases with an increase in the calcining temperature.
Figure 9 shows the SEM microphotographs of Co0.7Zn0.3LaxFe2−xO4 (x = 0.05, 0.10) annealed at 800 °C. The sample Co0.7Zn0.3La0.05Fe1.95O4 we prepared has a relatively uniform grain distribution and good crystallinity, while Co0.7Zn0.3La0.10Fe1.90O4 has lower crystallinity. Figure 10 shows the histogram of the grain size distribution of Co0.7Zn0.3LaxFe2−xO4 (x = 0.05, 0.10) ferrites. The significant decrease in the average grain size with increasing La content may be due to the growth of grain being inhibited with LaFeO3 [18]. The average grain size of Co0.7Zn0.3La0.05Fe1.95O4 and Co0.7Zn0.3La0.10Fe1.90O4 estimated by a statistical method is approximately 35.21 and 30.90 nm, respectively. The average grain size is slightly larger than the average crystallite size of X-ray, so each particle is composed of several crystallites [22].

2.3. Room Temperature Mössbauer Spectra

Figure 11 shows the room temperature Mössbauer spectroscopy curve for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C. We used the Mösswinn 3.0 program to fit the measured Mössbauer spectrum, which consists of two sets of six-line peaks.The six linear peaks with a large isomer shift (I.S.) correspond to the iron ions at position B, while the other set of six linear peaks with a smaller isomer shift (I.S.) correspond to the iron ions at position A. This difference is due to the varying internuclear separations between Fe3+ and O2− ions [23,24]. The changes in I.S. values due to RE3+ substitution are relatively small, and the effect of RE3+ doping on the charge distribution around Fe3+ is not significant [24]. According to other reports, the I.S. values for Fe3+ ions range from 0.1 mm/s to 0.5 mm/s [25]. Therefore, the I.S. values indicate that the iron in our sample is Fe3+ ions (Table 5). The magnetic hyperfine field (H) does not appear to be affected by RE3+ substitution. Figure 12 displays the spatial structure of spinel ferrite cells. It is possible that micro-substitution is insufficient to significantly alter the H. The quadrupole displacement value in the sample CoRExFe2−xO4 is the smallest, indicating good symmetry of the electric field around the atomic nucleus in cobalt ferrite.
Figure 13 displays the room temperature Mössbauer spectra of Co0.7Zn0.3LaxFe2−xO4. The spectra of Co0.7Zn0.3Fe2O4, fitted with two sextets of Zeeman-split, are attributed to Fe3+ at tetrahedral and octahedral sites, indicating the ferrimagnetic behavior of the samples. The six linear peaks with a large I.S. correspond to the iron ions at position B, while the other set of six linear peaks with a smaller I.S. correspond to the iron ions at position A [20]. When 0.01 ≤ x ≤ 0.15, the Mössbauer spectrum of Co0.7Zn0.3LaxFe2−xO4 only shows a set of magnetic six-line peaks, corresponding to the six-line peaks at the B site, indicating that Fe3+ ions only occupy the octahedral B site [44]. The Mössbauer spectrum of x = 0.20 was fitted with one single sextet and a central paramagnetic doublet, and it shows the relaxation effects features. The A–B exchange interaction decreases with the substitution of nonmagnetic La3+ ions, resulting in a decrease in the magnetic hyperfine field value [8]. The large isomer shift (I.S.) value of Fe2+ ions is 0.1–0.5 mm/s, while for Fe3+, it is 0.6–1.7 mm/s [44]. From the isomer shift (I.S.) values in Table 6, we know that the iron ions in our sample are in the Fe3+ state. Figure 14 displays the A site of the tetrahedral lattice and the B site of the octahedral lattice. The quadrupole shift of the magnetic sextet for A and B sites is very small, which indicates that the charge distribution of atomic nuclei is close to spherical symmetry. Mössbauer spectrum of Co0.7Zn0.3La0.2Fe1.8O4 was fitted with one single sextet and a central paramagnetic doublet, which shows the relaxation effects features [45].
Figure 15 depicts the room temperature (RT) Mössbauer spectroscopy curve for CoFe2O4 powders that were calcined at 400, 800, and 1000 °C. The spectra comprise two sets of Zeeman sextet splits. Table 7 indicates that there were no significant changes in Mössbauer data for CoFe2O4 samples that were calcined at various temperatures. The H remained constant with increasing annealing temperature. This may be due to the high crystallinity of the samples without calcination, as observed in the XRD patterns. However, there were changes in the Mössbauer absorption area, indicating that the calcination temperature affected iron ions at the tetrahedral A and octahedral B sites.
Figure 16 shows the room temperature Mössbauer spectrum of Co0.7Zn0.3La0.01Fe1.99O4 ferrite. The spectra of all ferrites sintered at different temperatures were fitted using a six-baryon pattern.Table 8 shows no significant change in the Mössbauer parameters of Co0.7Zn0.3La0.01Fe1.99O4 calcined at different temperatures (unsintered and sintered at 400, 800 °C). The magnetic hyperfine field does not vary with annealing temperature. From the XRD spectrum, this may be due to the good crystallinity of the uncalcined sample.

2.4. Magnetic Property of Particles

The RT magnetic hysteresis curves of CoFe2O4 calcined at 800 and 1000 °C are presented in Figure 17. Table 9 shows that the saturation magnetization of CoFe2O4 increases with the annealing temperature, which is a result of the effect of calcination temperature on particle size [4]. The SEM results reveal that the higher calcination temperature led to greater crystallinity and larger average grain size of the CoFe2O4 sample. The coercivity of CoFe2O4 decreases as the annealing temperature increases [10]. The coercivity (HC) of the sample is related to the grain size (D) of the sample. When the sample is a single domain, their relationship is HC = gh/D2, and when the sample is multi-domain, their relationship is HC = a + b/D. In the single-domain region, the coercive force increases with the increase in grain size, while in the multi-domain region, the coercive force decreases with the increase in particle diameter. The critical size of cobalt ferrite is approximately 70 nm [13,21]. Based on SEM analysis, the particle size of our sample belongs to a multi-domain region, and the coercive force decreases with the increase in annealing temperature.
Figure 18 displays the RT magnetic hysteresis curve of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) measured at 295K. The magnetization of CoRExFe2−xO4 approaches saturation at 10,000 Oe. The magnetism of rare-earth elements mainly comes from the magnetic moment of their 4f electrons. Due to the inner arrangement of 4f electrons, their magnetic moment is very stable, allowing the magnetism of rare-earth elements to remain stable over a wide temperature range [16,29]. The Gd element has a Curie temperature of 293.2 K, close to RT (295 K) [18]. The magnetic dipoles are arranged disorderly at the temperature of 295 K; therefore, RE ion doping does not contribute to magnetization.
From Table 10, the substitution of RE3+ ions results in a reduction in the saturation magnetization, which can be calculated using the following equation [26,27]:
σ s = 5585 × n B M
where M is the molecular mass, and nB is the Bohr magneton.As RE3+ was replaced, the relative molecular mass of CoRExFe2−xO4 increased. La3+, Sm3+, Gd3+, Co2+, and Fe3+ ions have magnetic moments of 0μB, 1.7μB, 7.94μB, 3μB, and 5μB [11,15,17,18,28], respectively.
As a result, RE ions (La3+, Sm3+ and Gd3+) are considered nonmagnetic at 295 K. Co2+ ions prefer to occupy the B sites, and RE3+ ions only occupy the B sites due to their large ionic radius [6,7,18,30,31]. According to the Neel theory, the magnetic moment nB of (Fe)A[CoRExFe1−x]BO4 can be expressed as [20,29]:
nB = MBMA = 3 + 5 (1 − x) − 5 = 3 − 5x
where MA and MB are the magnetic moments of the A site and B site, respectively. The substitution of RE3+ ions leads to a decrease in the magnetic moment. The theoretical saturation magnetization also decreases, as per Equation (2), which is consistent with experimental results. Table 10 indicates an increase in the coercivity of CoRExFe2−xO4 with the addition of RE ions. For CoFe2O4 ferrite, the large coercivity value is primarily at the B site due to the anisotropy of the cobalt ions [6]. Coercivity is influenced by microstrain, magnetic particle morphology, and magnetic domain size [16,32,33]. Rare-earth ions (RE = La, Sm, and Gd) exhibit stronger magnetocrystalline anisotropy; therefore, the coercivity of cobalt ferrite increases with the substitution of RE3+ ions [7,8,21,29]. Additionally, the crystallite size of RE-substituted ferrite decreases with RE3+ ion substitution [13,16,34,35]. In our research, the grain size of CoFe2O4 calcined at 800 and 1000 °C falls within the multi-domain region; the grain size of the sample increases with increasing calcination temperature, so the coercivity decreases as the annealing temperature increases.
Figure 19 shows the room temperature hysteresis loop of Co0.7Zn0.3LaxFe2−xO4. For all samples, the magnetization reached saturation in an external field of 10,000 Oe.
In Table 11, the saturation magnetization decreases with an increase in La content. The doping of La element affects the magnetic properties of cobalt ferrite.Co2+ prefers octahedral sites, and the Zn2+ ion prefers tetrahedral sites, which is due to Co–Zn ferrite being an inverted spinel and La3+ ions having strong B site preference [14,16], so the cation distribution is (Zn0.3Fe)A [Co0.7LaxFe1−x]BO4 [43,44].
According to the two sublattice model of Néel’s theory, the magnetic moment nB is expressed by [46]:
nB = MBMA = 3 × 0.7 + 5(1 − x) − 5 = 2.5 − 5x
where MB and MA are the sublattice magnetic moments of the B and A positions. Figure 20 shows the changes in theoretical and experimental magnetic moments as La concentration increases. According to Figure 20, the experimental and theoretical magnetic moments decrease with an increase in La content x. According to Equation (3), the theoretical saturation magnetization decreases with an increase in La content x.
For all samples, the variation in experimental saturation magnetization is well consistent with the theoretical saturation magnetization. As shown in Table 11, the coercivity of Co0.7Zn0.3LaxFe2−xO4 decreases with increasing La content x. All samples display low coercivity, which shows that they are typical soft ferrites [46,47]. The large coercivity essentially originates from the magnetocrystalline anisotropy of the Co2+ ions in the octahedral site [47,48]. The coercivity decreases with increasing lanthanum content, which can be due to the weakening of anisotropy when the Co2+ ions migrate to the tetrahedral site. While x = 0.05 and x = 0.10, the coercivity has no significant changes, which may be attributed to the appearance of impurity phases LaFeO3 [47,49,50].

3. Experimental Method

Magnetic Functional Nanomaterials Co1−yZnyRExFe2−xO4 (RE (rare-earth) = La,Sm,Gd) were synthesized by the sol–gel combustion method. The synthetic raw material of the sample is analytically pure nitrate Fe(NO3)3·9H2O, M(NO3)2·6H2O (M = Co, Zn, La, Sm, Gd), ammonia (NH3·H2O), and citric acid (C6H8O7·H2O). Add deionized water to citric acid and metal nitrate to form the solution and adjust the pH value to around 7 by adding C6H8O7·H2O and NH3·H2O. Place the mixed solution in a constant temperature water bath at 80 °C, stir it electrically until the gel is dried, dry it further in an oven at 120 °C, and ignite it in air with a small amount of alcohol as an oxidant. After self-ignition, the powder material is annealed in a muffle furnace at a specific temperature. Place the mixed solution in a constant temperature water bath at 80 °C, stir it electrically until the gel is dried, dry it further in an oven at 120 °C, and ignite it in air with a small amount of alcohol as an oxidant. After self-ignition, the powder material is annealed at a specific temperature in a muffle furnace. The various analytical techniques (XRD, SEM, Mössbauer, VSM) were used to determine the following features: the impact of different doping amounts of La, Sm, Gd ion, calcination temperature, and calcination time on the structure, chemical bonding, particle shape and size, magnetic performance, and hyperfine interaction of samples.

4. Conclusions

We used the sol–gel combustion method to synthesize Co1−yZnyRExFe2−xO4 (RE =La, Sm, Gd) nanoparticles. CoFe2O4 samples calcined at different temperatures (unsintered and sintered at 400 °C, 800 °C, 1000 °C) have good crystallinity. Furthermore, the XRD results indicate that CoRExFe2−xO4 ferrites are single spinel-structured. SEM images suggest that the samples are well crystallized, with homogeneously distributed grains and composed of nanoparticles. RT Mössbauer spectroscopy of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) demonstrates ferrimagnetic behavior. Mössbauer spectra of CoFe2O4 reveal that the magnetic properties are influenced by the calcination temperature. The magnetization curve results suggest that RE ion doping impacts the coercivity and saturation magnetization, allowing the regulation of the sample’s magnetism through RE ion doping. The XRD patterns of Co0.7Zn0.3LaxFe2−xO4 confirm the main phase is a cubic spinel phase structure along with the appearance of impurity phases LaFeO3. It indicates that in spinel lattice, lanthanum has very low solubility. The average grain size decreases significantly with increasing La content. As the La content increases, the average grain size significantly decreases. The Mössbauer spectrum indicates that with the doping of nonmagnetic ions, the sample transitions from a ferrous magnetic state to a relaxed state, while the iron ions in the sample are in the Fe3+ state. The saturation magnetization and coercivity decrease with the increase in La content x. The change in coercivity is attributed to the decrease in magnetic crystal anisotropy and the presence of impurity phases LaFeO3.

Author Contributions

Conceptualization, Q.L. and J.L.; validation, J.L., X.Y. and K.S.; formal analysis, J.L., F.Y., X.Y. and K.S.; investigation, J.L., Y.H. and Q.L.; writing—original draft preparation, F.Y., X.Y. and Q.L.; writing—review and editing, J.L., K.S. and Q.L.; supervision, Q.L. and F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (NO. 12164006, 11364004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

J.L., X.Y. and K.S. contributed equally to this work. All authors discussed the results and commented on the manuscript. Co-corresponding authors Q.L. and F.Y. contributed equally to this work.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Szatmari, A.; Bortnic, R.; Souca, G.; Hirian, R.; Barbu-Tudoran, L.; Nekvapil, F.; Iacovita, C.; Burzo, E.; Dudric, R.; Tetean, R. The Influence of Zn Substitution on Physical Properties of CoFe2O4 Nanoparticles. Nanomaterials 2023, 13, 189. [Google Scholar] [CrossRef] [PubMed]
  2. Li, S.; Wang, X.; Ouyang, F.; Liu, R.; Xiong, X. Novel Functional Soft Magnetic CoFe2O4/Fe Composites: Preparation, Characterization, and Low Core Loss. Materials 2023, 16, 3665. [Google Scholar] [CrossRef] [PubMed]
  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. Waini, I.; Khan, U.; Zaib, A.; Ishak, A.; Pop, I.; Akkurt, N. Time-Dependent Flow of Water-Based CoFe2O4-Mn-ZnFe2O4 Nanoparticles over a Shrinking Sheet with Mass Transfer Effect in Porous Media. Nanomaterials 2022, 12, 4102. [Google Scholar] [CrossRef]
  5. Dascalu, G.; Popescu, T.; Feder, M.; Caltun, O.F. Structural, electric and magnetic properties of CoFe1.8RE0.2O4 (RE = Dy, Gd, La) bulk materials. J. Magn. Magn. Mater. 2013, 333, 69–74. [Google Scholar] [CrossRef]
  6. Tahar, L.B.; Artus, M.; Ammar, S.; Smiri, L.S.; Herbst, F.; Vaulay, M.-J.; Richard, V.; Grenèche, J.-M.; Villain, F.; Fiévet, F. Magnetic properties of CoFe1.9RE0.1O4 nanoparticles (RE = La, Ce, Nd, Sm, Eu, Gd, Tb, Ho) prepared in polyol. J. Magn. Magn. Mater. 2008, 320, 3242–3250. [Google Scholar] [CrossRef]
  7. Amiri, S.; Shokrollahi, H. Magnetic and structural properties of RE doped Co-ferrite (RE = Nd, Eu, and Gd) nano-particles synthesized by co-precipitation. J. Magn. Magn. Mater. 2013, 345, 18–23. [Google Scholar] [CrossRef]
  8. Zhao, L.; Yang, H.; Yu, L.; Cui, Y.; Zhao, X.; Feng, S. Study on magnetic properties of nanocrystalline La-, Nd-, or Gd-substituted Ni-Mn ferrite at low temperatures. J. Magn. Magn. Mater. 2006, 305, 91–94. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Wen, D. Infrared emission properties of RE (RE = La, Ce, Pr, Nd, Sm, Eu, Gd, Tb, and Dy) and Mn co-doped Co0.6Zn0.4Fe2O4 ferrites. Mater. Chem. Phys. 2012, 131, 575–580. [Google Scholar] [CrossRef]
  10. Kumar, L.; Kar, M. Effect of La3+ substitution on the structural and magnetocrystalline anisotropy of nanocrystalline cobalt ferrite (CoFe2−xLaxO4). Ceram. Int. 2012, 38, 4771–4782. [Google Scholar] [CrossRef]
  11. Ruiz, M.M.; Mietta, J.L.; Antonel, P.S.; Pérez, O.E.; Negri, R.M.; Jorge, G. Structural and magnetic properties of Fe2−xCoSmxO4-nanoparticles and Fe2−xCoSmxO4-PDMS magnetoelastomers as a function of Sm content. J. Magn. Magn. Mater. 2013, 327, 11–19. [Google Scholar] [CrossRef]
  12. Rana, A.; Thakur, O.P.; Kumar, V. Effect of Gd3+ substitution on dielectric properties of nano cobalt ferrite. Mater. Lett. 2011, 65, 3191–3192. [Google Scholar] [CrossRef]
  13. Guo, L.; Shen, X.; Song, F.; Lin, L.; Zhu, Y. Structure and magnetic property of CoFe2−xSmxO4 (x = 0–0.2) nanofibers prepared by sol-gel route. Mater. Chem. Phys. 2011, 129, 943–947. [Google Scholar] [CrossRef]
  14. Guo, L.; Shen, X.; Meng, X.; Feng, Y. Effect of Sm3+ ions doping on structure and magnetic properties of nanocrystalline NiFe2O4 fibers. J. Alloys Compd. 2010, 490, 301–306. [Google Scholar] [CrossRef]
  15. Jiang, J.; Yang, Y.-M.; Li, L.-C. Synthesis and magnetic properties of lanthanum-substituted lithium-nickel ferrites via a soft chemistry route. Phys. B 2007, 399, 105–108. [Google Scholar] [CrossRef]
  16. Jiang, J.; Yang, Y.-M. Effect of Gd substitution on structural and magnetic properties of Zn-Cu-Cr ferrites prepared by novel rheological technique. Mater. Sci. Technol. 2009, 25, 415. [Google Scholar] [CrossRef]
  17. Meng, Y.Y.; Liu, Z.W.; Dai, H.C.; Yu, H.Y.; Zeng, D.C.; Shukla, S.; Ramanujan, R. Structure and magnetic properties of Mn(Zn)Fe2−xRExO4 ferrite nano-powders synthesized by co-precipitation and refluxing method. Powder Technol. 2012, 229, 270–275. [Google Scholar] [CrossRef]
  18. Zhao, J.; Cui, Y.; Yang, H.; Yu, L.; Jin, W.; Feng, S. The magnetic properties of Ni0.7Mn0.3GdxFe2−xO4 ferrite. Mater. Lett. 2006, 60, 104–108. [Google Scholar] [CrossRef]
  19. Iqbal, M.A.; Islam, M.-u.; Ashiq, M.N.; Ali, I.; Iftikhar, A.; Khan, H.M. Effect of Gd-substitiution on physical and magnetic properties of Li1.2Mg0.4GdxFe(2−x)O4 ferrites. J. Alloys Compd. 2013, 579, 181–186. [Google Scholar] [CrossRef]
  20. Chand, J.; Kumar, G.; Kumar, P.; Sharma, S.K.; Knobel, M.; Singh, M. Effect of Gd3+ doping on magnetic, electric and dielectric properties of MgGdxFe2−xO4 ferrites processed by solid state reaction technique. J. Alloys Compd. 2011, 509, 9638–9644. [Google Scholar] [CrossRef]
  21. Panda, R.N.; Shih, J.C.; Chin, T.S. Magnetic properties of nano-crystalline Gd- or Pr-substituted CoFe2O4 synthesized by the citrate precursor technique. J. Magn. Magn. Mater. 2003, 257, 79–86. [Google Scholar] [CrossRef]
  22. Borhan, A.I.; Slatineanu, T.; Iordan, A.R.; Palamaru, M.N. Influence of chromium ion substitution on the structure and properties of zinc ferrite synthesized by the sol-gel-combution route. Polyhedron 2013, 56, 82–89. [Google Scholar] [CrossRef]
  23. Zhao, L.; Han, Z.; Yang, H.; Yu, L.; Cui, Y.; Jin, W.; Feng, S. Magnetic properties of nanocrystalline Ni0.7Mn0.3Gd0.1Fe1.9O4 ferrite at low temperatures. J. Magn. Magn. Mater. 2007, 309, 11–14. [Google Scholar] [CrossRef]
  24. Inbanathan, S.S.R.; Vaithyanathan, V.; Chelvane, J.A.; Markandeyulu, G.; Bharathi, K.K. Mössbauer studies and enhanced electrical properties of R (R = Sm, Gd and Dy) doped Ni ferrite. J. Magn. Magn. Mater. 2014, 353, 41–46. [Google Scholar] [CrossRef]
  25. Kumar, S.; Farea, A.M.M.; Batoo, K.M.; Lee, C.G.; Koo, B.H.; Yousef, A. Mössbauer studies of Co0.5CdxFe2.5−xO4 (0.0–0.5) ferrite. Phys. B 2008, 403, 3604–3607. [Google Scholar] [CrossRef]
  26. Al-Hilli, M.F.; Li, S.; Kassim, K.S. Structural analysis, magnetic and electrical properties of samarium substituted lithium-nickel mixed ferrites. J. Magn. Magn. Mater. 2012, 324, 873–879. [Google Scholar] [CrossRef]
  27. Gadkari, A.B.; Shinde, T.J.; Vasambekar, P.N. Magnetic properties of rare earth ion (Sm3+) added nanocrystalline Mg-Cd ferrites, prepared by oxalate co-precipitation method. J. Magn. Magn. Mater. 2010, 322, 3823–3827. [Google Scholar] [CrossRef]
  28. 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]
  29. Nikumbh, A.K.; Pawar, R.A.; Nighot, D.V.; Gugale, G.S.; Sangale, M.D.; Khanvilkar, M.B.; Nagawade, A.V. Structural, electrical, magnetic and dielectric properties of rare-earth substituted cobalt ferrites nanoparticles synthesized by the co-precipitation method. J. Magn. Magn. Mater. 2014, 355, 201–209. [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, 2015, 854840. [Google Scholar] [CrossRef]
  31. MRashad, M.M.; Mohamed, R.M.; El-Shall, H. Magnetic properties of nanocrystalline Sm-substituted CoFe2O4 synthesized by citrate precursor method. J. Mater. Process. Technol. 2008, 198, 139–146. [Google Scholar] [CrossRef]
  32. 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]
  33. Abraham, A.G.; Manikandan, A.; Manikandan, E.; Jaganathan, S.K.; Baykal, A.; Renganathan, P. Enhanced Opto-Magneto Properties of NixMg1−xFe2O4 (0.0≤ x≤ 1.0) Ferrites Nano-Catalysts. J. Nanoelectron. Optoelectron. 2017, 12, 1326–1333. [Google Scholar] [CrossRef]
  34. Amiri, S.; Shokrollahi, H. The role of cobalt ferrite magnetic nanoparticles in medical science. Mater. Sci. Eng. C 2013, 33, 1–8. [Google Scholar] [CrossRef]
  35. Gimaev, R.R.; Komlev, A.S.; Davydov, A.S.; Kovalev, B.B.; Zverev, V.I. Magnetic and Electronic Properties of Heavy Lanthanides (Gd, Tb, Dy, Er, Ho, Tm). Crystals 2021, 11, 82. [Google Scholar] [CrossRef]
  36. Ahmed, M.A.; EL-Sayed, M.M.; EL-Desoky, M.M. The effect of highly activated hopping process on the physical properties of Co-Zn-La ferrite. Phys. B 2010, 405, 727–731. [Google Scholar] [CrossRef]
  37. Ishaque, M.; Islam, M.U.; Ali, I.; Khan, M.A.; Rahman, I.Z. Electrical transport properties of Co-Zn-Y-Fe-O system. Ceram. Int. 2012, 38, 3337–3342. [Google Scholar] [CrossRef]
  38. 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]
  39. Veverkaa, M.; Veverka, P.; Jirák, Z.; Kaman, O.; Knížek, K.; Maryško, M.; Pollert, E.; 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]
  40. Kulal, S.R.; Khetre, S.S.; Jagdale, P.N.; Gurame, V.M.; Waghmode, D.P.; Kolekar, G.B.; Sabale, S.R.; Bamane, S.R. Synthesis of Dy doped Co-Zn ferrite by sol-gel auto combustion method and its characterization. Mater. Lett. 2012, 84, 169–172. [Google Scholar] [CrossRef]
  41. Roy, P.K.; Bera, J. Enhancement of the magnetic properties of Ni-Cu-Zn ferrites with the substitution of a small fraction of lanthanum for iron. Mater. Res. Bull. 2007, 42, 77–83. [Google Scholar] [CrossRef]
  42. Reddy, M.P.; Kim, I.G.; Yoo, D.S.; Madhuri, W.; Venkata Ramana, A.; Shaaban, N.; Ramamanohar Reddy, K.V.; Siva Kumar, R. Ramakrishna Reddy. Effect of La substitution on structural and magnetic properties of microwave treated Mg0.35Cu0.05Zn0.60LaxFe2−xO4 ceramics. Superlattices Microstruct. 2013, 56, 99–106. [Google Scholar] [CrossRef]
  43. Gabal, M.A.; Asiri, A.M.; AlAngari, Y.M. On the structural and magnetic properties of La-substituted NiCuZn ferrites prepared using egg-white. Ceram. Int. 2011, 37, 2625–2630. [Google Scholar] [CrossRef]
  44. Bayoumi, W. Structural and electrical properties of zinc-substituted cobalt ferrite. J. Mater. Sci. 2007, 42, 8254–8261. [Google Scholar] [CrossRef]
  45. Bogacz, B.F.; Gargula, R.; Kurzydo, P.; Pdziwiatr, A.T.; Paliychuk, N. Two-Level Model Description of Superparamagnetic Relaxation in Nanoferrites (Co,Zn)Fe2O4. Acta Phys. Pol. A 2018, 134, 993–998. [Google Scholar] [CrossRef]
  46. Kumar, P.; Sharma, S.K.; Knobel, M.; Singh, M. Effect of La3+ doping on the electric, dielectric and magnetic properties of cobalt ferrite processed by co-precipitation technique. J. Alloys Compd. 2010, 508, 115–118. [Google Scholar] [CrossRef]
  47. Zhao, L.; Yang, H.; Zhao, X.; Yu, L.; Cui, Y.; Feng, S. Magnetic properties of CoFe2O4 ferrite doped with rare earth ion. Mater. Lett. 2006, 60, 1–6. [Google Scholar] [CrossRef]
  48. Kumar, A.; Shen, J.; Yang, W.; Zhao, H.; Sharma, P.; Varshney, D.; Li, Q. Impact of Rare Earth Gd3+ Ions on Structural and Magnetic Properties of Ni0.5Zn0.5Fe2−xGdxO4 Spinel Ferrite: Useful for Advanced Spintronic Technologies. J. Supercond. Nov. Magn. 2018, 31, 1173–1182. [Google Scholar] [CrossRef]
  49. Kumar, S.; Ahmed, F.; Shaalan, N.M.; Kumar, R.; Alshoaibi, A.; Arshi, N.; Dalela, S.; Sayeed, F.; Dwivedi, S.; Kumari, K. Structural, Magnetic, and Electrical Properties of CoFe2O4 Nanostructures Synthesized Using Microwave-Assisted Hydrothermal Method. Materials 2022, 15, 7955. [Google Scholar] [CrossRef]
  50. Yunasfi, Y.; Mashadi, M.; Winatapura, D.S.; Setiawan, J.; Taryana, Y.; Gunanto, Y.E.; Adi, W.A. Enhanced Magnetic and Microwave Absorbing Properties of Nd3+ Ion Doped CoFe2O4 by Solid-State Reaction MethodPhys. Phys. Status Solidi A Appl. Mater. Sci. 2023, 220, 2200718. [Google Scholar] [CrossRef]
Figure 1. XRD diffraction patterns of the CoFe2O4 unsintered and sintered at 400, 800, and 1000 °C.
Figure 1. XRD diffraction patterns of the CoFe2O4 unsintered and sintered at 400, 800, and 1000 °C.
Molecules 28 06280 g001
Figure 2. XRD diffraction patterns of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Figure 2. XRD diffraction patterns of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Molecules 28 06280 g002
Figure 3. XRD diffraction patterns of Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Figure 3. XRD diffraction patterns of Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Molecules 28 06280 g003
Figure 4. The variation of lattice parameter and average crystallite size for Co0.7Zn0.3LaxFe2−xO4.
Figure 4. The variation of lattice parameter and average crystallite size for Co0.7Zn0.3LaxFe2−xO4.
Molecules 28 06280 g004
Figure 5. X-ray diffraction patterns of Co0.7Zn0.3La0.01Fe1.99O4 calcined at different temperatures.
Figure 5. X-ray diffraction patterns of Co0.7Zn0.3La0.01Fe1.99O4 calcined at different temperatures.
Molecules 28 06280 g005
Figure 6. SEM micrographs of CoRExFe2−xO4 (x = 0, 0.02; Re = La, Sm, Gd) sintered at 800 °C.
Figure 6. SEM micrographs of CoRExFe2−xO4 (x = 0, 0.02; Re = La, Sm, Gd) sintered at 800 °C.
Molecules 28 06280 g006
Figure 7. Grain size of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Figure 7. Grain size of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Molecules 28 06280 g007
Figure 8. Scanning electron microscopy micrographs and grain size distribution diagram for CoFe2O4 sintered at 1000 °C.
Figure 8. Scanning electron microscopy micrographs and grain size distribution diagram for CoFe2O4 sintered at 1000 °C.
Molecules 28 06280 g008
Figure 9. SEM micrographs of Co0.7Zn0.3La0.05Fe1.95O4 (x = 0.05) and Co0.7Zn0.3La0.10Fe1.90O4 (x = 0.10) calcined at 800 °C.
Figure 9. SEM micrographs of Co0.7Zn0.3La0.05Fe1.95O4 (x = 0.05) and Co0.7Zn0.3La0.10Fe1.90O4 (x = 0.10) calcined at 800 °C.
Molecules 28 06280 g009
Figure 10. Histogram of grain size distribution of Co0.7Zn0.3La0.05Fe1.95O4 (x = 0.05) and Co0.7Zn0.3La0.10Fe1.90O4 (x = 0.10) calcined at 800 °C.
Figure 10. Histogram of grain size distribution of Co0.7Zn0.3La0.05Fe1.95O4 (x = 0.05) and Co0.7Zn0.3La0.10Fe1.90O4 (x = 0.10) calcined at 800 °C.
Molecules 28 06280 g010
Figure 11. Room temperature Mössbauer spectroscopy curve for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C. (The blue and purple line spectra represent the Mössbauer spectra of iron ions in the A and B lattice, while the red line spectra represent the total Mössbauer spectra).
Figure 11. Room temperature Mössbauer spectroscopy curve for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C. (The blue and purple line spectra represent the Mössbauer spectra of iron ions in the A and B lattice, while the red line spectra represent the total Mössbauer spectra).
Molecules 28 06280 g011
Figure 12. The spatial structure of spinel ferrite cells.
Figure 12. The spatial structure of spinel ferrite cells.
Molecules 28 06280 g012
Figure 13. Room temperature Mössbauer spectroscopy curve of Co0.7Zn0.3LaxFe2–xO4 calcined at 800 °C. (The blue and purple line spectra represent the Mössbauer spectra of iron ions in the A and B lattice, while the red line spectra represent the total Mössbauer spectra).
Figure 13. Room temperature Mössbauer spectroscopy curve of Co0.7Zn0.3LaxFe2–xO4 calcined at 800 °C. (The blue and purple line spectra represent the Mössbauer spectra of iron ions in the A and B lattice, while the red line spectra represent the total Mössbauer spectra).
Molecules 28 06280 g013
Figure 14. The A site of the tetrahedral lattice and the B site of the octahedral lattice. (1–6 and a–d represent the Oxygen ion).
Figure 14. The A site of the tetrahedral lattice and the B site of the octahedral lattice. (1–6 and a–d represent the Oxygen ion).
Molecules 28 06280 g014
Figure 15. Room temperature Mössbauer spectroscopy curve for CoFe2O4 unsintered and sintered at 400, 800, and 1000 °C. (The blue and purple line spectra represent the Mössbauer spectra of iron ions in the A and B lattice, while the red line spectra represent the total Mössbauer spectra).
Figure 15. Room temperature Mössbauer spectroscopy curve for CoFe2O4 unsintered and sintered at 400, 800, and 1000 °C. (The blue and purple line spectra represent the Mössbauer spectra of iron ions in the A and B lattice, while the red line spectra represent the total Mössbauer spectra).
Molecules 28 06280 g015
Figure 16. Room temperature Mössbauer spectroscopy curve for Co0.7Zn0.3La0.01Fe1.99O4 sintered at 0, 400, and 800 °C. (The red line spectra represent theMössbauer spectra of iron ions in the B lattice and the total Mössbauer spectra).
Figure 16. Room temperature Mössbauer spectroscopy curve for Co0.7Zn0.3La0.01Fe1.99O4 sintered at 0, 400, and 800 °C. (The red line spectra represent theMössbauer spectra of iron ions in the B lattice and the total Mössbauer spectra).
Molecules 28 06280 g016
Figure 17. Room temperature magnetic hysteresis curve of CoFe2O4 sintered at 800 and 1000 °C.
Figure 17. Room temperature magnetic hysteresis curve of CoFe2O4 sintered at 800 and 1000 °C.
Molecules 28 06280 g017
Figure 18. Room temperature magnetic hysteresis curve of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Figure 18. Room temperature magnetic hysteresis curve of CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Molecules 28 06280 g018
Figure 19. Hysteresis loops of Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Figure 19. Hysteresis loops of Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Molecules 28 06280 g019
Figure 20. Variation of theoretical and experimental magnetic moment with lanthanum substitution.
Figure 20. Variation of theoretical and experimental magnetic moment with lanthanum substitution.
Molecules 28 06280 g020
Table 1. XRD pattern data for CoFe2O4 unsintered and sintered at 400, 800, and 1000 °C.
Table 1. XRD pattern data for CoFe2O4 unsintered and sintered at 400, 800, and 1000 °C.
Reaction Temperature (Centigrade)Lattice Parameter (Å)Average Crystallite Size (Å)Density (g/cm3)
unsintered8.412803615.2347
4008.391494075.2747
8008.354975565.3468
10008.386155205.2847
Table 2. XRD pattern data for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd.) sintered at 800 °C.
Table 2. XRD pattern data for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd.) sintered at 800 °C.
Content (x)Lattice Parameter (Å)Average Crystallite Size (Å)Density (g/cm3)
none8.354975565.3468
La8.379883915.3341
Sm8.386293745.3279
Gd8.404574025.2954
Table 3. XRD data for Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Table 3. XRD data for Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Content(x)Lattice Parameter (Å)Average Crystallite Size (Å)Density (g/cm3)
08.411964555.2794
0.018.423013545.2769
0.038.391982905.3732
0.058.397292275.4003
0.078.390912335.4500
0.098.378991995.5108
0.108.389842155.5081
0.158.402062575.5771
0.208.403662425.6669
Table 4. XRD pattern data for Co0.7Zn0.3La0.01Fe1.99O4 calcined at different temperatures.
Table 4. XRD pattern data for Co0.7Zn0.3La0.01Fe1.99O4 calcined at different temperatures.
Reaction Temperature (Centigrade)Lattice Parameter (Å)Average Cryst Size (Å)Density (g/cm3)
unsintered8.427982365.2678
4008.435712275.2533
8008.423013545.2769
Table 5. The room temperature Mössbauer spectroscopy data of specimens CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Table 5. The room temperature Mössbauer spectroscopy data of specimens CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
ContentComponentI.S. (mm/s)Q.S. (mm/s)H (T)Γ (mm/s)A0 (mm/s)
noneSextet (1)0.237−0.00448.9460.36032.4
Sextet (2)0.375−0.02445.6950.32267.6
LaSextet (1)0.228−0.00549.0230.36324.5
Sextet (2))0.342−0.03546.3810.35675.5
SmSextet (1)0.229−0.00349.1460.40027.6
Sextet (2)0.368−0.00544.8270.32872.4
GdSextet (1)0.2310.01549.2010.35725.5
Sextet (2)0.349−0.02046.1730.34274.5
Table 6. The room temperature Mössbauer spectroscopy data of Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Table 6. The room temperature Mössbauer spectroscopy data of Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Content(x)ComponentI.S. (mm/s)Q.S. (mm/s)H(T)Γ (mm/s)A0 (%)
0Sextet (1)0.2350.05547.5080.42911
Sextet (2)0.306−0.05038.9460.33889
0.01Sextet (2)0.276−0.05539.4160.343100
0.03Sextet (2)0.306−0.00238.3960.346100
0.05Sextet (2)0.3160.00137.3220.424100
0.07Sextet (2)0.282−0.06936.7020.334100
0.09Sextet (2)0.296−0.10335.3110.320100
0.10Sextet (2)0.266−0.05834.7000.401100
0.15Sextet (2)0.285−0.05834.7800.332100
0.20Sextet (2)0.311−0.00534.4760.46997.2
Doublet0.3500.445-0.2962.8
Table 7. The Mössbauer spectroscopy data of CoFe2O4 unsintered and calcined at 400 °C, 800 °C, and 1000 °C.
Table 7. The Mössbauer spectroscopy data of CoFe2O4 unsintered and calcined at 400 °C, 800 °C, and 1000 °C.
Reaction Temperature (Centigrade)ComponentI.S. (mm/s)Q.S. (mm/s)H (T)Γ (mm/s)A0 (mm/s)
unsinteredSextet (1)0.246−0.02249.2770.40126.5
Sextet (2)0.357−0.00345.7560.34173.5
400Sextet (1)0.2410.00249.3230.42230.4
Sextet (2)0.367−0.01745.0780.31969.6
800Sextet (1)0.237−0.00448.9460.36032.4
Sextet (2)0.375−0.02445.6950.32267.6
1000Sextet (1)0.238−0.01148.8520.36628.4
Sextet (2)0.3550.000445.8890.33871.6
Table 8. The room temperature Mössbauer spectroscopy data of Co0.7Zn0.3La0.01Fe1.99O4 calcined at 800 °C.
Table 8. The room temperature Mössbauer spectroscopy data of Co0.7Zn0.3La0.01Fe1.99O4 calcined at 800 °C.
Reaction Temperature (Centigrade)ComponentI.S. (mm/s)Q.S. (mm/s)H (T)Γ (mm/s)A0 (mm/s)
unsinteredSextet (2)0.3080.00139.4250.363100
400Sextet (2)0.300−0.01739.1230.351100
800Sextet (2)0.276−0.05539.4160.343100
Table 9. Magnetic data for CoFe2O4 sintered at 800 °C and 1000 °C.
Table 9. Magnetic data for CoFe2O4 sintered at 800 °C and 1000 °C.
Reaction Temperature (Centigrade)Ms (emu/g)Hc (Oe)Mr (emu/g)nB
80072.581005.3334.713.05
100080.89802.7737.153.40
Table 10. Magnetic data for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
Table 10. Magnetic data for CoRExFe2−xO4 (x = 0, 0.02; RE = La, Sm, Gd) sintered at 800 °C.
SampleMs (emu/g)Hc (Oe)Mr (emu/g)nB
none72.581005.3334.713.05
La70.921254.0037.403.00
Sm71.501367.6638.163.03
Gd69.971351.2538.682.96
Table 11. Magnetic data for Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Table 11. Magnetic data for Co0.7Zn0.3LaxFe2−xO4 calcined at 800 °C.
Content (x)Ms (emu/g)Hc (Oe)Mr (emu/g)nB
083.51301.7525.003.54
0.0180.02250.9729.583.40
0.0565.22200.9019.632.81
0.1064.02200.8919.042.81
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lin, J.; Yang, X.; Su, K.; Yang, F.; He, Y.; Lin, Q. Effects of Nonmagnetic Zn2+ Ion and RE Ion Substitution on the Magnetic Properties of Functional Nanomaterials Co1−yZnyRExFe2−xO4 (RE = La, Sm, Gd) by Sol–Gel. Molecules 2023, 28, 6280. https://doi.org/10.3390/molecules28176280

AMA Style

Lin J, Yang X, Su K, Yang F, He Y, Lin Q. Effects of Nonmagnetic Zn2+ Ion and RE Ion Substitution on the Magnetic Properties of Functional Nanomaterials Co1−yZnyRExFe2−xO4 (RE = La, Sm, Gd) by Sol–Gel. Molecules. 2023; 28(17):6280. https://doi.org/10.3390/molecules28176280

Chicago/Turabian Style

Lin, Jinpei, Xingxing Yang, Kaimin Su, Fang Yang, Yun He, and Qing Lin. 2023. "Effects of Nonmagnetic Zn2+ Ion and RE Ion Substitution on the Magnetic Properties of Functional Nanomaterials Co1−yZnyRExFe2−xO4 (RE = La, Sm, Gd) by Sol–Gel" Molecules 28, no. 17: 6280. https://doi.org/10.3390/molecules28176280

Article Metrics

Back to TopTop