Next Article in Journal
Characterization of Enamel and Dentine about a White Spot Lesion: Mechanical Properties, Mineral Density, Microstructure and Molecular Composition
Previous Article in Journal
Ta Doping Effect on Structural and Optical Properties of InTe Thin Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic and Magneto-Optical Oroperties of Iron Oxides Nanoparticles Synthesized under Atmospheric Pressure

1
Department of Applied Physics, National Pingtung University, Pingtung County 90003, Taiwan
2
Research Institute of Physics, Southern Federal University, Rostov-on-Don 344090, Russia
3
Department of Chemical and Materials Engineering, Southern Taiwan University of Science and Technology, Tainan city 710, Taiwan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(9), 1888; https://doi.org/10.3390/nano10091888
Submission received: 1 September 2020 / Revised: 12 September 2020 / Accepted: 17 September 2020 / Published: 21 September 2020

Abstract

:
Magnetite nanoparticles were synthesized by a simple thermal decomposition process, involving only iron (III) nitrate nonahydrate as a precursor, and hexadecylamine as a solvent and stabilizer at reaction temperatures varied from 200 to 380 °C. The results of the structural analysis showed that the average crystallite size depends on the reaction temperature and increases from 4.8 to 13.3 nm. The behavior of the coercivity indicates that all synthesized samples are single domain; herewith, it was found that the critical size corresponding to the transition to the superparamagnetic state at room temperature is about 9 nm. The effect of the reaction temperature on changes in the saturation magnetization was studied. It was found that the size effect in the MCD spectra is observed for the IVCT transition and one ISCT transition, and the influence of the reaction temperature on the change in the MCD spectra was discussed.

Graphical Abstract

1. Introduction

Over the past decade, there has been an increasing interest in iron oxide nanoparticles (IONs), caused not only by fundamental scientific interest but also by their practical application. Due to the unique size-dependent magnetic properties, IONs are used in a wide variety of fields, including nanocatalysis, all-optical switching devices, microwave absorption, environmental remediation, chemical and biological sensors, energy generation and storage, biomedicine and biotechnologies, etc. [1,2,3,4,5,6,7,8,9,10,11,12]. Among various IONs, magnetite (Fe3O4) nanoparticles are one of the most promising, since they find application in such areas as high-density magnetic storage media, capacitor electrodes, development of sensors, ferrofluid, for absorption heavy metals [5,13,14,15,16,17,18,19], have wide application in biotechnologies and biomedicine, for example, magnetic resonance imaging, hyperthermia, targeted delivery of drugs, detect and treat cancer [20,21,22,23,24,25,26].
Many different methods for the synthesis of magnetite nanoparticles have been described in the literature, including co-precipitation [27,28,29], hydrothermal [30,31,32], sol-gel [33,34,35], thermal decomposition [36,37,38,39,40], among others [41,42,43]. Besides, the key role of the synthesis conditions, such as reaction temperature, reagents concentration, the effect of different additives, and others on the magnetic properties and size control have also been discussed in the literature. Such methods as sol-gel and hydrothermal have a long reaction time and/or occur at high pressure [44,45], which reduces the possibility of their widespread use. Among these techniques, the method of thermal decomposition (TD) of various iron complexes using solvents with high boiling points has demonstrated promising results for obtaining of size-controlled, superparamagnetic magnetite nanoparticles. The most common iron complex for the TD method is iron (III) acetylacetonate (Fe(acac)3) [39,46,47]; in addition, the use of other precursors [37,48,49] has also been described in the literature. However, most of the thermal decomposition processes use expensive reagent mixtures, for example, with 1,2-hexadecanediol or they proceed under complicated conditions. All this complicates the development of magnetite nanoparticles and their study for practical application. Therefore, the development of simple synthesis methods without the use of expensive reagents and the study of the influence of synthesis conditions on the physicochemical properties of magnetite nanoparticles is an important task contributing to the progress of their practical application.
In this regard, the use of other solvents and iron complexes is of particular interest and has been investigated in the literature [36,37,39,47]. Therefore, hexadecylamine (HDA) has been used as a solvent, stabilizing and reducing agent for the synthesis of nanoparticles of silver [50], nickel [51], bismuth [52], and others. Authors of work [53] have synthesized superparamagnetic magnetite nanoparticles via the solvothermal method by the reduction of Fe(acac)3 using a mixture with HDA as a stabilizer. However, the uses of other synthesis methods and precursors, as well as the influence of synthesis conditions on the formation of magnetite nanoparticles, have not been described in the literature.
In this work, we present a simple and low-cost method for the synthesis of magnetite nanoparticles by thermal decomposition process involving only two chemicals: iron (III) nitrate nonahydrate (Fe(NO3)3·9H2O) and HDA, which was used as a solvent and stabilizer. As far as we know, the thermal decomposition process of Fe(NO3)3·9H2O using HDA has not been described previously. Besides, the synthesis process proceeds without the use of protective gas (such as nitrogen or argon) and at atmospheric pressure, making it easily reproducible. The influence of the reaction temperature on the phase purity, magnetic and magneto-optical properties of magnetite nanoparticles has been investigated.

2. Materials and Methods

2.1. Synthesis of Magnetite Nanoparticles

Magnetite nanoparticles were synthesized by a thermal pyrolysis method, using HDA as a stabilizing agent. Analytically pure reagents were used as the starting materials without further purification. In a typical synthesis process, 12.2 g of 1–hexadecylamine (Fisher Scientific International, Inc., Pittsburgh, PA, USA) was melted at 80 °C in a round-bottom flask, and then 2.9 g of iron (III) nitrate nonahydrate (Merck Millipore, Burlington, MA, USA) (Fe(NO3)3·9H2O) was dissolved in it. To remove water from iron nitrate, the mixture was heated to 120 °C for 30 min under magnetic stirring. To obtain nanoparticles with different sizes, the mixture was heated to the final temperature of the reaction (TR), which was varied from 200 to 380 °C and held at appropriate temperatures for 1 h. After completion of the reaction, the resulting black solution was cooled to 70 °C and washed several times with toluene, which ensures complete removal of HDA from the samples.

2.2. Characterizations

The crystal structure and morphology of the nanoparticles were characterized by X-ray diffraction (XRD) measurements using a Bruker D8 Advance diffractometer (Bruker Corp., Billerica, MA, USA) (Cu Kα radiation, 40 kV, 25 mA, λ = 1.5418 Å) and transmission electron microscopy (TEM) (JEOL JEM-1230 (JEOL Ltd., Japan, Tokyo) microscope operated at an accelerating voltage of 80 kV). The Raman spectra were obtained using a Shamrock 750 spectrograph (Andor Technology Ltd., Belfast, Northern Ireland) equipped with a CCD detector. The 533-nm line from the CW He–Ne randomly polarized laser was used for excitation at 0.45 mW laser power. Magnetic properties, such as the saturation magnetization (Ms) and coercivity (Hc) were studied via a vibrating sample magnetometer (Lakeshore 7400 series VSM (Lake Shore Cryotronics Inc., Westerville, OH, USA)) in the applied field of H = ±15 kOe. Magnetic circular dichroism (MCD) spectra were recorded at room temperature in a magnetic field of 8 kOe with a JASCO J-820 spectropolarimeter (JASCO Inc., Mary’s CourtEaston, MD, USA) equipped with an electromagnet GMW Associates 5403. All measurements were carried out at ambient conditions.

3. Results and Discussions

3.1. Structural Characterization

The XRD patterns of the samples synthesized at different reaction temperatures are shown in Figure 1. It can be seen from the figure that with an increase in the reaction temperature, the peaks become narrower and sharper, which indicates an increase in crystallite size and better crystallinity. The reflections in the patterns are consistent with the cubic inverse spinel structure of magnetite (JCPDS 19-0629) [54] with space group F d 3 ¯ m and do not contain features of other oxides such as hematite, goethite or wustite. However, since magnetite and maghemite (γ-Fe2O3) have similar XRD patterns with close reflections, they are difficult to be distinguished by XRD. Therefore, to better distinguish the phases of magnetite and maghemite, Raman spectroscopy was used further.
The average crystallite sizes were calculated using Scherrer’s equation (1) [55] from the broadening of the most intense peak (311) and the results are listed in Table 1.
d X R D = 0.89 λ β c o s θ
where λ is the radiation wavelength (0.5418 nm for Cu Kα); β is the line broadening of a diffraction peak at angle θ.
The results show that an increase in the reaction temperature leads to a change in crystallite size from 4.8 to 13.3 nm. Thus, in this synthesis process, crystallite size control can be easily carried out by changing the reaction temperature.
The lattice parameter of the samples was calculated based on Bragg’s law by the following relation and obtained results are also presented in Table 1.
a = d h k l h 2 + k 2 + l 2
where dhkl—inter planar distance; (hkl) is the Miller indices of the planes.
The results show that the lattice constant is slightly smaller than that of bulk magnetite (abulk ≈ 8.4) [54]. The difference between these values can be explained by the presence of a small fraction of maghemite in the samples (amagh. = 8.35 [56]). Therein, with an increase in the reaction temperature, the values of the lattice parameter shift toward higher values, which may indicate a decrease in the fraction of γ-Fe2O3 in the samples. However, Lemine et al. [33] have synthesized pure magnetite nanoparticles, which was confirmed by Mössbauer spectroscopy, and have obtained the value of lattice parameter a = 8.3639 for nanoparticles with a size of 8 nm. This result has been explained by the size effect on the crystal lattice. Both these factors probably influence the change in the lattice parameter with the increase in the reaction temperature, and, as a consequence, with a change in the size of nanoparticles.
The TEM image of the sample synthesized at 380 °C is shown in Figure 1b and the inset demonstrates particle size distribution. It can be seen that nanoparticles have a quasi-spherical shape with a tendency to agglomerate. The particle size of the sample oscillates between 10 and 50 nm (insert in Figure 1b), with an average size of 24 ± 2 nm. This value is larger than calculated from the XRD data and the same tendency is observed for all samples. The difference between size obtained by TEM and XRD can be explained by Ostwald ripening process, wherein small nanoparticles dissolved in supersaturated solution and become larger since equilibrium of the solution is unstable [57]. In addition, a reasonable agreement of the obtained values indicates a balance between stabilization and crystal growth in HDA.
To better distinguish the phases of iron oxides in the samples, Raman spectroscopy was performed. Under normal conditions, magnetite has the space group F d 3 ¯ m and group theory analysis by White and DeAngelis [58] predicts five Raman active modes: A1g + Eg + 3T2g. The Raman spectra of the samples in the range of 150–2000 cm−1 are presented in Figure 2 and the identification of Raman modes was carried out based on literature data obtained for various iron oxides.
The Raman peak at ~667 cm−1 is assigned to A1g mode and corresponds to the symmetrical Fe-O stretch [59,60,61] and is a clear indication of the magnetite phase. The peaks observed at 190 and 570 cm−1, which is more clearly observed with an increase of particle size, confirm the phase of magnetite and are assigned to T2g(1) and T2g(3) modes [62,63], respectively. The position of the peak at 487 cm−1 corresponds to the T2g(2) mode of magnetite [64,65]; however, this peak should be weaker than that shown in Figure 3. We assume that the peak at 487 cm−1 is a mix of the T2g(2) mode of magnetite and Eg mode of maghemite. The band at about 712 cm−1 is the sign of the A1g mode of maghemite [64,66], which may be due to the oxidation of magnetite to maghemite, which occurs more actively at low laser intensity [64], as in the case of our experiment. Further, for the samples synthesized at high temperatures, the peak at 712 cm−1 is separated from the peak of magnetite and has a lower intensity. This fact, along with an increase in peak intensity at 570 cm−1, allows us to conclude that an increase in the reaction temperature leads to the formation of more phase-pure samples. We have semi-quantitatively estimated the ratio magnetite/maghemite based on the results of Raman spectroscopy. Figure 2b,c show the decomposition of the A1g peak for the samples synthesized at 200 and 380 °C into two Gaussian components, which correspond to the contributions of magnetite and maghemite to the peak. The fitting results showed that the contribution of the components (magnetite/maghemite) to the peak varies from ~54.83%/45.17% (for the sample with size dXRD = 4.8 nm) to 60.33%/39.67% (for the sample with size dXRD = 13.3 nm). Dubois and co-authors [67] have determined a correlation between the contribution of the magnetite and maghemite to the experimental Raman spectra and the weight fraction of each oxide in the samples. Based on the results obtained in work [67], we calculated the ratio magnetite/maghemite in the synthesized samples. This semi-quantitative estimate showed that the ratio is ~93%/7% (for the sample with dXRD = 4.8 nm) and ~94%/6% (for the sample with dXRD = 13.3 nm. The Raman peaks at about 1358 and 1582 cm−1 are in a good agreement with the positions of D and G bands of carbon. The D peak is associated with defects in the hexagonal sp2 C network or the finite particle size effect, whereas the G peak arises from the stretching of the C–C bond in graphitic materials, and is common to all sp2 carbon systems [68]. The presence of carbon in the samples can be explained by the synthesis process using HDA, during which carbon was formed. As a result, it accumulated on the surface of the nanoparticles and was recorded on the Raman spectra.

3.2. Magnetic Measurements

The room temperature magnetic hysteresis loops of the samples, synthesized at different reaction temperatures, are presented in Figure 3a,b, which show the size dependence of saturation magnetization and coercivity. The magnetic parameters, obtained from the curves, are summarized in Table 2, which also presents the data for magnetite nanoparticles synthesized by various methods.
As can be seen from Figure 3b, the HC curve demonstrates a growing trend and does not go through a maximum. It is known that the coercivity of nanoparticles is closely related to their size [69]. In a single domain region, the coercivity is given by
H C = g h D 3 / 2
Refs. [70,71,72], while in the multidomain region an experimentally found equation is written approximately as
H C = a + b / D
Refs. [71,72], where a, b, g, h are constants, and D is particle size. As follows from Equations (3) and (4), coercivity should have a maximum in the size range around Dcr, which corresponds to a transition from single- to multidomain states. Thus, it allows us to conclude that all synthesized samples have a single-domain structure.
The M–H curves of the samples with sizes 9 nm ≤ dXRD are typical magnetization loops for superparamagnetic nanoparticles with zero or almost zero remanence magnetization and coercivity at room temperature [31,73]. However, as particle size increases, the remanence and the coercivity increase significantly, which indicates that nanoparticles with sizes dXRD > 9 nm are not superparamagnetic at room temperature. The value of saturation magnetization for all samples are lower than the value MS for bulk magnetite (~96 emu/g) [74]; herewith, the saturation magnetization decreases with decreasing particle size. Such behavior is attributed to the lack of full alignment of the spins even in large applied fields [75]. It is assumed that the surface of the nanoparticles consists of canted or disordered spins that prevent the core spins from aligning along the field direction, resulting in a decrease in the saturation magnetization with a decrease in particle size [39,76]. In addition, a reduced in the saturation magnetization may be due to the presence in the samples of the maghemite phase identified by Raman measurements and which has MS (bulk) = 76 emu/g [77]. Another reason affecting the decrease in MS is the accumulation of carbon on the surface or near-surface layer of nanoparticles during the synthesis process. Moreover, as follows from the results of Raman spectroscopy, with an increase in the reaction temperature from 350 to 380 °C, an increase in the intensity of the carbon peaks is observed, which explains a slight decrease in the saturation magnetization with an increase in particle size from 11.2 to 13.3 nm.
As can be seen from Table 2, the values of saturation magnetization of synthesized nanoparticles are in agreement with the range of values reported in the literature for synthesis methods using iron (III) nitrate nonahydrate as a precursor. Moreover, the method demonstrates, on average, higher values compared to some other methods presented in Table 2, thus they can be used for various applications.

MCD Spectroscopy

The measured MCD spectra of the Fe3O4 nanoparticles with various sizes and Gaussian band fitting of the selected spectra are shown in Figure 4.
The obtained spectra have two main features: negative at energies ≤ ~2.6 eV and positive at ≥ ~2.6 eV. Besides, it can be seen that the intensity of the negative peak increases significantly when the crystallite size exceeds 9 nm. It can be explained by the loss of the superparamagnetic state by nanoparticles at room temperature, which is accompanied by an increase in magnetization. It should be noted that, although the Raman spectra revealed the presence of maghemite in the samples, two positive peaks: an intense peak at 2.8 eV and a weaker peak at around 1.8 eV [83] were not observed in our spectra. The absence of maghemite peaks in the MCD spectra indicates its low concentration in the samples and allows excluding the maghemite phase when interpreting the MCD signals. For deconvolution analysis of the MCD spectra, we have decomposed them into a set of Gaussian components. To obtain a good agreement between the experimental spectra and the sum of the components, seven Gaussian components were necessary for all samples. The fitting parameters for each component are listed in Table 3.
The obtained Gaussian components were identified with the electron transitions of different nature in accordance with theoretical calculations of electronic structures [84,85] and experimental studies of the magneto-optical Kerr effect [86,87] and optical magnetic circular dichroism [88] in magnetite. Based on these data from the literature, it can be concluded that the decomposition components correspond to the following transitions: peak 1 at ~2.05 corresponds to intervalence charge–transfer transition (IVCT) between two sites differing only in oxidation state, namely [Fe2+]t2g→[Fe3+]eg; peaks 2—[Fe2+]t2g→(Fe2+)e (~2.3 eV), 4—(Fe3+)t2→[Fe3+]t2g (~2.76 eV), and 6—[Fe3+]eg→(Fe2+)t2 (~3.44 eV) are associated with transitions between the same ions belonging to different magnetic sublattice (intersublattice charge-transfer transitions, ISCT); peaks located at ~2.53 eV (3), ~2.95 eV (5), and ~3.98 eV (7) are assigned to ligand-to-metal p—d charge transfer transitions (LMCT) across the optical gap of the spin minority involving polarized O(2p): (O2p→Fe3d). From the obtained results, it was found that the size effect is observed for peaks 1 and 4 (Figure 5) and both peaks show a growing trend of intensity with increasing crystallite size.
This behavior can be explained as follows: at low reaction temperatures, magnetite nanoparticles with smaller sizes are formed; as a consequence, they are more easily oxidized to maghemite, which is confirmed by the results of structural analysis. In this process, Fe o c t 2 + ions are oxidized to Fe3+ in octahedral sites [89,90], which leads to the decrease in the intensity of the IVCT peak [Fe2+]t2g→[Fe3+]eg. Furthermore, it was shown in [91,92] using the Mössbauer spectroscopy that the oxidation of magnetite to maghemite is accompanied by a decrease in the fraction of trivalent iron ions in tetrahedral sites. It is assumed that this effect leads to a decrease in the intensity of the intersublattice charge-transfer transition (Fe3+)t2→[Fe3+]t2g with the decrease in particle size.

4. Conclusions

A simple and inexpensive method for synthesis magnetite nanoparticles, based on thermal decomposition of iron (III) nitrate nonahydrate in HDA, is presented. Moreover, the process proceeds without the use of protective gas and at atmospheric pressure, which makes it easily reproducible. Results obtained showed that the average crystallite size increased with increasing reaction temperature. Based on the results of the analysis of the change in the lattice parameter and Raman spectroscopy, it was concluded that the samples contain a small fraction of maghemite, herewith an increase in the reaction temperature leads to the formation of more phase-pure samples. Magnetic measurements revealed that all synthesized samples are single domain, herewith when the crystallite size exceeds 9 nm, the nanoparticles cease to be superparamagnetic at room temperature. The saturation magnetization has a maximum MS = 56 emu/g for the sample synthesized at TR = 350 °C and the obtained values are agreement with the literature data for synthesis methods using iron (III) nitrate nonahydrate as a precursor. The MCD spectra of Fe3O4 nanoparticles were decomposed by a set of seven components that provides a good agreement between the experimental spectra and the sum of the components. The decomposition components were identified with a high degree of confidence with electronic transitions of various nature in accordance with theoretical calculations of electronic structures and experimental studies of magnetite. It was found that the intensities of the IVCT peak [Fe2+]t2g→[Fe3+]eg and the ISCT peak (Fe3+)t2→[Fe3+]t2g have a growing trend with increasing crystallite size, which is explained by the effect of reaction temperature on the size and phase composition of the samples.

Author Contributions

C.-R.L. and D.S. conceptualized the idea and together with Y.-C.C. and C.-C.W. designed the experiments; Y.-C.C., C.-C.W. and A.S. performed the experiments; A.S., Y.-C.C. and D.S. analyzed the data; C.-R.L.—funding acquisition; C.-C.W. contributed reagents and materials; A.S. explained and visualized the data; A.S. and C.-C.W. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology of Taiwan (Grant No 106-2112-M-153-001-MY3 and Grant No 107-2811-M-153-500).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dalpozzo, R. Magnetic nanoparticle supports for asymmetric catalysts. Green Chem. 2015, 17, 3671–3686. [Google Scholar] [CrossRef]
  2. Battiato, M.; Minar, J.; Wang, W.; Ndiaye, W.; Richter, M.C.; Heckmann, O.; Mariot, J.-M.; Parmigiani, F.; Hricovini, K.; Cacho, C. Distinctive Picosecond Spin Polarization Dynamics in Bulk Half Metals. Phys. Rev. Lett. 2018, 121, 077205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wu, T.; Liu, Y.; Zeng, X.; Cui, T.; Zhao, Y.; Li, Y.; Tong, G. Facile Hydrothermal Synthesis of Fe3O4/C Core–Shell Nanorings for Efficient Low-Frequency Microwave Absorption. ACS Appl. Mater. Interfaces 2016, 8, 7370–7380. [Google Scholar] [CrossRef] [PubMed]
  4. Rajabi, H.R.; Arjmand, H.; Hoseini, S.J.; Nasrabadi, H. Surface modified magnetic nanoparticles as efficient and green sorbents: Synthesis, characterization, and application for the removal of anionic dye. J. Magn. Magn. Mater. 2015, 394, 7–13. [Google Scholar] [CrossRef]
  5. Molina, L.; Gaete, J.; Alfaro, I.; Ide, V.; Valenzuela, F.; Parada, J.; Basualto, C. Synthesis and characterization of magnetite nanoparticles functionalized with organophosphorus compounds and its application as an adsorbent for La (III), Nd (III) and Pr (III) ions from aqueous solutions. J. Mol. Liq. 2019, 275, 178–191. [Google Scholar] [CrossRef]
  6. Wang, Y.; Ma, H.; Wang, X.; Pang, X.; Wu, D.; Du, B.; Wei, Q. Novel signal amplification strategy for ultrasensitive sandwich-type electrochemical immunosensor employing Pd-Fe3O4-GS as the matrix and SiO2 as the label. Biosens. Bioelectron. 2015, 74, 59–65. [Google Scholar] [CrossRef]
  7. Patra, M.; Manzoor, K.; Manoth, M.; Negi, S.; Vadera, S.; Kumar, N. Nanotechnology Applications for Chemical and Biological Sensors. Def. Sci. J. 2008, 58, 636–649. [Google Scholar] [CrossRef]
  8. Wang, Z.; Liu, C.-J. Preparation and application of iron oxide/graphene based composites for electrochemical energy storage and energy conversion devices: Current status and perspective. Nano Energy 2015, 11, 277–293. [Google Scholar] [CrossRef]
  9. Hussein, A.K. Applications of nanotechnology in renewable energies—A comprehensive overview and understanding. Renew. Sustain. Energy Rev. 2015, 42, 460–476. [Google Scholar] [CrossRef]
  10. Magro, M.; Baratella, D.; Bonaiuto, E.; Roger, J.D.A.; Vianello, F. New Perspectives on Biomedical Applications of Iron Oxide Nanoparticles. Curr. Med. Chem. 2018, 25, 540–555. [Google Scholar] [CrossRef]
  11. Laurent, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Elst, L.V.; Muller, R.N. Magnetic Iron Oxide Nanoparticles: Synthesis, Stabilization, Vectorization, Physicochemical Characterizations, and Biological Applications. Chem. Rev. 2008, 108, 2064–2110. [Google Scholar] [CrossRef] [PubMed]
  12. Song, C.; Sun, W.; Xiao, Y.; Shi, X. Ultrasmall iron oxide nanoparticles: Synthesis, surface modification, assembly, and biomedical applications. Drug Discov. Today 2019, 24, 835–844. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, L.; Mendoza-Garcia, A.; Li, Q.; Sun, S. Organic Phase Syntheses of Magnetic Nanoparticles and Their Applications. Chem. Rev. 2016, 116, 10473–10512. [Google Scholar] [CrossRef] [PubMed]
  14. Khan, U.; Nairan, A.; Irfan, M.; Ali, H.; Han, X. Temperature mediated morphological and magnetic phase transitions of iron/iron oxide Core/Shell nanostructures. J. Alloys Compd. 2017, 696, 362–368. [Google Scholar] [CrossRef]
  15. Kwon, Y.H.; Huie, M.M.; Choi, D.; Chang, M.; Marschilok, A.C.; Takeuchi, K.J.; Takeuchi, E.S.; Reichmanis, E. Toward Uniformly Dispersed Battery Electrode Composite Materials: Characteristics and Performance. ACS Appl. Mater. Interfaces 2016, 8, 3452–3463. [Google Scholar] [CrossRef]
  16. Bagherzadeh, M.; Pirmoradian, M.; Riahi, F. Electrochemical Detection of Pb and Cu by Using DTPA Functionalized Magnetic Nanoparticles. Electrochim. Acta 2014, 115, 573–580. [Google Scholar] [CrossRef]
  17. Reddy, K.R.; Lee, K.-P.; Saianand, G. Novel electrically conductive and ferromagnetic composites of poly(aniline-co-aminonaphthalenesulfonic acid) with iron oxide nanoparticles: Synthesis and characterization. J. Appl. Polym. Sci. 2007, 106, 1181–1191. [Google Scholar] [CrossRef]
  18. Shaterabadi, Z.; Nabiyouni, G.; Soleymani, M. Optimal size for heating efficiency of superparamagnetic dextran-coated magnetite nanoparticles for application in magnetic fluid hyperthermia. Phys. C Supercond. 2018, 549, 84–87. [Google Scholar] [CrossRef]
  19. Kalantari, K.; Ahmad, M.B.; Masoumi, H.R.F.; Shameli, K.; Basri, M.; Khandanlou, R. Rapid Adsorption of Heavy Metals by Fe3O4/Talc Nanocomposite and Optimization Study Using Response Surface Methodology. Int. J. Mol. Sci. 2014, 15, 12913–12927. [Google Scholar] [CrossRef] [Green Version]
  20. Jin, R.; Lin, B.; Li, D.; Ai, H. Superparamagnetic iron oxide nanoparticles for MR imaging and therapy: Design considerations and clinical applications. Curr. Opin. Pharmacol. 2014, 18, 18–27. [Google Scholar] [CrossRef]
  21. Rezayan, A.H.; Mousavi, M.; Kheirjou, S.; Amoabediny, G.; Ardestani, M.S.; Mohammadnejad, J. Monodisperse magnetite (Fe3O4) nanoparticles modified with water soluble polymers for the diagnosis of breast cancer by MRI method. J. Magn. Magn. Mater. 2016, 420, 210–217. [Google Scholar] [CrossRef] [Green Version]
  22. Hedayatnasab, Z.; Abnisa, F.; Daud, W.M.A.W. Review on magnetic nanoparticles for magnetic nanofluid hyperthermia application. Mater. Des. 2017, 123, 174–196. [Google Scholar] [CrossRef]
  23. Dukenbayev, K.; Korolkov, I.; Tishkevich, D.I.; Kozlovskiy, A.; Trukhanov, S.; Gorin, Y.G.; Shumskaya, A.; Kaniukov, E.Y.; Vinnik, D.; Zdorovets, M.V.; et al. Fe3O4 Nanoparticles for Complex Targeted Delivery and Boron Neutron Capture Therapy. Nanomaterials 2019, 9, 494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wong, J.; Prout, J.; Seifalian, A.M. Magnetic Nanoparticles: New Perspectives in Drug Delivery. Curr. Pharm. Des. 2017, 23, 2908–2917. [Google Scholar] [CrossRef] [PubMed]
  25. Ding, C.; Li, Z. A review of drug release mechanisms from nanocarrier systems. Mater. Sci. Eng. C 2017, 76, 1440–1453. [Google Scholar] [CrossRef]
  26. Revia, R.A.; Zhang, M. Magnetite nanoparticles for cancer diagnosis, treatment, and treatment monitoring: Recent advances. Mater. Today 2016, 19, 157–168. [Google Scholar] [CrossRef] [PubMed]
  27. Sari, A.Y.; Eko, A.S.; Candra, K.; Hasibuan, D.P.; Ginting, M.; Sebayang, P.; Simamora, P. Synthesis, Properties and Application of Glucose Coated Fe3O4Nanoparticles Prepared by Co-precipitation Method. IOP Conf. Ser. Mater. Sci. Eng. 2017, 214, 12021. [Google Scholar] [CrossRef] [Green Version]
  28. Petcharoen, K.; Sirivat, A. Synthesis and characterization of magnetite nanoparticles via the chemical co-precipitation method. Mater. Sci. Eng. B 2012, 177, 421–427. [Google Scholar] [CrossRef]
  29. Rani, S.; Varma, G. Superparamagnetism and metamagnetic transition in Fe3O4 nanoparticles synthesized via co-precipitation method at different pH. Phys. B Condens. Matter 2015, 472, 66–77. [Google Scholar] [CrossRef]
  30. Zhao, F.; Zhang, B.; Feng, L. Preparation and magnetic properties of magnetite nanoparticles. Mater. Lett. 2012, 68, 112–114. [Google Scholar] [CrossRef]
  31. Ahmadi, S.; Chia, C.H.; Zakaria, S.; Saeedfar, K.; Asim, N. Synthesis of Fe3O4 nanocrystals using hydrothermal approach. J. Magn. Magn. Mater. 2012, 324, 4147–4150. [Google Scholar] [CrossRef]
  32. Zhang, H.; Zhu, G. One-step hydrothermal synthesis of magnetic Fe3O4 nanoparticles immobilized on polyamide fabric. Appl. Surf. Sci. 2012, 258, 4952–4959. [Google Scholar] [CrossRef]
  33. Lemine, O.; Omri, K.; Zhang, B.; El Mir, L.; Sajieddine, M.; Alyamani, A.; Bououdina, M. Sol–gel synthesis of 8nm magnetite (Fe3O4) nanoparticles and their magnetic properties. Superlattices Microstruct. 2012, 52, 793–799. [Google Scholar] [CrossRef]
  34. Chomchoey, N.; Bhongsuwan, D.; Bhongsuwan, T. Magnetic Properties of Magnetite Nanoparticles Synthesized by Oxidative Alkaline Hydrolysis of Iron Powder. Nat. Sci. 2010, 44, 963–971. [Google Scholar]
  35. Worawong, A.; Jutarosaga, T.; Onreabroy, W. Influence of Calcination Temperature on Synthesis of Magnetite (Fe3O4) Nanoparticles by Sol-Gel Method. Adv. Mater. Res. 2014, 979, 208–211. [Google Scholar] [CrossRef]
  36. Belaïd, S.; Laurent, S.; Vermeersch, M.; Elst, L.V.; Pérez-Morga, D.; Muller, R.N. A new approach to follow the formation of iron oxide nanoparticles synthesized by thermal decomposition. Nanotechnology 2013, 24, 55705. [Google Scholar] [CrossRef]
  37. Amara, D.; Felner, I.; Nowik, I.; Margel, S. Synthesis and characterization of Fe and Fe3O4 nanoparticles by thermal decomposition of triiron dodecacarbonyl. Colloids Surf. A Physicochem. Eng. Asp. 2009, 339, 106–110. [Google Scholar] [CrossRef]
  38. Toyos-Rodríguez, C.; Calleja-García, J.; Torres-Sánchez, L.; López, A.; Abu-Dief, A.M.; Costa, A.; Elbaile, L.; Crespo, R.D.; Garitaonandia, J.S.; Lastra, E.; et al. A Simple and Reliable Synthesis of Superparamagnetic Magnetite Nanoparticles by Thermal Decomposition of Fe(acac)3. J. Nanomater. 2019, 2019, 1–10. [Google Scholar] [CrossRef]
  39. Roca, A.G.; Morales, M.P.; O’Grady, K.; Serna, C.J. Structural and magnetic properties of uniform magnetite nanoparticles prepared by high temperature decomposition of organic precursors. Nanotechnology 2006, 17, 2783–2788. [Google Scholar] [CrossRef]
  40. Li, D.; Jiang, D.; Jiang, D.; Xie, J.; Wu, Y.; Dang, S.; Zhang, J. An easy fabrication of monodisperse oleic acid-coated Fe3O4 nanoparticles. Mater. Lett. 2010, 64, 2462–2464. [Google Scholar] [CrossRef]
  41. Majewski, P.; Thierry, B. Functionalized Magnetite Nanoparticles—Synthesis, Properties, and Bio-Applications. Crit. Rev. Solid State Mater. Sci. 2007, 32, 203–215. [Google Scholar] [CrossRef]
  42. Salazar-Alvarez, G.; Muhammed, M.; Zagorodni, A.A. Novel flow injection synthesis of iron oxide nanoparticles with narrow size distribution. Chem. Eng. Sci. 2006, 61, 4625–4633. [Google Scholar] [CrossRef]
  43. Abbas, M.; Rao, B.P.; Naga, S.; Takahashi, M.; Kim, C. Synthesis of high magnetization hydrophilic magnetite (Fe3O4) nanoparticles in single reaction—Surfactantless polyol process. Ceram. Int. 2013, 39, 7605–7611. [Google Scholar] [CrossRef]
  44. Wu, W.; He, Q.; Jiang, C. Magnetic Iron Oxide Nanoparticles: Synthesis and Surface Functionalization Strategies. Nanoscale Res. Lett. 2008, 3, 397–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Fan, R.; Chen, X.; Gui, Z.; Liu, L.; Chen, Z. A new simple hydrothermal preparation of nanocrystalline magnetite Fe3O4. Mater. Res. Bull. 2001, 36, 497–502. [Google Scholar] [CrossRef]
  46. Sun, S.; Zeng, H. Size-Controlled Synthesis of Magnetite Nanoparticles. J. Am. Chem. Soc. 2002, 124, 8204–8205. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, J.; Zhang, B.; Wang, L.; Wang, M.; Gao, H. One-pot synthesis of water-soluble superparamagnetic iron oxide nanoparticles and their MRI contrast effects in the mouse brains. Mater. Sci. Eng. C 2015, 48, 416–423. [Google Scholar] [CrossRef]
  48. Naidek, K.P.; Bianconi, F.; Rocha, T.C.R.; Zanchet, D.; Bonacin, J.A.; Novak, M.A.; Vaz, M.G.F.; Winnischofer, H. Structure and morphology of spinel MFe2O4 (M = Fe, Co, Ni) nanoparticles chemically synthesized from heterometallic complexes. J. Colloid Interface Sci. 2011, 358, 39–46. [Google Scholar] [CrossRef]
  49. Kwon, S.G.; Piao, Y.; Park, J.; Angappane, S.; Jo, Y.; Hwang, N.-M.; Park, J.-G.; Hyeon, T. Kinetics of Monodisperse Iron Oxide Nanocrystal Formation by “Heating-Up” Process. J. Am. Chem. Soc. 2007, 129, 12571–12584. [Google Scholar] [CrossRef]
  50. Riaz, N.; Faheem, M.; Riaz, A. Surfactant-modified silver nanoparticle ink for high-resolution ink-jet printed narrow-gaped organic electrodes. Mater. Express 2017, 7, 113–122. [Google Scholar] [CrossRef]
  51. Wang, H.; Jiao, X.; Chen, D. Monodispersed Nickel Nanoparticles with Tunable Phase and Size: Synthesis, Characterization, and Magnetic Properties. J. Phys. Chem. C 2008, 112, 18793–18797. [Google Scholar] [CrossRef]
  52. Winter, H.; Christopher-Allison, E.; Brown, A.L.; Goforth, A.M. Aerobic method for the synthesis of nearly size-monodisperse bismuth nanoparticles from a redox non-innocent precursor. Nanotechnology 2018, 29, 155603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Hou, Y.; Yu, J.; Gao, S. Solvothermal reduction synthesis and characterization of superparamagnetic magnetite nanoparticlesElectronic supplementary information (ESI) available: Size distributions of samples modified with TOPO + PVP, HDA + PVP, and PVP only. J. Mater. Chem. 2003, 13, 1983. [Google Scholar] [CrossRef]
  54. He, H.; Zhong, Y.; Liang, X.; Tan, W.; Zhu, J.; Wang, C.Y. Natural Magnetite: An efficient catalyst for the degradation of organic contaminant. Sci. Rep. 2015, 5, 10139. [Google Scholar] [CrossRef] [Green Version]
  55. Patterson, A.L. The Scherrer Formula for X-ray Particle Size Determination. Phys. Rev. 1939, 56, 978–982. [Google Scholar] [CrossRef]
  56. Goss, C. Saturation magnetisation, coercivity and lattice parameter changes in the system Fe3O4-Fe2O3, and their relationship to structure. Phys. Chem. Miner. 1988, 16, 164–171. [Google Scholar] [CrossRef]
  57. Cabe, W.M.; Smith, J.; Harriott, P. Unit Operations of Chemical Engineering, 5th ed.; McGraw-Hill: New York, NY, USA, 1993; Volume 1130, p. 895. [Google Scholar]
  58. White, W.; De Angelis, B. Interpretation of the vibrational spectra of spinels. Spectrochim. Acta Part A Mol. Spectrosc. 1967, 23, 985–995. [Google Scholar] [CrossRef]
  59. De Faria, D.L.A.; Silva, S.V.; de Oliveira, M.T. Raman Microspectroscopy of Some Iron Oxides and Oxyhydroxides. J. Raman Spectr. 1997, 28, 873–878. [Google Scholar] [CrossRef]
  60. Shebanova, O.N.; Lazor, P. Raman spectroscopic study of magnetite (FeFe2O4): A new assignment for the vibrational spectrum. J. Solid State Chem. 2003, 174, 424–430. [Google Scholar] [CrossRef]
  61. Jacintho, G.V.M.; Brolo, A.G.; Corio, P.; Suarez, P.A.Z.; Rubim, J.C. Structural Investigation of MFe2O4(M=Fe, Co) Magnetic Fluids. J. Phys. Chem. C 2009, 113, 7684–7691. [Google Scholar] [CrossRef]
  62. Verble, J. Temperature-dependent light-scattering studies of the Verwey transition and electronic disorder in magnetite. Phys. Rev. B 1974, 9, 5236–5248. [Google Scholar] [CrossRef]
  63. Gasparov, L.V.; Romero, D.; Tanner, D.B.; Berger, H.; Margaritondo, G.; Forró, L. Infrared and Raman studies of the Verwey transition in magnetite. Phys. Rev. B 2000, 62, 7939–7944. [Google Scholar] [CrossRef] [Green Version]
  64. Letti, C.; Paterno, L.G.; Pereira-Da-Silva, M.A.A.; Morais, P.; Soler, M. The role of polymer films on the oxidation of magnetite nanoparticles. J. Solid State Chem. 2017, 246, 57–64. [Google Scholar] [CrossRef]
  65. Graves, P.; Johnston, C.; Campaniello, J. Raman scattering in spinel structure ferrites. Mater. Res. Bull. 1988, 23, 1651–1660. [Google Scholar] [CrossRef]
  66. Legodi, M.; DeWaal, D. The preparation of magnetite, goethite, hematite and maghemite of pigment quality from mill scale iron waste. Dyes Pigments 2007, 74, 161–168. [Google Scholar] [CrossRef]
  67. Dubois, F.; Mendibide, C.; Pagnier, T.; Perrard, F.; Duret, C. Raman mapping of corrosion products formed onto spring steels during salt spray experiments. A correlation between the scale composition and the corrosion resistance. Corros. Sci. 2008, 50, 3401–3409. [Google Scholar] [CrossRef]
  68. Ferrari, A.; Robertson, J. Resonant Raman spectroscopy of disordered, amorphous, and diamondlike carbon. Phys. Rev. B 2001, 64, 075414. [Google Scholar] [CrossRef] [Green Version]
  69. Stoner, E.C.; Wohlfarth, E.P. A mechanism of magnetic hysteresis in heterogeneous alloys. IEEE Trans. Magn. 1991, 27, 3475–3518. [Google Scholar] [CrossRef]
  70. Bean, C.P.; Livingston, J.D. Superparamagnetism. J. Appl. Phys. 1959, 30, S120–S129. [Google Scholar] [CrossRef]
  71. Kneller, E.F.; Luborsky, F.E. Particle Size Dependence of Coercivity and Remanence of Single-Domain Particles. J. Appl. Phys. 1963, 34, 656. [Google Scholar] [CrossRef]
  72. Cullity, B.D.; Graham, C.D. Introduction to Magnetic Materials, 2nd ed.; Wiley-IEEE Press: Hoboken, NJ, USA, 2008; p. 568. ISBN 978-0-471-47741-9. [Google Scholar]
  73. Caruntu, D.; Caruntu, G.; O’Connor, C.J. Magnetic properties of variable-sized Fe3O4 nanoparticles synthesized from non-aqueous homogeneous solutions of polyols. J. Phys. D Appl. Phys. 2007, 40, 5801–5809. [Google Scholar] [CrossRef] [Green Version]
  74. Goya, G.; Berquoó, T.S.; Fonseca, F.C.; Morales, M.P. Static and dynamic magnetic properties of spherical magnetite nanoparticles. J. Appl. Phys. 2003, 94, 3520. [Google Scholar] [CrossRef] [Green Version]
  75. Coey, J.M.D. Noncollinear Spin Arrangement in Ultrafine Ferrimagnetic Crystallites. Phys. Rev. Lett. 1971, 27, 1140–1142. [Google Scholar] [CrossRef] [Green Version]
  76. Berkowitz, A.E.; Lahut, J.A.; Jacobs, I.S.; Levinson, L.M.; Forester, D.W. Spin Pinning at Ferrite-Organic Interfaces. Phys. Rev. Lett. 1975, 34, 594–597. [Google Scholar] [CrossRef]
  77. Cao, D.; Li, H.; Pan, L.; Li, J.; Wang, X.; Jing, P.; Cheng, X.; Wang, W.; Wang, J.; Liu, Q. High saturation magnetization of γ-Fe2O3 nano-particles by a facile one-step synthesis approach. Sci. Rep. 2016, 6, 32360. [Google Scholar] [CrossRef] [Green Version]
  78. Vega-Chacón, J.; Picasso, G.; Félix, L.A.; Jafelicci, M., Jr. Influence of synthesis experimental parameters on the formation of magnetite nanoparticles prepared by polyol method. Adv. Nat. Sci. Nanosci. Nanotechnol. 2016, 7, 15014. [Google Scholar] [CrossRef]
  79. Gee, S.H.; Hong, Y.-K.; Erickson, D.W.; Park, M.H.; Sur, J.C. Synthesis and aging effect of spherical magnetite (Fe3O4) nanoparticles for biosensor applications. J. Appl. Phys. 2003, 93, 7560. [Google Scholar] [CrossRef]
  80. Ou, P.; Xu, G.; Xu, C.; Zhang, Y.; Hou, X.; Han, G. Synthesis and characterization of magnetite nanoparticles by a simple solvothermal method. Mater. Sci. 2010, 28, 817–822. [Google Scholar]
  81. Lyubutin, I.; Lin, C.-R.; Tseng, Y.-T.; Spivakov, A.; Baskakov, A.; Starchikov, S.; Funtov, K.; Jhang, C.-J.; Tsai, Y.-J.; Hsu, H.-S. Structural and magnetic evolution of FexOy@carbon core-shell nanoparticles synthesized by a one-step thermal pyrolysis. Mater. Charact. 2019, 150, 213–219. [Google Scholar] [CrossRef]
  82. Xu, G.; Zhang, M.; Ou, P.; Zhang, Y.; Han, G.R. Synthesis of Monodispersed Fe3O4 Magnetite Nanoparticles by Ethylene Glycol Solvothermal Method. Adv. Mater. Res. 2013, 634, 2276–2279. [Google Scholar] [CrossRef]
  83. Fantechi, E.; Campo, G.; Carta, D.; Corrias, A.; Fernández, C.D.J.; Gatteschi, D.; Innocenti, C.; Pineider, F.; Rugi, F.; Sangregorio, C. Exploring the Effect of Co Doping in Fine Maghemite Nanoparticles. J. Phys. Chem. C 2012, 116, 8261–8270. [Google Scholar] [CrossRef]
  84. Antonov, V.N.; Harmon, B.N.; Antropov, V.P.; Perlov, A.Y.; Yaresko, A.N. Electronic structure and magneto-optical Kerr effect of Fe3O4 and Mg2+- orAl3+-substituted Fe3O4. Phys. Rev. B 2001, 64. [Google Scholar] [CrossRef]
  85. Fontijn, W.F.J.; Van Der Zaag, P.; Feiner, L.F.; Metselaar, R.; Devillers, M.A.C. A consistent interpretation of the magneto-optical spectra of spinel type ferrites (invited). J. Appl. Phys. 1999, 85, 5100–5105. [Google Scholar] [CrossRef] [Green Version]
  86. Kim, K.J.; Lee, H.-S.; Lee, M.H.; Lee, S.H. Comparative magneto-optical investigation of d–d charge–transfer transitions in Fe3O4, CoFe2O4, and NiFe2O4. J. Appl. Phys. 2002, 91, 9974. [Google Scholar] [CrossRef]
  87. Fontijn, W.F.J.; Van Der Zaag, P.J.; Devillers, M.A.C.; Brabers, V.A.M.; Metselaar, R. Optical and magneto-optical polar Kerr spectra of Fe3O4 and Mg2+- or Al3+-substituted Fe3O4. Phys. Rev. B 1997, 56, 5432–5442. [Google Scholar] [CrossRef] [Green Version]
  88. Chen, J.; Hsu, H.; Huang, Y.-H.; Huang, D.-J. Spin-dependent optical charge transfer in magnetite from transmitting optical magnetic circular dichroism. Phys. Rev. B 2018, 98, 085141. [Google Scholar] [CrossRef]
  89. Jolivet, J.-P.; Tronc, E. Interfacial electron transfer in colloidal spinel iron oxide. Conversion of Fe3O4-γFe2O3 in aqueous medium. J. Colloid Interface Sci. 1988, 125, 688–701. [Google Scholar] [CrossRef]
  90. Rebodos, R.L.; Vikesland, P.J. Effects of Oxidation on the Magnetization of Nanoparticulate Magnetite. Langmuir 2010, 26, 16745–16753. [Google Scholar] [CrossRef]
  91. Joos, A.; Rümenapp, C.; Wagner, F.E.; Gleich, B. Characterisation of iron oxide nanoparticles by Mössbauer spectroscopy at ambient temperature. J. Magn. Magn. Mater. 2016, 399, 123–129. [Google Scholar] [CrossRef]
  92. Lengyel, A.; Tolnai, G.; Klencsar, Z.; Garg, V.K.; De Oliveira, A.C.; Singh, L.H.; Homonnay, Z.; Szalay, R.; Nemeth, P.; Szabolcs, B.; et al. The effect of carboxylic acids on the oxidation of coated iron oxide nanoparticles. J. Nanopart. Res. 2018, 20, 137. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the samples synthesized at different reaction temperatures and (b) TEM micrograph of magnetite nanoparticles synthesized at 380 °C. The inset shows the histogram of particle size distribution.
Figure 1. (a) XRD patterns of the samples synthesized at different reaction temperatures and (b) TEM micrograph of magnetite nanoparticles synthesized at 380 °C. The inset shows the histogram of particle size distribution.
Nanomaterials 10 01888 g001
Figure 2. Raman spectra of magnetite nanoparticles synthesized at different reaction temperatures (a) and the best Gaussian fitting for the A1g peaks of the samples with d = 4.8 nm (b) and 13.3 nm (c); *—artifacts.
Figure 2. Raman spectra of magnetite nanoparticles synthesized at different reaction temperatures (a) and the best Gaussian fitting for the A1g peaks of the samples with d = 4.8 nm (b) and 13.3 nm (c); *—artifacts.
Nanomaterials 10 01888 g002
Figure 3. (a) Magnetic hysteresis loops of the magnetite nanoparticles; (b) size dependence of saturation magnetization and coercivity. The inset shows the hysteresis loops in an enlarged scale.
Figure 3. (a) Magnetic hysteresis loops of the magnetite nanoparticles; (b) size dependence of saturation magnetization and coercivity. The inset shows the hysteresis loops in an enlarged scale.
Nanomaterials 10 01888 g003
Figure 4. Experimental MCD spectra of the magnetite nanoparticles measured at room temperature and the best Gaussian fitting of the spectra with d = 4.8, 9, and 13.3 nm. Red dotted curves correspond the sum of the deconvoluted spectra.
Figure 4. Experimental MCD spectra of the magnetite nanoparticles measured at room temperature and the best Gaussian fitting of the spectra with d = 4.8, 9, and 13.3 nm. Red dotted curves correspond the sum of the deconvoluted spectra.
Nanomaterials 10 01888 g004
Figure 5. The change in the intensity of peaks 1 and 4 with increasing average crystallite size of Fe3O4 nanoparticles.
Figure 5. The change in the intensity of peaks 1 and 4 with increasing average crystallite size of Fe3O4 nanoparticles.
Nanomaterials 10 01888 g005
Table 1. The average crystallite size and lattice constants (a) of the samples synthesized at temperatures 200 ≤ TR ≤ 380 °C.
Table 1. The average crystallite size and lattice constants (a) of the samples synthesized at temperatures 200 ≤ TR ≤ 380 °C.
TR, °C200260300350380
dXRD, nm4.86.8911.213.3
a, Å8.368.3798.3738.3868.392
Table 2. Comparison of magnetic properties at room temperature and sizes of magnetite nanoparticles synthesized by various methods.
Table 2. Comparison of magnetic properties at room temperature and sizes of magnetite nanoparticles synthesized by various methods.
Method of SynthesisdXRD, nmReferenceMS, emu/gMR, emu/gHC, Oe
Thermal decomposition of iron (III) nitrate nonahydrate using HDA4.8 3600.6
6.8 4301
9This work460.32
11.2 561.514
13.3 553.947
Co-precipitation14.7[27]35.414.6283.58
15.1[28]58.722-27.076
20.3[29]53-0
Hydrothermal9.3[30]4500
13.4[31]27.2-58.4
25.5[32]50895
Sol–gel8[33]47-0.655
12.23[34]52.21.321.1
13[35]35-17
Thermal decomposition using various reagents5.5[36]43.7--
7.4[37]41.700
9[38]6501
13.6[39]72-0
15[38]70-12
24.2[40]78.68-0
Various methods using iron (III) nitrate nonahydrate as a precursor4.4 *[78]39.2-0
6 *[78]52-0
7 *[79]49-0
12 *[80]43.600
13[35]35-17
~45 *[81]90944
~75 *[82]68.812.9138.5
* The values obtained from TEM data.
Table 3. Results of the fit analysis of the MCD spectra of the Fe3O4 nanoparticles. The tetrahedral and octahedral sublattices are denoted as () and [], respectively. E—peak position (eV); I—peak intensity (arb. unit).
Table 3. Results of the fit analysis of the MCD spectra of the Fe3O4 nanoparticles. The tetrahedral and octahedral sublattices are denoted as () and [], respectively. E—peak position (eV); I—peak intensity (arb. unit).
ComponentTypeTransitiond = 4.8 nmd = 6.8 nmd = 9 nm
EIEIEI
1IVCT[Fe2+]t2g→[Fe3+]eg2.01−6.592.07−4.952.04−13.15
2ISCT[Fe2+]t2g→(Fe2+)e2.30−7.232.34−20.472.27−12.96
3LMCTO(2p)→Fe(3d)2.52−3.42.56−18.672.53−11.17
4ISCT(Fe3+)t2→[Fe3+]t2g2.7414.622.822.792.7319.02
5LMCTO(2p)→Fe(3d)2.9228.862.9875.642.9161.0
6ISCT[Fe3+]eg→(Fe2+)t23.4570.43.49230.193.45157.44
7LMCTO(2p)→Fe(3d)3.9852.294.02219.854.099.16
ComponentTypeTransitiond = 11.2 nmd = 13.3 nm
EIEI
1IVCT[Fe2+]t2g→[Fe3+]eg2.04−33.022.07−47.91
2ISCT[Fe2+]t2g→Fe2+)e2.32−43.282.26−13.16
3LMCTO(2p)→Fe(3d)2.54−19.632.51−2.42
4ISCT(Fe3+)t2→[Fe3+]t2g2.847.072.7554.24
5LMCTO(2p)→Fe(3d)2.99113.512.9395.95
6ISCT[Fe3+]eg→(Fe2+)t23.45258.063.37186.92
7LMCTO(2p)→Fe(3d)3.97148.433.9569.77

Share and Cite

MDPI and ACS Style

Spivakov, A.; Lin, C.-R.; Chang, Y.-C.; Wang, C.-C.; Sarychev, D. Magnetic and Magneto-Optical Oroperties of Iron Oxides Nanoparticles Synthesized under Atmospheric Pressure. Nanomaterials 2020, 10, 1888. https://doi.org/10.3390/nano10091888

AMA Style

Spivakov A, Lin C-R, Chang Y-C, Wang C-C, Sarychev D. Magnetic and Magneto-Optical Oroperties of Iron Oxides Nanoparticles Synthesized under Atmospheric Pressure. Nanomaterials. 2020; 10(9):1888. https://doi.org/10.3390/nano10091888

Chicago/Turabian Style

Spivakov, Aleksandr, Chun-Rong Lin, Yu-Chuan Chang, Cheng-Chien Wang, and Dmitriy Sarychev. 2020. "Magnetic and Magneto-Optical Oroperties of Iron Oxides Nanoparticles Synthesized under Atmospheric Pressure" Nanomaterials 10, no. 9: 1888. https://doi.org/10.3390/nano10091888

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