Iwona
Rogalska
ab,
Bogumił
Cieniek
a,
Anna
Tomaszewska
c,
Magdalena
Kulpa-Greszta
c,
Piotr
Krzemiński
a,
Bartosz
Zarychta
d,
Ireneusz
Stefaniuk
*a and
Robert
Pązik
*c
aInstitute of Materials Engineering, College of Natural Sciences, University of Rzeszow, Pigonia 1, Rzeszow PL-5310, Poland. E-mail: istef@ur.edu.pl
bInternational Research Centre MagTop, Institute of Physics, Polish Academy of Sciences, Aleja Lotników 32/46, Warsaw PL-02668, Poland
cDepartment of Biotechnology, Institute of Biotechnology, College of Natural Sciences, University of Rzeszow, Pigonia 1, Rzeszow PL35-310, Poland. E-mail: rpazik@ur.edu.pl
dFaculty of Chemistry, University of Opole, Oleska 48, Opole PL-45052, Poland
First published on 16th December 2024
In this work, we are showing the results of the X- and Q-band electron magnetic resonance measurements of ultra-small Zn1−xMnxFe2O4 nanoparticles (ca. 8 nm) with a very narrow size distribution. The chosen synthetic route allows for precise structural modifications with a broad concentration range (x = 0, 0.2, 0.5, 0.8, 1). The crystal structure was evaluated by means of X-ray diffraction, while cell parameters were calculated using Rietveld refinement. EMR spectral studies indicated that the prepared nanoparticles were superparamagnetic. The linewidth of EMR signal for any ferrite material generally originates from two sources: (a) magnetic dipole–dipole interactions among particles and (b) interparticle superexchange interactions between magnetic ions through oxygen ions. Observed effects are more complex interactions than in pure zinc and manganese ferrites. As a result of the study, a relationship was observed between the composition of the material and the magnetic properties with striking antiferromagnetism and ferrimagnetism change. Hence, by structural modification of materials, the magnetic character (FM–AFM–FiM) can be controlled.
The most well known representative of the nanoferrite family is the magnetite Fe3O4 (FeFe2O4) where iron has a mixed Fe2+ and Fe3+ valency. However, unless it is surface-protected, magnetite is highly prone to oxidation into α-Fe2O3.10 This chemical transformation results in a drastic change in magnetic properties and can have a detrimental effect on certain bio-related applications due to a change of physical properties characteristic.11 To avoid that, one of the possibilities relies on the usage of more chemically stable AFe2O4 ferrites, for instance ZnFe2O4 or MnFe2O4. Both compounds are generally defined by the normal spinel structure.5,9 However, upon particle size reduction, some cationic disorder can be observed, especially at the nanoscale. The degree of inversion in ZnFe2O4 and MnFe2O4 was shown to depend quite strongly on the fabrication technique.6,7,12 What makes them particularly interesting is that they offer extraordinary particle size-dependent physicochemical properties that can be fine-tuned by chemical approaches.13–16 The range of applications is wide and includes electronic devices,17 telecommunications,18 hyperthermia,19 virus sensor detection,20 batteries,21 catalysis,22 contrast agents for magnetic imaging resonance (MRI),23,24 drug delivery platforms,25 gas sensing,26etc. Mixed zinc–manganese ferrites (Zn1−xMnxFe2O4) are of great interest, as their properties can be shaped by structural modification.16,25,27 A careful reader can find numerous reports on the synthetic routes towards ferrite compounds such as coprecipitation,28 hydrothermal,19 sol–gel,29 solid-state,30 rapid hot-injection,31 thermal decomposition,32etc.
Many antiferromagnetic materials have been reported to have a positive Curie–Weiss temperature, which is a sign of ferromagnetic interactions.33–36 This behaviour was explained by the coexistence of competing antiferromagnetic and ferromagnetic interactions. In fact, ZnFe2O4 is a model system for studying competing interactions. The normal (Zn)a(Fe2)bO4 is antiferromagnetic, but the exchange of only a few Zn–Fe cations between sites gives place to (Zn1-δFe)a[ZnFe2-δ]bO4. Recent evidence for such coexistence of ferrimagnetic clusters within an antiferromagnetic matrix37–39 allows for correlation of Zn–Fe exchange with the anomalous increase in positive Curie–Weiss temperature despite the antiferromagnetic character of sample.36 If the A ion is magnetic, AFe2O4 can be decomposed into two inequivalent magnetic sublattices, denoted as O (octahedral) and T (tetrahedral), respectively.40 The diamond-like T sublattice is not frustrated within the nearest neighbour (NN) interaction, whereas in the presence of antiferromagnetic (AFM) interactions within the sublattice, the pyrochlore-like O sublattice can host a large NN frustration.41 The ions within a sublattice often interact ferromagnetically (FM), and the ions between two sublattices interact antiferromagnetically.40,42,43 Some spinel compounds can cause the interaction of AFM between ions at O sites, which leads to geometric frustration.44,45 Consequently, the general magnetic behaviour depends on the strength of the exchange interactions between O–O (JOO), T–T (JTT) and O–T (JOT).44–46
All Fe3+ ions are coupled to each other via a superexchange pathway through A sites, and their interactions are very weak. The normal ZnFe2O4 spinel shows long-range antiferromagnetic ordering at TN = 9–11 K.47,48 Magnetization was found to increase with grain size reduction. This characteristic is generally associated with an increase in cation inversion and a decrease in grain size.43,49,50
Electron magnetic resonance spectroscopy (EMR) is a technique used to study chemical species with unpaired electrons. EMR spectroscopy also plays an important role in understanding organic and inorganic radicals, transition-metal complexes, and biomolecules. Among many experimental research methods used to study the magnetic properties of ferrite nanoparticles, EMR proves to be a very sensitive and provides information on local magnetic properties, the correlation between particle moments and the internal field distribution.51–55 The properties and range of EMR applications depend mainly on the frequency band used. The maximum geometrical size of the object that can be examined by the EMR method is directly proportional to the wavelength of the microwave. As the frequency of the microwave signal increases, the ability to detect closely located resonance lines increases. EMR is a universal technique that provides valuable information on a variety of paramagnetic systems. This is favoured by high measurement sensitivity, selectivity, the ability to measure a specific compound in the presence of many other unidentified compounds, relatively short measurement time, or uncomplicated sample preparation. The information extracted from EMR spectroscopy includes the identity and atomic oxidation state of atoms, properties of free radicals, the local environment around the unpaired electrons, and the hyperfine interaction of nuclei with unpaired electrons.
When Zn1−xMnxFe2O4 is subject to a strong magnetic field, an induced magnetic potential energy splits the spin state of unpaired electrons (Zeeman effect) accordingly. Afterwards, simultaneously applied microwave frequencies are absorbed by the unpaired electrons, which can cause the transition from one spin state to another. This mechanism can be affected by several interactions such as the hyperfine interaction and the zero-field interaction, etc. We focused on investigating the influence of the Zn/Mn content on the microstructure, magnetic properties, and EMR characteristics of Mn–Zn ferrite nanoparticles, without neglecting the effect of the contribution of Fe ions. It is also worth recalling that in 1947 Néel was interested in the so-called spinel ferrites of formula Fe2O3MO, where M is a 3d divalent metal cation (e.g. Fe2+). The magnetic properties of these materials, with spontaneous magnetization, were not understood in detail. Néel called them non-compensated antiferromagnets “ferrimagnets”.56
In this work, we conducted an in-depth characterisation of the chosen physical properties of ultrasmall (ca. 8 nm) Zn1−xMnFe2O4 nanoparticles with narrow size distribution fabricated by the non-hydrolytic approach. The main focus was on detailed EMR characteristics. We observed that the positive Curie–Weiss temperature obtained from macroscopic measurements of the inverse susceptibility, at temperatures well below the Curie temperature of the ferrimagnetic clusters, is assigned to the superparamagnetic behaviour of clusters of atoms with a net magnetic moment originated through A–B superexchange interactions. The linear thermal dependence of the inverse of susceptibility appears only at temperatures above the blocking temperatures of these superparamagnetic clusters. Thus, the apparent Curie–Weiss temperature was expected to be in the range of the blocking temperatures corresponding to the cluster size distribution. It is important to note that this temperature is not the critical one for a phase transition, but it corresponds to the temperature at which the relaxation time of the magnetic moment is similar to the measurement time.36 The EMR method makes it possible to measure the magnetic properties of materials in a way that other methods cannot. In some respects, the EMR can be more sensitive than SQUID because it responds only to magnetic ions with unpaired spin, whereas SQUID responds to both magnetic ions (seen in the EMR) and elements (atoms).
A Tecnai Osiris X-FEG (FEI Company, USA) transmission electron microscope (TEM) operating at 200 kV was used for nanoparticle (NP) imaging and evaluation of particle size, distribution, and morphology. The sample preparation involved the sonication of ferrite-containing ethanol suspensions (ca. 25 μg ml−1) and placing a droplet of colloids on a 200 mesh carbon-coated copper grid (EM Resolutions, United Kingdom). The grids with the deposited materials were dried overnight under dust protection. Image analysis was performed in ImageJ freeware software (v. 1.8.0_1720). The concentration of particles in stock solution was determined using the Radwag MYA 5.4Y scale through three independent repetitions and final mass averaging.
Elemental analysis of mixed zinc manganese ferrites was performed using a Tescan Vega 3 scanning electron microscope (SEM) equipped with a Bruker Nano XFlash 6130 detector for energy-dispersive X-ray spectroscopy analysis (EDS). All samples for SEM-EDS analysis were prepared by placing a dry powder on a carbon tape attached to the aluminium-made holder. Data analysis was performed in a dedicated Bruker Esprit software (v. 2.1.1.1.17430) provided by the manufacturer.
Magnetic resonance measurements were performed on X-band (ν ≈ 9.4 GHz) and Q-band (ν ≈ 34 GHz) with modulation of the magnetic field at 100 kHz and 50 kHz, respectively, by a Bruker multifrequency and multiresonance FT-EPR ELEXSYS E580 spectrometer with X-band with a sensitivity of 7 × 109 spins per 0.1 mT having a resolution of 2.35 (micro) T or even better.61 Temperature dependence measurements were carried out at room temperature, and as a function of temperature in the X-band in the range from 95 K to 300 K using a Bruker liquid nitrogen gas-flow cryostat with 41131 VT digital controller, and in the Q-band in the range from 4 K to 300 K, using an ESR 900 liquid helium gas-flow cryostat from Oxford Instruments with Mercury iTC digital controller. Samples containing approximately 20 mg of sample powder were placed in 4 mm diameter quartz tubes.
![]() | (1) |
![]() | ||
Fig. 1 Graphical representation of the recorded X-ray patterns of pure and mixed zinc–manganese ferrites along with fitting results obtained through Rietveld refinement. |
Crystal data | |||||
---|---|---|---|---|---|
Chemical Formula | ZnFe2O4 | Zn0.8Mn0.2Fe2O4 | Zn0.5Mn0.5Fe2O4 | Zn0.2Mn0.8Fe2O4 | MnFe2O4 |
Crystal system, space group | Regular, Fd![]() |
||||
a, (Å) | 8.4388 | 8.4259 | 8.4370 | 8.4355 | 8.4473 |
V (Å3) | 600.95 | 598.20 | 600.57 | 600.25 | 602.77 |
Z | 8 | ||||
Radiation type | Cu Kα | ||||
Refinement | |||||
R | 0.0208 | 0.0216 | 0.0218 | 0.0210 | 0.0206 |
wR | 0.0261 | 0.0269 | 0.0273 | 0.0261 | 0.0258 |
S | 0.32 | 0.32 | 0.33 | 0.32 | 0.32 |
The size of the pure and mixed zinc–manganese ferrite particles, their distribution, and morphology were estimated using the TEM technique (Fig. 2 and 3). It should be noted that the fabricated nanomaterials did not differ in size, which was estimated to be 7.5 in all cases with a narrow distribution (monodisperse particles). This is beneficial since all of the observed effects will be associated only with that small particle size. The morphology of all ferrites is close to spherical with the presence of less regular particles with a tendency to form agglomerates. This behaviour is very typical for the synthesis of nanomaterials without the addition of any surface-blocking agents that will lead to a reduction of the surface energy responsible for agglomeration. Concerning particle size, good correspondence was found with the crystallite size estimated by the Scherrer formula, meaning that individual particles seen in TEM are represented as single crystallites.
![]() | ||
Fig. 2 TEM images of the pure and mixed zinc–manganese ferrites obtained by the non-hydrolytic approach. |
SEM images as well as element maps of the Zn1−xMnxFe2O4 nanoferrites are shown in Fig. 4. It is evident that after manganese doping, the Mn2+ signal increases, while Zn2+ decreases significantly. It is, together with the XRD data, a straightforward proof of the Mn2+ incorporation into the structure of zinc ferrite for a broad concentration range. The content of the manganese ratio was calculated from the EDS data to be sufficiently close to the desired final composition, i.e., 19% for x = 0.2, 46% for x = 0.5, and 71% for the x = 0.8, respectively. Some deviation from the expected content of Mn2+ might be caused by the standard error of the EDS analysis (5%) and losses during substrate transfer.
![]() | ||
Fig. 4 EDS and SEM elemental mapping of the Zn1−xMnxFe2O4 nanoparticles. EDS was performed covering a sample area of 250 μm × 250 μm in each case. |
The EMR characteristics of Zn1−xMnxFe2O4 ferrites samples were studied by electron magnetic resonance spectroscopy. The authors determined the resonance field (Bres) and resonance linewidth (Bpp) from measured EMR spectra to understand the homogeneity and crystalline anisotropy of Zn1−xMnxFe2O4 ferrites. Representative EMR spectra of Zn1−xMnxFe2O4 ferrites are plotted in Fig. 5.
Various magnetic parameters such as resonance linewidth (BPP), Lande's splitting factor (effective g-value) of Zn1−xMnxFe2O4 ferrite nanoparticles are listed in Table 2. The g-value is a function of the molecular motion, the symmetry of the ions, and the paramagnetic properties, which is also a constant of proportionality between the field and frequency. The effective g-value is calculated using the following relation:64
![]() | (2) |
ZnFe2O4 | Zn0.8Mn0.2Fe2O4 | Zn0.5Mn0.5Fe2O4 | Zn0.2Mn0.8Fe2O4 | MnFe2O4 | |
---|---|---|---|---|---|
B pp – resonance linewidth, QCW – paramagnetic Curie temperature, C – Curie constant. | |||||
Effective g-factor | 2.010 | 2.014 | 2.124 | 2.115 s | 2.253 |
B pp (mT) | 33.60 | 56.03 | 113.46 | 133.67 | 108.40 |
Q CW (K) | 89.1 | 75.6 | 5.6 | −112.0 | −183.6 |
C | 1.57 × 10−10 | 3.88 × 10−10 | 8.32 × 10−10 | 14.58 × 10−10 | 71.43 × 10−10 |
The material studied Zn1−xMnxFe2O4 is a complex structure with various magnetic interactions. It is a kind of combination of two materials, MnFe2O4 and ZnFe2O4. A detailed analysis of the line parameters is planned in a separate article, including the calculation of the spin Hamiltonian. In the case of MnFe2O4, Fe-3d electrons occupy the crowded B-site while Mn-3d occupy the less electrons crowded A-site. From this anomalous behavior for the Fe 3d energy states position, electrons are not able to share the conduction mechanism (−8 eV away from Fermi level). The conduction mechanism in the spinel depends mainly on the super exchange between sites (A and B, for example between Mn and Fe cations through oxygen). Therefore, the Fe-3d electrons site for this cycle of conduction (Fe, O, and Mn) is missing. From the above anomalous explanation of density of states figures, it can be observed that a ferrimagnetic spin current between Fe-3d and Mn-3d through O-2p is yielded according to spin open or closed loop.42 Spin open loop:
(1) Fe–O spin exchange interaction in minority channel between −10 and −8 eV;
Fe-3d↓ + O-2p → Fe-3d + O-2p↓, | (3) |
(2) O–O spin exchange interaction between minority and majority;
O-2p↓(minority) + O-2p(majority) → O-2p↑, | (4) |
(3) O–Mn spin exchange interaction in majority channel between −2 and 0 eV.
O-2p↑, Mn-3d → O-2p, Mn-3d↑, | (5) |
Spin closed loop: this p–d overlap is the origin of the ferrimagnetic superexchange interaction between tetra and octahedral sites in ferrites, which are characterized by high hardness, further confirming the p–d overlapping for oxygen, and that TM is too much strong.42 In the case of ZnFe2O4, the susceptibility χ is the sum of all magnetic contributions, paramagnetic (PM), diamagnetic (DM), AFM, and FM-like. It can be described as:
χ = χPM + χDM + χAFM + χFM. | (6) |
On the other hand, at high fields, only the PM, DM, and AFM orderings vary with the field since the others are saturated.66 Therefore, FM-like contributions can be discarded at HF, and high-field susceptibility χHF can be described as:
χHF = χDM + χPM + χAFM. | (7) |
Consequently, the ferri-, ferro-, and/or superparamagnetic contributions can be evaluated by calculating χHF and subtracting it from the entire magnetization curve.37 Theoretical calculations show the influence of inversion within a single unit cell, the smallest local inversion (δc) is either 0 or 1/8, the last corresponding to the exchange of a single Zn/Fe pair in the unit cell. Assuming that only a single Zn/Fe exchange can occur in a cell, around 40% of the unit cells suffer a single Zn/Fe cations exchange when the macroscopic δ is as small as 0.05. Unit cells with δc = 1/8 are FiM (40%) with a magnetic moment of 5.9μB,37 while for δc = 0 the cells are AFM (60%). This illustrates the dramatic effect that δ ≈ 0 can have on the magnetic and calorimetric properties.
Cobos et al.38 show for ZnFe2O4, that the magnetic entropy increment can be associated with the AFM to PM transition, and the small experimental value of 8.7 J mol−1 K−1 indicates that only 29% of the sample has evolved from AFM to the paramagnetic phase. In summary, instead of the expected 60%, only 29% of the sample appears to be AFM ordered.38
To briefly recap, the contribution of ZnFe2O4 gives a narrow line, while the line of MnFe2O4 is several times wider. The observed spectrum consists of a line from the Fe3+ ion (a line of about 150 mT), and a broad line that is a superposition of at least two components. In addition, a narrow line (about 350 mT) is also visible. As the concentration of Mn ions increases, the width of the resonance line also increases. The only exception is MnFe2O4, due to the fact that the effective g-factor comes from the superposition of the line. The spectrum of the Fe3+-coupled pair (Fe3+–O–Fe3+) in Fe2O3 is known to give a resonance line with an effective g-factor around 2.0. So, the effective g-value, 2.169, for the MnFe2O3 sample can be attributed to the superexchange interaction between the Fe3+ ions. A weak signal observed at about effective g-factor 4.2, in addition to the intense spectral line, can be attributed to an isolated Fe3+ in the orthorhombic field.67,68
As the temperature increases, the signals gradually get sharper and more symmetrical, with a shift to a high field (see Fig. 6). Furthermore, with increasing temperature, a sharp decrease in the line width and an increase in the intensity of the EMR signal is observed (Fig. 9). This is because the thermal fluctuation of magnetic moments can reduce the extent of the anisotropic contributions to the free energy density and the effective magnetic field (the superparamagnetic effect).69 In this event, at elevated temperatures very narrow resonance spectra are observed, while at low temperatures the thermal fluctuations are gradually frozen out, and the magnetization direction is blocked, resulting in the broadening of the resonance line.
Broadening and shift to lower magnetic fields of magnetic resonance spectra with increasing temperature are typically observed for superparamagnetic (SPM) nanoparticles, and analogous cases can be found in a number of systems.69 However, for SPM particles, whose direction of magnetization fluctuates at a rate faster than the Larmor frequency, this results in a narrow resonance line because of an averaging effect of this fluctuation on the magnetocrystalline anisotropy. With a decrease in the temperature, the resonance line of the SPM particle broadens as the averaging effect of thermal fluctuations is reduced and the direction of magnetization is blocked, first in bigger and progressively in smaller particles. Thus, the narrow resonance line at room temperature, which progressively broadens with a reduction in temperature, once again confirms the superparamagnetic behavior of the prepared samples.68
In ferromagnetic materials, magnetic domains are in a fragile state of equilibrium, and the Bloch wall, which is a narrow transition region at the boundary between magnetic domains, moves with very low applied fields. In fact, LFMA is associated with the dynamics of magnetic domains in material.70 The existence of this absorption at room temperature is an indication of the ferromagnetic state of the material used to detect magnetic order. For bulk samples, the LFMA signal is used to determine the Curie temperature of ferromagnetic compounds.71,72 Above 28 nm, the existence of LFMA shows that these compounds are in a magnetic multi-domain state, and a flat response for compounds with smaller size shows that they are in a single-domain state. The absence of the LFMA signal in ferromagnetic compounds is a good indication of the superparamagnetic state in samples.73 By size reduction, the nanocrystalline passes from a particle with several magnetic domains to a monodomain particle; the latter is either in a single-domain state or in a superparamagnetic state. Low-field absorption cannot determine the intermediate state. Particles belonging to the single-domain state are characterized by the maximum magnetocrystalline anisotropy energy, consequently the direction of magnetization is ‘frozen’. This characteristic has an effect on the linewidth of resonant absorption; a comparative analysis can reveal the critical size of changes in magnetic states.
In the higher temperature range, the paramagnetic region, the line is relatively narrow and more regular. When the temperature is lower, we observe a widening of the line along with its deformation (the upper part), the centre moves along with the lower part of the line. The lower the temperature, the deformation increases. In the Q-band we observe a clear separation of the line (Fig. 7).
Increasing the content of manganese the value of the effective factor g increases, from 2.008 to 2.169 (for 100% Mn ions). A similar relation occurs for the width of the line. For ZnFe2O4, the line is the narrowest (33.6 mT). We observe an increase in line width to 130.6 mT for 80% manganese. Finally, for 100% Mn ion, the width is 108 mT, which is less than described above, due to the fact that we are facing two components of the resonance line that are offset from each other. Still, as a result of the summation, we will eventually get the widest recorded line. If the ZnFe2O4 component is missing, the line from the MnF2O4 component becomes narrower. The superposition of the two components of the line causes the relationship shown in Fig. 8 to change rapidly at temperatures around 120 K. This is particularly visible at contents of 50% or more manganese.
![]() | ||
Fig. 8 Temperature dependence of the effective g-factor for Zn1−xMnxFe2O4 nanoparticles obtained in the X-band. |
We used the Curie–Weiss law to analyze the temperature dependences of the integral intensity, which is directly proportional to the magnetic susceptibility χ. A linear increase of χ−1 (T) at higher temperatures can be fitted to the Curie–Weiss law:
(χ − χ0)−1 = (T − QCW)/C, | (8) |
Q CW is the paramagnetic Curie temperature, χ0 temperature independent coefficient to calculate the magnetic permeability, and the Pauli term for paramagnetism.74 Based on the theory, the inverse relationship of intensity (1/Int) is shown as a function of temperature for Zn1−xMnxFe2O4 nanoparticles samples (Fig. 9). Obtained values of Curie temperatures and Curie constants for all samples shown in Table 2.
![]() | ||
Fig. 9 Temperature dependence of peak to peak line width for Zn1−xMnxFe2O4 nanoparticles obtained in X-band. |
Fig. 10(a–e) show the dependence of the inverse intensity as a function of temperature for Zn1−xMnxFe2O4 nanoparticles obtained in the X-band for EMR lines in a higher magnetic field (about 350 mT). Due to the antiferromagnetic character of the B–B and A–B superexchange interactions in spinel ferrites, the experimental finding of a positive value of the apparent Curie–Weiss temperature may be surprising. As Lotgering75 indicated, the measuring temperature should increase to 800 K to observe the straight line with negative Curie–Weiss temperature due to the antiferromagnetic A–B coupling, which gives rise to the ferrimagnetic clusters. On the other hand, the negative Curie–Weiss temperature associated with the B–B antiferromagnetic coupling could only be observed for δ = 0 (for example, with ideal ZnFe2O4 samples, in which there is no trace of ferrimagnetic clusters). However, since this ideal case is difficult to achieve, the previously reported anomalous sign for the Curie–Weiss temperature can be understood as a consequence of superparamagnetism associated with the almost unavoidable presence of a few ferrimagnetic clusters.36 We observed a non-linear increase in Qcw values for the increase in Zn content (Fig. 10f) as well as deviations from the designated Qcw from the literature. We identified the following reasons for these differences (1) only values for pure compounds are given in the literature; (2) particle size affects the Curie temperature76 that is especially manifested for nanoparticles with dimensions in the range of a few to several nanometers, thus observed changes in Qcw of several tens of K; (3) it is not insignificant that our measurements involve materials being a combination of Zn and Mn, so more complex magnetic interactions are to be expected than for single compounds.
Observed effects are more complex interactions than in pure zinc an manganese ferrites. As a result of the study, a relationship was observed between the composition of the material and the magnetic properties with striking antiferromagnetism to ferrimagnetism change (see Fig. 10f). Hence, by modifying the composition of materials, we have the possibility of controlling the magnetic properties (FM–AFM–FiM).
EMR spectral studies indicated that the prepared nanoparticles were superparamagnetic. The linewidth of EMR signal for any ferrite material generally originated from two sources: (a) magnetic dipole–dipole interactions among particles and (b) interparticle superexchange interactions between magnetic ions through oxygen ions. The presented EMR lines are obtained from contributions from ZnFe2O4 and MnFe2O4. The contribution of ZnFe2O4 gives a narrow line, while the line of MnFe2O4 is several times wider. The observed spectrum is composed of a line originating from the Fe3+ ion (a line around 150 mT) and a broad line that is a superposition of at least two components. A narrow line (around 350 mT) was also analyzed. An increase in the width of the resonant line was observed, proportional to the increase in Mn concentration in the sample. The only exception is MnFe2O4, due to the fact that the effective g-factor comes from the superposition of the line.
In the higher temperature range, the paramagnetic region, the observed lines are relatively narrow and regular. With a decrease in the temperature, a broadening of the line is observed along with a deformation of the upper part, with the centre of the line moving along with the lower part. The lower the temperature, the greater the deformation becomes. Performing measurements in the Q-band allowed us to observe separated resonant lines. Both ZnFe2O4 and MnFe2O4 can be decomposed into two inequivalent magnetic sublattices, denoted as O and T, respectively. The T sublattice is not frustrated in the NN interaction, whereas in the presence of AFM interactions within the sublattice, the O sublattice can host a large NN frustration. While the ions within a sublattice often interact FM and the ions between two sublattices interact AFM, some spinel compounds can give rise to the interaction of AFM between ions at the O sites that leads to geometric frustration. As a result, the overall magnetic behavior depends on the strength of the exchange interaction between O–O (JOO), T–T (JTT) and O–T (JOT). We observe a non-linear increase in Qcw values with an increase of Zn content. Such complex behavior is caused by the size of the particles as well as chemical structural modification that leads to more complex magnetic interactions than that expected for pure compounds.
This journal is © The Royal Society of Chemistry 2025 |