Next Article in Journal
Definition, Fabrication, and Compression Testing of Sandwich Structures with Novel TPMS-Based Cores
Previous Article in Journal
A Study on Pigment Composition of Buddhist Cave Paintings Based on Hyperspectral Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Structural, Elastic and Magnetic Properties of CoCr2−xZrxO4 Nanoparticles

by
Mai M. E. Barakat
1,2,* and
Doaa El-Said Bakeer
3
1
Department of Physics, Faculty of Science, Taibah University, Yanbu 46423, Saudi Arabia
2
Department of Physics, Faculty of Science, Alexandria University, Alexandria 21568, Egypt
3
Department of Physics, Faculty of Science, Damanhour University, Damanhour 22511, Egypt
*
Author to whom correspondence should be addressed.
Materials 2024, 17(21), 5149; https://doi.org/10.3390/ma17215149
Submission received: 27 September 2024 / Revised: 17 October 2024 / Accepted: 18 October 2024 / Published: 22 October 2024

Abstract

:
This study investigates the impact of zirconium substitution on the structural, elastic and magnetic properties of CoCr2O4 nanoparticles. A series of CoCr2−xZrxO4 nanoparticles, x = 0.00, 0.05, 0.10, 0.15 and 0.20, are synthesized via the co-precipitation method. X-ray diffraction (XRD) patterns affirm the formation of single-phase cubic structure with the space group Fd3m. Special attention is given to accurately calculating the average crystallite size (D) and lattice parameter (a) using Williamson–Hall (W–H) analysis and the Nelson–Riley (N–R) extrapolation function, respectively. The increase in Zr4+ content leads to a reduction in crystallite size and an increase in the lattice parameter. Elastic properties are estimated from force constants and the lattice constant, determined from FTIR and XRD, respectively. The observed changes in the elastic constants are attributed to the strength of interatomic bonding. The stiffness constants decrease, while Poisson’s ratio increases with increasing Zr4+ content, reflecting the increase in the ductility of the prepared samples. As the Zr4+ content increases, the stiffness constants decrease, and Poisson’s ratio increases, reflecting enhanced ductility of the samples. Furthermore, as Zr4+ content rises, Young’s modulus, the rigidity modulus and Debye temperature decrease. The magnetic hysteresis loop measurements are carried out at room temperature using a vibrating sample magnetometer (VSM) over a field range of 25 kg. Unsubstituted CoCr2O4 exhibits ferrimagnetic behavior. As Zr4+ content increases, saturation magnetization (Ms) and magnetic moment decrease, while remanent magnetization (Mr) and coercivity (Hc) initially decrease up to x = 0.10, then increase with further increases in x. The novel key of this study is how Zr4+ substitution in CoCr2O4 nanoparticles can effectively modify their elastic moduli and magnetic properties, making them suitable for various applications such as flexible electronics, protective coatings, energy storage components and biomedical implants.

1. Introduction

Chromates (MCr2X4; M = Co, Fe, etc., X = O, S ions) are members of the multiferroic materials class, exhibiting uniform magnetization as huge ferrimagnetism, spatially modulated magnetism and magneto resistance [1,2]. The normal cubic spinel structure for cobalt chromites, CoCr2O4, arises from the occupation of tetrahedral sites by Co2+ divalent cations and octahedral sites by Cr3+ + trivalent cations [3,4,5,6]. Three magnetic transitions are sustained by CoCr2O4 as a result of the exchange interaction competing for Cr–Cr, Co-Cr and Co–Co. (1) The existence of long-range non-collinear spiral and ferroelectric magnetic transition occurs at TS ≈ 26 K. (2) Above TS, the material undergoes a transition from paramagnetic to collinear short-range ferrimagnetism, reaching Curie temperature TC ≈ 95 K. CoCr2O4 has a multiferroic property, resulting from the onset of non-collinear spiral magnetic transition to the onset of shortrange ferrimagnetic transition [7,8]. (3) Below TS, a lock-in transition appears at TL ≈ 14 K. Both ferromagnetic and ferroelectric orders give cobalt chromates great attention in practical applications such as high temperature ceramics, catalysis, semiconductors, biomedical materials, electrochemical sensors, telecommunication systems, data storage media and various nanotechnology technological applications [9,10,11,12,13,14,15].
The properties of magnetic materials, such as permanent magnetization and magnetic saturation, are strongly influenced by particle size. Therefore, the preparation methods of magnetic materials play an important role in controlling the particle size and shape. Several methods have been used to prepare CoCr2O4 nanoparticles, such as the hydrothermal method [16], the sonochemical technique [17], thermolysis of the polymer metal complex [18,19], co-precipitation [20], the low-temperature combustion method by citric acid [21,22] and the sole gel route with propylene oxide as a gelation agent [23]. Among all these techniques, the co-precipitation method stands out for nanoparticle synthesis due to its simplicity, cost-effectiveness and scalability. It offers precise control over particle size, shape and composition by adjusting reaction conditions, producing uniform and high-purity nanoparticles. Its versatility allows for the synthesis of a wide range of materials, with the potential for doping and functionalization. Additionally, co-precipitation delivers reproducible results, making it suitable for both industrial and research applications in catalysis, biomedical fields and environmental remediation. The average particle size of CoCr2O4 ranges from 8 to 40 nm, depending on the preparation method and combustion or annealing temperatures [17,18,19,20,21,22,23]. The hydrothermal preparation for CoCr2O4 tends to produce particle sizes greater than 90 nm [16].
The effect of the partial ionic substitution for CoCr2O4 nanoparticles brings great attention to different electrical and magnetic applications [24]. Modifying the structural and magnetic properties of CoCr2O4 nanoparticles can be achieved through changing the dopant nature, dopant concentration, sintering temperature and dopant particle size. The increase in Mg concentration into Co1−xMgxCr2O4 nanoparticles suppresses both TC and TS transition states [25]. The decrease in TC value was reported with increasing Zn concentration into Co1−xZnxCr2O4 nanoparticles [26]. The effects of various dopants on Co1−xMxCr2O4 (M = Zn, Cu, Mg) were studied by Pankaj and Dinesh [27], who found a change in the structure of CoCr2O4 nanoparticles with different doping. Moreover, the doping with Zn and Mg made CoCr2O4 nanoparticles more suitable for high frequency microelectronic applications in comparison with those doped by Cu. The increase in Mn doping into Co1−xMnxCr2O4 nanoparticles resulted in lower TC, TS and TL values but enhanced the saturation magnetization [28]. The temperature-dependent magnetic response of CoCr2O4 nanoparticles warrants emphasis [29,30,31]. Researchers have demonstrated that the ZFC/FC curves indicate a transition from paramagnetic to ferromagnetic behavior at Tc = 100 K. As the temperature decreased, the magnetization ( M s ) reached a maximum at 75 K before gradually declining to 5 K [29]. Additionally, a sharp increase in coercivity (HC) at low temperatures was observed, consistent with the modified Kneller’s law [29].
Numerous studies have examined the impact of Zr4+ ion substitution on the structural and physical properties of oxide nanomaterials [32,33,34]. Kavitha et al. [32] reported that in Co1−xZrxFe2O4, well-defined reflections of spinel lattice planes were observed at lower zirconium concentrations. However, at high Zr content for x ≥ 0.75, an impurity peak corresponding to Fe2O3 appears. Additionally, the (311) peaks shifted slightly to higher angles, indicating strain induced by the larger Zr ions due to ionic size mismatch. Naik et al. [33] showed that Zn1−xZrxO (0 ≤ x ≤ 0.05) nanoparticles showed no impurity phase in both undoped and Zr-doped ZnO. Monaji et al. [34] confirmed that for Co1+xZrxCo2−2xO4 (0 ≤ x ≤ 0.4), X-ray spectral analysis revealed a single-phase cubic structure up to x = 0.2 with no impurity phases. In addition, coercivity ( H c ) was found to increase significantly with decreasing temperature for all Zr-substituted samples. Based on the earlier findings, it is expected that substituting with a high concentration of Zr4+ ions could destabilize the structure, potentially leading to the formation of secondary phases and introducing strain due to the larger size of the Zr ions. The substitution of Zr in BaZxTi1−xO3 enhanced the chemical stability and reduced the dielectric loss [35]. The magnetic properties of Mg0.5Cu0.5Fe2−2xNixZrxO4 were significantly affected by the Zr4+ dopant, as it weakened the material’s magnetic properties [36]. Reda et al. [37] reported that doping BaTi1xZrxO3 with Zr greatly improved its ferroelectric properties. Furthermore, a decrease in the energy band gap was observed in LaFe1−xZrxO3 as the Zr content increased [38].
FTIR and the elastic properties of solids are linked through molecular vibrations and structural characteristics. FTIR measures infrared radiation absorption, revealing a material’s molecular composition and bonding. These molecular features significantly influence elastic properties, such as Young’s modulus and shear modulus. By analyzing FTIR spectra, researchers can deduce how molecular interactions affect physical properties, including force constants, elastic moduli, Debye temperature and molar heat capacity [39,40,41,42].
This study mainly focuses on synthesizing CoCr2−xZrxO4 nanoparticles through the best conditions of the co-precipitation technique and studying the effects of Zr4+ substitution on the structural, elastic and magnetic properties of cobalt chromates nanoparticles. It is important to note that a comprehensive study on the structural, elastic and magnetic properties of Zr4+ substituted CoCr2O4 nanoparticles has not yet been reported. The elastic properties of cobalt chromates are a powerful tool for determining its suitability for industrial applications. Furthermore, understanding the elastic behavior is crucial for gaining insights into the thermal properties of chromates and revealing the nature of their interionic bonding. FTIR spectroscopy is often used to estimate the elastic properties of such materials.

2. Experimental Techniques

A series of CoCr2−xZrxO4 nanoparticles with varying zirconium content (x = 0.00, 0.05, 0.10, 0.15 and 0.20) were synthesized using a chemical co-precipitation method. Stoichiometric amounts of analytical-grade chemical reagents, cobalt chloride hexahydrate (CoCl2·6H2O, Sigma-Aldrich, St. Louis, MO, USA, ≥98%), chromium chloride hexahydrate (CrCl3·6H2O, Sigma-Aldrich, ≥98%) and zirconium chloride octahydrate (ZrOCl2·8H2O, Sigma-Aldrich, 98%) were dissolved in distilled water to form homogeneous aqueous solutions. The solutions were thoroughly mixed using a magnetic stirrer and heated to 40 °C (Ceramic Plate LCD Digital Magnetic Hotplate Stirrer, Klang, Selangor, Malaysia). Sodium hydroxide (NaOH) was then added to the salt solution to adjust the pH to 12.0, resulting in the formation of a precipitate. The mixture was continuously stirred and heated to 60 °C, where it was maintained for 2 h. The precipitate was washed with deionized water to remove any residual salts and then dried at 100 °C for 24 h to eliminate any remaining water. Finally, the dried sample was calcined at 600 °C for 2 h to enhance its crystalline properties.
The structural and phase purity of CoCr2−xZrxO4 nanoparticles was analyzed using X-ray powder diffraction (XRD) with a Bruker D8 Advance powder diffractometer (Billerica, MA, USA), utilizing Cu-Kα radiation (λ = 1.54056 Å) over 2θ ranges from 10° to 80°. The morphology and size of the synthesized nanoparticles were examined using a Jeol JEM-2100 transmission electron microscope (TEM) operating at 200 kV (Tokyo, Japan). Elemental composition of cobalt chromate was determined via scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM/EDS; JSM-7200F). Fourier transform infrared (FTIR) spectroscopy was conducted on a Shimadzu FTIR 8400S spectrophotometer with KBr pellets (Kyoto, Japan). The magnetic properties of the CoCr2−xZrxO4 nanoparticles (0.00 ≤ x ≤ 0.20) were evaluated at room temperature using a Lakeshore 7410 vibrating sample magnetometer (VSM) (Westerville, OH, USA).

3. Results and Discussion

To study the effect of Zirconium substitution on the phase formation and the crystallite size of CoCr2O4, XRD analysis is performed on the prepared samples. Figure 1 illustrates the XRD patterns for CoCr2−xZrxO4, where x varies from 0.0 to 0.20. The prepared samples display well-defined Bragg reflection planes, indicating robust crystallization. Structural analysis affirms a single-phase cubic spinel structure (JCPDS card 001-1122) [43] with the space group Fd3m, devoid of additional diffraction lines associated with other crystallographic phases or unreacted components. Previous works reported similar behavior, where no impurity phases appeared with increasing Zr4+ content in other oxide nanomaterials [32,33,34]. As observed in Figure 1, the diffraction peaks of CoCr2O4 spinel phase broaden and lose sharpness with increasing Zr4+ content, leading to a decrease in crystallite size.
The experimentally determined lattice parameter, a e x p , is calculated using the following formula:
a e x p = d h k l h 2 + k 2 + l 2 1 / 2
where dhkl is plane spacing, and h, k and l are Miller indices.
The theoretical lattice parameter a t h of the investigated samples is calculated using the following equation [44]:
a t h = 8 3 3 r a + R 0 + 3     ( r b + R 0 )
where R 0 is the radius of the oxygen ion (1.32 Å), and r a and r b are the ionic radii of the tetrahedral and octahedral sites, respectively. Table 1 clearly demonstrates a match between the theoretical and experimental lattice parameters, confirming the expansion of the lattice.
The Nelson–Riley (N–R) method [45] provides a precise determination of lattice parameters through the following function:
F θ = 1 2 C o s 2 θ S i n θ + C o s 2 θ θ
Figure 2 displays the linear fit of a e x p versus F θ for the prepared samples. A graph is plotted with the calculated values of lattice parameters for each peak of a sample against F θ . By extrapolating these plots at F θ 0 , accurate values of the lattice constant are obtained, as listed in Table 1. The inset of Figure 2 illustrates that the lattice parameter adheres to Vegard’s law [46], showing a linear increase with the rise in Zr4+ content. This increase is attributed to the larger ionic radius of Zr4+ (0.86 Å) compared to Cr3+ (0.755 Å), with both ions having a coordination number of 6.
The percentage porosity ( P % ) of the samples is estimated using the following relation [47]:
P % = 1 ρ b ρ x × 100
where ρ b and ρ x are the experimental and theoretical (X-ray) densities, respectively.
ρ x = Z M w t N A a 3
ρ b = m π r 2 t
where Z is the number of molecules per primitive unit cell ( Z   = 8), M w t is the molecular weight of the material, NA is Avogadro’s number and m ,   r and t are the mass, radius and thickness of the pellet, respectively.
As observed in Table 1, the value of ρ x for CoCr2O4 is consistent with the value reported in literature [48]. The X-ray density ( ρ x ) increases with higher Zr4+ content. This increase in ρ x is likely due to the rise in molecular weight ( M w t ) associated with increasing Zr4+ content, which outweighs the effect of the expanding lattice constant. Additionally, ρ x is higher than the bulk density ( ρ b ) due to the presence of pores, which are influenced by both the sintering temperature and the preparation conditions [49,50]. The inverse behavior between the porosity and ρ x is observed, as it decreases with increasing Zr4+ content. The reduction in porosity is due to the reduction in crystallite sizes caused by increasing Zr4+ content, which results in a denser material structure with fewer voids.
The distances between magnetic ions for the tetrahedral ( L A ) and octahedral ( L B ) sites are determined by calculating the hopping lengths using the following equations [51]:
L A = a 3 4   L B = a 2 4  
Co2+ ions predominantly occupy the tetrahedral positions, while Cr3+ ions preferentially occupy the octahedral positions. The cation distribution [Co2+ ↔ Co3+] suggests that the probability of hopping is higher at octahedral sites, as the hopping length L B is shorter than that at the tetrahedral sites, L A , as shown in Table 1. The calculated values of L A and L B are consistent with those reported in literature [51]. The increase in both A-site ( L A ) and B-site ( L B ) hopping lengths with increasing Zr4+ content is primarily attributed to lattice expansion.
The Scherrer method [52] is widely regarded as the most common technique for estimating crystallite size, utilizing the broadening of the prominent (311) reflection line.
D = k λ β C o s θ
where k is the shape factor (usually assumed to be 0.9), λ is the wavelength of Cu Kα radiation, β is the full width of diffraction reflection at half the maximum intensity, and θ is the diffraction angle.
Another approach is the Williamson–Hall method, which provides more accurate insights into the crystallite size and the contributions of micro-strains observed within the crystal lattice [53].
β C o s θ = k λ D + 4 ε S i n θ
Here, D and ε correspond to the crystallite size value and microstrain, respectively.
The most intense reflection lines, namely (220), (311), (400), (511) and (400), are selected to create the linear plots. Plotting β C o s θ against 4 S i n θ allows for estimation of the average crystallite size and strain from the line’s extrapolation and slope, respectively, as shown in Figure 3. Crystallite sizes of CoCr2−xZrxO4 nanoparticles calculated by the Scherrer equation and the W–H method, along with the strain induced due to crystal imperfection, are listed in Table 1. The crystallite size calculated by both methods shows a consistent tendency, decreasing with increasing Zr4+ content, but the obtained values differ. The reduction in crystallite size may be attributed to the substitution of Cr3+ ions by Zr4+ ions, which introduces lattice distortion due to the differences in ionic radii between Zr4+ and Cr3+ ions. This distortion can hinder the growth of crystallites, resulting in smaller nanoparticle sizes. The substitution of Zr4+ affects the lattice parameters of the CoCr2O4 structure, impacting both the stability and the growth of the crystallites. An increase in lattice parameters can introduce internal negative stress, further limiting crystallite growth [54]. Strain values increase from 0.0013 to 0.0077 as Zr4+ content rises, which correlates with the observed decrease in particle size [55].
Figure 4a–d show transmission electron micrographs, with a scale bar of 100 nm, for CoCr2−xZrxO4 at x = 0.0, 0.05, 0.10 and 0.20, respectively. TEM images reveal spherical-like particles, with some agglomeration due to the magnetic nature and interaction between the nanoparticles [55]. Due to the non-uniform particle size, the average size of several particles is calculated. For CoCr2O4, the mean particle size is approximately 17.0 nm, and it decreases with increasing Zr4+ content. The particle size determined by TEM and the crystallite size calculated from the XRD results exhibit a similar trend, both decreasing as the Zr4+ content increases, as shown in Table 1. The particle size estimated from TEM images may encompass one or more crystallites, but it is consistently smaller than the crystallite sizes calculated using Scherrer’s equation from XRD spectra. This discrepancy is likely because TEM samples are prepared directly from the reaction solution, while XRD measurements are performed on powdered samples. Aging in the solution influences particle size, making particles more susceptible to aging effects that could increase band gap energy [56]. Thus, the aging process in the reaction solution is likely a key factor contributing to the observed differences between XRD and TEM measurements.
Spinel structure of multiferroic CoCr2O4 can be confirmed by infrared spectra (IR), Figure 5 presents the IR spectra of the synthesized nanoparticles within the range of 500–3500 cm⁻1. The characteristics of CoCr2O4 nanoparticles are influenced by the cation distribution at tetrahedral and octahedral sites. Two distinct bands at 629 cm⁻1 and 523 cm⁻1 confirm the formation of CoCr2O4 nanoparticles, attributed to the stretching vibrations of Cr3+-O2⁻ at tetrahedral and octahedral sites, respectively. Additionally, a minor band around 950 cm⁻1 corresponds to the Co(II)–O bond, aligning with previous findings [57,58]. A weak absorption band at 1631.58 cm⁻1 is due to atmospheric CO2, linked to the O–C vibration of the carboxyl group [59]. The final absorption peak observed at 3422 cm⁻1 is likely associated with the stretching vibrations of hydroxyl groups (OH) or H–O–H in H2O [60]. The values of ν1 and ν2, summarized in Table 2, show a shift to lower values with increasing Zr4+ substitution. The observed reduction in ν2 can be explained by the replacement of Cr3+ ions with the larger Zr4+ ions, which increases the metal-oxygen bond length within the octahedral site, leading to a decrease in the ν2 band. It is observed that the ν1 values are higher than ν2, which can be attributed to the shorter bond lengths in tetrahedral sites, resulting in stronger metal-oxygen bonds and higher vibrational frequencies.
IR data can be used to infer the elastic constants and Debye temperature of the spinel cobalt chromate system.
Force constants of CoCr2O4 nanoparticles at the tetrahedral ( K t ) and octahedral ( K O ) sites are calculated using the following formula [51]:
k = 4 π 2 c 2 v 2 μ
where c is the speed of light (~2.99 × 10 10 cm/s), v is the IR (vibrational) band frequency of the tetrahedral (A) and octahedral (B) sites and μ is the reduced mass of the Cr3+ and O2− ions (~2.03 × 10−23 g).
Table 2 displays the estimated values of K t and K O . It is noted that the force constant at the A-site ( K t ) is greater than that at the B-site ( K O ). The higher force constant at the tetrahedral site (A-site), compared to the octahedral site (B-site) in CoCr2O4, is due to the shorter bond lengths at the A-site and the stronger metal-oxygen bonds in the tetrahedral coordination (where the cation is surrounded by four oxygen atoms). In contrast, the B-site has an octahedral coordination, with six oxygen atoms surrounding the cation [61].
The lattice constant a and average force constant K a v are used to calculate the stiffness constant C 11 using the following relation [62]:
C 11 = K a v a
whereas the stiffness constant C 12 is given by the following [62]:
C 12 = C 11   σ ( 1 σ )
where σ is Poisson’s ratio and is given by the following [63]:
σ = 0.324 ( 1 1.043 P )
Table 2 shows that the stiffness constant C 11 reduces, whereas C 12 remains nearly constant with increasing Zr4+ content in CoCr2−xZrxO4. The decrease in C 11 can be attributed to the expansion of the lattice and the increased hopping length, which weaken the interatomic forces. According to Frabtsevich et al. [64], Poisson’s ratio is an indicator of a material’s brittleness or ductility. If Poisson’s ratio is less than 0.26, the material tends to be brittle; if it is greater than 0.26, the material tends to be ductile. It is clear from Table 2 that the ductility of the as-prepared samples increases with increasing Zr4+ content. This increase in ductility can be attributed to Zr4+ substitution, which influences grain size and the distribution of grain boundaries. A more refined grain structure with well-distributed boundaries enhances ductility, as these boundaries act as barriers to crack propagation.
For isotropic and homogeneous materials like spinel chromates, important elastic constants such as Young’s modulus ( E ), the bulk modulus ( B ) and the rigidity modulus ( G ) are critical in engineering applications.
The values of these moduli for all the samples are calculated by using the following formulas [65,66,67]:
Y o u n g   m o d u l u s       ( E ) = ( C 11 C 12 ) ( C 11 + 2 C 12 ) ( C 11 + C 12 )
B u l k   m o d u l u s   B   = 1 3 C 11 + 2 C 12
R i g i d i t y   m o d u l u s   G   = E 2 ( σ + 1 )
The values of all elastic moduli, E, B and G, are presented in Figure 6. These moduli show decreasing trends as the Zr4+ content increases. Comparable results have been reported for Al-substituted nickel ferrites, where the substitution was suggested to enhance interatomic bonding [67]. The decrease in elastic moduli is primarily due to the substitution of Cr3+ by Zr4+, which causes lattice expansion and weakens atomic bonding, resulting in lower elastic moduli. The increased ductility and decreased rigidity modulus in CoCr2−xZrxO4 materials make them well-suited for applications in flexible electronics, protective coatings, energy storage components, biomedical implants, and catalytic processes. In energy storage systems, improved ductility allows CoCr2−xZrxO4 electrodes to better accommodate mechanical stresses during repeated charge-discharge cycles, preventing cracking and increasing the system’s longevity and reliability. Additionally, reduced stiffness aids in distributing mechanical stress evenly, lowering the risk of failure. This also boosts ionic conductivity, leading to faster charge/discharge rates, while facilitating better thermal management by allowing slight deformation under thermal stress, which reduces the risk of thermal cracking in high-power applications.
For coatings, increased ductility enables CoCr2−xZrxO4 to flex with substrates undergoing mechanical deformation or thermal expansion, maintaining strong adhesion and preventing cracking. This enhanced ductility also improves impact resistance by absorbing and dissipating mechanical forces. The reduced stiffness allows the coating to conform to complex surface geometries, ensuring uniform protection and minimizing residual stresses that could lead to defects or cracks.
In flexible electronics, enhanced ductility allows CoCr2−xZrxO4 to withstand repeated bending and flexing while maintaining structural integrity and performance, thus improving fatigue resistance. The reduction in stiffness also makes the material easier to bend, which is essential for wearable technology, foldable screens, and other devices that require integration with soft substrates, ultimately enhancing both functionality and user comfort.
The longitudinal elastic wave velocity ( V l ), shear wave velocity ( V s ) and mean wave velocity V m are calculated using the following equations [65]. All velocities show a decreasing trend as the Zr4+ content increases, as illustrated in Figure 7.
V l = ( C 11 ρ x ) 0.5
V s = ( G 0 ρ x ) 0.5
1   V m 3 = 1 V l 3 + 2 V s 3
where G 0 is the rigidity modulus with zero pore fraction.
The decrease in shear velocity ( V s ) with increasing Zr4+ content is likely due to the larger ionic radius of Zr4+ compared to Cr3+. This leads to an expansion in the lattice parameters, which in turn reduces the material’s ability to transmit shear waves. Additionally, the substitution of Zr4+ (atomic mass ~91.22 u) for Cr3+ (atomic mass ~51.996 u) increases the atomic mass, thereby slowing the movement of atoms and reducing the shear velocity.
The Debye temperature is regarded as a fundamental property of solids and can be used to correlate the elastic properties with the thermodynamic properties of solids.
Two methods have been used to determine the Debye temperature: In the first method, the average value of the wavenumbers of bands in the FTIR spectrum is used to calculate the Debye temperature ( θ D I R ) using Waldron’s formula [68].
θ D I R = h c ν a v k B
where h is Planck’s constant, k B is Boltzmann’s constant, c is the velocity of light and ν a v is the average value of wavenumbers of bands, ν a v = v 1 + v 2 / 2 .
The Debye temperature ( θ D A ) is calculated in the second method using the Anderson Formula [69]:
θ D A = h k B 3 ρ x n N A 4 Π M w t 1 / 3 V m
Figure 8 demonstrates the good agreement between the Debye temperature computed from the elastic data ( θ D A ) and infrared data ( θ D I R ).
As shown in Figure 8, the Debye temperature decreases with increasing Zr4+ content. This is due to Zr4+’s higher atomic mass compared to Co2+ and its tendency to form weaker bonds than Cr3+, resulting in reduced lattice stiffness and, consequently, a lower Debye temperature. Similar findings have been reported for Zn-substituted cobalt ferrite [44]. Additionally, the substitution of Cr3+ with Zr4+ introduces n-type carriers, which may shift the material’s electrical conductivity and create oxygen vacancies to maintain charge neutrality. These additional conduction electrons could interact with phonons, leading to changes in the specific heat and a further reduction in θ D , consistent with specific heat theory [63]. Lower Debye temperatures enhance thermal shock resistance but may also accelerate material degradation due to weaker atomic bonds. As a result, these materials are well-suited for intermediate-temperature applications where both flexibility and moderate thermal management are essential.
Typical plots of the hysteresis loops for various Zr4+ doped cobalt chromate nanoparticles samples are shown in Figure 9. CoCr2O4 exhibits ferrimagnetic behavior at room temperature, as evidenced by its hysteresis loop. The net magnetization arises from the antiparallel alignment of Co2+↓ and Cr3+↑ ions within the spinel structure. The saturation magnetization ( M s ) of CoCr2−xZrxO4 is determined by an approximation of Stoner–Wohlfarth (S-W) theory by extrapolating the plot of M versus 1 / H 2 to approach zero according to the following equation [70], as displayed in the insets of Figure 9.
M = M s 1 α H 2
The estimated values of M s are displayed in Table 3. It is observed that M s reduces with increasing Zr4+ content, whereas the remnant magnetization ( M r ) and coercivity ( H c ) both decrease with increasing x and attain a minimum value at x = 0.10, after which they increase. The reduction in M s is consistent with previous findings for Zr-substituted Sr hexaferrite [71] and BiFeO3 [72]. Meanwhile, the increase in coercivity at higher Zr4+ content aligns with observations for zirconium-substituted CoFe2O4 [32]. The decrease in M s is primarily due to the substitution of non-magnetic Zr4+ ions for magnetic Cr3+ ions, which dilutes the magnetic ion network and weakens the superexchange interactions. Additionally, the larger ionic radius of Zr4+ induces lattice distortions, disrupts magnetic interactions and may cause spin canting, further reducing the overall magnetization. The reduction in coercivity with increasing Zr4+ concentration is attributed to the decrease in crystallite size. In the single-domain region, coercivity is related to crystallite size by the following equation [73]:
H c = g C D 2
where g and C are constants. However, the increase in coercivity with higher Zr4+ content is due to the creation of more pinning sites for magnetic domain walls, caused by increased lattice strain and defects from the substitution of Zr4+ ions. This substitution distorts the lattice, enhancing magnetic anisotropy and making domain wall movement more difficult, thus raising coercivity.
The magnetic moment ( μ m ) is determined from saturation magnetization by utilizing the following formula:
μ m = ( M w t × M s ) / 5585
where M w t = molecular weight, and 5585 = magnetic factor.
It is seen from Table 3 that the magnetic moment values gradually decrease from 0.021 to 0.018. The reduction in magnetic moment with increasing Zr4+ content is mainly due to the replacement of the magnetic Cr3+ ions with non-magnetic Zr4+ ions, which lacks unpaired d-electrons and therefore does not contribute to the magnetic moment. Moreover, a high Zr4+ concentration can cause disorder, potentially resulting in spin-glass-like behavior that further reduces the observable magnetic moment. CoCr2−xZrxO4 nanoparticles, with tunable magnetic properties, have a wide range of applications. Their lower coercivity makes them suitable for magnetic data storage, enabling easier writing and reading of data. They can also be used in catalysis for efficient separation processes and to enhance electrochemical performance in batteries and supercapacitors. Additionally, these nanoparticles have potential applications in sensors for detecting magnetic field changes and in composite materials where reduced magnetic strength is beneficial.

4. Conclusions

An efficient co-precipitation technique was used to synthesize Zr4+-substituted cobalt chromates nanoparticles, CoCr2−xZrxO4 (0.00 < x ≤ 0.20). X-ray diffraction (XRD) analysis confirmed the formation of a cubic spinel structure in all samples without any secondary phases. The lattice parameter, determined using the Nelson–Riley (N–R) extrapolation method, showed an increasing trend with higher Zr4+ content, which can be attributed to the larger ionic radius of Zr4+ (0.86 Å) compared to Cr3+ (0.755 Å). The broadening of XRD peaks for CoCr2−xZrxO4 nanoparticles was analyzed using Scherrer’s equation and Williamson–Hall (W–H) analysis, consistent with TEM results. These analyses revealed a gradual decrease in crystallite size as Zr4+ content increased. The increase in lattice parameters likely induces internal negative stress, which restricts crystallite growth. Fourier-transform infrared (FTIR) spectra revealed two characteristic absorption bands between 629 cm⁻1 and 523 cm⁻1, typical of spinel cobalt chromates. Notably, the absorption bands for the Zr4+-substituted samples shifted to lower wavenumbers compared to the as-prepared sample. This shift is likely due to the increase in lattice parameter, which leads to a longer metal-oxygen bond in the octahedral site and a consequent reduction in ν2. Through infrared spectral analysis, the stiffness constant, elastic moduli, and wave velocity values were determined. A decrease in the stiffness constant and all elastic moduli was observed with the substitution of Cr3+ ions by Zr4+ ions, attributed to lattice expansion and increased hopping length, both of which weaken interatomic forces. Additionally, the Debye temperature decreased with increasing Zr4+ content, linked to the introduction of n-type carriers due to the substitution of Cr3+ with Zr4+ ions. This process potentially alters electrical conductivity and generates oxygen vacancies to maintain charge neutrality. The magnetic properties of CoCr2−xZrxO4 (0.00 ≤ x ≤ 0.20) nanoparticles were investigated at room temperature using VSM. Analysis of the M ( H ) curve revealed that the nanoparticles exhibit nearly ferrimagnetic behavior. Magnetic measurements showed a decrease in saturation magnetization ( M s ) and magnetic moment ( μ m ) with increasing Zr4+ content. This reduction is due to the substitution of non-magnetic Zr4+ ions for magnetic Cr3+ ions, which dilutes the magnetic ion network and weakens superexchange interactions. Meanwhile, coercivity ( H c ) and remnant magnetization ( M r ) decreased with rising Zr4+ content up to x = 0.10 but increased beyond that. This rise in coercivity with higher Zr4+ content is attributed to the formation of additional pinning sites for magnetic domain walls, caused by increased lattice strain and defects from Zr4+ substitution. This makes domain wall movement more difficult and, as a result, increases coercivity.

Author Contributions

Methodology, D.E.-S.B.; Investigation, D.E.-S.B. and M.M.E.B.; Data analysis, M.M.E.B.; Writing—original draft, D.E.-S.B.; Writing—review & editing, M.M.E.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia through the project number 442/95.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research and Innovation, Ministry of Education in Saudi Arabia for funding this research work the project number 442/95. Also, the authors would like to extend their appreciation to Taibah University for its supervision support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lyons, D.H.; Kaplan, T.A.; Dwight, K.; Menyuk, N. Classical theory of the ground spin-state in cubic spinels. Phys. Rev. B 1962, 126, 540–555. [Google Scholar] [CrossRef]
  2. van Groenou, A.B.; Bongers, P.F.; Stuyts, A.L. Magnetism, microstructure and crystal chemistry of spinel ferrites. Mater. Sci. Eng. 1969, 3, 317–392. [Google Scholar] [CrossRef]
  3. Tomiyasu, K.; Fukunaga, J.; Suzuki, H. Magnetic short-range order and reentrant-spin-glass-like behavior in CoCr2O4 and MnCr2O4 by means of neutron scattering and magnetization measurements. Phys. Rev. B—Condens. Matter Mater. Phys. 2004, 70, 214434. [Google Scholar] [CrossRef]
  4. Lawes, G.; Melot, B.; Page, K.; Ederer, C.; Hayward, M.A.; Proffen, T.; Seshadri, R. Dielectric anomalies and spiral magnetic order in CoCr2O4. Phys. Rev. B—Condens. Matter Mater. Phys. 2006, 74, 024413. [Google Scholar] [CrossRef]
  5. Ishibashi, H.; Yasumi, T. Structural transition of spinel compound NiCr2O4 at ferrimagnetic transition temperature. J. Magn. Magn. Mater. 2007, 310, e610–e612. [Google Scholar] [CrossRef]
  6. Bush, A.A.; Shkuratov, V.Y.; Kamentsev, K.E.; Cherepanov, V.M. Preparation and X-Ray diffraction, dielectric, and Mössbauer characterization of Co1−xNix Cr2O4 solid solutions. Inorg. Mater. 2013, 49, 296–302. [Google Scholar] [CrossRef]
  7. Kim, I.; Oh, Y.S.; Liu, Y.; Chun, S.H.; Lee, J.-S.; Ko, K.-T.; Park, J.-H.; Chung, J.-H.; Kim, K.H. Electric polarization enhancement in multiferroic CoCr2O4 crystals with Cr-site mixing. Appl. Phys. Lett. 2009, 94, 042505. [Google Scholar] [CrossRef]
  8. Choi, Y.J.; Okamoto, J.; Huang, D.J.; Chao, K.S.; Lin, H.J.; Chen, C.T.; Van Veenendaal, M.; Kaplan, F.T.; Cheong, S.W. Thermally or magnetically induced polarization reversal in the multiferroic CoCr2O4. Phys. Rev. Lett. 2009, 102, 067601. [Google Scholar] [CrossRef] [PubMed]
  9. Köseoğlu, Y.; Baykal, A.; Toprak, M.S.; Gözüak, F.; Başaran, A.C.; Aktaş, B. Synthesis and characterization of ZnFe2O4 magnetic nanoparticles via a PEG-assisted route. J. Alloys Compd. 2008, 462, 209–213. [Google Scholar] [CrossRef]
  10. Iqbal, M.J.; Siddiquah, M.R. Electrical and magnetic properties of chromium-substituted cobalt ferrite nanomaterials. J. Alloys Compd. 2008, 453, 513–518. [Google Scholar] [CrossRef]
  11. He, H.Y. Structural and Magnetic Property of Co1−x Nix Fe2O4 Nanoparticles Synthesized by Hydrothermal Method. Int. J. Appl. Ceram. Technol. 2014, 11, 626–636. [Google Scholar] [CrossRef]
  12. Zakrzewska, K. Mixed oxides as gas sensors. Thin Solid Films 2001, 391, 229–238. [Google Scholar] [CrossRef]
  13. Kim, B.-N.; Hiraga, K.; Morita, K.; Sakka, Y. A high-strain-rate superplastic ceramic. Nature 2001, 413, 288–291. [Google Scholar] [CrossRef] [PubMed]
  14. Galdikas, A.; Martūnas, Z.; Šetkus, A. SnInO-based chlorine gas sensor. Sens. Actuators B Chem. 1992, 7, 633–636. [Google Scholar] [CrossRef]
  15. Reddy CV, G.; Manorama, S.V.; Rao, V.J. Semiconducting gas sensor for chlorine based on inverse spinel nickel ferrite. Sens. Actuators B Chem. 1999, 55, 90–95. [Google Scholar] [CrossRef]
  16. Durrani, S.K.; Hussain, S.Z.; Saeed, K.; Khan, Y.; Arif, M.; Ahmed, N. Hydrothermal synthesis and characterization of nanosized transition metal chromite spinels. Turk. J. Chem. 2012, 36, 111–120. [Google Scholar] [CrossRef]
  17. Dutta, D.P.; Manjanna, J.; Tyagi, A.K. Magnetic properties of sonochemically synthesized CoCr2O4 nanoparticles. J. Appl. Phys. 2009, 106, 043915. [Google Scholar] [CrossRef]
  18. Li, S.; Zhao, G.; Bi, H.; Huang, Z.; Lai, H.; Gai, R.; Du, Y. Synthesis and anomalous magnetic properties of CoCr2O4 nanocrystallites with lattice distortion. J. Magn. Magn. Mater. 2006, 305, 448–451. [Google Scholar] [CrossRef]
  19. Li, S.; Bi, H.; Tian, Z.; Xu, F.; Gu, B.; Lu, M.; Du, Y. Surface spin pinning effect of polymer decomposition residues in CoCr2O4 nanocrystallites system. J. Magn. Magn. Mater. 2004, 281, 11–16. [Google Scholar] [CrossRef]
  20. Rath, C.; Mohanty, P. Magnetic phase transitions in cobalt chromite nanoparticles. J. Supercond. Nov. Magn. 2011, 24, 629–633. [Google Scholar] [CrossRef]
  21. Hu, D.-S.; Han, A.-J.; Ye, M.-Q.; Chen, H.-H.; Zhang, W. Preparation and Spectroscopy Analysis of Spinel CoCr2−xAlxO4 by Low-temperature Combustion Synthesis. J. Inorg. Mater. 2011, 26, 285–289. [Google Scholar] [CrossRef]
  22. Singh, R.K.; Yadav, A.; Narayan, A.; Singh, A.K.; Verma, L.; Verma, R.K. Thermal, structural and magnetic studies on chromite spinel synthesized using citrate precursor method and annealed at 450 and 650 °C. J. Therm. Anal. Calorim. 2012, 107, 197–204. [Google Scholar] [CrossRef]
  23. Cui, H.; Zayat, M.; Levy, D. Sol-gel synthesis of nanoscaled spinels using propylene oxide as a gelation agent. J. Sol-Gel Sci. Technol. 2005, 35, 175–181. [Google Scholar] [CrossRef]
  24. Das, D.; Biswas, R.; Ghosh, S. Systematic analysis of structural and magnetic properties of spinel CoB2O4 (B = Cr, Mn and Fe) compounds from their electronic structures. J. Phys. Condens. Matter 2016, 28, 446001. [Google Scholar] [CrossRef] [PubMed]
  25. Kamran, M.; Ullah, A.; Rahman, S.; Tahir, A.; Nadeem, K.; ur Rehman, M.A.; Hussain, S. Structural, magnetic, and dielectric properties of multiferroic Co1−xMgxCr2O4 nanoparticles. J. Magn. Magn. Mater. 2017, 433, 178–186. [Google Scholar] [CrossRef]
  26. Kumar, N.; Sundaresan, A. On the observation of negative magnetization under zero-field-cooled process. Solid State Commun. 2010, 150, 1162–1164. [Google Scholar] [CrossRef]
  27. Suchomski, C.; Reitz, C.; Brezesinski, K.; de Sousa, C.T.; Rohnke, M.; Iimura, K.-I.; de Araujo, J.P.E.; Brezesinski, T. Structural, Optical, and Magnetic Properties of Highly Ordered Mesoporous MCr2O4 and MCr2–xFexO4 (M = Co, Zn) Spinel Thin Films with Uniform 15 nm Diameter Pores and Tunable Nanocrystalline Domain Sizes. Chem. Mater. 2012, 24, 155–165. [Google Scholar] [CrossRef]
  28. Nadeem, K.; Rehman, H.U.; Zeb, F.; Ali, E.; Kamran, M.; Noshahi, N.; Abbas, H. Magnetic phase diagram and dielectric properties of Mn doped CoCr2O4 nanoparticles. J. Alloys Compd. 2020, 832, 155031. [Google Scholar] [CrossRef]
  29. Kamran, M.; Nadeem, K.; Mumtaz, M. Negative and anomalous T-dependent magnetization trend in CoCr2O4 nanoparticles. Solid State Sci. 2017, 72, 21–27. [Google Scholar] [CrossRef]
  30. Kumar, G.J.; Rath, C. Study of exchange bias and memory effect in core-shell CoCr2O4 nanoparticles. J. Magn. Magn. Mater. 2018, 466, 69–74. [Google Scholar] [CrossRef]
  31. Tian, Z.; Zhu, C.; Wang, J.; Xia, Z.; Liu, Y.; Yuan, S. Size dependence of structure and magnetic properties of CoCr2O4 nanoparticles synthesized by hydrothermal technique. J. Magn. Magn. Mater. 2015, 377, 176–182. [Google Scholar] [CrossRef]
  32. Kavitha, S.; Kurian, M. Effect of zirconium doping in the microstructure, magnetic and dielectric properties of cobalt ferrite nanoparticles. J. Alloys Compd. 2019, 799, 147–159. [Google Scholar] [CrossRef]
  33. Naik, E.I.; Naik, H.B.; Viswanath, R.; Kirthan, B.R.; Prabhakara, M.C. Effect of zirconium doping on the structural, optical, electrochemical and antibacterial properties of ZnO nanoparticles prepared by sol-gel method. Chem. Data Collect. 2020, 29, 100505. [Google Scholar] [CrossRef]
  34. Monaji, V.R.; Indla, S.; Rayaprol, S.; Sowmya, S.; Srinivas, A.; Das, D. Temperature dependent magnetic properties of Co1+xTxFe2−2xO4 (T = Zr, Ti). J. Alloys Compd. 2017, 700, 92–97. [Google Scholar] [CrossRef]
  35. Xu, Q.; Zhan, D.; Huang, D.-P.; Liu, H.-X.; Chen, W.; Zhang, F. Dielectric inspection of BaZr0.2Ti0.8O3 ceramics under bias electric field: A survey of polar nano-regions. Mater. Res. Bull. 2012, 47, 1674–1679. [Google Scholar] [CrossRef]
  36. Faraz, A. Effect of Concentration of Zr4+ and Ni2+ Dopants on Electrical, Magnetic and Y–K Angle of Mg–Cu Complex Spinel Nanoferrites. J. Supercond. Nov. Magn. 2012, 25, 1055–1063. [Google Scholar] [CrossRef]
  37. Reda, M.; El-Dek, S.I.; Arman, M.M. Improvement of ferroelectric properties via Zr doping in barium titanate nanoparticles. J. Mater. Sci. Mater. Electron. 2022, 33, 16753–16776. [Google Scholar] [CrossRef]
  38. Purnamasari, I.; Triyono, D. Effect of zirconium substitution on structural and optical properties of lanthanum orthoferrite. IOP Conf. Ser. Mater. Sci. Eng. 2020, 902, 012031. [Google Scholar] [CrossRef]
  39. Hashim, M.; Kumar, S.; Shirsath, S.E.; Kotnala, R.K.; Shah, J.; Kumar, R. Influence of Cr3+ ion on the structural, ac conductivity and magnetic properties of nanocrystalline Ni–Mg ferrite. Ceram. Int. 2013, 39, 1807–1819. [Google Scholar] [CrossRef]
  40. Dawood, M.S.; Elmosalami, T.; Desoky, W. Enhancement of elastic, optical and opto-electrical properties of Ni-Substituted CoFe2O4 nanoparticles with different concentrations. Opt. Mater. 2021, 117, 111101. [Google Scholar] [CrossRef]
  41. Chandekar, K.V.; Shkir, M.; AlFaify, S. A structural, elastic, mechanical, spectroscopic, thermodynamic, and magnetic properties of polymer coated CoFe2O4 nanostructures for various applications. J. Mol. Struct. 2020, 1205, 127681. [Google Scholar] [CrossRef]
  42. Babu, B.R.; Tatarchuk, T. Elastic properties and antistructural modeling for nickel-zinc ferrite-aluminates. Mater. Chem. Phys. 2018, 207, 534–541. [Google Scholar] [CrossRef]
  43. Manjunatha, K.; Jagadeesha Angadi, V.; Srinivasamurthy, K.M.; Matteppanavar, S.; Pattar, V.K.; Mahaboob Pasha, U. Exploring the structural, dielectric and magnetic properties of 5 Mol% Bi3+-substituted CoCr2O4 nanoparticles. J. Supercond. Nov. Magn. 2020, 33, 1747–1757. [Google Scholar] [CrossRef]
  44. Patil, V.; Shirsath, S.E.; More, S.; Shukla, S.; Jadhav, K. Effect of zinc substitution on structural and elastic properties of cobalt ferrite. J. Alloys Compd. 2009, 488, 199–203. [Google Scholar] [CrossRef]
  45. Debnath, S.; Das, R. Cobalt doping on nickel ferrite nanocrystals enhances the micro-structural and magnetic properties: Shows a correlation between them. J. Alloys Compd. 2021, 852, 156884. [Google Scholar] [CrossRef]
  46. Denton, A.R.; Ashcroft, N.W. Vegard’s law. Phys. Rev. A 1991, 43, 3161. [Google Scholar] [CrossRef]
  47. Safaan, S.; El Ata, A.A.; El Messeery, M. Study of some structural and magnetic properties of Mn-substituted SrCu hexagonal ferrites. J. Magn. Magn. Mater. 2006, 302, 362–367. [Google Scholar] [CrossRef]
  48. Akyol, M.; Adanur, I.; Ayaş, A.O.; Ekicibil, A. Magnetic field dependence of magnetic coupling in CoCr2O4 nanoparticles. Phys. B Condens. Matter 2017, 525, 144–148. [Google Scholar] [CrossRef]
  49. Islam, M.U.; Abbas, T.; Niazi, S.B.; Ahmad, Z.; Sabeen, S.; Chaudhry, M.A. Electrical behaviour of fine particle, co-precipitation prepared Ni–Zn ferrites. Solid State Commun. 2004, 130, 353–356. [Google Scholar] [CrossRef]
  50. Deraz, N.; Alarifi, A. Preparation and characterization of nano-magnetic Mn0.5Zn0.5Fe2O4 system. Int. J. Electrochem. Sci. 2012, 7, 5828–5836. [Google Scholar] [CrossRef]
  51. Choudhary, P.; Saxena, P.; Yadav, A.; Sinha, A.K.; Rai, V.N.; Varshney, M.D.; Mishra, A. Weak ferroelectricity and leakage current behavior of multiferroic CoCr2O4 nanomaterials. J. Supercond. Nov. Magn. 2019, 32, 2639–2645. [Google Scholar] [CrossRef]
  52. López-Ortega, A.; Lottini, E.; Fernández, C.D.J.; Sangregorio, C. Exploring the magnetic properties of cobalt-ferrite nanoparticles for the development of a rare-earth-free permanent magnet. Chem. Mater. 2015, 27, 4048–4056. [Google Scholar] [CrossRef]
  53. Nair, D.S.; Kurian, M. Highly selective synthesis of diphenyl methane via liquid phase benzylation of benzene over cobalt doped zinc nanoferrite catalysts at mild conditions. J. Saudi Chem. Soc. 2019, 23, 127–132. [Google Scholar] [CrossRef]
  54. El-Said Bakeer, D. Investigation on Optical, Dielectric, and Magnetic Properties of CoAl2−x FexO4 Nanoparticles. J. Supercond. Nov. Magn. 2020, 33, 1789–1801. [Google Scholar] [CrossRef]
  55. Rikamukti, N.; Purnama, B. Effect of doping Strontium ions in co-precipitated cobalt ferrite. J. Phys. Conf. Ser. 2017, 909, 012012. [Google Scholar] [CrossRef]
  56. Bitar, Z.; Isber, S.; Noureddine, S.; Bakeer, D.E.-S.; Awad, R. Synthesis, Characterization, Optical Properties, and Electron Paramagnetic Resonance for Nano Zn0.5Co0.5Fe2−xPrxO4. J. Supercond. Nov. Magn. 2017, 30, 3603–3609. [Google Scholar] [CrossRef]
  57. Hu, J.; Zhao, W.; Hu, R.; Chang, G.; Li, C.; Wang, L. Catalytic activity of spinel oxides MgCr2O4 and CoCr2O4 for methane combustion. Mater. Res. Bull. 2014, 57, 268–273. [Google Scholar] [CrossRef]
  58. Choudhary, P.; Saxena, P.; Yadav, A.; Rai, V.N.; Mishra, A. Dielectric and ferroelectric properties of CoCr2O4 nanoceramics. J. Adv. Dielectr. 2019, 9, 1950015. [Google Scholar] [CrossRef]
  59. Khattab, R.; Sadek, H.; Gaber, A. Synthesis of CoxMg1−xAl2O4 nanospinel pigments by microwave combustion method. Ceram. Int. 2017, 43, 234–243. [Google Scholar] [CrossRef]
  60. Deepty, M.; Srinivas, C.; Kumar, E.R.; Mohan, N.K.; Prajapat, C.L.; Rao, T.C.; Meena, S.S.; Verma, A.K.; Sastry, D.L. XRD, EDX, FTIR and ESR spectroscopic studies of co-precipitated Mn–substituted Zn–ferrite nanoparticles. Ceram. Int. 2019, 45, 8037–8044. [Google Scholar] [CrossRef]
  61. El-Said Bakeer, D. Elastic study and optical dispersion characterization of Fe-substituted cobalt aluminate nanoparticles. Appl. Phys. A 2020, 126, 443. [Google Scholar] [CrossRef]
  62. Tatarchuk, T.; Paliychuk, N.; Bououdina, M.; Al-Najar, B.; Pacia, M.; Macyk, W.; Shyichuk, A. Effect of cobalt substitution on structural, elastic, magnetic and optical properties of zinc ferrite nanoparticles. J. Alloys Compd. 2018, 731, 1256–1266. [Google Scholar] [CrossRef]
  63. El-Ghazzawy, E.H.; Amer, M.A. Structural, elastic and magnetic studies of the as-synthesized Co1−xSrxFe2O4 nanoparticles. J. Alloys Compd. 2017, 690, 293–303. [Google Scholar] [CrossRef]
  64. Frantsevich, S.A.B.I.N.; Voronov, F.F.; Frant, I.N. Elastic Constants and Elastic Moduli of Metals and Insulators Handbook; Naukova Dumka: Kiev, Ukraine, 1983. [Google Scholar]
  65. Patange, S.; Shirsath, S.E.; Lohar, K.; Algude, S.; Kamble, S.; Kulkarni, N.; Mane, D.; Jadhav, K. Infrared spectral and elastic moduli study of NiFe2−xCrxO4 nanocrystalline ferrites. J. Magn. Magn. Mater. 2013, 325, 107–111. [Google Scholar] [CrossRef]
  66. Modi, K.B.; Gajera, J.D.; Pandya, M.P.; Vora, G.; Joshi, H.H. Far-infrared spectral studies of magnesium and aluminum co-substituted lithium ferrites. Pramana 2004, 62, 1173–1180. [Google Scholar] [CrossRef]
  67. Patange, S.; Shirsath, S.E.; Jadhav, S.; Hogade, V.; Kamble, S.; Jadhav, K. Elastic properties of nanocrystalline aluminum substituted nickel ferrites prepared by co-precipitation method. J. Mol. Struct. 2013, 1038, 40–44. [Google Scholar] [CrossRef]
  68. Waldron, R.D. Infrared spectra of ferrites. Phys. Rev. B 1955, 99, 1727–1735. [Google Scholar] [CrossRef]
  69. Anderson, O.L.; Mason, W.P. (Eds.) Physics Acoustics; Academic Press: New York, NY, USA, 1965; Volume 3BC. [Google Scholar]
  70. Amir; Gungunes, H.; Slimani, Y.; Tashkandi, N.; El Sayed, H.S.; Aldakheel, F.; Sertkol, M.; Sozeri, H.; Manikandan, A.; Ercan, I.; et al. Mössbauer studies and magnetic properties of cubic CuFe2O4 nanoparticles. J. Supercond. Nov. Magn. 2019, 32, 557–564. [Google Scholar] [CrossRef]
  71. Almessiere, M.; Slimani, Y.; Sertkol, M.; Nawaz, M.; Baykal, A.; Ercan, I. The impact of Zr substituted Sr hexaferrite: Investigation on structure, optic and magnetic properties. Results Phys. 2019, 13, 102244. [Google Scholar] [CrossRef]
  72. Arora, M.; Chauhan, S.; Sati, P.C.; Kumar, M. Effect of non-magnetic ions substitution on structural, magnetic and optical properties of BiFeO3 nanoparticles. J. Supercond. Nov. Magn. 2014, 27, 1867–1871. [Google Scholar] [CrossRef]
  73. Culity, B.D.; Graham, C.D. Introduction to Magnetic Materials; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009. [Google Scholar]
Figure 1. XRD patterns of CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20.
Figure 1. XRD patterns of CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20.
Materials 17 05149 g001
Figure 2. The variation of lattice parameter vs. Nelson–Riley (N–R) function for CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20. The inset shows the variation of lattice parameter with Zr4+ content, x.
Figure 2. The variation of lattice parameter vs. Nelson–Riley (N–R) function for CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20. The inset shows the variation of lattice parameter with Zr4+ content, x.
Materials 17 05149 g002
Figure 3. The plots of β C o s θ versus 4 S i n θ according to the W–H method for CoCr2−xZrxO4 nanoparticles.
Figure 3. The plots of β C o s θ versus 4 S i n θ according to the W–H method for CoCr2−xZrxO4 nanoparticles.
Materials 17 05149 g003
Figure 4. TEM images, with a scale bar of 100 nm, for CoCr2−xZrxO4 nanoparticles, (a) x = 0.00, (b) x = 0.05, (c) x = 0.10 and (d) x = 0.20.
Figure 4. TEM images, with a scale bar of 100 nm, for CoCr2−xZrxO4 nanoparticles, (a) x = 0.00, (b) x = 0.05, (c) x = 0.10 and (d) x = 0.20.
Materials 17 05149 g004aMaterials 17 05149 g004b
Figure 5. FTIR spectra for CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20.
Figure 5. FTIR spectra for CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20.
Materials 17 05149 g005
Figure 6. Variation of elastic moduli constants with Zr4+ content.
Figure 6. Variation of elastic moduli constants with Zr4+ content.
Materials 17 05149 g006
Figure 7. Variation of longitudinal ( V l ), shearing ( V s ) and mean ( V m ) wave velocities with Zr4+ content.
Figure 7. Variation of longitudinal ( V l ), shearing ( V s ) and mean ( V m ) wave velocities with Zr4+ content.
Materials 17 05149 g007
Figure 8. The variation of the Debye temperature θ D I R and θ D A with Zr4+ content.
Figure 8. The variation of the Debye temperature θ D I R and θ D A with Zr4+ content.
Materials 17 05149 g008
Figure 9. Room temperature magnetic hysteresis loops of BaFe12−xHgxO19 nanoparticles, x = 0.00, 0.05, 0.10, 0.15 and 0.20. The insets show the plots of M versus 1/ H 2 .
Figure 9. Room temperature magnetic hysteresis loops of BaFe12−xHgxO19 nanoparticles, x = 0.00, 0.05, 0.10, 0.15 and 0.20. The insets show the plots of M versus 1/ H 2 .
Materials 17 05149 g009aMaterials 17 05149 g009b
Table 1. XRD parameters, crystallite size and micro strain of CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20.
Table 1. XRD parameters, crystallite size and micro strain of CoCr2−xZrxO4 nanoparticles, 0.00 ≤ x ≤ 0.20.
Zr4+ Content, x a
(Å)
a t h
(Å)
ρ x
(g/cm3)
ρ b
(g/cm3)
P%Hopping LengthsScherrer
Equation
W-H MethodTEM
L A
(Å)
L B
(Å)
D (nm)D (nm) ε × 10−3L (nm)
0.008.3118.7815.2514.21819.6723.5982.93818.50022.011.3016.998
0.058.3148.7885.2914.29918.7403.6002.93914.99920.423.1411.172
0.108.3188.7955.3284.35718.2243.6022.94114.29419.533.219.346
0.158.3198.8065.3724.48816.4623.6022.94111.17917.033.327.073
0.208.3208.8105.4154.64914.1513.6032.9429.26511.917.716.120
Table 2. The absorption bands (ν1 and ν2), force constant ( K t and K O ) and Poisson’s ratio of CoCr2−xZrxO4 nanoparticles.
Table 2. The absorption bands (ν1 and ν2), force constant ( K t and K O ) and Poisson’s ratio of CoCr2−xZrxO4 nanoparticles.
xν1
(cm−1)
ν2
(cm−1)
K t × 10 5
(dyne/cm)
K O × 10 5
(dyne/cm)
C 11
GPa
C 12
GPa
σ
0.00629.66523.082.8371.958288.538103.2980.2636
0.05626.56522.792.8091.956286.627102.8440.2641
0.10625.47521.572.8001.947285.3545103.3200.2658
0.15624.81519.642.7941.932284.101102.8740.2659
0.20623.74517.922.7841.919282.724102.7750.2666
Table 3. Magnetic parameters of CoCr2−xZrxO4 nanoparticles.
Table 3. Magnetic parameters of CoCr2−xZrxO4 nanoparticles.
x M s (emu/g) M r × 10−2
(emu/g)
H c (G) μ m
0.000.5201.36362.070.0209
0.050.5151.10278.920.0211
0.100.5060.82213.250.0209
0.150.4660.851209.20.0194
0.200.4320.931357.10.0181
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

Barakat, M.M.E.; El-Said Bakeer, D. Investigation of Structural, Elastic and Magnetic Properties of CoCr2−xZrxO4 Nanoparticles. Materials 2024, 17, 5149. https://doi.org/10.3390/ma17215149

AMA Style

Barakat MME, El-Said Bakeer D. Investigation of Structural, Elastic and Magnetic Properties of CoCr2−xZrxO4 Nanoparticles. Materials. 2024; 17(21):5149. https://doi.org/10.3390/ma17215149

Chicago/Turabian Style

Barakat, Mai M. E., and Doaa El-Said Bakeer. 2024. "Investigation of Structural, Elastic and Magnetic Properties of CoCr2−xZrxO4 Nanoparticles" Materials 17, no. 21: 5149. https://doi.org/10.3390/ma17215149

APA Style

Barakat, M. M. E., & El-Said Bakeer, D. (2024). Investigation of Structural, Elastic and Magnetic Properties of CoCr2−xZrxO4 Nanoparticles. Materials, 17(21), 5149. https://doi.org/10.3390/ma17215149

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