Next Article in Journal
Temperature-Dependent Phase Evolution in FePt-Based Nanocomposite Multiple-Phased Magnetic Alloys
Next Article in Special Issue
Mechanical Reinforcement of ABS with Optimized Nano Titanium Nitride Content for Material Extrusion 3D Printing
Previous Article in Journal
Waste Citrus reticulata Assisted Preparation of Cobalt Oxide Nanoparticles for Supercapacitors
Previous Article in Special Issue
Study of the Reinforcement Effect in (0.5–x)TeO2–0.2WO3–0.1Bi2O3–0.1MoO3–0.1SiO2–xCNDs Glasses Doped with Carbon Nanodiamonds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of Phase Transformations and Hyperfine Interactions in Fe3O4 and Fe3O4@Au Nanoparticles

by
Vyacheslav S. Rusakov
1,
Artem L. Kozlovskiy
2,3,*,
Maxim S. Fadeev
1,
Kamila B. Egizbek
2,3,
Assel Nazarova
2,
Kayrat K. Kadyrzhanov
2,
Dmitriy I. Shlimas
2,3 and
Maxim V. Zdorovets
2,3
1
Faculty of Physics, M.V. Lomonosov Moscow State University, 119991 Moscow, Russia
2
Engineering Profile Laboratory, L.N. Gumilyov Eurasian National University, Nur-Sultan 010008, Kazakhstan
3
Laboratory of Solid State Physics, The Institute of Nuclear Physics, Almaty 050032, Kazakhstan
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(23), 4121; https://doi.org/10.3390/nano12234121
Submission received: 31 October 2022 / Revised: 18 November 2022 / Accepted: 20 November 2022 / Published: 22 November 2022
(This article belongs to the Special Issue Nanostructured Ceramics in Modern Materials Science)

Abstract

:
The paper presents the results of a study of iron oxide nanoparticles obtained by chemical coprecipitation, coated (Fe3O4@Au) and not coated (Fe3O4) with gold, which were subjected to thermal annealing. To characterize the nanoparticles under study, scanning and transmission electron microscopy, X-ray diffraction, and Mössbauer spectroscopy on 57Fe nuclei were used, the combination of which made it possible to establish a sequence of phase transformations, changes in morphological and structural characteristics, as well as parameters of hyperfine interactions. During the studies, it was found that thermal annealing of nanoparticles leads to phase transformation processes in the following sequence: nonstoichiometric magnetite (Fe3−γO4) → maghemite (γ-Fe2O3) → hematite (α-Fe2O3), followed by structural ordering and coarsening of nanoparticles. It is shown that nanoparticles of nonstoichiometric magnetite with and without gold coating are in the superparamagnetic state with a slow relaxation rate. The magnetic anisotropy energy of nonstoichiometric magnetite is determined as a function of the annealing temperature. An estimate was made of the average size of the region of magnetic ordering of Fe atoms in nonstoichiometric magnetite, which is in good agreement with the data on the average sizes of nanoparticles determined by scanning electron microscopy.

1. Introduction

In recent years, among the variety of known nanomaterials, much attention has been paid to iron-containing oxide nanoparticles, the interest in which is due to their wide range of practical applications as well as the potential for their use in almost all fields of science and technology [1,2,3]. Iron-containing nanoparticles also have great potential for catalysis, where they can be used as an absorbent in the catalytic or photocatalytic decomposition of organic dyes or as heavy metal absorbents to extract them from aqueous media, etc. [2,3,4,5,6,7]. An important role is played by iron-containing oxide nanoparticles in the creation of magnetic sensors, as well as in the energy sector, where they are used as the basis for anode materials for lithium-ion batteries [8,9]. In the past few years, areas of application of nanoparticles in medicine have been actively developed, where they can be used as carriers for targeted delivery of drugs, contrast markers for MRI, materials for hyperthermia exposure, etc. At the same time, the emphasis on the use of nanoparticles is shifting more and more in the biomedical direction, which imposes additional requirements on nanostructures related not only to their structural and magnetic properties, but also to resistance to external influences, toxicity, and biological activity [10,11,12,13]. These developments, despite the rather large number of experimental works and reviews, require more and more scientific research in this direction [14,15,16,17]. An analysis of publication activity for such keywords as “magnetic nanoparticles”, “magnetic nanoparticles for biomedical application”, and ““core-shell” magnetic nanoparticles” in the Web of Science Core Collection database showed a growth trend in the number of publications in these areas [18]. The total number of publications over the past two years has more than doubled compared to the same number of publications over the period from 2012 to 2014. At the same time, more than 35% of scientific articles and reviews out of the total number of published works are related to the biomedical application of nanoparticles. Notably, the annual increase in scientific articles and reviews is more than 10%, which indicates an increase in research in this direction.
Separately, it should be noted that one of the most promising areas of research related to iron-containing oxide materials is the creation of core-shell nanostructures or complex nanocomposites [19,20,21,22]. As can be seen from the analysis of publication activity, interest in these types of nanomaterials is increasing every year. The total number of published articles and reviews in this area is more than 30% of the total number of scientific articles related to magnetic nanoparticles.
Interest in core-shell nanostructures is primarily due to the possibilities for their application, which are significantly expanded via shell-coating with noble metals, such as Au and Ag, or any polymer or organosilicon compound on magnetic nanoparticles [23,24,25]. Coatings on magnetic nanoparticles not only increase their resistance to external factors that can lead to corrosion and subsequent degradation, but also increase the possibility of binding various complex compounds or drugs to the surface of nanoparticles [26,27]. Such structures make it possible to create nanocontainers with a magnetic core, due to which it is possible to control nanoparticles using external magnetic fields. In this case, the nonmagnetic shell reduces the effect of agglomeration of magnetic nanoparticles into larger complexes. These properties, as well as a wide variety of other properties and their variations, make core-shell nanoparticles very promising objects for research [28,29,30].
The purpose of this work is to study the effect of thermal annealing of Fe3O4 and Fe3O4@Au nanoparticles on changes in their morphology and structural properties, on the kinetics of temperature phase transformations, and on hyperfine interactions of Fe57 nuclei. The choice of Fe3O4 and Fe3O4@Au nanoparticles as objects of study is due to the prospects for practical application in microelectronics, batteries, catalysis, purification of aqueous media, and biomedicine [31,32,33]. At the same time, the previously proposed method for modifying Fe3O4 nanoparticles by depositing a gold shell on them makes it possible to obtain core-shell structures [34]. However, to date, little is known about the effect of coating with a gold shell on the change in structural parameters and the dynamics of phase transformations during thermal action on nanoparticles, as a result of which, phase transformation processes are initiated in iron-containing oxide nanoparticles. Much attention is paid to the analysis of the parameters of magnetic hyperfine interactions, as well as their connection with data on changes in structural parameters and phase transformations. It should be noted that the use of Mössbauer spectroscopy, in contrast to classical X-ray methods, makes it possible to answer several questions related to the phase analysis and nonstoichiometry of iron-containing oxide phases in nanoparticles during their thermal annealing.

2. Experimental Part

The initial nanoparticles were synthesized using the method of chemical coprecipitation from 2M FeCl2 and 1M FeCl3 salts with the addition of 50 mL of ammonium hydroxide (NH4OH) to the solution. The deposition of the gold shell was carried out in two stages, including the dispersion of nanoparticles in a solution of citric acid (with a concentration of 0.1 g/mL) followed by treatment in a solution of gold chloride and sodium citrate. A detailed description of the procedure for the synthesis of iron-containing nanoparticles, as well as the deposition of a gold shell on them, is presented in [34].
To initialize the processes of phase transformations in nanoparticles, the method of thermal annealing in an air atmosphere was applied. The samples were annealed in a SNOL muffle furnace in a temperature range of 100–800 °C with a step of 100 °C for 5 h, and followed by cooling the samples to room temperature for 10–24 h, depending on the annealing temperature.
The morphological features of the synthesized nanoparticles were studied using high-resolution scanning electron microscopy (SEM) and high-resolution transmission electron microscopy (TEM). For these studies, a Jeol 7500F scanning electron microscope (Jeol, Tokyo, Japan) and a Jeol JEM1400 Plus transmission electron microscope (Jeol, Tokyo, Japan) were used. SEM and TEM images were processed using ImageJ software V.2.0 (Github Inc, San Francisco, CA, USA).
The analysis of structural changes and phase transformations as a result of thermal annealing was carried out using the X-ray diffraction (XRD) method implemented on D8 Advance ECO (Bruker, Berlin, Germany) and MiniFlex 600 (Rigaku Corporation, Tokyo, Japan) diffractometers. The diffraction patterns were taken in the Bragg–Brentano geometry, in the angular range 2θ = 25–85°, with a step of 0.03°. The diffraction patterns were processed using the SmartLab Studio II software and the ICDD PDF-2 database.
Investigations of hyperfine interactions of 57Fe nuclei were carried out at room temperature by Mössbauer spectroscopy (MS) on an MS1104Em spectrometer (Research Institute of Physics, Rostov State University. Rostov-on-Don, Russia) operating in the constant acceleration mode with a triangular change in the Doppler velocity of the source relative to the absorber. The 57Co nuclei in the Rh matrix acted as a source of resonant γ-quanta. The Mössbauer spectrometer was calibrated at room temperature using an α-Fe reference absorber. The obtained Mössbauer spectra were analyzed using the SpectrRelax software (MSU, Moscow, Russia) [35].

3. Results and Discussion

The results of the study of the morphological features of synthesized and annealed Fe3O4 and Fe3O4@Au nanoparticles, obtained using SEM, are shown in Figure 1. Analysis of SEM images of nanoparticles and their size distributions shows that the shape of nanoparticles at low annealing temperatures is close to spherical, both in the case of Fe3O4 nanoparticles, and Fe3O4@Au nanoparticles.
Figure 2 shows the results of estimating the average size of nanoparticles obtained from the SEM image analysis data, depending on the annealing temperature tann. For Fe3O4@Au nanoparticles after synthesis and subsequent annealing at temperatures of 100–300 °C, the average particle sizes exceed the average sizes of Fe3O4 nanoparticles by 4–6 nm, which indicates the presence of a shell with a thickness of 2–3 nm.
At temperatures of 100–400 °C, undispersed Fe3O4 nanoparticles with an average size of ~15–20 nm stick together, forming agglomerates of particles with an average size of ~45 nm, which are easily dispersed under ultrasonic treatment. A further increase in the annealing temperature leads to sticking and combination of the initial dispersed and non-dispersed nanoparticles with the formation of larger nanoparticles, up to ~90 nm at tann = 800 °C.
It should be noted that the process of particle association begins for dispersed Fe3O4 nanoparticles at ~400 °C, and for Fe3O4@Au nanoparticles, at a higher temperature of ~550 °C (Figure 2a). Such association of nanoparticles occurs primarily due to particles with the most probable and largest sizes, as evidenced by the increase in the width of the size distributions of formed nanoparticles with increasing annealing temperature (see Figure 2b). At the same time, the average size and width of the size distribution at the same annealing temperature are noticeably larger for Fe3O4 nanoparticles than for Fe3O4@Au nanoparticles, for which the presence of a shell prevents their association.
Figure 3 shows TEM images (including high resolution) of the investigated Initial and Annealed at temperatures of 400 °C and 600 °C Fe3O4 nanoparticles. According to the presented data, unannealed Fe3O4 nanoparticles consist of several crystallites (regions of structural ordering). At the same time, the interplanar distances of the main part of the crystallites are characteristic of magnetite with an unknown stoichiometry degree (d(Fe3O4)(022) = 3.00 ± 0.13 Å). For nanoparticles annealed at a temperature of 400 °C, there are two different orientations of parallel atomic planes with different interplanar distances, presumably corresponding to nonstoichiometric magnetite (d(Fe3O4)(111) = 4.41 ± 0.23 Å) and hematite (d(α-Fe2O3)(012) = 3.59 ± 0.16Å). At the same time, for nanoparticles annealed at a temperature of 600 °C, only one system of parallel atomic planes with an interplanar spacing characteristic of hematite is observed.
A TEM image of Fe3O4@Au nanoparticles and high-resolution image of one Fe3O4@Au nanoparticle, as well as mapping results of this nanoparticle, are shown in Figure 4. Analysis of the obtained TEM images (Figure 4a) shows that the studied Fe3O4@Au nanoparticles have a core-shell structure, where the core is a nanoparticle consisting of iron oxide (dark central area in the high-resolution image), and its lighter shell, according to the mapping results, is a shell of gold (see Figure 4b). The shell thickness was estimated to be 3–5 nm, which is in good agreement with the values obtained by scanning electron microscopy (SEM) as a result of comparing the average sizes of Fe3O4 and Fe3O4@Au nanoparticles.
The TEM image of Fe3O4@Au nanoparticles (Figure 4a) shows a single large particle, which is very different from the rest of the nanoparticles. According to the result of elemental analysis, this particle is a gold particle, the presence of which may be associated with the process of gold agglomeration during synthesis.
The results of X-ray diffraction, showing the dynamics of phase transformations in the studied nanoparticles depending on the annealing temperature, are shown in Figure 5. An analysis of the obtained data made it possible to establish that the initial nanoparticles are nanoparticles of nonstoichiometric magnetite Fe3−γO4 with an inverse spinel structure (sp. gr. F d 3 ¯ m ). Along with an increase in the annealing temperature of nanoparticles, magnetite is oxidized, as a result of which the diffraction patterns reveal reflections characteristic of the α-Fe2O3 hematite phase with a rhombohedral structure (sp. gr. R 3 ¯ c ); the contribution of such reflections increases with an increase in the annealing temperature (Figure 5). As annealing temperature increases, a decrease in the widths of the diffraction reflections of the nonstoichiometric magnetite and hematite phases is observed, which indicates the ordering of the crystal structures and an increase in the sizes of the regions of structural ordering (coherence lengths) for both phases. In the case of Fe3O4@Au nanoparticles, the diffraction patterns contain reflections (Figure 5) characteristic of gold (Au) nanoparticles with a face-centered cubic crystal lattice (sp. gr. F m 3 ¯ m ), the presence of which was established using transmission electron microscopy (Figure 4).
Figure 6a shows the dependences of the relative intensities of the diffraction patterns of the established phases on the annealing temperature tann. For Fe3O4 nanoparticles, already at tann = 300 °C, along with nonstoichiometric Fe3−γO4 magnetite, the appearance of the α-Fe2O3 hematite phase (I ~ 3%) is fixed, which becomes dominant (I ~ 100%) at annealing temperatures tann ≥ 500 °C. In the case of Fe3O4@Au nanoparticles, the oxidation of nonstoichiometric magnetite (Fe3−γO4) and its transformation into hematite (α-Fe2O3) occurs at tann > 450 °C. At the same time, for Fe3O4@Au, the contribution of the diffraction pattern of Au nanoparticles (I ~ 4%) and the unit cell parameter of Au (a ~ 4.082 Å) remain unchanged during annealing, which indicates the absence of decomposition (peeling) of the shell and the formation of Au nanoparticles, as well as the appearance of a substitution phase or introductions.
Depending on the annealing temperature, Figure 6b shows the average values d of the sizes of regions of structural ordering (coherence lengths/crystallite sizes) of nanoparticles, determined using the Scherrer formula. As can be seen from the data presented, up to an annealing temperature of 400 °C, only a slight increase in the coherence length is observed, which is associated mainly with the oxidation and ordering of the crystal structure of nonstoichiometric magnetite. In this case, the association of nanoparticles, and as a result a sharp increase in the coherence length, occurs at higher temperatures, when they are mainly hematite particles. At the same time, as we see in Figure 6b, for Fe3O4@Au nanoparticles, the Au shell makes it difficult for them to combine; the size of coated nanoparticles is, on average, ~5 nm smaller than that of Fe3O4 nanoparticles after combining.
As a result of the processing of X-ray diffraction patterns by the Rietveld method for the studied Fe3O4 and Fe3O4@Au nanoparticles, the unit cell parameters of the crystal lattice for nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 were determined depending on the annealing temperature tann (Figure 7a). In the case of nonstoichiometric magnetite Fe3−γO4, the presence of which is typical for tann ≤ 400 °C, a noticeable decrease in the unit cell parameter is observed with an increase in the annealing temperature, which indicates an increase in the degree of its nonstoichiometry γ [36,37]. For hematite, the unit cell parameters slightly decrease with the annealing temperature. At the same time, a slight decrease in the ratio of unit cell parameters c/a indicates the improvement of its crystal structure (Figure 7b) [38]. Since in the case of nanoparticles of nonstoichiometric magnetite, it is not possible to accurately determine the degree of its nonstoichiometry γ by X-ray diffraction, then, including for this purpose, the method of Mössbauer spectroscopy was applied.
Figure 8 and Figure 9 show the most characteristic Mössbauer spectra of the studied Fe3O4 and Fe3O4@Au nanoparticles. As can be seen, these spectra, especially for nanoparticles annealed at low temperatures (100–300 °C), are poorly resolved and show signs of the relaxation behavior of nanoparticles. Therefore, for the model fitting and interpretation of these spectra, a priori information about the properties of nanoparticles obtained using electron microscopy and X-ray diffraction methods was used and reasonable physical assumptions were made.
In the general case, a magnetite nanoparticle can be represented as a particle with a gradient change in the degree of magnetite nonstoichiometry. The limiting options for representing such a particle are either a particle of homogeneous composition with an average value of the nonstoichiometry degree γ, i.e., a solid solution of magnetite Fe3O4 and maghemite γ-Fe2O3, or a mixture of magnetite and maghemite, for example, in the center and on the surface of the particle, respectively (see Figure 10).
In the presence of fast electron exchange, regardless of the type of representation of the crystal chemical formula of the studied oxide, it contains trivalent iron ions in the tetrahedral (A) position ( Fe A 3 + ) , trivalent iron ions in the octahedral (B) position ( Fe B 3 + ) , as well as ions with intermediate valence in the octahedral position ( Fe B 2.5 + ), participating in the Verwey mechanism of electron exchange [39,40].
The crystal chemical formula in the presence of a fast electron exchange between Fe atoms in the case of a solid solution of magnetite and maghemite (nonstoichiometric magnetite Fe3−γO4) is written as
Fe 3 γ O 4 = ( Fe 3 + ) A [ Fe 2 ( 1 3 γ ) 2.5 + Fe 5 γ 3 + γ ] B O 4 2 ,
and in the case of a mixture of magnetite and maghemite phases as
( 1 b ) · ( Fe 3 + ) A [ Fe 2 2.5 + ] B O 4 2 + b · ( Fe 3 + ) A [ Fe 5 3 3 + 1 3 ] B O 4 2 ,
where γ is the number of vacancies of Fe atoms ( ) per formula unit (degree of nonstoichiometry), 0 ≤ b ≤ 1 is the molar concentration of maghemite, while b = 3γ.
According to crystal chemical Formulas (1) and (2), it is possible to determine the molar concentration of maghemite (b) and the number of vacancies of iron atoms per formula unit (γ) using the intensity ratios of the subspectra corresponding to different iron ions, taking into account the known probability ratio of the Mössbauer effect for Fe atoms in octahedral and tetrahedral positions (fB/fA = 0.94 ± 0.02 [41]):
b = 3 γ = 3 5 · f A f B · I ( Fe B 3 + ) I ( Fe A 3 + ) = 1 1 2 · f A f B · I ( Fe B 2.5 + ) I ( Fe A 3 + ) .
Since the obtained Mössbauer spectra are characteristic of the relaxation behavior of the studied nanoparticles, when fitting them, it is necessary to use the model of multilevel superparamagnetic relaxation [42]. Some of its main parameters are the relaxation rate (R) and the ratio of the magnetic anisotropy energy ( E ma = K eff V ) to the thermal energy ( k B T ):
α = E ma k B T = K eff V k B T .
As can be seen, using the relaxation model, it is possible to estimate the volumes (V), and hence the characteristic sizes of the magnetic ordering regions, if we use the values of the effective magnetic anisotropy coefficients Keff of nanoparticles at different annealing temperatures. The values of the coefficients Keff were estimated using the literature data on the magnetic anisotropy coefficients for stoichiometric magnetite (Fe3O4) and maghemite (γ-Fe2O3), as well as the values of the molar concentration of maghemite b obtained from the model fitting of the Mössbauer spectra [34].
Using SpectrRelax, a program for processing and analyzing Mössbauer spectra [35], a model was implemented for deciphering the Mössbauer spectra of iron oxides in the form of nanoparticles of a mixture of magnetite (Fe3O4) and maghemite (γ-Fe2O3) or nanoparticles of nonstoichiometric magnetite (Fe3–γO4) in the presence of fast electron exchange, taking into account multilevel superparamagnetic relaxation for Fe atoms in various structural and charge states [34]. Thus, when fitting the spectra of nanoparticles, we used a model consisting of three interconnected relaxation subspectra of nonstoichiometric magnetite (corresponding to Fe A 3 + , Fe B 3 + and Fe B 2.5 + ), to which an independent Zeeman sextet corresponding to hematite was added. The use of this model for fitting the spectra made it possible to establish changes in the following characteristics with the annealing temperature: the molar concentration of maghemite b, the magnetite nonstoichiometry degree γ, the magnetic anisotropy energy Ema, the magnetic anisotropy coefficient Keff, and the size of the region of magnetic ordering of iron atoms d in nonstoichiometric magnetite. As can be seen in Figure 8 and Figure 9, all experimental spectra are well-described within the model used: the normalized chi-squared values for Fe3O4 nanoparticles range from 0.96 to 1.36, and for Fe3O4@Au nanoparticles, from 0.97 to 1.20. It should be noted that slow superparamagnetic relaxation was observed for all nanoparticles when the relaxation time was noticeably longer (by about 2 orders of magnitude) than the lifetime of the 57Fe nucleus in the excited state (the relaxation rate R was noticeably smaller than the natural width of the excited state level).
Figure 11 shows the relative intensities of subspectra corresponding to different states of Fe atoms in Fe3O4 and Fe3O4@Au nanoparticles, depending on the annealing temperature. An analysis of the obtained data showed that at annealing temperatures below 300 °C, Fe3O4 nanoparticles are non-stoichiometric magnetite Fe3−γO4, while with an increase in the annealing temperature in the octahedral position (B) of the inverse spinel structure, oxidation of iron atoms occurs—an increase in the relative number of Fe3+ ions due to a decrease in the number of Fe2.5+ ions. Above the annealing temperature of 300 °C, nonstoichiometric magnetite in Fe3O4 nanoparticles transforms into hematite (α-Fe2O3).
For Fe3O4@Au nanoparticles, the same sequence of phase changes is observed, only at higher annealing temperatures, by approximately 150 °C. It should be noted that the dependences of the relative contributions to the Mössbauer spectrum (MS) and X-ray diffraction pattern (XRD) of magnetite and hematite in nanoparticles on the annealing temperature are in good agreement with each other (see Figure 12).
Figure 13 shows data on non-stoichiometric magnetite depending on the annealing temperature, obtained as a result of model fitting of the Mössbauer spectra of the studied nanoparticles. For the initial Fe3O4 nanoparticles, the molar concentration of maghemite γ-Fe2O3 was b = 0.49 ± 0.01, and the magnetite nonstoichiometry degree was γ = 0.165 ± 0.004. Nonstoichiometric magnetite in the initial Fe3O4 nanoparticles is completely oxidized to maghemite γ-Fe2O3 at tann ≥ 200 °C. The core of the initial Fe3O4@Au nanoparticles is mainly maghemite (b = 0.98 ± 0.02, γ = 0.326 ± 0.008) (Figure 13a), since the oxidation of magnetite in it occurs already in the synthesis process.
The parameter of the multilevel superparamagnetic relaxation model α, equal to the ratio of the magnetic anisotropy energy to the thermal energy (4), and the magnetic anisotropy energy Ema as a function of the annealing temperature are shown in Figure 13b. It can be seen that for Fe3O4@Au nanoparticles, the energy Ema increases monotonically, while for Fe3O4 nanoparticles, a rather “convoluted” dependence is observed at low annealing temperatures. This behavior of the dependence Ema(tann) can be explained if we take into account the oxidation of nanoparticles not coated with gold. In accordance with the data on the molar concentration of maghemite b, the effective coefficients of magnetic anisotropy of nanoparticles Keff were determined, the results of which are shown in Figure 13c. As can be seen, for Fe3O4@Au nanoparticles, the Keff coefficient practically remains unchanged, while for Fe3O4 nanoparticles it sharply decreases with temperature in accordance with the oxidation of nonstoichiometric magnetite to maghemite (Figure 13a). If this value differs for magnetic nanoparticles from the bulk samples of magnetite presented in [43,44,45], this may be due to size factors, as well as structural disordering of the initial magnetite nanoparticles, which also leads to a change in the magnetic parameters.
Using the results of estimates of the effective coefficient of magnetic anisotropy, in accordance with (4), the average sizes of the magnetic ordering regions of Fe atoms in nonstoichiometric magnetite for Fe3O4 and Fe3O4@Au nanoparticles were calculated; they are presented in Figure 13d. The assessment of the “magnetic ordering region” for uncoated Fe3O4 samples annealed at 500–800 °C and for coated Fe3O4@Au samples at 600–800 °C was not carried out due to the small contribution of Fe3−γO4 to the experimental spectrum (see Figure 9e,f), making it impossible to reliably find the values of the relaxation model parameter α (see Formula (4) and Figure 9e,f). It is evident that with an increase in the annealing temperature, the average sizes of these regions gradually increase both due to the oxidation of nanoparticles and due to their structural and magnetic ordering. Note that the average size of the magnetic ordering region increases from 16.4 ± 0.4 to 18.1 ± 0.9 nm as a result of gold coating of the initial nanoparticles, which is associated with oxidation during gold coating. At the same time, the obtained values of the average size of the magnetic ordering region are in good agreement with the data of scanning electron microscopy (Figure 2a).
As a result of model fitting of the Mössbauer spectra of Fe3O4 and Fe3O4@Au nanoparticles subjected to thermal annealing, the values of hyperfine parameters of subspectra (hyperfine magnetic fields Hn, isomer shifts δ, and quadrupole shifts ɛ of resonance lines) were obtained, which made it possible to unambiguously identify subspectra and analyze their dependences on the annealing temperature.
Figure 14, which demonstrates the hyperfine parameters of the subspectra of nonstoichiometric magnetite Fe 3 γ O 4 , shows that the isomer shifts δ for trivalent iron ions vary in the ranges of 0.19–0.26 mm/s and 0.39–0.43 mm/s (Figure 14b), which is typical for tetrahedral (A) and octahedral (B) oxygen environments in the Fe 3 γ O 4 structure.
The hyperfine magnetic fields Hn on the 57Fe nuclei for trivalent iron ions in the tetra- and octahedral positions are quite close ( H n = H n B H n A 4 kOe) and, unlike the other hyperfine parameters for these ions, at tann ≤ 400 °C, they noticeably increase with increasing annealing temperature, which indicates both improvement of the structure as well as an increase in the magnetite nonstoichiometry degree. This is also evidenced by the observed small changes in the isomer shifts of the subspectra of nonstoichiometric magnetite (Figure 14b). As for the quadrupole shifts, their values turned out to be close to zero (Figure 14c).
Dependences of the hyperfine parameters of the subspectrum of hematite α-Fe2O3 in Fe3O4 and Fe3O4@Au nanoparticles on the annealing temperature are shown in Figure 15.
It can be seen that the isomer shift and the quadrupole shift of the resonance lines practically do not change, and the hyperfine magnetic field somewhat increases with an increase in the annealing temperature, approaching a value corresponding to the literature data for pure bulk hematite. Thus, we can conclude that with an increase in the annealing temperature, the crystal and magnetic structure of hematite is improved.

4. Conclusions

Fe3O4 and Fe3O4@Au nanoparticles were studied using scanning and transmission electron microscopy, X-ray diffraction, and Mössbauer spectroscopy, as a result of which a sequence of phase transformations and a change in the morphology and structure of nanoparticles depending on the annealing temperature were established. Using the Mössbauer spectroscopy method, the nonstoichiometry degree γ of nonstoichiometric magnetite Fe3−γO4 in the studied nanoparticles was determined. It has been established that the initial Fe3O4 nanoparticles with sizes of 10–18 nm and a shape close to spherical are nonstoichiometric magnetite nanoparticles with a stoichiometry degree γ = 0.165 ± 0.004. Initial Fe3O4@Au nanoparticles have a core-shell structure with a 2–3 nm thick gold shell and a core consisting mainly of extremely oxidized magnetite—maghemite γ-Fe2O3 (γ = 0.326 ± 0.008). At low annealing temperatures tann (100–400 °C), Fe3O4 nanoparticles stick together, forming particle agglomerates with an average size of ~45 nm that are easily dispersed under ultrasonic treatment. Higher annealing temperatures lead to sticking and coalescence of the initial dispersed and non-dispersed nanoparticles with the formation of larger nanoparticles (up to ~90 nm at 800 °C) consisting of hematite α-Fe2O3. The presence of a gold shell in Fe3O4@Au nanoparticles prevents their association. In this case, the transformation of nonstoichiometric magnetite into hematite for Fe3O4@Au nanoparticles occurs at annealing temperatures ~150 °C higher than for Fe3O4 nanoparticles (at ~300 °C).
It has been shown by Mössbauer spectroscopy that nonstoichiometric magnetite nanoparticles with and without gold coating are in the superparamagnetic state with a slow relaxation rate compared to the reciprocal lifetime of the 57Fe core in the excited state. Depending on the annealing temperature, the energy of the magnetic anisotropy of nonstoichiometric magnetite was determined and an estimate was made of the average size of the region of magnetic ordering of Fe atoms in nonstoichiometric magnetite, which is in good agreement with the data on the sizes of nanoparticles determined by scanning electron microscopy.

Author Contributions

Conceptualization, V.S.R., A.L.K., M.S.F., K.B.E., A.N., K.K.K., D.I.S., M.V.Z.; methodology V.S.R., A.L.K., M.S.F., K.B.E., A.N., K.K.K., D.I.S., M.V.Z.; formal analysis, V.S.R., A.L.K., M.S.F., K.B.E., A.N., K.K.K., D.I.S., M.V.Z.; investigation, K.K.K., A.L.K., V.S.R., M.S.F., K.B.E., A.N.; resources, K.K.K., A.L.K.; writing—original draft preparation, review and editing, V.S.R., M.S.F., A.L.K.; visualization, K.K.K., A.L.K., V.S.R., M.S.F.; supervision, K.K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Ministry of Education and Science of the Republic of Kazakhstan (grant AP09259184). The authors V.S.R. and M.S.F. thank M.V. Lomonosov Moscow State University for updating the equipment for scientific research used in the performance of the work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, S.; Sun, A.; Zhai, F.; Wang, J.; Xu, W.; Zhang, Q.; Volinsky, A.A. Fe3O4 magnetic nanoparticles synthesis from tailings by ultrasonic chemical co-precipitation. Mater. Lett. 2011, 65, 1882–1884. [Google Scholar] [CrossRef]
  2. Ali, A.; Shah, T.; Ullah, R.; Zhou, P.; Guo, M.; Ovais, M.; Rui, Y. Review on recent progress in magnetic nanoparticles: Synthesis, characterization, and diverse applications. Front. Chem. 2021, 9, 629054. [Google Scholar] [CrossRef] [PubMed]
  3. Shukla, S.; Khan, R.; Daverey, A. Synthesis and characterization of magnetic nanoparticles, and their applications in wastewater treatment: A review. Environ. Technol. Innov. 2021, 24, 101924. [Google Scholar] [CrossRef]
  4. Siavashy, S.; Soltani, M.; Ghorbani-Bidkorbeh, F.; Fallah, N.; Farnam, G.; Mortazavi, S.A.; Hamedi, M.H. Microfluidic platform for synthesis and optimization of chitosan-coated magnetic nanoparticles in cisplatin delivery. Carbohydr. Polym. 2021, 265, 118027. [Google Scholar] [CrossRef]
  5. Irfan, M.; Dogan, N.; Bingolbali, A.; Aliew, F. Synthesis and characterization of NiFe2O4 magnetic nanoparticles with different coating materials for magnetic particle imaging (MPI). J. Magn. Magn. Mater. 2021, 537, 168150. [Google Scholar] [CrossRef]
  6. Adewunmi, A.A.; Kamal, M.S.; Solling, T.I. Application of magnetic nanoparticles in demulsification: A review on synthesis, performance, recyclability, and challenges. J. Pet. Sci. Eng. 2021, 196, 107680. [Google Scholar] [CrossRef]
  7. Mollarasouli, F.; Zor, E.; Ozcelikay, G.; Ozkan, S.A. Magnetic nanoparticles in developing electrochemical sensors for pharmaceutical and biomedical applications. Talanta 2021, 226, 122108. [Google Scholar] [CrossRef]
  8. Sanad, M.M.S.; Azab, A.A.; Taha, T.A. Inducing lattice defects in calcium ferrite anode materials for improved electrochemical performance in lithium-ion batteries. Ceram. Int. 2022, 48, 12537–12548. [Google Scholar] [CrossRef]
  9. Golmohammad, M.; Mirhabibi, A.; Golestanifard, F.; Kelder, E.M. Optimizing synthesis of maghemite nanoparticles as an anode for Li-ion batteries by exploiting design of experiment. J. Electron. Mater. 2016, 45, 426–434. [Google Scholar] [CrossRef]
  10. Esmi, F.; Nematian, T.; Salehi, Z.; Khodadadi, A.A.; Dalai, A.K. Amine and aldehyde functionalized mesoporous silica on magnetic nanoparticles for enhanced lipase immobilization, biodiesel production, and facile separation. Fuel 2021, 291, 120126. [Google Scholar] [CrossRef]
  11. Anik, M.I.; Hossain, M.K.; Hossain, I.; Mahfuz, A.M.U.B.; Rahman, M.T.; Ahmed, I. Recent progress of magnetic nanoparticles in biomedical applications: A review. Nano Sel. 2021, 2, 1146–1186. [Google Scholar] [CrossRef]
  12. Li, R.N.; Da, X.H.; Li, X.; Lu, Y.S.; Gu, F.F.; Liu, Y. Functionalized magnetic nanoparticles for drug delivery in tumor therapy. Chin. Phys. B 2021, 30, 017502. [Google Scholar] [CrossRef]
  13. Keçili, R.; Ghorbani-Bidkorbeh, F.; Dolak, İ.; Canpolat, G.; Karabörk, M.; Hussain, C.M. Functionalized magnetic nanoparticles as powerful sorbents and stationary phases for the extraction and chromatographic applications. TrAC Trends Anal. Chem. 2021, 143, 116380. [Google Scholar] [CrossRef]
  14. Flores-Rojas, G.G.; López-Saucedo, F.; Vera-Graziano, R.; Mendizabal, E.; Bucio, E. Magnetic Nanoparticles for Medical Applications: Updated Review. Macromol 2022, 2, 374–390. [Google Scholar] [CrossRef]
  15. Ivanova, A.V.; Nikitin, A.A.; Gabashvily, A.N.; Vishnevskiy, D.A.; Abakumov, M.A. Synthesis and intensive analysis of antibody labeled single core magnetic nanoparticles for targeted delivery to the cell membrane. J. Magn. Magn. Mater. 2021, 521, 167487. [Google Scholar] [CrossRef]
  16. Lavorato, G.C.; Das, R.; Masa, J.A.; Phan, M.H.; Srikanth, H. Hybrid magnetic nanoparticles as efficient nanoheaters in biomedical applications. Nanoscale Adv. 2021, 3, 867–888. [Google Scholar] [CrossRef]
  17. Ilosvai, A.M.; Dojcsak, D.; Váradi, C.; Nagy, M.; Kristály, F.; Fiser, B.; Vanyorek, L. Sonochemical Combined Synthesis of Nickel Ferrite and Cobalt Ferrite Magnetic Nanoparticles and Their Application in Glycan Analysis. Int. J. Mol. Sci. 2022, 23, 5081. [Google Scholar] [CrossRef]
  18. Available online: https://www.webofscience.com/ (accessed on 20 September 2022).
  19. Kolosnjaj-Tabi, J.; Javed, Y.; Lartigue, L.; Volatron, J.; Elgrabli, D.; Marangon, I.; Gazeau, F. The one year fate of iron oxide coated gold nanoparticles in mice. ACS Nano 2015, 9, 7925–7939. [Google Scholar] [CrossRef] [Green Version]
  20. Pajor-Świerzy, A.; Szczepanowicz, K.; Kamyshny, A.; Magdassi, S. Metallic core-shell nanoparticles for conductive coatings and printing. Adv. Colloid Interface Sci. 2022, 299, 102578. [Google Scholar] [CrossRef]
  21. Bayles, A.; Tian, S.; Zhou, J.; Yuan, L.; Yuan, Y.; Jacobson, C.R.; Halas, N.J. Al@TiO2 Core–Shell Nanoparticles for Plasmonic Photocatalysis. ACS Nano 2022, 16, 5839–5850. [Google Scholar] [CrossRef]
  22. Ovchinnikov, O.V.; Smirnov, M.S.; Chevychelova, T.A.; Zvyagin, A.I.; Selyukov, A.S. Nonlinear absorption enhancement of Methylene Blue in the presence of Au/SiO2 core/shell nanoparticles. Dye. Pigment. 2022, 197, 109829. [Google Scholar] [CrossRef]
  23. Hodak, J.H.; Henglein, A.; Giersig, M.; Hartland, G.V. Laser-induced inter-diffusion in AuAg core−shell nanoparticles. J. Phys. Chem. B 2000, 104, 11708–11718. [Google Scholar] [CrossRef]
  24. Billen, A.; de Cattelle, A.; Jochum, J.K.; Van Bael, M.J.; Billen, J.; Seo, J.W.; Verbiest, T. Novel synthesis of superparamagnetic plasmonic core-shell iron oxide-gold nanoparticles. Phys. B Condens. Matter 2019, 560, 85–90. [Google Scholar] [CrossRef]
  25. Beik, J.; Asadi, M.; Khoei, S.; Laurent, S.; Abed, Z.; Mirrahimi, M.; Shakeri-Zadeh, A. Simulation-guided photothermal therapy using MRI-traceable iron oxide-gold nanoparticle. J. Photochem. Photobiol. B Biol. 2019, 199, 111599. [Google Scholar] [CrossRef] [PubMed]
  26. Li, J.F.; Zhang, Y.J.; Ding, S.Y.; Panneerselvam, R.; Tian, Z.Q. Core–shell nanoparticle-enhanced Raman spectroscopy. Chem. Rev. 2017, 117, 5002–5069. [Google Scholar] [CrossRef]
  27. Ding, H.L.; Zhang, Y.X.; Wang, S.; Xu, J.M.; Xu, S.C.; Li, G.H. Fe3O4@SiO2 core/shell nanoparticles: The silica coating regulations with a single core for different core sizes and shell thicknesses. Chem. Mater. 2012, 24, 4572–4580. [Google Scholar] [CrossRef]
  28. Llamosa, D.; Ruano, M.; Martínez, L.; Mayoral, A.; Roman, E.; García-Hernández, M.; Huttel, Y. The ultimate step towards a tailored engineering of core@ shell and core@shell@shell nanoparticles. Nanoscale 2014, 6, 13483–13486. [Google Scholar] [CrossRef]
  29. Wei, S.; Wang, Q.; Zhu, J.; Sun, L.; Lin, H.; Guo, Z. Multifunctional composite core–shell nanoparticles. Nanoscale 2011, 3, 4474–4502. [Google Scholar] [CrossRef]
  30. Ghosh Chaudhuri, R.; Paria, S. Core/shell nanoparticles: Classes, properties, synthesis mechanisms, characterization, and applications. Chem. Rev. 2012, 112, 2373–2433. [Google Scholar] [CrossRef]
  31. Atar, N.; Eren, T.; Yola, M.L.; Karimi-Maleh, H.; Demirdögen, B. Magnetic iron oxide and iron oxide@gold nanoparticle anchored nitrogen and sulfur-functionalized reduced graphene oxide electrocatalyst for methanol oxidation. RSC Adv. 2015, 5, 26402–26409. [Google Scholar] [CrossRef]
  32. Ali Dheyab, M.; Aziz, A.A.; Jameel, M.S. Recent advances in inorganic nanomaterials synthesis using sonochemistry: A comprehensive review on iron oxide, gold and iron oxide coated gold nanoparticles. Molecules 2021, 26, 2453. [Google Scholar] [CrossRef] [PubMed]
  33. Dheyab, M.A.; Aziz, A.A.; Jameel, M.S.; Noqta, O.A.; Mehrdel, B. Synthesis and coating methods of biocompatible iron oxide/gold nanoparticle and nanocomposite for biomedical applications. Chin. J. Phys. 2020, 64, 305–325. [Google Scholar] [CrossRef]
  34. Fadeev, M.; Kozlovskiy, A.; Korolkov, I.; Egizbek, K.; Nazarova, A.; Chudoba, D.; Zdorovets, M. Iron oxide@gold nanoparticles: Synthesis, properties and potential use as anode materials for lithium-ion batteries. Colloids Surf. A Physicochem. Eng. Asp. 2020, 603, 125178. [Google Scholar] [CrossRef]
  35. Matsnev, M.E.; Rusakov, V.S. SpectrRelax: An application for Mössbauer spectra modeling and fitting. AIP Conf. Proc. 2012, 1489, 178–185. [Google Scholar]
  36. Yang, J.B.; Zhou, X.D.; Yelon, W.B.; James, W.J.; Cai, Q.; Gopalakrishnan, K.V.; Nikles, D.E. Magnetic and structural studies of the Verwey transition in Fe3−δO4 nanoparticles. J. Appl. Phys. 2004, 95, 7540–7542. [Google Scholar] [CrossRef]
  37. Schmidbauer, E.; Keller, M. Magnetic hysteresis properties, Mossbauer spectra and structural data of spherical 250 nm particles of solid solutions Fe3O4−γ-Fe2O3. Magn. Magn. Mater. 2006, 297, 107–117. [Google Scholar] [CrossRef]
  38. Rozenberg, G.K.; Dubrovinsky, L.S.; Pasternak, M.P.; Naaman, O.; Le Bihan, T.; Ahuja, R. High-pressure structural studies of hematite Fe2O3. Phys. Rev. B 2002, 65, 064112. [Google Scholar] [CrossRef]
  39. Stary, O.; Malyshev, A.V.; Lysenko, E.N.; Petrova, A. Formation of magnetic properties of ferrites during radiation-thermal sintering. Eurasian Phys. Tech. J. 2021, 17, 6–10. [Google Scholar] [CrossRef]
  40. Verwey, E.J.W. Electronic conduction of magnetite (Fe3O4) and its transition point at low temperatures. Nature 1939, 144, 327. [Google Scholar] [CrossRef]
  41. Sawatzky, G.A.; Van Der Woude, F.; Morrish, A.H. Recoilless-fraction ratios for Fe57 in octahedral and tetrahedral sites of a spinel and a garnet. Phys. Rev. 1969, 183, 383–386. [Google Scholar] [CrossRef]
  42. Jones, D.H.; Srivastava, K.K.P. Many-state relaxation model for the mossbauer spectra of superparamagnets. Phys. Rev. B 1986, 34, 7542–7548. [Google Scholar] [CrossRef] [PubMed]
  43. Berger, R.; Kliava, J.; Bissey, J.C.; Baïetto, V. Magnetic resonance of superparamagnetic iron-containing nanoparticles in annealed glass. J. Appl. Phys. 2000, 87, 7389–7396. [Google Scholar] [CrossRef]
  44. Lin, C.R.; Chiang, R.K.; Wang, J.S.; Sung, T.W. Magnetic properties of monodisperse iron oxide nanoparticles. J. Appl. Phys. 2006, 99, 08N710. [Google Scholar] [CrossRef]
  45. Balaev, D.A.; Semenov, S.V.; Dubrovskiy, A.A.; Yakushkin, S.S.; Kirillov, V.L.; Martyanov, O.N. Superparamagnetic blocking of an ensemble of magnetite nanoparticles upon interparticle interactions. J. Magn. Magn. Mater. 2017, 440, 199–202. [Google Scholar] [CrossRef]
Figure 1. SEM images and size distributions of the studied Initial (initial samples) and Annealed (annealing at tann) Fe3O4 and Fe3O4@Au nanoparticles.
Figure 1. SEM images and size distributions of the studied Initial (initial samples) and Annealed (annealing at tann) Fe3O4 and Fe3O4@Au nanoparticles.
Nanomaterials 12 04121 g001aNanomaterials 12 04121 g001b
Figure 2. Results of evaluation of average nanoparticle sizes (a) and standard deviation of nanoparticle size distribution (b) as a function of annealing temperature tann.
Figure 2. Results of evaluation of average nanoparticle sizes (a) and standard deviation of nanoparticle size distribution (b) as a function of annealing temperature tann.
Nanomaterials 12 04121 g002
Figure 3. (a) TEM images of studied initial (Initial) and annealed (Annealing) (a) Fe3O4 nanoparticles at temperatures of 400 °C (b) and 600 °C (c).
Figure 3. (a) TEM images of studied initial (Initial) and annealed (Annealing) (a) Fe3O4 nanoparticles at temperatures of 400 °C (b) and 600 °C (c).
Nanomaterials 12 04121 g003
Figure 4. (a) TEM image of Fe3O4@Au nanoparticles (inset shows high-resolution TEM image of Fe3O4@Au nanoparticles); (b) mapping results of Fe3O4@Au nanoparticles.
Figure 4. (a) TEM image of Fe3O4@Au nanoparticles (inset shows high-resolution TEM image of Fe3O4@Au nanoparticles); (b) mapping results of Fe3O4@Au nanoparticles.
Nanomaterials 12 04121 g004
Figure 5. X-ray diffraction patterns of initial and annealed at temperatures of 400 °C, 500 °C, 600 °C, and 800 °C Fe3O4 and Fe3O4@Au nanoparticles.
Figure 5. X-ray diffraction patterns of initial and annealed at temperatures of 400 °C, 500 °C, 600 °C, and 800 °C Fe3O4 and Fe3O4@Au nanoparticles.
Nanomaterials 12 04121 g005
Figure 6. Dependences of relative intensities I of diffraction patterns of nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 (a), as well as the average size d of the regions of structural ordering of nanoparticles of nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 (b) on the annealing temperature tann.
Figure 6. Dependences of relative intensities I of diffraction patterns of nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 (a), as well as the average size d of the regions of structural ordering of nanoparticles of nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 (b) on the annealing temperature tann.
Nanomaterials 12 04121 g006
Figure 7. Unit cell parameters for nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 (a), as well as ratios of unit cell parameters a/c of hematite α-Fe2O3 (b) depending on the annealing temperature tann.
Figure 7. Unit cell parameters for nonstoichiometric magnetite Fe3−γO4 and hematite α-Fe2O3 (a), as well as ratios of unit cell parameters a/c of hematite α-Fe2O3 (b) depending on the annealing temperature tann.
Nanomaterials 12 04121 g007
Figure 8. Mössbauer spectra of the initial (a) and annealed at temperatures of 100 °C (b), 200 °C (c), 400 °C (d), 600 °C (e), 800 °C (f) Fe3O4 nanoparticles (on each spectrum the positions of the resonance lines of the subspectra used for model fitting are shown; the lines corresponding to magnetite are shown in blue, and the lines corresponding to hematite are shown in orange).
Figure 8. Mössbauer spectra of the initial (a) and annealed at temperatures of 100 °C (b), 200 °C (c), 400 °C (d), 600 °C (e), 800 °C (f) Fe3O4 nanoparticles (on each spectrum the positions of the resonance lines of the subspectra used for model fitting are shown; the lines corresponding to magnetite are shown in blue, and the lines corresponding to hematite are shown in orange).
Nanomaterials 12 04121 g008aNanomaterials 12 04121 g008b
Figure 9. Mössbauer spectra of the initial (a) and annealed at temperatures of 100 °C (b), 200 °C (c), 400 °C (d), 600 °C (e), 800 °C (f) Fe3O4@Au nanoparticles (on each spectrum shows the positions of the resonance lines of the subspectra used for model fitting; the lines corresponding to magnetite are shown in blue, and the lines corresponding to hematite are shown in orange).
Figure 9. Mössbauer spectra of the initial (a) and annealed at temperatures of 100 °C (b), 200 °C (c), 400 °C (d), 600 °C (e), 800 °C (f) Fe3O4@Au nanoparticles (on each spectrum shows the positions of the resonance lines of the subspectra used for model fitting; the lines corresponding to magnetite are shown in blue, and the lines corresponding to hematite are shown in orange).
Nanomaterials 12 04121 g009aNanomaterials 12 04121 g009b
Figure 10. Diagram of limiting options for representing a magnetite nanoparticle with a spatially heterogeneous nonstoichiometry degree.
Figure 10. Diagram of limiting options for representing a magnetite nanoparticle with a spatially heterogeneous nonstoichiometry degree.
Nanomaterials 12 04121 g010
Figure 11. Relative intensities of subspectra corresponding to different states of Fe atoms in Fe3O4 and Fe3O4@Au nanoparticles, depending on the annealing temperature tann.
Figure 11. Relative intensities of subspectra corresponding to different states of Fe atoms in Fe3O4 and Fe3O4@Au nanoparticles, depending on the annealing temperature tann.
Nanomaterials 12 04121 g011
Figure 12. Dependences of the relative contributions to the Mössbauer spectrum (MS) and X-ray diffraction pattern (XRD) of nonstoichiometric magnetite and hematite in Fe3O4 and Fe3O4@Au nanoparticles on the annealing temperature tann.
Figure 12. Dependences of the relative contributions to the Mössbauer spectrum (MS) and X-ray diffraction pattern (XRD) of nonstoichiometric magnetite and hematite in Fe3O4 and Fe3O4@Au nanoparticles on the annealing temperature tann.
Nanomaterials 12 04121 g012
Figure 13. Dependences on the annealing temperature tann of the molar concentration of maghemite b and the nonstoichiometry degree γ (a), the parameter of the multilevel superparamagnetic relaxation model α and the magnetic anisotropy energy Ema (b), the magnetic anisotropy coefficient Keff (c), and the average size of the magnetic ordering region d (d) for nonstoichiometric magnetite in Fe3O4 and Fe3O4@Au nanoparticles.
Figure 13. Dependences on the annealing temperature tann of the molar concentration of maghemite b and the nonstoichiometry degree γ (a), the parameter of the multilevel superparamagnetic relaxation model α and the magnetic anisotropy energy Ema (b), the magnetic anisotropy coefficient Keff (c), and the average size of the magnetic ordering region d (d) for nonstoichiometric magnetite in Fe3O4 and Fe3O4@Au nanoparticles.
Nanomaterials 12 04121 g013aNanomaterials 12 04121 g013b
Figure 14. Dependences on the annealing temperature tann of the hyperfine magnetic field Hn (a), the isomer shift δ (b), and the quadrupole shift ɛ (c) for nonstoichiometric magnetite in Fe3O4 and Fe3O4@Au nanoparticles.
Figure 14. Dependences on the annealing temperature tann of the hyperfine magnetic field Hn (a), the isomer shift δ (b), and the quadrupole shift ɛ (c) for nonstoichiometric magnetite in Fe3O4 and Fe3O4@Au nanoparticles.
Nanomaterials 12 04121 g014aNanomaterials 12 04121 g014b
Figure 15. Dependences on the annealing temperature tann of the hyperfine magnetic field Hn (a), the isomer shift δ, and the quadrupole shift ɛ (b) for hematite α-Fe2O3 in Fe3O4 and Fe3O4@Au nanoparticles.
Figure 15. Dependences on the annealing temperature tann of the hyperfine magnetic field Hn (a), the isomer shift δ, and the quadrupole shift ɛ (b) for hematite α-Fe2O3 in Fe3O4 and Fe3O4@Au nanoparticles.
Nanomaterials 12 04121 g015
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rusakov, V.S.; Kozlovskiy, A.L.; Fadeev, M.S.; Egizbek, K.B.; Nazarova, A.; Kadyrzhanov, K.K.; Shlimas, D.I.; Zdorovets, M.V. Study of Phase Transformations and Hyperfine Interactions in Fe3O4 and Fe3O4@Au Nanoparticles. Nanomaterials 2022, 12, 4121. https://doi.org/10.3390/nano12234121

AMA Style

Rusakov VS, Kozlovskiy AL, Fadeev MS, Egizbek KB, Nazarova A, Kadyrzhanov KK, Shlimas DI, Zdorovets MV. Study of Phase Transformations and Hyperfine Interactions in Fe3O4 and Fe3O4@Au Nanoparticles. Nanomaterials. 2022; 12(23):4121. https://doi.org/10.3390/nano12234121

Chicago/Turabian Style

Rusakov, Vyacheslav S., Artem L. Kozlovskiy, Maxim S. Fadeev, Kamila B. Egizbek, Assel Nazarova, Kayrat K. Kadyrzhanov, Dmitriy I. Shlimas, and Maxim V. Zdorovets. 2022. "Study of Phase Transformations and Hyperfine Interactions in Fe3O4 and Fe3O4@Au Nanoparticles" Nanomaterials 12, no. 23: 4121. https://doi.org/10.3390/nano12234121

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