Tran Thi Ngoc Nhaa,
Sankar Hari Prakashb,
Selvaraj Mohana Roopan
b,
James Jebaseelan Samuelc,
Dang Ngoc Toan
de,
Dinh Thanh Khanf,
Do Danh Bichg,
Tran Dang Thanh
h,
Le Thi Tuyet Ngan
i,
Do Hung Manh
h and
Pham Thanh Phong
*jk
aGraduate University of Science and Technology, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet, Hanoi, Vietnam
bChemistry of Heterocycles & Natural Product Research Laboratory, Department of Chemistry, School of Advanced Sciences, Vellore Institute of Technology, Vellore, Tamil Nadu, India
cMedical Gel Dosimetry Lab, Department of Physics, School of Advanced Sciences, Vellore Institute of Technology, Vellore, Tamil Nadu, India
dInstitute of Research and Development, Duy Tan University, Danang 550000, Vietnam
eFaculty of Natural Sciences, Duy Tan University, Danang 550000, Vietnam
fThe University of Danang-University of Science and Education, Danang 550000, Vietnam
gFaculty of Physics, Hanoi National University of Education, 136 Xuan Thuy, Cau Giay, Hanoi 100000, Vietnam
hInstitute of Materials Science, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet, Hanoi, Vietnam
iFaculty of Engineering and Technology, Thai Nguyen University of Information and Communication Technology, Z115 Street, Quyet Thang Ward, Thai Nguyen Province, Vietnam
jLaboratory of Magnetism and Magnetic Materials, Science and Technology Advanced Institute, Van Lang University, Ho Chi Minh City, Vietnam. E-mail: phamthanhphong@vlu.edu.vn
kFaculty of Applied Technology, School of Technology, Van Lang University, Ho Chi Minh City, Vietnam
First published on 21st August 2025
MFe2O4 (M = Co and Mn) nanoparticles were synthesized from coconut coir extract using a microwave-assisted co-precipitation method, representing a green and sustainable approach for ferrite nanomaterial preparation. The physical properties of the samples were characterized using X-ray diffraction, scanning electron microscopy, ultraviolet-visible spectroscopy, photoluminescence, Raman spectroscopy, and vibrating sample magnetometry. Scanning electron micrographs revealed nanoscale morphology with evidence of polymorphism. Rietveld refinement confirmed the formation of single-phase spinel ferrites with lattice constants ranging from 8.4224 Å to 8.4782 Å for CoFe2O4 and MnFe2O4, respectively. The distribution of metal cations at the tetrahedral and octahedral sites in the AB2O4 spinel lattice was found to depend on the synthesis route and significantly influenced the magnetic and optical behaviors of the materials. Raman spectra exhibited characteristic peaks corresponding to a mixed spinel structure. The optical band gaps estimated from the UV-vis spectra were 2.66 eV for CoFe2O4 and 2.64 eV for MnFe2O4. PL spectra showed four distinct emission peaks at 458, 692, 758, and 871 nm. Based on UV-vis and photoluminescence spectral results, a schematic energy band structure was constructed. Magnetic measurements, analyzed using the “law of approach” to saturation, revealed saturation magnetizations of 70 emu g−1 (CoFe2O4) and 49 emu g−1 (MnFe2O4) at 55 K—values that are among the highest reported for these systems; the squareness ratios were 0.58 and 0.12, respectively. The CoFe2O4 sample exhibited high effective anisotropy due to surface spin contributions, resulting in high coercivity and squareness. In contrast, the enhanced dipolar interactions in MnFe2O4 reduced coercivity and squareness. These magnetic behaviors were interpreted within the frameworks of the Stoner–Wohlfarth and superparamagnetic models that account for interparticle interactions.
Manganese ferrite (MnFe2O4) in its bulk form is generally classified as a normal spinel9 or, in some cases, a mixed spinel.10 However, at the nanoscale, MnFe2O4 nanoparticles (NPs) typically exhibit a mixed or partially inverted spinel structure, which can be described by the formula (Mn1−x2+Fex3+)[Mnx2+Fe2−x3+]O42−.4 This cation redistribution enhances the strength of the superexchange interactions between the A and B sublattices, potentially resulting in a higher saturation magnetization compared to their bulk counterparts.9 These distinctive structural and magnetic features render MnFe2O4-based nanoparticles highly suitable for a range of applications, including magnetic data storage, catalysis, ferrofluids, sensors, and actuators.11 Furthermore, MnFe2O4 NPs have demonstrated significant promise in magneto-optical and optoelectronic devices due to their capability to absorb low-energy photons, making them ideal for photocatalytic applications.12,13 Their inherently low coercivity, high magnetic permeability, and minimal energy loss also position them as attractive candidates for biomedical technologies, such as targeted drug delivery, magnetic hyperthermia, and magnetic resonance imaging contrast enhancement.14,15
In contrast, cobalt ferrite (CoFe2O4) exhibits an inverse spinel structure in its bulk form, a configuration that is largely preserved even at the nanoscale.16 CoFe2O4 NPs are characterized by a high Curie temperature (TC ≈ 793 K), large coercivity, moderate saturation magnetization (∼80 emu g−1 at room temperature), and pronounced magnetocrystalline anisotropy. In addition, they possess exceptional chemical and electrochemical stability, high mechanical hardness, and excellent magnetic permeability.17 These properties have enabled their deployment across a diverse array of biomedical and technological domains, including antibacterial, anticancer, antioxidant, and enzyme-inhibitory applications, as well as in energy storage systems, catalytic processes, and components for sensors, transistors, solar cells, and photocatalytic devices.18–23
The diverse applications of ferrites necessitate the selection of appropriate synthesis techniques to ensure optimal performance. A variety of physical and chemical methods have been employed for the fabrication of ferrite NPs, including sol–gel,24 hydrothermal,25 co-precipitation,26 thermal decomposition,27 and high-energy ball milling.28 While these approaches offer flexibility in process control and enable the rapid production of large quantities of ferrite NPs, many rely on the use of toxic solvents, reducing agents, and stabilizers. Such chemicals pose significant environmental and health hazards, especially in long-term exposure. Consequently, there is an increasing demand for alternative, environmentally benign synthesis routes—particularly green methodologies—that minimize toxic by-products while maintaining or enhancing the desired functional properties of nanoferrites.
Among the various techniques, co-precipitation is widely favored due to its simplicity and capacity for precise control over key synthesis parameters, such as particle size, morphology, temperature, pH, precursor concentration, and alkali addition rate.29 However, achieving high crystallinity in the resulting ferrite NPs typically requires a post-synthesis calcination step at elevated temperatures.30 Moreover, maintaining consistent product quality and stability throughout the synthesis process can be challenging, particularly when the precipitation reaction lacks continuous control.31 To address these limitations, we propose a green, microwave-assisted co-precipitation method that incorporates a biocompatible and non-toxic natural surfactant derived from coconut coir extract. This approach integrates the advantages of microwave heating with sustainable synthesis practices. Microwave irradiation significantly reduces reaction time and enhances reaction efficiency by directly coupling with materials through mechanisms, such as dielectric and magnetic interactions, in contrast to the indirect convective heating used in conventional methods.25,32 As a result, this method facilitates rapid and uniform heating of reactants, promotes homogeneous nucleation, and enables the formation of well-defined, crystalline MFe2O4 nanoparticles. The proposed technique thus represents a promising, eco-friendly strategy for the synthesis of high-performance ferrite nanomaterials.
MFe2O4 (M = Co, Mn, Ni, Zn, etc.) NPs have been successfully synthesized using phytochemicals extracted from various plant components, including flowers, leaves, stems, seeds, fruits, and roots.33 These biological sources, often derived from agricultural waste or naturally occurring plants, are typically washed with distilled water to remove surface impurities. The extracts are then obtained either directly from fresh biomass or through drying and grinding processes. Among these sources, coconut stands out as one of the most abundant and readily available plants in Southeast Asia and India, making it an ideal candidate for synthesizing green nanomaterials. In particular, coconut-coir extract has been shown to contain various phytochemicals, such as polyphenols, tannins, and flavonoids, which act as natural reducing and stabilizing agents during the formation of nanoparticles.34 These compounds facilitate the formation of metal oxide nanoparticles by reducing metal ions and preventing agglomeration, thereby enabling a sustainable and non-toxic synthesis route. The reducing capability of these phytochemicals has been previously verified using FTIR spectroscopy. For example, Elango et al.35,36 reported the disappearance or shift of the –OH and CO stretching bands after the synthesis of Ni and Pd nanoparticles with coconut coir extract, suggesting that these functional groups are directly involved in the redox reaction. Although our current study did not include FTIR or LC-MS analysis, the presence of such functional groups in the extract—as confirmed in earlier studies—strongly supports its role in nanoparticle formation. The effective control over particle size and phase purity observed in our work is thus consistent with the phytochemical-mediated synthesis mechanism proposed in the literature. Despite its availability, only a limited number of studies have employed coconut-based bio-extracts in the synthesis of nanoparticle formulations. Existing reports have focused primarily on the green synthesis of metal and metal oxide nanoparticles—such as silver, lead, palladium, copper oxide, and zinc oxide—using coconut coir as a reducing and stabilizing agent.34–38 To date, however, no reports have been published on the one-step green synthesis of CoFe2O4 and MnFe2O4 NPs via microwave-assisted co-precipitation (MAP) using coconut coir extract. The present study addresses this gap by developing a sustainable MAP approach for the synthesis of MFe2O4 (M = Co, Mn) NPs. We further investigate the influence of morphology, crystal structure, and cation distribution on the optical and magnetic properties of the resulting materials. In particular, the UV-vis absorption data were used to estimate the optical bandgap, while photoluminescence (PL) measurements supported the construction of energy band structure diagrams. These spectral results also allowed us to assess the photothermal response of the samples using theoretical parameters derived from Mie theory. Beyond the novelty of the green synthesis approach, this work reveals notable differences in the magnetic behavior between CoFe2O4 and MnFe2O4 NPs. While the magnetism in CoFe2O4 is predominantly governed by surface spin contributions, the magnetic properties of MnFe2O4 are better described by a modified superparamagnetic interaction model.
Importantly, the synthesis process described here does not require external catalysts, is performed at ambient temperature and atmospheric pressure, and relies entirely on non-toxic, plant-derived materials—highlighting its potential for eco-friendly and scalable production of functional ferrite nanomaterials. Unlike previous reports that use either microwave heating or plant extracts separately, this study uniquely combines both, with coconut-coir extract serving as the biogenic complexing agent—an approach not yet reported in the literature for transition-metal ferrites.
In this study, samples were fabricated at Vellore Institute of Technology, India. The X-ray diffraction (XRD), Raman, and PL measurements were conducted at the University of Science and Education, Danang University. The SEM and magnetic measurements were performed at the Institute of Materials Science, Vietnam Academy of Science and Technology, and UV-vis measurements were conducted at Hanoi National University of Education.
Furthermore, the reproducibility of the synthesis protocol was evaluated by repeating the microwave-assisted synthesis under the same fixed conditions (power, time, and precursor concentration). The resulting XRD patterns showed negligible variation across batches, confirming the robustness of the protocol for both CoFe2O4 and MnFe2O4 ferrites. Although the in situ monitoring of pH or temperature was not conducted, the consistent phase formation across multiple experiments indicates that the reaction proceeds reliably once the optimal conditions are established. This observation is consistent with prior reports that employed fixed microwave parameters to achieve reproducible spinel ferrite phases without the need for kinetic monitoring.39–41
To obtain precise crystallographic information, Rietveld refinement was performed using the FullProf software suite. The fitted profiles (red lines) exhibit excellent agreement with the experimental data (black dots), as shown in Fig. 1(a and b). The refinement quality was quantitatively assessed using the goodness-of-fit parameter (χ2), calculated from the final weighted profile (Rwp) and expected R-factor (RExp). In general, if χ2 lies within the reliability criterion of 1 ≤ χ2 < 2 or Rwp is less than 10%, refinement can be considered to be of a high degree of refinement accuracy.42
It is worth noting that the reduced χ2 value obtained from the Rietveld refinement is relatively low (see Table 1), approximately 0.035. Although values in the range 1 < χ2 < 2 are often considered typical, significantly lower values may occur when the dataset has very low statistical variance—particularly in cases of high-quality experimental data, low background noise, and a large number of refined points. As discussed by Brian H. Toby,42 such conditions can reduce the denominator of the χ2 equation (i.e., the variance of the intensity), thereby lowering the overall value without implying overfitting or an artificially perfect model. This interpretation is consistent with our results: the difference plot (Yobs. − Ycalc.) shown in Fig. 1 exhibits random, non-systematic deviations around zero, further confirming the validity of the model. Additionally, the refinement achieved Rwp values below 10%, which—together with the low χ2—provides strong evidence for the reliability and consistency of the extracted structural parameters within the spinel phase.
Sample | Cation distribution | aexp (Å) | V (Å3) | Rp (%) | Rwp (%) | χ2 | D (nm) | ε (10−3) | d (nm) | |
---|---|---|---|---|---|---|---|---|---|---|
A site | B site | |||||||||
CoFe2O4 | Co0.6582+Fe0.3423+ | Co0.3422+Fe1.6583+ | 8.4224 | 597.45 | 5.94 | 7.54 | 0.035 | 15.74 | 0.52 | 17.21 |
MnFe2O4 | Mn0.6702+Fe0.3303+ | Mn0.3302+Fe1.6703+ | 8.4782 | 609.41 | 8.06 | 10.3 | 0.029 | 29.51 | 2.48 | 29.83 |
The experimentally determined lattice parameters (aexp) and unit cell volumes are listed in Table 1. These values are consistent with those reported in prior studies on spinel ferrite systems.4,6,43,44 Notably, the lattice parameter of MnFe2O4 was found to be larger than that of CoFe2O4. This difference can be attributed to the larger ionic radii of Mn2+ ions compared to Co2+ ions at both tetrahedral (A) and octahedral (B) sites (0.66 Å vs. 0.58 Å at A sites; 0.83 Å vs. 0.745 Å at B sites),45 resulting in a greater lattice expansion for MnFe2O4. In addition to lattice constants, the cation distribution in the spinel lattice was estimated from Rietveld refinement using a crystallographic model constrained by the Fdm space group symmetry. Although conventional XRD lacks the sensitivity to unambiguously distinguish between transition metal cations with similar atomic numbers (e.g., Co2+, Mn2+, Fe3+), the refinement provides semi-quantitative insight into site occupancies when constrained by physically meaningful structural models. The obtained inversion parameters for CoFe2O4 and MnFe2O4 are consistent with literature values reported for materials synthesized under comparable conditions.46 The reliability of the refinement is supported by low Rwp and χ2 values, as well as by well-distributed residual plots (Fig. 1), indicating good agreement between the experimental and calculated diffraction profiles.
While more element-specific techniques such as Mössbauer spectroscopy or X-ray absorption spectroscopy (XAS) are typically required for a definitive determination of site occupancies, the present refinement yields values that are in good agreement with those obtained from such methods in previous studies.47–49 Accordingly, the cation distributions discussed herein should be regarded as refined estimations based on structural constraints rather than absolute determinations. The refined site occupancies and atomic coordinates are summarized in Table 1, and the corresponding crystallographic models are illustrated in Fig. 1(c and d). For CoFe2O4, the tetrahedral 8a site was occupied by 34% Co2+ and 66% Fe3+, while the octahedral 16d site was composed of 17% Co2+ and 83% Fe3+. In the case of MnFe2O4, the 8a site was occupied by 33% Mn2+ and 67% Fe3+, whereas the 16d site was composed of 32% Mn2+ and 68% Fe3+. In both samples, the oxygen anions fully occupied the 32e site. The observed distribution of M2+ and Fe3+ ions across both tetrahedral and octahedral sites confirms the formation of mixed spinel structures in CoFe2O4 and MnFe2O4 nanoparticles. This cationic disorder is known to significantly influence the magnetic and optical properties of spinel ferrites and is consistent with the results reported in the literature.
Based on the Rietveld refinement results, the full width at half maximum (FWHM) values for all diffraction peaks in the XRD patterns of CoFe2O4 and MnFe2O4 nanoparticles were determined and subsequently used to estimate the average crystallite size and lattice strain through Williamson–Hall (W–H) analysis. The W–H equation is expressed as:
![]() | (1) |
Fig. 2 illustrates the W–H plots of βhklcos
θ versus 4ε
sin
θ for both CoFe2O4 and MnFe2O4 samples. The fitted linear trends follow eqn (1), with the slope corresponding to the lattice strain and the Y-intercept (at sin
θ = 0) was used to calculate the average crystallite size. As summarized in Table 1, the estimated average crystallite sizes were 15.7 nm for CoFe2O4 and 29.5 nm for MnFe2O4. Interestingly, the lattice strain was found to increase from 0.52 × 10−3 in CoFe2O4 to 2.48 × 10−3 in MnFe2O4, despite the concurrent increase in crystallite size. This behavior contrasts with the typical inverse relationship between crystallite size and lattice strain, where smaller crystallites tend to exhibit higher strain due to enhanced lattice distortion at reduced dimensions.50 In this case, however, the trend reversal is attributed to the larger ionic radius of Mn2+ compared to Co2+, which leads to an expansion of lattice constant and unit cell volume, thereby increasing internal strain within the crystal lattice.51,52
An increase in crystallite size was generally accompanied by an increase in particle size, as revealed by the field-emission scanning electron microscopy (FE-SEM) images presented in Fig. 3(a and c). In these images, the MFe2O4 nanoparticles were observed to form nearly spherical morphologies with relatively uniform sizes, although some extent of agglomeration was also evident. Particle size distribution histograms, constructed using ImageJ software and shown in the inset of Fig. 3(a and c), were employed to statistically determine the average particle diameters (d) of the samples.
As summarized in Table 1, the mean particle size was found to increase from 17.21 nm for CoFe2O4 to 29.83 nm for MnFe2O4. This difference may be attributed to the variation in ionic radii between Mn2+ and Co2+ ions, which are known to affect the crystallization dynamics and grain growth rates during synthesis. The grain sizes estimated from SEM analysis were found to be in close agreement with the crystallite sizes obtained from X-ray diffraction (XRD) using the Williamson–Hall method, with minimal deviation. This consistency suggests that the particles are likely to be monocrystalline or consist of highly crystalline domains. Furthermore, no sub-grain boundaries or distinct contrast variations were observed within individual particles in the SEM micrographs, which would otherwise indicate internal polycrystallinity. Although the possibility of inter-crystallite agglomeration cannot be entirely excluded, the absence of inter-particle fringes, combined with the close correspondence to the XRD-derived sizes, supports the assumption that the grains are structurally coherent and predominantly single-domain. Nevertheless, it is acknowledged that more conclusive identification of crystallinity would require high-resolution transmission electron microscopy (HRTEM) or selected area electron diffraction (SAED), which were beyond the scope of the current study.
As shown in the insets of Fig. 3(b and d), the atomic percentages of metal cations and anions, determined from the EDX spectra averaged over three different positions for each sample, were found to be close to the ideal cation-to-anion stoichiometric ratio of 3:
4 for spinel ferrites. This observation further confirmed that the metal precursors were effectively incorporated into the ferrite lattice, forming compositionally homogeneous MFe2O4 nanoparticles without detectable residual or impurity elements.
In addition, the lattice strain (ε), determined via Williamson–Hall analysis, was found to increase from 0.52 × 10−3 for CoFe2O4 to 2.48 × 10−3 for MnFe2O4. This trend indicates that the incorporation of Mn ions not only expanded the unit cell but also introduced greater internal stress, which could further influence both the microstructural evolution and the functional properties of the materials.
Fig. 4 displays the Raman spectra of CoFe2O4 and MnFe2O4 nanoparticles, recorded in the spectral range of 70–1000 cm−1. Peak assignment was performed based on literature ref. 55–59. The A1g mode was observed at 633 cm−1 for CoFe2O4 and 672 cm−1 for MnFe2O4, attributed to the symmetric stretching vibration of oxygen atoms in the AO4 tetrahedra. The T2g(3), T2g(2), and T2g(1) modes, corresponding to various symmetric and asymmetric bending vibrations of the metal–oxygen bonds, were observed at approximately 480, 304, and 125 cm−1 for CoFe2O4 and at 459, 308, and 125 cm−1 for MnFe2O4, respectively. Most of the Raman bands appeared asymmetric, indicating overlapping vibrational contributions from multiple lattice sites. To resolve these overlapping features, spectral deconvolution was performed. The fitting parameters obtained are summarized in Table 2. The presence of multiple sub-peaks, often represented as doublets,16,57–59 further supports the existence of a mixed or partially inverted spinel structure in both samples, consistent with previous reports. This splitting of Raman modes can be attributed to variations in cation occupancy between tetrahedral (A) and octahedral (B) sites, leading to local symmetry distortions and non-equivalent vibrational environments. Such structural disorder, arising from partial inversion and site-sharing between Co2+ and Fe3+ ions, directly influences the observed optical phonon modes and results in the complex spectral features recorded. Therefore, the Raman spectra serve as a sensitive probe of the cation distribution and local structural symmetry in the spinel lattice. These findings are consistent with the cation distribution obtained from Rietveld refinement, demonstrating the complementary nature of Raman spectroscopy and crystallographic analysis in elucidating the structural properties of ferrite nanoparticles.
Sample | T2g(1) (cm−1) | T2g(1) (cm−1) | T2g(2) (cm−1) | T2g(2) (cm−1) | T2g(3) (cm−1) | T2g(3) (cm−1) | A1g (cm−1) | A1g (cm−1) |
---|---|---|---|---|---|---|---|---|
CoFe2O4 | 117.25(0) | 155.33(4) | 287.98(5) | 310.09(3) | 482.62(1) | — | 617.33(8) | 667.54(7) |
MnFe2O4 | 80.86(2) | 109.78(5) | 337.33(4) | 413.11(9) | 453.70(9) | 595.30(1) | 619.36(7) | 666.88(2) |
The optical bandgap energy (Eg) of the samples was estimated using the Tauc relation:62
αhν = A(hν − Eg)n, | (2) |
In the case of spinel ferrites, such as CoFe2O4 and MnFe2O4, numerous theoretical and experimental studies have confirmed that they exhibit direct allowed transitions.63–65 Therefore, the Tauc plots in Fig. 6(a and b) were constructed using n = 1/2, corresponding to a direct allowed transition, and the optical bandgap was extracted from the linear extrapolation of (αhν)2 versus hν.
The Eg values were determined by extrapolating the linear portion of the plots to the photon energy axis (hν), where (αhν)2 = 0. The estimated bandgap energies were 2.66 eV for CoFe2O4 and 2.64 eV for MnFe2O4, in close agreement with previously reported values for similar nanocrystalline ferrites.6,64,65 These results confirm that the optical properties of the synthesized nanoparticles are consistent with those expected for spinel-type MFe2O4 materials.
The presence of structural defects in a material introduces localized states within the bandgap, which can significantly alter the electronic band structure by disrupting the long-range periodic potential. These defects impede the direct excitation of electrons into the conduction band and give rise to an exponential absorption tail in the sub-bandgap region of the UV-vis spectrum, commonly known as the “Urbach tail”.66 The characteristic energy associated with this tail is termed the Urbach energy (Eu), which provides a quantitative measure of the degree of structural disorder in a material.67
The Urbach energy can be derived from the absorption coefficient (α) using the relation:
![]() | (3) |
Fig. 6(c and d) present the plots of ln(α) versus photon energy (E) for CoFe2O4 and MnFe2O4 NPs, respectively. The Eu values, determined from the inverse slopes of the linear fits in the low-energy absorption tail, were found to be 0.83 eV for CoFe2O4 and 0.78 eV for MnFe2O4. Although the MnFe2O4 sample exhibited a larger grain size, its Eu value was slightly lower than that of CoFe2O4. This observation suggests that the disorder in these spinel ferrites is not dominated solely by grain boundary density but also significantly influenced by oxygen vacancies, surface disorder, and cationic redistribution at the tetrahedral and octahedral sites.
To further explore the origin of this optical disorder, we examined its correlation with microstructural strain. The lattice strain, estimated via the Williamson–Hall method, increased markedly from 0.52 × 10−3 in CoFe2O4 to 2.48 × 10−3 in MnFe2O4. Despite this significant rise in strain, the corresponding decrease in Eu implies that lattice strain alone does not dominate the density of localized states in the band tail. This inverse trend reinforces the view that Eu is not governed solely by macroscopic strain but rather by a combination of structural imperfections—including cation disorder and local bonding distortions—that are not fully captured by average strain measurements. These moderate Eu values in both systems reflect a notable degree of structural disorder, consistent with previous reports on spinel ferrites.68–71
Photoluminescence (PL) spectroscopy is a sensitive and non-destructive technique for investigating electron–hole recombination dynamics and evaluating the influence of surface and structural defects on the energy-band structure of ferrite nanomaterials.72 In nanoscale systems, the high surface-to-volume ratio results in a large density of defect states—such as oxygen vacancies, surface dangling bonds, and cationic disorder—that serve as recombination centers. These defect states can significantly broaden the PL emission peaks and reduce the overall intensity due to enhanced non-radiative recombination pathways, where excited carriers relax via thermal dissipation rather than photon emission.69
Fig. 7 illustrates the room-temperature PL spectra of CoFe2O4 and MnFe2O4 NPs in the 400–900 nm wavelength range. Both samples exhibited emission bands centered at approximately 458, 692, 753/758, and 871/882 nm, indicating multiple defect-related recombination pathways.
The PL spectra of both samples exhibit multiple emission bands arising from defect-related recombination pathways. The broad red emission band centered around 692 nm, observed in both CoFe2O4 and MnFe2O4, along with near-infrared peaks at ∼871 nm and ∼882 nm, respectively, can be attributed to electron transitions from deep-level defect states—such as oxygen vacancies or antisite defects—to the valence band. The higher-wavelength emission in MnFe2O4 is likely related to its narrower bandgap and greater degree of lattice disorder, which aligns with its larger Urbach energy. Additionally, the blue emission at ∼458 nm and the red band near 753/758 nm are assigned to transitions involving shallow trap states and surface defects, in agreement with the typical PL behavior reported for spinel ferrites. These peak assignments are consistent with previous reports on CoFe2O4 and MnFe2O4 nanoparticles,8,13,16,17,25,68,69 which associate the observed PL emissions with intrinsic defects, such as oxygen vacancies, Fe2+/Fe3+ transitions, and antisite disorder in the spinel lattice. While the PL intensity is generally expected to decrease with increasing defect density due to enhanced non-radiative recombination, some studies—such as that by Tongay et al.73 on MoS2, MoSe2, and WSe2—have reported the opposite trend. This highlights the complex interplay between radiative and non-radiative pathways in nanomaterials and reinforces the importance of interpreting PL results in conjunction with structural and optical analyses.
Therefore, to draw meaningful interpretations regarding electronic transitions and defect states in MFe2O4 (M = Co, Mn) NPs, the PL spectra must be considered in the broader context of the band structures of samples, which are derived from both UV-vis, PL and Urbach analyses.
The interpretation of the energy band structure and crystal field splitting energies (Δcf,O and Δtf,O) in spinel ferrites containing 3d transition metal ions (e.g., Fe3+, Co2+) remains a complex issue due to variability in synthesis methods and measurement techniques.74 Consequently, reported values of crystal field splitting energies often show discrepancies across studies. For instance, Fontijn et al.75 determined the octahedral crystal field splitting energy (Δcf,O) for Fe3+ ions to be approximately 1.3 eV and the tetrahedral splitting energy (Δtf,O) to be around 0.8 eV. Camphausen et al.76 reported Δcf,O values between 1.7–2.0 eV and Δtf,O values from 0.86 to 1.17 eV. Boxall et al.,77 through photoelectrophoretic measurements, found Δcf,O = 2.2 eV and Δtf,O = 0.8 eV. In contrast, Alvarado et al.,78 using spin polarization and energy distribution techniques, obtained Δcf,O = 1.75 eV and Δtf,O = 1.55 eV. For Co2+ ions, Fantechi et al.79 reported Δcf,O = 1.22 eV and Δtf,O = 1.77 eV via Kerr photo-magnetic spectroscopy, while Papalardo et al.80 and Kim et al.74 found Δtf,O values of 0.83 eV and 2.2 eV, respectively. Overall, the literature suggests that Δcf,O for 3d ions generally lies in the range of 1.2–2.2 eV, whereas Δtf,O typically falls between 0.8 and 2.2 eV.74
Based on these findings and the results obtained in this work from UV-vis absorption, PL spectroscopy, and Urbach energy analyses, a comprehensive electronic band structure diagram was proposed for MFe2O4 (M = Co, Mn) nanoparticles (Fig. 8). The energy gap between the O(2p) valence band and the M(4s) conduction band is typically in the range of 4–6 eV for spinel ferrites.76,81,82 Within this range, the 3d orbitals of transition metal ions split into distinct energy levels under crystal field interactions: for octahedral (B) sites, the splitting between t2g and eg levels is defined by Δcf,O ≈ 1.75 eV, and for tetrahedral (A) sites, the splitting between e and t2 levels is characterized by Δtf,O ≈ 0.8 eV.74 Additional splitting (∼0.9 eV) between the O(2p) valence band and the t2g/e levels further modulates the band structure, as observed in various experimental reports.75,77 The bandgap energy extracted from Tauc plots (Fig. 6(a and b)) was 2.66 eV for CoFe2O4 and 2.64 eV for MnFe2O4, corresponding well to the transition from the O(2p) level to the eg level in the B crystal field. Moreover, the onset of absorption observed in the Urbach plots (Fig. 6c and d) appears to correspond closely with Δcf,O (∼1.75 eV) or transitions from the O(2p) band to the t2 level of the A field (∼1.8 eV). These values support the involvement of localized sub-bandgap states in the optical absorption process.
The presence of oxygen vacancies and lattice distortions, as previously evidenced by Williamson–Hall analysis and FTIR spectral features, suggests the existence of localised defect states within the bandgap of both CoFe2O4 and MnFe2O4. These structural imperfections are expected to play a significant role in charge carrier dynamics. Under excitation at 310 nm (∼4.0 eV), electrons are promoted to high-energy states such as the eg orbital at the B-site cations, followed by radiative or non-radiative relaxation. Initially, both samples exhibited a broad PL emission peak centred at 458 nm (2.70 eV), which coincides with the optical bandgap obtained from UV-vis absorption, suggesting a possible direct transition from the conduction band to the O(2p) valence band.
However, Gaussian deconvolution of the PL spectra revealed that this apparent single peak comprises two components: 443 nm (2.80 eV) and 485 nm (2.56 eV) in CoFe2O4, and 443 nm (2.80 eV) and 509 nm (2.44 eV) in MnFe2O4. The higher-energy emission at 2.80 eV may be attributed to excitonic transitions or recombination via shallow donor states, while the lower-energy bands are ascribed to deep-level defect emissions, possibly involving oxygen vacancies or Fe3+/Co2+ trap states. These observations indicate that, although near-band-edge transitions are present, defect-assisted recombination processes dominate the photoluminescence behaviour of both ferrite systems. The red emission at 692 nm (1.79 eV) can be attributed either to t2g → eg transitions in the B site or to electron recombination from the e orbital in the A site to the O(2p) valence band. The weaker red emission at 753 nm (1.65 eV) likely originates from e → t2 transitions within the A site. Lastly, the near-infrared emissions at 871 nm (1.42 eV) for CoFe2O4 and 882 nm (1.40 eV) for MnFe2O4 are attributed to deep trap states associated with oxygen vacancies and cation disorder in the A site lattice environment. These energy transitions are coherent with the proposed schematic (Fig. 8), demonstrating the strong correlation between electronic structure, defect states, and the optical behavior of spinel ferrite nanoparticles.
The combined structural and spectroscopic analyses of CoFe2O4 and MnFe2O4 nanoparticles reveal a coherent picture of how cation distribution, crystallite size, and surface defect states modulate their optical and electronic properties. The constructed energy band schematic captures the essential transitions involving the O(2p) → 3d crystal field levels, explaining both the band-edge absorption and PL emission behaviors. Notably, blue and red emissions were matched with recombinations involving eg → O(2p) and t2g → eg transitions, while near-infrared emissions were attributed to electron traps associated with oxygen vacancies. Together, these findings underscore the importance of controlled cation distribution and structural order in tailoring the optical and electronic performance of spinel ferrite nanoparticles for applications in magneto-optical, photocatalytic, and biomedical technologies.
The proposed energy band diagrams for MFe2O4 nanoparticles were constructed based on the experimental UV-vis absorption and PL data. The band edge positions and defect-related emission features were interpreted in conjunction with literature reports on similar spinel ferrites,69,75,77 facilitating a qualitative but coherent description of electronic transitions and energy transfer mechanisms. Although direct probing of the electronic structure through techniques such as X-ray photoelectron spectroscopy (XPS), ultraviolet photoelectron spectroscopy (UPS), or density functional theory (DFT) calculations was not performed, theoretical data from previous studies were incorporated to enhance the credibility of the proposed models.
For CoFe2O4, the direct optical band gap obtained from the UV-vis spectra (2.66 eV) was found to be consistent with previous experimental reports (∼2.7 eV), and lies within the range of DFT-calculated values (1.0–2.3 eV), depending on the functional employed.83 This transition has generally been attributed to the O(2p) → Co(3d) charge transfer at the band edges.83
In the case of MnFe2O4, a direct band gap of 2.64 eV was determined, which agrees with prior optical measurements on nanoscale MnFe2O4 (∼2.6 eV).84 In contrast, standard DFT approaches have been shown to underestimate the band gap, yielding values as low as 0.4 eV.85 More accurate GGA + U calculations have been reported to yield values around 1.3 eV,86 and other studies have suggested values ranging from 1.3 to 1.6 eV depending on the degree of cation ordering and particle size.84,87 The larger experimental band gap observed in this study may be attributed to quantum confinement and surface effects, which have been well documented in spinel ferrites.88,89
Overall, the experimentally determined band gaps were found to align well with both theoretical predictions and previously published results. The PL emission bands, which were associated with deep-level defect states, were observed to correlate with features in the calculated density of states reported in the literature. Accordingly, even in the absence of direct DFT computations, the proposed energy band diagrams are supported by a consistent and converging body of prior theoretical and experimental evidence.
Mie theory offers a robust framework for quantitatively assessing the photothermal performance of nanomaterials by analyzing their light–matter interactions. Specifically, the wavelength dependence of three key optical parameters—absorption efficiency (Qabs), extinction efficiency (Qext), and scattering efficiency (Qsca)—was investigated for the CoFe2O4 and MnFe2O4 NPs. These parameters are defined by the following equations:69,81
![]() | (4a) |
![]() | (4b) |
Qabs = Qext − Qsca, | (4c) |
Numerical computations of these parameters as functions of wavelength were carried out using the “MiePlot” simulation software developed by Philip Laven.90 As shown in Fig. 9, CoFe2O4 NPs exhibited a dominant absorption contribution to extinction efficiency in the longer wavelength region (>650 nm), suggesting effective conversion of light into heat in this spectral range. In contrast, MnFe2O4 NPs showed negligible absorption contribution across the entire studied wavelength range, implying low photothermal conversion efficiency.
![]() | ||
Fig. 9 The wavelength dependence of extinction (Qext), absorption (Qabs), and scattering (Qsca) efficiencies for (a) CoFe2O4 NPs and (b) MnFe2O4 NPs. The parameters are determined from the Mie theory using eqn 4(a)–(c). |
The ratio of absorption to extinction efficiency (Qabs/Qext)—a key indicator of photothermal conversion capability—was substantially higher for CoFe2O4 than for MnFe2O4. In the context of Mie theory, the absorption efficiency (Qabs) quantifies the fraction of incident light that is absorbed by the nanoparticle and subsequently converted into other forms of energy, such as heat. Meanwhile, scattering efficiency (Qsca) represents the portion of light that is elastically scattered by the particle without absorption. The total extinction efficiency (Qext) is the sum of these two contributions, reflecting the overall interaction of light with the particle. A higher Qabs/Qext ratio indicates that a greater fraction of the light interacting with the particle is being absorbed rather than scattered, which is a desirable feature for efficient photothermal conversion. This observation aligns with the smaller particle size and higher specific surface area of CoFe2O4, both of which are known to enhance light absorption and heat generation.69 Additionally, the scattering efficiency was significantly higher for MnFe2O4, which can be attributed to its larger particle size and is consistent with classical Mie scattering predictions. While the current analysis relies on Mie theory to estimate photothermal behavior, it offers valuable insight into the interplay between nanoparticle size, composition, and optical response. We acknowledge that the absence of direct experimental photothermal measurements is a limitation of the present work. Nevertheless, theoretical modeling serves as a predictive tool for evaluating wavelength- and size-dependent photothermal efficiency. Future studies will aim to experimentally validate these findings and benchmark the photothermal performance of CoFe2O4 and MnFe2O4 nanoparticles against established photothermal agents such as gold, silver, or copper sulfide nanostructures.
![]() | (5) |
![]() | (6) |
![]() | ||
Fig. 10 The typical hysteresis loop of CoFe2O4 NPs (a) and MnFe2O4 NPs (b) for different temperatures. The insets are an enlargement of the magnetization at low field. |
Sample | Temp. (K) | Ms (emu g−1) | Ms (μB) | Mr (emu g−1) | Hc (Oe) | Mr/Ms | b × 106 (Oe2) | K1 × 106 (erg cm−3) |
---|---|---|---|---|---|---|---|---|
CoFe2O4 | 55 | 70.10 | 2.94 | 40.32 | 4333.43 | 0.58 | 17.42 | 5.53 |
100 | 67.96 | 2.85 | 35.15 | 2809.84 | 0.52 | 12.92 | 4.62 | |
150 | 65.67 | 2.76 | 28.12 | 1972.03 | 0.43 | 10.79 | 4.08 | |
200 | 61.85 | 2.60 | 20.01 | 996.65 | 0.32 | 8.80 | 3.47 | |
250 | 57.05 | 2.40 | 12.5 | 522.20 | 0.22 | 7.27 | 2.91 | |
300 | 51.60 | 2.18 | 7.24 | 146.01 | 0.14 | 6.14 | 2.42 | |
MnFe2O4 | 55 | 49.90 | 2.06 | 5.87 | 126.65 | 0.12 | 5.00 | 2.03 |
100 | 46.86 | 1.93 | 2.73 | 51.95 | 0.06 | 4.79 | 1.87 | |
150 | 42.81 | 1.77 | 1.64 | 28.33 | 0.04 | 4.29 | 1.62 | |
200 | 38.57 | 1.59 | 0.636 | 11.84 | 0.02 | 3.92 | 1.39 | |
250 | 34.36 | 1.42 | 0.24 | 5.12 | 0.01 | 3.52 | 1.17 | |
300 | 30.25 | 1.25 | 0.85 | 15.10 | 0.03 | 3.18 | 0.98 |
The fitting of magnetization M(H) data to the Law of Approach to Saturation (LAS) model reached convergence only when the sum of squared residuals (SSR) was minimized. This condition was satisfied by selectively adjusting the parameters b and Ms within the high-field region (H > 1.0 T), where the approach to magnetic saturation becomes more pronounced. Fig. 11 displays both the experimental M(H) curves and the theoretical fits derived from the LAS model for CoFe2O4 and MnFe2O4 samples. The optimized values of the fitting parameters b and Ms at different temperatures are summarized in Table 3.
Using eqn (6), the magnetocrystalline anisotropy constant K1 was calculated by substituting the obtained values of b and Ms. The resulting K1 values, corresponding to the range of studied temperatures, are also listed in Table 3. Notably, at all temperatures, the K1 value for CoFe2O4 was significantly higher than that for MnFe2O4, consistent with the known strong anisotropy of cobalt ferrites. Furthermore, the variation of K1 with temperature for both samples followed the expected trends, reflecting the thermal dependence of magnetic anisotropy in spinel ferrites. These findings confirm the strong dependence of magnetic anisotropy on the type of transition metal cation occupying the octahedral (B) site in the spinel structure. The higher K1 values observed for CoFe2O4 across all temperatures reflect the dominant contribution of Co2+ ions, which exhibit significant spin–orbit coupling and strong magnetocrystalline anisotropy. In contrast, the relatively lower K1 values for MnFe2O4 are attributed to the weaker anisotropy of Mn2+ ions in similar coordination environments. Additionally, the saturation magnetization Ms of CoFe2O4 was consistently higher than that of MnFe2O4, which can be explained by differences in cation distribution and magnetic interactions between the A- and B-site ions. The partially inverse spinel structure of CoFe2O4 likely facilitates enhanced superexchange interactions between Fe3+ ions at the tetrahedral (A) and octahedral (B) sites, thereby increasing Ms. Meanwhile, MnFe2O4 exhibits lower Ms, possibly due to spin canting or surface disorder effects that are more pronounced in cobalt-based ferrites. The temperature dependence of K1 for both CoFe2O4 and MnFe2O4 was consistent with previously reported values in the literature.93–95 The observed increase in K1 can be attributed to a combination of (i) surface anisotropy, (ii) dipolar interactions, and (iii) the nanocrystalline nature of the particles with small grain sizes.96 Furthermore, both samples exhibited similar magnetic trends with temperature: saturation magnetization and coercivity increased as the temperature decreased. The monotonic increase in Ms is primarily due to the enhancement of exchange interactions between neighboring magnetic moments at lower temperatures, consistent with the suppression of spin-wave excitations. This behavior follows Bloch's T3/2 law :92,97
Ms(T) = Ms(0)(1 − AT3/2), | (7) |
To investigate the role of surface anisotropy, dipolar interactions, and nanocrystalline grain effects on the magnetic behavior of CoFe2O4 and MnFe2O4 nanoparticles, the squareness ratio (R = Mr/Ms) was determined at various temperatures and is summarized in Table 3. According to the Stoner–Wohlfarth (SW) model, which assumes an ensemble of randomly oriented, non-interacting, single-domain particles, the ideal squareness ratio is R = 0.5. For uniaxially aligned particles without thermal agitation, the model predicts a higher ratio of R = 0.832. It is important to note that both values neglect thermal fluctuation effects. As shown in Table 3, the CoFe2O4 sample exhibits an R value of approximately 0.6 at 55 K, which decreases to 0.14 at 300 K. This trend suggests that the CoFe2O4 nanoparticles possess cubic magnetocrystalline anisotropy at low temperatures, consistent with previous reports on monodisperse CoFe2O4 NPs,100 single-domain CoFe2O4 systems,101 and CoFe2O4/CTAB nanocomposites.102 In contrast, the MnFe2O4 sample exhibits a very low squareness ratio (R < 0.1) across all temperatures studied. The low squareness ratio observed in MnFe2O4 can be attributed to several factors, including its relatively larger average particle size, broader size distribution, and possible surface spin disorder. These structural characteristics are known to enhance magnetostatic (dipolar) interactions and magnetic relaxation phenomena, which collectively suppress remanent magnetization. The notably small squareness value also suggests the presence of strong interparticle magnetic interactions, thereby violating the assumptions of the Stoner–Wohlfarth model, which is valid only for non-interacting, single-domain particles with uniaxial anisotropy.103 This behavior is consistent with the soft magnetic nature and possible superspin glass (SSG) tendencies observed in the MnFe2O4 sample. For both samples, the observed R values are lower than the theoretical limit of 0.5, which is frequently observed in magnetic nanostructures and generally attributed to surface spin disorder and frustration effects. Consequently, while the SW model can still provide qualitative insights into the behavior of CoFe2O4, it appears unsuitable for describing the MnFe2O4 system. The relatively high squareness ratio (up to 0.6 at 55 K) for CoFe2O4 further supports the presence of significant surface anisotropy, influencing its magnetocrystalline anisotropy constant K1. This is consistent with the known high cubic anisotropy of CoFe2O4, which has a theoretical R ≈ 0.832 and K1 > 0. The maximum value of K1 observed in this study was 5.37 × 106 erg cm−3 at 55 K—substantially higher than that reported for bulk CoFe2O4 (typically in the range of 1.8–3.0 × 106 erg cm−3).104 Such enhanced values can be attributed to increased surface contributions, reduced grain size, and strong dipolar interactions. Conversely, lower K1 values are generally indicative of weaker intrinsic anisotropy and enhanced thermal agitation at higher temperatures.
Additional evidence for the presence of magnetic interactions in both nanoparticle systems can be derived from their temperature-dependent coercivity behavior. According to Kneller's law,105 the coercive field Hc(T) follows a square-root temperature dependence in single-domain systems:
Hc(T) = Hc(0)[1 − (T/TB)1/2], | (8) |
To gain further insight into the magnetic transition behavior and confirm the thermal stability of the nanoparticles, temperature-dependent magnetization measurements were conducted in the range of 55–350 K. Fig. 14 shows the temperature-dependent magnetization curves measured under zero-field-cooled (ZFC) and field-cooled (FC) conditions and an applied field of 10 Oe for the CoFe2O4 and MnFe2O4 samples, respectively. For MnFe2O4, both FC and ZFC curves exhibit a gradual decrease with decreasing temperature. The FC curve decreases slowly, while the ZFC curve shows a more pronounced drop at low temperatures, which is commonly associated with surface spin freezing. This behavior has been previously reported107 and is often attributed to an SSG state, induced by strong interparticle dipolar interactions. Notably, the FC and ZFC curves nearly converge above 325 K, suggesting the onset of reversible superparamagnetic behavior beyond the blocking temperature (TB), which is not reached within the measured range (TB > 350 K). This trend is consistent with the very low coercivity observed at room temperature (Hc ∼15 Oe at 300 K). In contrast, the CoFe2O4 sample exhibits clear separation between the ZFC and FC curves up to 310 K, with a distinct TB observed near 276 K. Below this point, the FC magnetization increases steadily with decreasing temperature, while the ZFC curve exhibits a peak and then decreases due to magnetic relaxation. The larger separation between the FC and ZFC curves reflects stronger magnetic anisotropy and more stable blocked states. This correlates well with the higher coercivity of CoFe2O4 (Hc ∼522 Oe at 250 K) and suggests minimal influence of dipolar interactions, as reported in similar systems, such as Fe3O4 nanocrystals108 and others.109
It is important to distinguish between superparamagnetic (SPM) and SSG behaviors, both of which can manifest in nanoparticle systems depending on size distribution and interparticle interactions. In the SPM regime, each nanoparticle behaves like a single magnetic domain whose net magnetic moment can thermally fluctuate and align with the external field above the blocking temperature (TB), leading to reversible FC and ZFC magnetization curves. In contrast, the SSG state arises from strong dipolar interactions among nanoparticles, which lead to frustrated collective spin freezing at low temperatures, analogous to atomic spin glasses. This behavior is typically characterized by a bifurcation between the FC and ZFC curves that does not converge even at high temperatures, broad peaks or the absence of peaks in the ZFC curves, and low coercivity at room temperature. In our study, the observed magnetic features of MnFe2O4—such as the slow decay of FC magnetization, suppressed ZFC peak, and high-temperature bifurcation—are consistent with the SSG-like behavior. Meanwhile, CoFe2O4 exhibits typical SPM features with a clear TB and larger coercivity due to stronger magnetic anisotropy.
The magnetic behavior of an ideal, non-interacting superparamagnetic system is characterized by the collapse of all M/Ms curves versus H/T curves onto a single universal curve. These normalized magnetization curves should follow the classical Langevin function:
![]() | (9a) |
However, in systems where dipolar interactions are present, deviations from this ideal behavior are often observed. To account for such interactions, Allia et al.110 proposed a modified Langevin function that incorporates an effective interaction temperature T*, which is related to the dipolar interaction energy εD by the expression:
εD = kBT*, | (9b) |
The modified expression becomes:
![]() | (9c) |
For the MnFe2O4 nanoparticles studied here, the experimental (M/Ms) versus H/T curves shown in Fig. 15(a) deviate from the universal behavior predicted by the classical Langevin model, indicating the presence of interparticle magnetic interactions. However, when the data are replotted as a function of H/Ms, the curves collapse onto a single master curve that follows the modified Langevin behavior, as illustrated in Fig. 15(b). This observation supports the existence of an interacting superparamagnetic regime.
To provide a quantitative measure of dipolar interactions, the experimental magnetization curves were fitted using the modified Langevin model. The fitting yielded an effective interaction temperature of approximately T* ≈ 120 K. From this value, the dipolar interaction energy was estimated as: εD = kBT* ≈ 1.034 × 10−2 eV. Although specific reports on dipolar interaction energies in MnFe2O4 are scarce, this value is comparable to those observed in magnetite (Fe3O4) nanoparticles dispersed in polymeric matrices, such as photoreticulated PEGDA-600.110 This supports the presence of moderate interparticle interactions in the studied MnFe2O4 system. Consequently, the observed low coercivity and small squareness ratio can be ascribed to the combined effects of reduced magnetic anisotropy and notable dipolar interactions, both of which inhibit the alignment and stability of magnetic moments under an applied field.
This journal is © The Royal Society of Chemistry 2025 |