Next Article in Journal
Review of Creep-Thermomechanical Fatigue Behavior of Austenitic Stainless Steel
Next Article in Special Issue
Magnetic and Electronic Properties of Edge-Modified Triangular WS2 and MoS2 Quantum Dots
Previous Article in Journal
Formation and Morphological Transition of Diversified Eutectic Structures in Ag-Based Brazing Filler Metals with Sn Addition
Previous Article in Special Issue
Tunable Sensing and Transport Properties of Doped Hexagonal Boron Nitride Quantum Dots for Efficient Gas Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Behavior of Virgin and Lithiated NiFe2O4 Nanoparticles

by
Ghadah M. Al-Senani
1,
Foziah F. Al-Fawzan
1,
Rasmiah S. Almufarij
1,
Omar H. Abd-Elkader
2,3,* and
Nasrallah M. Deraz
4
1
Department of Chemistry, College of Science, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
Physics and Astronomy Department, Science College, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
3
Electron Microscope and Thin Films Department, Physics Research Institute, National Research Centre, P.O. Box 21111, Giza 12622, Egypt
4
Physical Chemistry Department, National Research Centre, P.O. Box 21111, Giza 12622, Egypt
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(1), 69; https://doi.org/10.3390/cryst13010069
Submission received: 22 November 2022 / Revised: 23 December 2022 / Accepted: 28 December 2022 / Published: 31 December 2022
(This article belongs to the Special Issue Crystalline Magnetic Compounds)

Abstract

:
A series of virgin and lithia-doped Ni ferrites was synthesized using egg-white-mediated combustion. Characterization of the investigated ferrites was performed using several techniques, specifically, X-ray Powder Diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), and High-resolution transmission electron microscopy (HRTEM). XRD-based structural parameters were determined. A closer look at these characteristics reveals that lithia doping enhanced the nickel ferrite lattice constant (a), unit cell volume (V), stress (ε), microstrain (σ), and dislocation density (δ). It also enhanced the separation between magnetic ions (LA and LB), ionic radii (rA, rB), and bond lengths (A-O and B-O) between tetrahedral (A) and octahedral (B) locations. Furthermore, it enhanced the X-ray density (Dx) and crystallite size (d) of random spinel nickel ferrite displaying opposing patterns of behavior. FTIR-based functional groups of random spinel nickel ferrite were determined. HRTEM-based morphological properties of the synthesized ferrite were investigated. These characteristics of NiFe2O4 particles, such as their size, shape, and crystallinity, demonstrate that these manufactured particles are present at the nanoscale and that lithia doping caused shape modification of the particles. Additionally, the prepared ferrite’s surface area and total pore volume marginally increased after being treated with lithia, depending on the visibility of the grain boundaries. Last, but not least, as the dopant content was increased through a variety of methods, the magnetization of virgin nickel ferrite fell with a corresponding increase in coercivity. Uniaxial anisotropy, rather than cubic anisotropy, and antisite and cation excess defects developed in virgin and lithia-doped nickel ferrites because the squareness ratio (Mr/Ms) was less than 0.5. Small squareness values strongly recommend using the assessed ferrites in high-frequency applications.

1. Introduction

Most ferrite-based nano particles (NPs) result from the solid-state reaction between the oxides of their constituent elements. The preparation process, which is among the most crucial of these variables, determines the type, structure, and behavior of these ferrites. One of the most crucial techniques for improving and controlling ferrite structures, as well as their optical, morphological, electrical, and magnetic properties, is the doping of ferrites with specific elements.
Doping of ferrites could result in many effects, including the following: (i) alterations in the cation distribution between the ferrite structure’s tetrahedral (A) and octahedral (B) locations [1,2,3,4,5,6,7,8]; (ii) Alteration in the energy of the ferrite grain boundaries, which drives the evolution of the grain [1]; (iii) Alternation in the size of ferrite particle by contraction or expansion of the ferrite lattice; and (iv) Creation of oxygen vacancies with a subsequent change in the ferrite surface. In other words, the control in the microstructures of the ferrites can be achieved by doping of the host matrices with smaller cations [2]. In fact, the magnetic properties of cobalt ferrite were improved by doping it with alumina or zincate [3,4]. The magnetic qualities of zinc ferrite can be controlled by varying the dopant content. The addition of 12–46 weight percent of Li to zinc ferrite when the dopant content was increased inhibited magnetization after being enhanced by Li2O [5]. However, doping zinc and nickel ferrites with alumina and magnesia led to a reduction in magnetism [6,7,8].
The preparation method is one of several variables that determine the physicochemical characteristics of ferrites and is expected to be essential for obtaining high-purity, nanoscale ferrites with particular attributes. Nanostructure ferrites can all be prepared using a wide range of techniques, including sol–gel, co-precipitation, refluxing, solvothermal, mechanochemical, ceramic, combustion, microwave plasma, micro emulsion, and mechanical alloying [1,4,5,6,7,8,9,10]. The main draw-back of the majority of earlier techniques is the low crystallinity of the resulting ferrite nanoparticles, which can be improved by further heat treatment [11].
Among the preparation routes of nano crystalline ferrites is the auto-combustion method, which has various advantages. Firstly, it is a simple and low-cost process depending upon a short reaction time and low processing temperature with subsequent low energy consumption [12,13]. Secondly, this approach possesses a high degree of purity, a homogeneous chemical composition, and good sinterability [13]. Thirdly, the structure, physicochemical properties, surface properties, and magnetic behavior are all under good control [13]. To put it another way, it demonstrates control over particle growth, reduces agglomeration of nanoparticles, increases their stability, and optimizes their magnetic buildability. Finally, by selecting a suitable aqueous or non-aqueous solvent mixture, by adjusting the kind and content of fuel utilized, it enables the large-scale manufacture of ferrites with controllable size and shape [4,5,6,7,8,14]. In fact, the nitrate salts are frequently used in preparation of spinel ferrites by the combustion method because these salts are water-soluble and low-temperature oxidizing agents [4,5,6,7,8].
In reality, ferrite-based nanoparticles (NPs) constitute a burgeoning technological and economic sector with complete expansion in a wide range of industrial and biomedical application domains. It is possible to attribute the popularity of nano sized ferrites in technological breakthroughs to their tunable physicochemical properties, which include structural, microstructural, surface, electrical and thermal conductivity, catalytic activity, and light absorption and scattering, and result in enhanced performance over their bulk counterparts. Ferrite-based materials are versatile and effective in the industry sector, for example, in magnetic recording media, photoelectric devices, sensors, magnetic pigments, storage devices, and batteries and solar cells. Ferrites can also be applied in the biomedical field, such as in controlled drug delivery, tumor treatment, magnetic resonance imaging, biomagnetic separation, cellular therapy, tissue repair, cell separation, and bio sensing [1,2,3,9,10,12,15,16,17,18,19,20].
This study aims the green synthesis of virgin spinel nickel ferrite nano particles (NiFe2O4 NPs) and nano particles doped with different amounts of lithia via the egg-white-associated auto-combustion method. This investigation is particularly interesting because there is a dearth of knowledge regarding the effects of lithium doping on the development and various aspects of virgin spinel nickel ferrite. We, therefore, investigated the effects of lithium doping on the formation, structural, morphological, surface, and magnetic characteristics of NiFe2O4 NPs.

2. Materials and Methods

2.1. Materials

Lithium nitrate, ferric nitrate hydrate, and nickel (II) nitrate all have linear formulae. The chemical ingredients employed were LiNO3, Fe (NO3)3.9H2O, and Ni (NO3)2.6H2O, respectively. These materials were supplied by the Sigma-Aldrich Company (Darmstadt, Taufkirchen, Germany). These reagents were applied quantitatively and required no further processing. Egg white was sourced from the raw eggs of local hens.

2.2. Preparation Method

One sample (S1) of virgin NiFe2O4 and two samples (S2 and S3) were fabricated by using the egg-white-mediated combustion route. Taking the stoichiometric ratio of Fe/Ni = 2 into consideration, the first sample (S1) was made by completely combining an equimolar combination of ferric nitrate hydrate and nickel nitrate hexahydrate with 6 mL of egg white in a crucible. The resulting liquid was first swirled at 60 °C in order to allow the water to evaporate and to boost viscosity. The mixture was subsequently converted into a gel by raising the temperature to 120 °C. The created precursor gel was calcined at 300 °C for 15 min to raise the temperature to a crucible level. A large amount of foam had already started to form when a spark erupted in one corner and immediately spread throughout the mass; the final product was a dense, fluffy solid. Two samples (S2 and S3) were prepared following the same procedure but in the presence of 6 mole % and 12 mole % lithium nitrate, respectively.

2.3. Characterization Systems

Using a BRUKER D8 advance diffractometer (Karlsruhe, Germany) and X-ray diffraction technology, measurements of diverse nanoparticles were carried out. The patterns were operated with Cu Kα radiation at 40 kV, 40 mA, and a scanning rate of 2° per minute. In order to calculate the mean crystallite size (d), dislocation density (δ), stress (ε), and strain (σ) of NiFe2O4 involved in the investigated product, and based on calculations using the Scherrer equation and X-ray diffraction line widening, Equations (1) through (4) were used [21,22]:
d = (B λ)/β cos θ
δ = 1/d2
ε = β cos θ/4
ơ = ε Y
where d is the average crystallite size of the phase being studied, B is the Scherrer constant (0.89), λ is the employed X-ray beam wavelength, β is the full-width half maximum (FWHM) of diffraction, θ is the Bragg’s angle, and Y is the Young’s modulus.
Fourier-transmission infrared spectra (FTIR) of different materials were measured using a Perkin-Elmer Spectrophotometer (type 1430). Two milligrams of each solid sample were mixed with 200 mg of vacuum-dried IR-grade KBr. A range of 1000–4000 cm−1 was used to measure the FTIR spectra. A steel die with a 13 mm diameter was used to scatter the mixture after it had been processed in a vibrating ball mill for three minutes. The identical disks were placed in the double grating FTIR spectrophotometer holder.
Utilizing JEOL JAX-840A and JEOL Model 1230 transmission electron microscopes (TEM) operating at 100 Kev, the specimens were placed in a specific solvent, and a deep coating of copper grid with carbon film was applied before the grid was allowed to completely evaporate the solvent before examination. The sample was briefly ultrasonically dispersed to spread individual particles over the mount arrangement and copper grids.
Utilizing both standard volumetric instruments (Brunauer–Emmett–Teller method) and surface area analyzers from the Micromeritics Gemini VII 2390 V1.03 series, the specific surface area (SBET), total pore volume (VP), and mean pore radius (ȓ) of various materials were evaluated (Microtrac, Alpharetta, GA, USA). Prior to the measurements, each sample was out-gassed for two hours at the lower pressure of 105 Torr at a temperature of 200 °C. A vibrating sample magnetometer (VSM; 9600-1 LDJ, USA) with a 20 kG maximum applied field was used to investigate the magnetic properties of the under-recognized solids.

3. Results

3.1. XRD Analysis

XRD analysis as a structural technique enabled us to study the purity and formation phase with subsequent investigation of different structural properties for the undoped and Li-doped NiFe2O4 studied. Figure 1 displays XRD patterns of virgin NiFe2O4 NPs (S1) and those doped by (S2 and S3) at two lithium concentrations (3 and 6 mole%, respectively). Table 1 lists the Miller indices, d spacing, and values of 2θ for the crystalline phases that are present in the as-synthesized solids. In contrast, the whole spectrum of diffraction peaks of the virgin sample is indexed as (220), (311), (222), (400), (422), (511), (440), (620), (533), (522), and (444) crystal planes at 2θ = 30.33°, 35.40°, 37.18°, 43.10°, 53.80°, 57.46°, 63.01°, 71.55°, 74.55°, and 75° (NiFe2O4-PDF: 44-1485). From this figure, it is evident that Li-doped samples do not show peaks related to any phases other than those belonging to the NiFe2O4-based spinel type structure. Pure sample lithia doping resulted in a shift in all diffraction peaks, as well as a reduction in the height and FWHM of various peaks. For all diffraction peaks, an increase in lithia concentration would, in actuality, lead to a greater in shift and a reduction in both the intensity and FWHM. The cation distribution on the tetrahedral and octahedral sites included in the spinel structure could be estimated from the literature by tracing the peak height at the (220) and (440) reflection planes [8]. The cations on the tetrahedral, octahedral, and oxygen ion parameters, respectively, have a strength and durability on the intensities of the (220), (440), and (511) planes. Therefore, the comparison of the peak height at these planes is very necessary and important, especially after doping of virgin NiFe2O4 with lithia. It was found that the peak height at the (440) reflection plane was greater than that at the (220) reflection plane with an increase in the content of lithium.
It was essential to calculate various parameters that depended on the XRD data in order to complete the majority of the information regarding the structural properties of the analyzed NiFe2O4. The following were some of these criteria: crystallite size (d), lattice constant (a), unit cell volume (V), X-ray density (Dx), stress (ε), microstrain (σ), and dislocation density (δ), which are the variables listed in Table 2. Additional criteria were the separation between the ionic radii (rA, rB), bond lengths (A-O and B-O), and magnetic ions (LA and LB) on the tetrahedral (A) and octahedral (B) sites, whereas based on the XRD data, Table 2 already included all the calculated values for these parameters, this table shows that the values of δ, ε, σ, a, LA, LB, rA, rB, A-O, and B-O of NiFe2O4 NPs grew noticeably with the increase in the dopant content. Although the behavior varied depending on the values of “d” and “Dx”, crystallinity declined, and the crystal grain size increased as the dopant content rose. The production of an intrinsic microstrain and a consequent reduction in crystallite size may be responsible for the broadening of the XRD peak. Due to changes in the lattice parameter imposed by crystal defects and dislocations, the microstrain was established.

3.2. FTIR Analysis

The majority of the features of spinel materials are based on their cation distribution; understanding this is, therefore, a requirement of this study. By using infrared (IR) analysis, the cation distribution of cubic spinel structures can be investigated. This method frequently reveals two fundamental vibration modes at 600 cm−1 and 400 cm−1, respectively, which are related to the distribution of cations at tetrahedral (A-) and octahedral (B-) sites. The S1, S2, and S3 sample room-temperature FTIR spectra, in the 4000–400 cm−1 region, are shown in Figure 2. Study of this figure reveals that the virgin sample exhibited two main absorption bands (υ1 and υ2) at 867–595 cm−1 and 550–400 cm−1. These bands confirm the formation of the spinel-type structure, as shown as in Table 3. However, these bands could be attributed to the metal–oxygen (M–O) bond-stretching vibrations, i.e., Fe-O and Ni-O vibrations at both A- and B-sites involved in the spinel ferrite. In other words, these bands confirm that Ni and Fe cations can be distributed between A- and B-sites. In this regard, the authors expect that two small bands (υ1* and υ2*) located at 867 cm−1 and 550 cm−1, may be related to Ni–O vibrations at both A- and B-sites, indicating the formation of random spinel structures. These findings, therefore, refer to the fabrication of random spinel structures. Furthermore, the position of these bands shifted when lithium, in varying amounts, was added to the virgin sample, showing that lithium ions were incorporated into the nickel ferrite lattice. Due to the doping with 3 and 6 mol% Li2O, respectively, two shoulder bands (υ1**) were observed at 760 cm−1 and 767 cm−1. These bands may indicate Li–O vibration. In addition, the stretching and bending vibrations of the hydroxyl groups (O-H) for the adsorbed water molecules on the sample surfaces may be responsible for the bands at 3433.6–3428.8 cm−1 and 1636.3–1633.4 cm−1. The blending of the samples with KBr media causes the adsorption of water molecules on their surfaces [22]. Nevertheless, the bands at 2924.6–2923.52, 1385.60–1384.64, and 1049.10–1044.26 cm−1 could be attributable to the stretching vibrations of hydrogen carbon and hydroxyl carbon (C-H and C-O-H). These bands are dependent on the existence of carbon traces that originated from the egg-white internal combustion process [23].

3.3. TEM Analysis

Transmission electron microscopy, TEM, must be employed to establish the impact of lithia doping on the microstructure of virgin nickel ferrite. The particle-size distribution of pure (S1) and 6 mol% Li2O-doped NiFe2O4 particles (S2) are shown in TEM, high resolution (HR)-TEM, and selected area electron diffraction (SAED) pictures, respectively. Analysis of these graphs show that the morphology of the S3 sample changed significantly compared with that of the pristine S1 sample, as received (TEM photographs, Figure 3a and Figure 4a). The microstructural analysis of the S1 sample shows that as-received particles are almost spherical with some agglomerations. On the other hand, irregular spherically shaped particles wholly disappeared, and instead, irregularly cubic particles emerged as shown in the S3 sample. This observation confirms significant changes in the microstructure of NiFe2O4 as a result of the doping process. Generally, the TEM micrographs of the investigated samples indicate that the distribution of grains is of nonuniform size. Different factors, including the diffusion coefficient and the concentration of dissimilar ions, might be responsible for this fluctuation in grain diameter [24]. Furthermore, using HR-TEM images obtained from one of numerous crystallites, it is possible to estimate the interplanar spacing of 0.25 nm between adjacent planes, which corresponds to the NiFe2O4 (311) lattice plane as illustrated in Figure 3b and Figure 4b. Moreover, Figure 3c and Figure 4c show the SAED pattern which consists entirely of concentric rings containing many differently oriented crystallites. The diffractogram depicts a regular pattern of bright spots indicating polycrystalline samples, powders or nanoparticles. Indeed, the S1 sample is characteristic, as demonstrated by the uniform intensity distribution along the ring circumference yielding smooth rings, as shown in Figure 3c. However, despite having sufficient crystallinity to produce smooth rings, as seen in Figure 4c, the S3 sample exhibits a non-uniform intensity distribution along the produced rings. Finally, Figure 3d and Figure 4d, respectively, show the histograms of the crystallite size statistical distribution (CSD) for the S1 and S3 samples. It can be seen from these figures that the grain diameter lies in the range of 3 nm–31 nm for the S1 sample and 35 nm–100 nm for the S3 sample, with average particles sizes located at 17 nm and 45 nm for these samples, respectively. In fact, there are differences in size between the XRD and TEM results, particularly in the case of the S1 sample, because the crystallite size is assumed to represent the size of a coherently diffracting domain and is not precisely the same as the particle size. However, due to several agglomerations present, the crystallite size of the S1 sample as determined by XRD patterns is significantly larger than the particle diameter detected by TEM. On the other hand, the S3 sample lithia doping caused the grain boundaries to be visible, indicating processes of redistribution and recrystallization. The high porosity and abundance of pores which formed on the S3 sample grain boundary may be responsible for this behavior.

3.4. Textural Characteristics of Virgin and Lithia-Doped Nickel Ferrites

We were able to distinguish between the various textural characteristics of samples S1 and S3 using N2-adsorption/desorption isotherms at 77 K. SBET, VP, Vm, and ȓ are shown in these properties for both the S1 and S3 samples. These type II isotherms with a type H3 hysteresis loop comprise the entirety of Figure 5. Table 4 contains values for SBET, VP, Vm, and ȓ; the table shows a modest rise in SBET, Vm, and the other values as a result of the 6 mol% Li2O doping, as seen by the S3 sample textural characteristics. The total pore volume may have somewhat increased, which would explain this rise. Additionally, Figure 6 displays the pore-size distribution of the S2 and S3 samples. According to this figure, the majority of the holes in the S1 sample are between 2 and 7 nm in size, with a 4.5 nm average. As can be seen, most of the pores in the S2 sample are between 0.5 and 3 nm, with an average pore size of 1 nm.

3.5. Magnetic Properties

The size and shape of the hysteresis loop are known to be significantly influenced by the grain size, porosity, and exchange interaction caused by cation dispersion. The magnetic domain is known to be reoriented at the start of an applied field, which causes a rapid increase in magnetization, then becomes sluggish and eventually saturated as a result of the spin revolution. The hysteresis loop used to compute magnetic properties, such as the coercive field (Hc), remanent magnetization (Mr), saturation magnetization (Ms), squareness (Mr/Ms), anisotropy constant (Ka), initial permeability (μi), and magnetic moment (μm) per unit formula in Bohr magnetrons for the tested materials (S1, S2, and S3). These variables are displayed in Table 5. These numbers, corresponding to the properties of magnetism, were extrapolated from the magnetic curves produced in Figure 7. At room temperature, a magnetic field ranging from -20 to +20 kG was applied while these graphs were produced using the VSM method. Based on these data, ferromagnetic property in lithium-doped and undoped Ni ferrite nanoparticles was established. This table demonstrates that as the lithia concentration increases, the values of Mr, Ms, μm, and μi decrease. On the other hand, the Mr/Ms value decreases in the presence of 3 mol% Li2O and then increases with 6 mol% Li2O. The opposite behavior was observed for the Ka value with the previous content of lithia.

4. Discussion

4.1. Formation of NiFe2O4-Based Mixed or Random Spinel Structures

In this study, fabrication of NiFe2O4 NPs using 6 mol egg-white assisted combustion resulted in the formation of random spinel structures. On the other hand, our previous work reported that using 3 mol or 10 mol egg white resulted in the production of inverse spinel nickel ferrite [23]. In other words, the preparation method, especially the egg-white content, affects the cation distribution in spinel nickel ferrite. Indeed, substitution of some Fe3+ ions by some Ni2+ ions at the A-site involved in NiFe2O4 crystallites resulted in the formation of random spinel structures. Because the atomic mass of iron is lower than that of nickel, this displacement led to an increase in the value of Dx from 5.4019 or 5.4769 to 5.4857 for NiFe2O4 crystallites prepared by 10, 3, and 6 mol egg white, respectively [23]. The peak height at the (4 4 0) reflection plane was equal to that at the (2 2 0) reflection plane indicating distribution of both Ni2+ and Fe3+ ions in the A- and B-sites, yielding partially inverse (random) spinel nickel ferrite. In our previous work, the peak heights at these planes were not equal due to the formation of inverse spinel nickel ferrite containing all Ni2+ ions at the A-site [24]. However, the FTIR spectra for the virgin sample shows two small bands (υ1* and υ2*) located at 867 cm−1 and 550 cm−1 that may be related to Ni–O vibrations
Nevertheless, we must note the following; firstly, the solid-state reaction between Ni and Fe oxides resulted in the formation of NiFe2O4 NPs at suitable preparation conditions. In other words, fabrication of NiFe2O4 NPs is due to the solid–solid interaction between Ni2+ and Fe3+ ions in the presence of oxygen. Secondly, one cannot ignore that both NiO and Fe2O3 include lattice defects, namely, Fe2+ and Ni3+ ions, respectively. Both Fe2+ and Ni3+ ions can be observed at the NiO and Fe2O3 interfaces, respectively. These cations cannot participate in the formation of NiFe2O4 NPs. The presence of both Fe2+ and Ni3+ ions must, therefore, be overcome to enhance nickel ferrite formation. The problem of the presence of the Fe2+ ions diminish and is almost negligible; this is because of the ease with which these ions pass via contact with, and diffusion through, the NiO surface, and their subsequent interaction with NiO in the presence of oxygen, to form nickel ferrites depending upon the preparation method [23]. The main problem, however, is the presence of nickel (Ni3+) ions that do not participate and even inhibit the formation of nickel ferrites. Hence, this problem being clear, the aim of the current study was to overcome the presence of Ni3+ ions which was achieved by Li2O doping. The treatment of the precursor containing nickel and ferric nitrates with small amounts of lithium nitrate in the presence of a certain amount of egg white resulted in the conversion of Ni3+ ions to Ni2+ ions which participated in the formation of NiFe2O4 NPs. This will be explained in detail based on our results.

4.2. Cation Distribution of Lithium-Doped NiFe2O4 NPs

Most scientists consider the fact that cation distribution affects the distinctive properties of each material. Because of this, they sought to control cation distribution using many different mechanisms, including preparation methods, doping, precursors, and the stoichiometries of constituents. To complement this goal, we studied the effect of Li2O doping on the type and formation of NiFe2O4 NP spinel structure. In fact, Li2O doping of virgin NiFe2O4 crystallites resulted in the appearance of the same diffractograms of NiFe2O4 NPs in XRD patterns, with a change in the intensity and position of the diffraction peaks. Furthermore, it was found that the XRD peak can be widened by defects or internal stress and strain.
The Li2O-based doping process was widely applied for a large number of solids, depending on the history of lithia and preparation conditions [25,26,27]. Lithium exhibits some affinity for dissolution in both Fe2O3 and NiO lattices due to the similarity of its ionic radius with the radii of both Fe3+ and Ni2+ ions [28,29,30]. This dissolution can proceed by incorporation or location of Li+ ions in an interstitial position and/or replacement of a host cation. In fact, the maximum quantity of lithia that can be dissolved in NiO can attain a value of 11 mol% before a lithium–nickel compound is formed [31,32]. It is well known that non-stoichiometric NiO contains Ni3+ ions in interstitial positions as lattice defects. In this study, Li2O doping of Ni and Fe oxides resulted in the conversion of some Ni3+ ions into Ni2+ ions which took part in the formation of nickel ferrite, with subsequent lattice expansion. This expansion could be attributed to the difference in the ionic radii of both Ni3+ (0.062 nm) and Ni2+ (0.078 nm) ions. Accordingly, lithia doping of the precursors containing a mixture of Ni and Fe nitrates with a certain amount of egg white enhances the formation of lithium-doped nickel ferrites.
Furthermore, Li2O doping increased the unit cell volume and crystallite size of NiFe2O4 NPs because it increased the interatomic space, which is dependent on an increase in the lattice according to Vegard’s law [33]. The X-ray density decreased when lithia was pre-sent and the unit cell capacity increased. The variation in X-ray density occurs according to the fluctuation in cation distribution within A- and B-sites [34]. The hopping length (LA and LB) between the magnetic ions within A- and B-sites increases in the presence of dopant due to substitution of some Fe3+ ions by some Ni2+ ions at the A-site, with subsequent migration of the substituted Fe3+ ions, that have an ionic radius lower than that of Ni2+ ions, to the B-site. Similar behavior was observed with the values of ionic radii A-O, B-O, rA, and rB. Moreover, the increase in the rA value is greater than that of the rB value due to lithia doping depending on the incorporation of Ni2+ ions within the A-site, besides its presence in the B-site [34]. Bearing in mind that the Li+ and Ni2+ ions have a high tendency for octahedral (B)-site occupancy, while Fe3+ ions are unevenly divided between tetrahedral (A) and octahedral (B) sites, several authors reported the octahedral-site preference of Ni2+, Fe3+, and Li+ ions in lithium-doped NiFe2O4 as Ni2+ > Li+ > Fe3+ [35]. However, the obtained results imply that Li ions play important roles at A- and B-sites, yielding redistribution of cations at these sites. The replacement of certain host Ni2+ and Fe3+ ions and their placement in interstitial sites producing solid solution can both contribute to the dissolution of Li+ ions in the NiO and Fe2O3 lattices participating in the production of nickel ferrite. Kroger’s notations [5] can be used in the following ways to streamline the dissolving procedure.
2Li+ + Ni3+ → 2Li + 2Ni2+ + 0.5O2
2Li+ + Ni2+ → 2Li(Ni2+) + C. V
Li+ + Fe3+ → Li(Fe3+) + C. V
2Li+ + 2Fe2+ + O2 → 2Li(Fe3+) + 2 Fe3+
Li(Ni2+) and Li(Fe3+) are the monovalent lithium ions located in the positions of host NiO and Fe2O3, respectively; Li represents lithium ions located in the interstitial positions of nickel and ferric oxide lattices; C.V. represents created cationic vacancies. The cationic vacancies that were created by Reactions (6) and (7) in the lattices of the reacting oxides may increase the mobility of the reacting oxide cations (Ni2+ and Fe3+), improving the cation distribution and leading to the production of nickel ferrite. On the other hand, dissolution of lithium ions in nickel and ferric oxide lattices according to Reactions (5) and (8) is expected to increase the number of Ni2+ and Fe3+ ions via converting Ni3+ and Fe2+ ions that may be present as lattice defects. Finally, the incorporation of Li+ ions at A- and B-sites augments the cation distribution based on the previous equations as follows. First, at the B-site, cationic vacancy stimulates the migration of the substituted Ni2+ ion by 2Li+ ions to the A-site, as shown in Equation (6). In addition, the transformation of the Fe2+ ion, which is expected to be present, to an Fe3+ ion enhances the ferrite formation. Second, at the A-site, cationic vacancy promotes the migration of the substituted Fe3+ ion by Li+ ion to the A-site, as shown in Equation (7).
In fact, most ferrite-based semiconductors contain antisite defects (ADs), which are greatly affected by the precursor history, doping, and preparation method. The change in ADs density resulted in significant modifications to the electronic, morphological, surface, and magnetic properties of ferrites. In NiFe2O4, although the total Fe and Ni atomic concentrations are well controlled, it is much more difficult to control the distribution of Fe and Ni atoms in the two interleaving lattices (A- and B-sites) during synthesis. In this study, all samples contained a certain number of ADs, where some of the Fe atoms exchanged positions with Ni atoms. Typically, these ADs are randomly distributed, and their number is characterized by the degree of Fe/Ni ordering in the sample, which depends on the content of Li dopant. Indeed, when NiFe2O4 contains ADs, some Fe occupy Ni sites (FeNi) and the same number of Ni atoms occupy Fe sites (NiFe), which produces new FeNi-O-Fe and NiFe-O-Ni bonds as a consequence of this disorder. We observed their effects on the distribution of iron and nickel ions, which was demonstrated through the displacement in the strength and location of the main bands, as shown in Figure 2. Additionally, the formation of oxygen vacancies (O*) in NiFe2O4, corresponding to the removal of neutral O atoms from the lattice, alters the number of delocalized electrons and also affects the magnitude of the magnetic moments of surrounding atoms.
The population of lithium ions at A- and B-sites involved in spinel nickel ferrite was confirmed by the appearance of two shoulder bands (υ1**) at 760 cm−1 and 767 cm−1 due to lithia doping with 3 and 6 mol%, respectively. Moreover, it was seen from the S1 XRD results and FTIR spectra that both Ni and Fe ions were unevenly divided between the A-and B-sites yielding random spinel structure. In addition, Li2O doping of virgin NiFe2O4 crystallites resulted in the appearance of the same functional groups, with a change in the intensity and frequency of the bands. On the other hand, the solubility of lithium ions in different cations (Ni2+ and Fe3+) at A- and B-sites resulted in redistribution of these cations, with a subsequent increase in the dislocation density, stress, and microstrain of the NiFe2O4 lattice, as observed in Table 2. These changes resulted in slight increases in the morphological and surface properties of NiFe2O4 NPs. Indeed, the change in the surface properties of nickel ferrite was monitored due to doping with lithium, depending on the change in total pore volume and the transformation of pore type from mesopores to micro/mesopores. In other words, the majority of pores in the S1 sample are meso-type pores, while the majority of pores in the S3 sample are micro-type pores. One cannot ignore the presence of some mesopores in the S3 sample. These modifications depend on the cation redistribution caused by lithia doping, which was confirmed by FTIR results, as shown in Figure 2.

4.3. Effect of Lithia Doping on Magnetization of NiFe2O4 NPs

Variation in the magnetic properties of nano particle ferrite materials depends entirely on the change in cation distribution which can be tuned by preparation method, doping, microstructures, and textural and lattice parameters. Moreover, the magnetization of different spinel ferrites results from variation in the magnetic moments of the ions between and at two sub lattices (A- and B-sites). In other words, the Neel model, which emphasizes A-B exchange interactions over A-A and B-B super-exchange ion interactions, can account for the fluctuation in the Ms value. Virgin and lithium-doped Ni ferrite nanoparticles exhibit different ferromagnetic behaviors, as shown by the S-shaped hysteresis loop [36]. Examination of the magnetic properties of the S1, S2, and S3 samples, tabulated in Table 4, revealed the following. Firstly, lithia doping of virgin nickel ferrite results in a slight decrease in both the value of Mr and Ms due to a decrease in particle size. However, the magnetic dilution of spinel nickel ferrite by lithia doping could be attributed to dissolution of nonmagnetic lithium ions in the crystal lattice. Additionally, because the net magnetic moment of the spinel lattice is equal to the difference between the magnetic moments of the A- and B-sites, the net magnetic moment of virgin nickel ferrite decreases as the amount of lithia increases [37]. Consequently, this decrease could be attributed to the replacement of Fe3+ ions having magnetic moments of 5 μB with the Ni2+ ions having lower magnetic moments of 2 μB at the tetrahedral site, with the subsequent migration of Fe3+ ions from the tetrahedral site to the octahedral site depending on the dissolution of lithium ions in the crystal lattice of nickel ferrite. Secondly, the treatment with Li2O created an increase in the coercivity of NiFe2O4 NPs. The range of Hc values depends on numerous factors, such as morphology, magneto crystalline anisotropy, size distributions, and exchange coupling. In this study, an increase in coercivity is the result of the change in the magnetic anisotropic constant, which may be due to the higher magneto anisotropy of Ni2+ ions [38]. Due to Jahn–Teller distortions brought on by Ni2+ ions at the A-site of the spinel ferrite system under the influence of 3 mol% Li2O doping, this behavior was more pronounced in the case of the S2 sample. Thirdly, the investigated ferrites display uniaxial anisotropy rather than cubic anisotropy because the squareness ratio (Mr/Ms) is less than 0.5. In addition, the small squareness values refer to the surface spin-disorder effects resulting from the canted spin on the surface of nanoparticles. Hence, the lower value of this ratio strongly promotes the studied ferrites for their use in high-frequency applications [38]. Finally, the measured total saturation magnetization of NiFe2O4 is interpreted as being due to the presence of antisite and cation-excess defects.

5. Conclusions

The authors concentrated their efforts on preparing pure and similar nickel ferrites using an easy, simple, and inexpensive method. In their quest to prepare these ferrites, their aim was to study the effect of doping with lithium oxide on the structural, surface, morphological, and finally, magnetic properties. Accordingly, and from the results obtained, the following conclusions were made:
  • There is a slight decrease in the crystallite size of nickel ferrite, although there is an increase in the lattice constant. This increase could be attributed to the replacement of some ferric cation at the spinel structure-based tetrahedral site with some nickel cations, with the subsequent migration of ferric cations to the octahedral site. In addition, inconsistencies and differences may exist between the size of both the crystallite and the particles because size was determined by Scherrer’s formula, the typical or “apparent” crystallite size, which is not always the same as the particle size.
  • XRD results confirm that the as-prepared ferrites belong to the cubic spinel structure NiFe2O4 of random type. However, lithia doping of this ferrite led to the shape transformation of particles, becoming cubic shape. Selected area electron diffraction (SAED) images depict a regular pattern of bright spots, indicating polycrystalline samples, powders, or nanoparticles.
  • Lithia doping of virgin nickel ferrite resulted in change in microstructure, with subsequent changes in stress, strain, and dislocation. Slight increases in the surface area and total pore volume were observed with increasing lithia content.
  • Magnetization and coercivity of NiFe2O4 increase by increasing the content of lithia-based dopant. The magnetic parameters were strongly affected by particle size and shape, as well as by their distribution, crystallinity and magnetic domain sizes, and micro-strains induced by Li2O doping. In addition, this doping led to changes in the antisite and cation excess defects, which enhanced the magnetization nickel ferrite.

Author Contributions

Validation was carried out by N.M.D., G.M.A.-S., F.F.A.-F., R.S.A. and O.H.A.-E.; formal analysis, N.M.D., G.M.A.-S. and O.H.A.-E.; systematic evaluation, N.M.D., G.M.A.-S. and O.H.A.-E.; writing—original preparation of the draft, N.M.D. and O.H.A.-E.; resources, N.M.D., G.M.A.-S. and O.H.A.-E.; curation of results, N.M.D., G.M.A.-S. and O.H.A.-E.; visualization; N.M.D. and O.H.A.-E.; supervision; G.M.A.-S., F.F.A.-F. and R.S.A.; project management; and the acquisition of funding. The final, published version of the paper has been read and approved by all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2023R67), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors would like to extend their sincere appreciation to Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2023R67), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abushad, A.; Arshad, M.; Naseem, S.; Ahmed, H.; Ansari, A.; Chakradhary, V.K.; Husain, S.; Khan, W. Synthesis and role of structural disorder on the optical, magnetic and dielectric properties of Zn doped NiFe2O4 nanoferrites. J. Mol. Struct. 2022, 1253, 132205. [Google Scholar]
  2. Dippong, T.; Levei, E.A.; Goga, F.; Petean, I.; Avram, A.; Cadar, O. The impact of polyol structure on the formation of Zn0.6Co0.4Fe2O4 spinel-based pigments. J. Sol.-Gel. Sci. Technol. 2019, 92, 736–744. [Google Scholar] [CrossRef]
  3. Bhame, S.D.; Joy, P.A. Enhanced strain sensitivity in magnetostrictive spinel ferrite Co1-xZnxFe2O4. J. Magn. Magn. Mater. 2018, 447, 150–154. [Google Scholar] [CrossRef]
  4. Deraz, N.M. Production, physicochemical characterization and magnetic behavior of nano-crystalline alumina doped cobalt ferrite system. Int. J. Electrochem. 2012, 7, 4596–4607. [Google Scholar]
  5. Deraz, N.M.; Alarifi, A. Synthesis and characterization of pure and Li2O doped ZnFe2O4 nanoparticles via glycine assisted route. Polyhedron 2009, 28, 4122–4130. [Google Scholar] [CrossRef]
  6. Deraz, N.M. Fabrication and characterization of magnetic alumina-doped zinc ferrite nanoparticles. J. Anal. Appl. Pyrolysis. 2011, 91, 48–54. [Google Scholar] [CrossRef]
  7. Deraz, N.M. Effects of magnesia addition on structural, morphological and magnetic properties of nano-crystalline nickel ferrite system. Ceram. Int. 2012, 38, 511–516. [Google Scholar] [CrossRef]
  8. Deraz, N.M.; Alarifi, A. Processing and Evaluation of Alumina Doped Nickel Ferrite Nano-Particles. Int. J. Electrochem. 2012, 7, 4585–4595. [Google Scholar]
  9. Kaur, A.; Bhargava, G.K. Review paper on nickel-zinc nano ferrite. Mater. Today Proc. 2021, 37, 3082–3086. [Google Scholar] [CrossRef]
  10. Pottker, W.E.; Ono, R.; Cobos, M.A.; Hernando, A.; Araujo, J.F.D.F.; Bruno, A.C.O.; Lourenço, S.A.; Longo, E.; La Porta, F.A. Influence of order-disorder effects on the magnetic and optical properties of NiFe2O4 nanoparticles. Ceram. Int. 2018, 44, 17290–17297. [Google Scholar] [CrossRef]
  11. Rama, G.; Dhiman, P.; Kumar, A.; Vo, D.V.N.; Sharma, G.; Sharma, S.; Naushad, M. Recent advances on nickel nano-ferrite: A review on processing techniques, properties and diverse applications. Chem. Eng. Res. Des. 2021, 175, 182–208. [Google Scholar]
  12. Samieemehr, M.; Arab, A.; Kiani, E. Influence of two-step sintering on power loss and permeability dispersion of MnZnNi ferrite. J. Magn. Magn. Mater. 2022, 553, 169269. [Google Scholar] [CrossRef]
  13. Salunkhe, A.B.; Khot, V.M.; Phadatare, M.R.; Thorat, N.D.; Joshi, R.S.; Yadav, H.M.; Pawar, S.H. Low temperature combustion synthesis and magnetostructural properties of Co-Mn nanoferrites. J. Magn. Magn. Mater. 2014, 352, 91–98. [Google Scholar] [CrossRef]
  14. Reddy, M.P.; Zhou, X.; Yann, A.; Du, S.; Huang, Q.; Mohamed, A. Low temperature hydrothermal synthesis, structural investigation and functional properties of CoxMn1-xFe2O4 (0 ≤ x ≤ 1) nanoferrites. Superlattices Microst. 2015, 81, 233–242. [Google Scholar] [CrossRef]
  15. Samson, V.A.; Bernadsha, S.B.; Xavier, R.; Rueshwin, C.S.T.; Prathap, S.; Madhavan, J.; Raj, M.V.A. One pot hydrothermal synthesis and characterization of NiFe2O4 nanoparticles. Mater. Today Proc. 2022, 50, 2665–2667. [Google Scholar] [CrossRef]
  16. Vinosha, P.A.; Xavier, B.; Krishnan, S.; Das, S.J. Investigation on zinc substituted highly porous improved catalytic activity of NiFe2O4 nanocrystal by co-precipitation method. Mater. Res. Bull. 2018, 101, 190–198. [Google Scholar] [CrossRef]
  17. Patil, K.; Phadke, S.; Mishra, A. A study of structural and dielectric properties of Zn2+ doped MnFe2O4 and NiFe2O4 spinel ferrites. Mater. Today Proc. 2021, 46, 2226–2228. [Google Scholar] [CrossRef]
  18. Sharifianjazi, F.; Moradi, M.; Parvin, N.; Nemati, A.; Rad, A.J.; Sheysi, N.; Abouchenari, A.; Mohammadi, A.; Karbasi, S.; Ahmadi, Z.; et al. Magnetic CoFe2O4 nanoparticles doped with metal ions: A review. Ceram. Int. 2020, 46, 18391–18412. [Google Scholar] [CrossRef]
  19. Lushchik, A.; Feldbach, E.; Kotomin, E.A.; Kudryavtseva, I.; Kuzovkov, V.N.; Popov, A.I.; Seeman, V.; Shablonin, E. Distinctive features of diffusion-controlled radiation defect recombination in stoichiometric magnesium aluminate spinel single crystals and transparent polycrystalline ceramics. Sci. Rep. 2020, 10, 7810. [Google Scholar] [CrossRef]
  20. Klym, H.; Hadzaman, I.; Ingram, A.; Shpotyuk, O.; Karbovnyk, I.; Kostiv, Y.; Vasylchyshyn, I.; Chalyy, D. Positron annihilation characterization of free volume in micro-and macro-modified Cu0.4Co0.4Ni0.4Mn1.8O4 ceramics. Low Temp. Phys 2016, 42, 601–605. [Google Scholar] [CrossRef] [Green Version]
  21. Cullity, B.D. Elements of X-Ray Diffraction; Chapter 14; Addison-Wesly Publishing Co., Inc.: Singapore, 1976. [Google Scholar]
  22. Al-Senani, G.M.; Al-Saeedi, S.I.; Al-Kadhi, N.S.; Abd-Elkader, O.H.; Deraz, N.M. Green Synthesis and Pinning Behavior of Fe-Doped CuO/Cu2O/Cu4O3 Nanocomposites. Processes 2022, 10, 729. [Google Scholar] [CrossRef]
  23. Al-Senani, G.M.; Foziah, F.; Al-Fawzan, F.F.; Rasmiah, S.; Almufarij, R.S.; Abd-Elkader, O.H.; Deraz, N.M. Biosynthesis, Physicochemical and Magnetic Properties of Inverse Spinel Nickel Ferrite System. Crystals 2022, 12, 1542. [Google Scholar] [CrossRef]
  24. Watawe, S.C.; Sarwade, B.D.; Bellad, S.S.; Sutar, B.D.; Chaugule, B.K. Microstructure and magnetic properties of Li-Co ferrites. Mater. Chem. Phys. 2000, 65, 173–177. [Google Scholar] [CrossRef]
  25. Deraz, N.M. Solid Solution and Crystal Engineering. Int. J. Nanomater. Mol. Nanotechnol. 2019, 1, 102. [Google Scholar]
  26. Deraz, N.M.; Abd-Elkader, O.H. Synthesis and Characterization of Nano- crystalline BixbyiteHopcalite Solids. Int. J. Electrochem. Sci. 2013, 8, 10112–10120. [Google Scholar]
  27. Begum, R.S.; Xavier, R.J. Effect of Lithium doping on Structural, Optical, Eletrochemical and Magnetic Properties of NiO Nanoparticles. Elixir Nanotechnol. 2016, 96, 41689–41695. [Google Scholar]
  28. Kanno, R.; Shirane, T.; Inaba, Y.; Kawamotoe, Y. Synthesis and electrochemical properties of lithium iron oxides with layer-related structures. J. Power Sources 1997, 68, 145–152. [Google Scholar] [CrossRef]
  29. Trivedi, U.N.; Jani, K.H.; Modi, K.B. Study of cation distribution in lithium doped nickel ferrite. J. Mater. Sci. Lett. 2000, 19, 1271–1273. [Google Scholar] [CrossRef]
  30. Cochran, S.J.; Larkins, F.P. An X-ray photoelectron study of doped and supported nickel oxide. Aust. J. Chem. 1985, 38, 1293–1300. [Google Scholar] [CrossRef]
  31. Zhu, W.; Wang, J.; Wu, D.; Li, X.; Luo, Y.; Han, C.; Ma, W.; He, S. Investigating the Heavy Metal Adsorption of Mesoporous Silica Materials Prepared by Microwave Synthesis. Nanoscale Res. Lett. 2017, 12, 323. [Google Scholar] [CrossRef] [Green Version]
  32. Yoshio, I.; Naoyoshi, H. A Note on the Solid Chemisty of the NiO-Li2O Solid Solution. Bull. Chem. Soc. Jpn. 1964, 37, 659–662. [Google Scholar]
  33. Denton, A.R.; Ashcroft, N.W. Vegard’s Law. Phys. Rev. A 1991, 43, 3161–3164. [Google Scholar] [CrossRef] [PubMed]
  34. Sontu, U.B.; Rao, N.; Reddy, V.R.M. Temperature dependent and applied field strength dependent magnetic study of cobalt nickel ferrite nano particles: Synthesized by an environmentally benign method. J. Magn. Magn. Mater. 2018, 352, 398–406. [Google Scholar] [CrossRef]
  35. Balamurugan, S.; Manimekalai, S.; Prakash, I. Synthesis, characterization and electrical properties of Li2NiFe2O4/NiFe2O4 nanocomposites. J. Mater. Sci. Mater. Electron. 2017, 28, 18610–18619. [Google Scholar] [CrossRef]
  36. Yu, R.; Zhao, J.; Zhao, Z.; Cui, F. Copper substituted zinc ferrite with abundant oxygen vacancies for enhanced ciprofloxacin degradation via peroxymonosulfate activation. J. Hazard. Mater. 2020, 390, 121998. [Google Scholar] [CrossRef] [PubMed]
  37. Batoo, K.M.; Kumar, S.; Lee, C.G. Influence of Al doping on electrical properties of Ni–Cd nano ferrites. Curr. Appl. Phys. 2009, 9, 826–832. [Google Scholar] [CrossRef]
  38. Rao, P.A.; Rao, K.S.; Raju, T.P.; Kapusetti, G.; Choppadandi, M.; Varma, M.C.; Rao, K.H. A systematic study of cobalt-zinc ferrite nanoparticles for self-regulated magnetic hyperthermia. J. Alloy. Compd. 2019, 794, 60–67. [Google Scholar]
Figure 1. XRD patterns of S1, S2 and S3 samples.
Figure 1. XRD patterns of S1, S2 and S3 samples.
Crystals 13 00069 g001
Figure 2. FTIR of the S1and S2 and S3 samples.
Figure 2. FTIR of the S1and S2 and S3 samples.
Crystals 13 00069 g002
Figure 3. (a) TEM, (b) HRTEM, (c) SAED, and (d) particle size distribution of the S1 sample.
Figure 3. (a) TEM, (b) HRTEM, (c) SAED, and (d) particle size distribution of the S1 sample.
Crystals 13 00069 g003
Figure 4. (a) TEM, (b) HRTEM, (c) SAED, and (d) particle size distribution of the S2 sample.
Figure 4. (a) TEM, (b) HRTEM, (c) SAED, and (d) particle size distribution of the S2 sample.
Crystals 13 00069 g004
Figure 5. N2- adsorption/desorption isotherm and pore size distribution of the S1 sample.
Figure 5. N2- adsorption/desorption isotherm and pore size distribution of the S1 sample.
Crystals 13 00069 g005
Figure 6. N2- adsorption/desorption isotherm and pore size distribution of the S3 sample.
Figure 6. N2- adsorption/desorption isotherm and pore size distribution of the S3 sample.
Crystals 13 00069 g006
Figure 7. The magnetic hysteresis loops of S1, S2 and S3 at room temperature.
Figure 7. The magnetic hysteresis loops of S1, S2 and S3 at room temperature.
Crystals 13 00069 g007
Table 1. Miller indices, and the values of 2θ for the crystalline phases (NiFe2O4) in the as-synthesized solids.
Table 1. Miller indices, and the values of 2θ for the crystalline phases (NiFe2O4) in the as-synthesized solids.
Samplehkl2θ Obs.2θ Calc.DifferenceError %PhaseSpace Group
+0%Li2O11118.470018.5433−0.0733−0.39686NiFe2O4FD3M
02230.430030.5082−0.0782−0.25698NiFe2O4FD3M
11335.810035.9392−0.1292−0.36079NiFe2O4FD3M
22237.350037.5958−0.2458−0.6581NiFe2O4FD3M
00443.530043.6879−0.1579−0.36274NiFe2O4FD3M
31348.100047.84050.25950.5395NiFe2O4FD3M
22454.080054.2204−0.1404−0.25962NiFe2O4FD3M
33357.580057.8084−0.2284−0.39667NiFe2O4FD3M
04463.020063.4982−0.4782−0.75881NiFe2O4FD3M
31567.049066.77710.27190.40552NiFe2O4FD3M
24468.129067.85180.27720.40688NiFe2O4FD3M
20672.375072.07440.30060.41534NiFe2O4FD3M
53375.493075.17510.31790.4211NiFe2O4FD3M
22676.522076.19840.32360.42288NiFe2O4FD3M
44479.620080.2507−0.6307−0.79214NiFe2O4FD3M
+3%Li2O02230.430030.4727−0.0427−0.14032NiFe2O4FD3M
11335.810035.8970−0.0870−0.24295NiFe2O4FD3M
22237.650037.55140.09860.26189NiFe2O4FD3M
00443.580043.6357−0.0557−0.12781NiFe2O4FD3M
31347.870047.78280.08720.18216NiFe2O4FD3M
22454.190054.15370.03630.06699NiFe2O4FD3M
11557.520057.7364−0.2164−0.37622NiFe2O4FD3M
04463.370063.4176−0.0476−0.07511NiFe2O4FD3M
31566.780066.69130.08870.13282NiFe2O4FD3M
24467.980067.76420.21580.31745NiFe2O4FD3M
20672.160071.97960.18040.25NiFe2O4FD3M
33575.240075.07490.16510.21943NiFe2O4FD3M
22676.340076.09630.24370.31923NiFe2O4FD3M
44479.710080.1409−0.4309−0.54058NiFe2O4FD3M
+6%Li2O02230.430030.4999−0.0699−0.22971NiFe2O4FD3M
11335.730035.9294−0.1994−0.55807NiFe2O4FD3M
22237.260037.5854−0.3254−0.87332NiFe2O4FD3M
40043.310043.6757−0.3657−0.84438NiFe2O4FD3M
31347.870047.82700.04300.08983NiFe2O4FD3M
42254.190054.2048−0.0148−0.02731NiFe2O4FD3M
11557.520057.7916−0.2716−0.47218NiFe2O4FD3M
04463.090063.4794−0.3894−0.61721NiFe2O4FD3M
31566.780066.75710.02290.03429NiFe2O4FD3M
24467.980067.83130.14870.21874NiFe2O4FD3M
60272.400072.05230.34770.48025NiFe2O4FD3M
53375.220075.15170.06830.0908NiFe2O4FD3M
22676.660076.17460.48540.63319NiFe2O4FD3M
44480.070080.2250−0.1550−0.19358NiFe2O4FD3M
Table 2. Lattice parameters of NiFe2O4 nanoparticles included in the S1, S2 and S3 samples.
Table 2. Lattice parameters of NiFe2O4 nanoparticles included in the S1, S2 and S3 samples.
ParametersS1Error %S2Error %S3Error %
d, nm49.090.20347.960.08346.920.170
a, nm0.827970.1280.829480.1830.83100.192
V, nm30.5676−0.5460.5707−0.5570.5739−0.565
Dx, g/cm35.48530.5415.45560.5475.42560.549
LA, nm0.3585−0.1950.3592−0.0550.35900.139
LB, nm0.2926−0.1700.29310.0340.29300.136
A-O, nm0.1893−0.1580.18960.0520.18950.101
B-O, nm0.2136−0.1870.21400.0460.21390.140
rA, nm0.0573−0.5230.05760.1730.05750.347
rB, nm0.0816−0.4900.08200.1210.08190.366
σ (Ν/m2)4.79 × 1070.5425.05 × 1070.4155.26 × 1070.409
ε7.06 × 10−40.4817.26 × 10−40.5477.3 × 10−40.507
δ, Lines/nm4.1 × 10−40.5364.32 × 10−40.4164.5 × 10−40.317
Table 3. FTIR bands of NiFe2O4 nanoparticles included in the S1, S2 and S3 samples.
Table 3. FTIR bands of NiFe2O4 nanoparticles included in the S1, S2 and S3 samples.
S1S2S3
ν cm−1T %ν cm−1T %ν cm−1T %
Peak No 13433.6484.973423.9960.643428.8168.37
Peak No 22923.5696.172925.4888.232924.5290.21
Peak No 31633.4193.181636.3087.241636.3187.81
Peak No 41385.6095.451385.6090.431384.6488.32
Peak No 51044.2697.581049.1083.131046.1991.93
Peak No 6890.2096.20880.3494.53875.2197.20
Peak No 7640.1392.01620.0990.13582.3982.20
Peak No 8400.2186.11450.3485.53413.6583.99
Table 4. Surface properties of the S1 and S3 samples.
Table 4. Surface properties of the S1 and S3 samples.
SamplesSBET (m2/g)Error %Vp (cm3/g)Error %ȓ(nm)Error %Vm (nm)Error %
S119.900.1010.03220.6216.4650.0774.5720.043
S320.830.1440.04050.4937.7690.0254.7860.062
Table 5. The magnetic properties of the S1, S2 and S3 samples.
Table 5. The magnetic properties of the S1, S2 and S3 samples.
S1Error %S2Error %S3Error %
Ms (emu/g)31.5440.01227.4470.01425.5100.019
Mr (emu/g)6.6130.0605.4060.0735.6180.053
Mr/Ms (emu/g)0.20960.2380.19690.4060.22020.090
Hc (G)133.990.074175.270.039163.780.036
μm1.32490.0111.15270.0521.07140.046
μi59.5250.0057.0540.0287.3400.027
Ka (erg/cm3)4402.690.0135011.080.0014352.110.014
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

Al-Senani, G.M.; Al-Fawzan, F.F.; Almufarij, R.S.; Abd-Elkader, O.H.; Deraz, N.M. Magnetic Behavior of Virgin and Lithiated NiFe2O4 Nanoparticles. Crystals 2023, 13, 69. https://doi.org/10.3390/cryst13010069

AMA Style

Al-Senani GM, Al-Fawzan FF, Almufarij RS, Abd-Elkader OH, Deraz NM. Magnetic Behavior of Virgin and Lithiated NiFe2O4 Nanoparticles. Crystals. 2023; 13(1):69. https://doi.org/10.3390/cryst13010069

Chicago/Turabian Style

Al-Senani, Ghadah M., Foziah F. Al-Fawzan, Rasmiah S. Almufarij, Omar H. Abd-Elkader, and Nasrallah M. Deraz. 2023. "Magnetic Behavior of Virgin and Lithiated NiFe2O4 Nanoparticles" Crystals 13, no. 1: 69. https://doi.org/10.3390/cryst13010069

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