Next Article in Journal
Development and Characterizations of Novel Aqueous-Based Ceramic Inks for Inkjet Printing
Next Article in Special Issue
Mordenite-Supported Ag+-Cu2+-Zn2+ Trimetallic System: A Variety of Nanospecies Obtained via Thermal Reduction in Hydrogen Followed by Cooling in an Air or Hydrogen Atmosphere
Previous Article in Journal
Investigation of the Flexural Behavior of Preloaded and Pre-Cracked Reinforced Concrete Beams Strengthened with CFRP Plates
Previous Article in Special Issue
Thermal Properties of Porous Silicon Nanomaterials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Core–Shell Fe3O4@C Nanoparticles for the Organic Dye Adsorption and Targeted Magneto-Mechanical Destruction of Ehrlich Ascites Carcinoma Cells

by
Oxana S. Ivanova
1,2,*,
Irina S. Edelman
1,2,
Chun-Rong Lin
3,*,
Evgeniy S. Svetlitsky
1,
Alexey E. Sokolov
1,2,
Kirill A. Lukyanenko
2,4,5,
Alexander L. Sukhachev
1,
Nikolay P. Shestakov
1,
Ying-Zhen Chen
3 and
Aleksandr A. Spivakov
3
1
Kirensky Institute of Physics, Federal Research Center KSC Siberian Branch, Russian Academy of Sciences, Krasnoyarsk 660036, Russia
2
Institute of Engineering Physics and Radioelectronics, Siberian Federal University, Krasnoyarsk 660041, Russia
3
Department of Applied Physics, National Pingtung University, Pingtung City 90003, Taiwan
4
Laboratory of Biomolecular and Medical Technologies, Krasnoyarsk State Medical University Named after Prof. V.F. Voino-Yasenetsky, Krasnoyarsk 660022, Russia
5
Laboratory for Digital Controlled Drugs and Theranostics, Federal Research Center KSC Siberian Branch, Russian Academy of Sciences, Krasnoyarsk 660036, Russia
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(1), 23; https://doi.org/10.3390/ma16010023
Submission received: 9 November 2022 / Revised: 14 December 2022 / Accepted: 15 December 2022 / Published: 20 December 2022
(This article belongs to the Special Issue Functional Nanomaterials for a Better Life)

Abstract

:
The morphology, structure, and magnetic properties of Fe3O4 and Fe3O4@C nanoparticles, as well their effectiveness for organic dye adsorption and targeted destruction of carcinoma cells, were studied. The nanoparticles exhibited a high magnetic saturation value (79.4 and 63.8 emu/g, correspondingly) to facilitate magnetic separation. It has been shown that surface properties play a key role in the adsorption process. Both types of organic dyes—cationic (Rhodomine C) and anionic (Congo Red and Eosine)—were well adsorbed by the Fe3O4 nanoparticles’ surface, and the adsorption process was described by the polymolecular adsorption model with a maximum adsorption capacity of 58, 22, and 14 mg/g for Congo Red, Eosine, and Rhodomine C, correspondingly. In this case, the kinetic data were described well by the pseudo-first-order model. Carbon-coated particles selectively adsorbed only cationic dyes, and the adsorption process for Methylene Blue was described by the Freundlich model, with a maximum adsorption capacity of 14 mg/g. For the case of Rhodomine C, the adsorption isotherm has a polymolecular character with a maximum adsorption capacity of 34 mg/g. To realize the targeted destruction of the carcinoma cells, the Fe3O4@C nanoparticles were functionalized with aptamers, and an experiment on the Ehrlich ascetic carcinoma cells’ destruction was carried out successively using a low-frequency alternating magnetic field. The number of cells destroyed as a result of their interaction with Fe3O4@C nanoparticles in an alternating magnetic field was 27%, compared with the number of naturally dead control cells of 6%.

1. Introduction

In recent decades, nanoscale materials have taken an increasingly important place, not only in engineering and technology but also in such areas of human activity as environmental protection or medical applications. Now, it is already impossible to solve, for example, the problems of water purification from a wide variety of pollutants without the use of nanoscale materials [1,2,3]. The adsorption of pollutants by the surface of nanoparticles is one of the main methods for purifying liquid substances, including water. Nanoparticles made of various materials and structures based on them showed high adsorption efficiency with respect to a large number of pollutants. Among other nanomaterials, tri-metallic layered double hydroxide (NiZnAl-LDH) nanosheets [4,5] and novel bimetallic (ZIF-8@ZIF-67) zeolitic imidazolate [6] were successively applied as adsorbents of rhodamine B and methyl orange. Natural montmorillonite modified with cetyl trimethyl ammonium bromide was used as an adsorbent for the removal of orange G anionic dye from aqueous solutions [7]. Chitosan modificated with 2,3-dihydroxy-benzaldehyde was successfully used for removing Rhodamine B dye from aqueous solution [8]. The authors of ref. [9] successively used the microorganism–graphene oxide (GO) composites for Uranium (VI) adsorption from aqueous solutions and revealed that Lysinibacillus-GO exhibited superior U (VI) removal ability. A Raphia–microorganism composite consisting of immobilized yeast cells Saccharomyces cerevisiae on Raphia farinifera fibers was studied in [10] as a sorbent for removing lead ions from aqueous solutions.
Considerable attention has recently been paid to magnetic nanoparticles (NPs). Being sometimes inferior to other types of nanoparticles in absorption capacity, magnetic NPs have the advantage that they are easily removed from a liquid medium using a magnetic field. In those situations, where the rapid removal of nanoparticles with adsorbed pollutants from water is critical, magnetic nanoparticles have no competitors [11,12,13]. An application of magnetic NPs as adsorbents is due to the ability of their surface to attach various compounds to themselves [1,14]. Having an indisputable number of advantages, the use of NPs faces certain difficulties, such as aggregation, surface contamination with substances from the atmosphere, adaptation of a laboratory experiment to large volumes, etc., and for the successful use of NPs on an industrial scale, all these problems must be solved.
It is known that the ability of magnetic NPs, in particular magnetite, to adsorb substances of different nature is due to both the presence of hydroxyl groups on their surface and to coordinately unsaturated Fe3+ and Fe2+ ions. Depending on the nature and structure of the adsorbate molecules, they are easily retained on the active sites of the magnetite surface due to electrostatic, donor–acceptor, hydrophobic interactions, and/or the formation of hydrogen bonds. In the process of covalent immobilization, strong covalent bonds are formed between the modifier and the oxide surface. For these reasons, magnetite NPs are very convenient objects for covering them with various shells, creating core–shell structures, and functionalizing their surface [15,16,17,18]. Carbon is often used to synthesize core–shell magnetic structures. It is a classic material that has long been used as a sorbent, and an inert shell of various modifications of carbon not only protects the magnetic core from oxidation and degradation but also gives better adsorption properties to NPs’ structure. A number of works present data on the study of the adsorption of typical organic dyes of cationic and anionic types by magnetite–carbon nanostructures [19,20,21,22,23,24,25,26]. Note that in all the cases described, the adsorption capacity of magnetite–carbon nanostructured materials is less than that of activated carbon 190–260 mg/g [27], and they adsorb cationic dyes better.
The wide spread in the adsorption capacity of magnetite nanoparticles and core–shell structure Fe3O4@C is due to individual surface properties depending critically on the synthesis technology. Usually, the nanoparticles’ adsorption is studied in relation to one type of pollutant, for example, to one type of dye. Study of the adsorption characteristics of the nanoparticles produced in one technological cycle with respect to various dyes, both cationic and anionic, undertaken in the first part of the work, seems to be a very useful task. Such a study could find the best ways to activate the NPs’ surface and select the proper candidates for specific applications.
The second part of the work is devoted to studying the possibility of using the synthesized Fe3O4@C NPs with the modified surface in biomedical applications. Surface modification is widely used in biomedical applications, for example, surface-modified NPs are being studied for use as nanocarriers for targeted drug delivery [18,28,29], as sensors for detecting biomolecules [30], for antibacterial therapy [17,31], for the treatment of cancer and iron deficiency anemia [32], and for magneto-mechanical destruction of cancer cells in low-frequency alternating or rotating magnetic fields [33,34,35]. The latter application is being actively developed, and the selective attachment of nanoparticles to cells that need to be destroyed is facilitated by their functionalization with an aptamer [34,36,37,38]. Aptamers are short single-stranded DNA or RNA oligonucleotides and are often referred to as synthetic analogs of antibodies [39]. They are also widely used in cancer diagnostics and therapy [40].
Cultures of Ehrlich carcinoma cancer cells are the most accessible for experiments, and recently, they have become model objects for studying the destruction of cancer cells under various influences, including the action of a magnetic field [34,41]. So, in ref. [34], aptamer-conjugated superparamagnetic iron NPs coated with ferroarabinogalactan were used for targeted magneto-dynamic therapy of an Ehrlich carcinoma tumor in vivo. Ref. [41] presents magneto-dynamic therapy using the gold-coated magnetic NPs functionalized with DNA aptamers to selectively kill tumor cells in vivo. The search for various biocompatible shells that homogeneously cover the magnetic core and the development of methods for attaching specific aptamers to them in order to influence cancer cells is an urgent task of modern research.
The present work is related to both lines of research described above. We investigated the adsorption of two organic dyes belonging to the cationic group and two dyes of the anionic group by Fe3O4 and core–shell Fe3O4@C NPs, having previously studied their morphology, structure, and magnetic properties. Typical cationic (Methylene Blue (MB), rhodomine C (RhC)) and anionic (Congo Red (CR), and Eosin Y (EoY)) were chosen for investigation. We also functionalized Fe3O4@C NPs with aptamers and used them to carry out experiments on the magneto-mechanical disruption of Ehrlich ascite carcinoma cell cultures.

2. Materials and Methods

2.1. Synthesis of Nanoparticles

Fe3O4 NPs were obtained as a result of the thermal decomposition of the iron–oleate complex. First, 120 mmol sodium oleate and 40 mmol iron (III) chloride hexahydrate were dissolved in a mixture of solvents (80 mL alcohol, 60 mL deionized water, and 140 mL n-hexane, C6H14) at 70 °C in air for 4 h and then cooled to room temperature. After cooling, the reaction product was divided into two layers. The top layer was an organic solution, the bottom layer was water. While mixing the two layers, 30 mL of deionized water was added to the mixture and the process was repeated three times, separating the sticky iron–oleate complex. Second, 40 mmol of iron–oleate complex and 20 mmole oleic acid (OA) were dissolved in 200 g of 1-Octadecene (ODE) at 320 °C under air for 3 h, and then cooled down to room temperature. After the mixture had cooled, 500 mL of ethanol was added to it. As a result, it again separated into two layers, the top layer was removed, and n-Hexane was added. A black material was precipitated and separated via applied magnetic field and washed several times with n-hexane. Finally, precipitate was dried at 30 °C for 6 h. The described process is schematically presented in Figure 1, top panel.
To prepare carbon-coated NPs, 0.2 g of Fe3O4 NPs and glucose (1 g) were dispersed in distilled water (30 mL) by ultrasonication for 15 min. The mixture was stirred for 30 min and then sealed in a Teflon-lined stainless-steel autoclave (50 mL capacity). The autoclave was heated to and maintained at 200 °C for 12 h, and then was allowed to cool to room temperature. The black products were separated by externally applied magnetic field and washed several times with water and ethanol. Finally, they were dried at 60 °C for 6 h. This method of carbon coating was already used by us earlier to create a second layer of carbon on particles that already have a carbon shell [20]. The scheme of NPs’ carbon coating is shown in Figure 1, bottom panel.

2.2. Functionalization of Fe3O4@C NPs by Aptamers

To use the obtained Fe3O4@C NPs in experiments with carcinoma cells, the carbon shell was chemically activated. First, the particles were treated in nitric acid to create carboxyl groups, similarly to the procedure used in [42] for the chemical oxidation of carbon-based shell magnetic NPs to improve colloidal stability. Then, they were activated similarly to the procedure used in [30] for the activation of the carboxyl group of carbon-coated nanotubes carboxylated by π–π stacking interaction, which were subsequently used for conjugation of aptamers. Next, the AS-42 aptamer was used, with the amino-primer AmPrcom5 complementary to the 5’-tail. This aptamer binds to Ehrlich ascitic carcinoma cells with high specificity and affinity [38]. The concentration of the aptamer (-nh2) in the buffer solution was 10 mM, and the incubation time was 30 min. NPs functionalized in this way with the aptamer were mixed with Ehrlich ascitic carcinoma cells for 30 min. To visualize the experiment, living cells were preliminarily tinted with trepan blue. The scheme of the functionalization of Fe3O4@C NPs by aptamers is shown in Figure 2.

2.3. Dyes

Methylene Blue (C16H18ClN3S) is a thiazine cationic dye soluble in water and ethanol but insoluble in ether, benzene, and CHCl3. Its electronic and structural characteristics are close to those of the active groups of flavin enzymes. Therefore, experiments with MB can serve to model the functions of the active groups of enzymes.
Rhodamine C (C28H31ClN2O3) is the cationic fluoronic dye. Rhodamine dyes are generally toxic and soluble in water, methanol, and ethanol.
Congo Red (C32H22N6Na2O6S2) is the disodium salt of 4,4′-bis-(1-amino-4-sulfo-2-naphthylazo)biphenyl, a typical chemical anionic azo dye and indicator, slightly soluble in cold water, and easily soluble in hot water, ammonia, and ethyl alcohol. It is a chemical dye that binds to the β-amyloid protein that accumulates in neurons in Alzheimer’s disease.
Eosine Y (C20H6Br4O5Na2) is the xanthene anionic dye, which is soluble in water and is obtained by the action of bromine on fluorescein. It is dyed pink. It is believed that the Eosine molecules are able to absorb light in the visible region of the spectrum (spectral sensitization), which contributes to a significant improvement in the photovoltaic characteristics of solar cells coated with this dye.

2.4. Characteristic Methods

The synthesized NPs were examined with transmission electron microscope (TEM) JEOL JEM-1230 (JEOL Ltd., Tokyo, Japan) operated at an accelerating voltage of 80 kV (Precision Instruments Center of NPUST). The X-ray diffraction measurements were performed using a Bruker D8 Advance diffractometer (Bruker Corp., Billerica, MA, USA) with Cu Kα radiation, 40 kV, 25 mA, λ = 1.5418 Å.
Fourier transform infrared absorption (FT-IR) spectra were recorded with a VERTEX 70 (Bruker Optik GMBH, Leipzig, Germany) spectrometer in the spectral region of 400 ÷ 4000 cm−1 with the resolution 4 cm−1. The spectrometer was equipped with a Globar as the light source and a wideband KBr beam splitter and RT-DLaTG as the detector (Bruker Optik GMBH, Leipzig, Germany). For the measurements, round tablet samples of about 0.5 mm thick and of 13 mm in diameter containing NPs were prepared as follows: nanopowders in an amount lower than 0.001 g were thoroughly ground with 0.14 g of KBr; the mixtures were formed into tablets, which were subjected to cold pressing at 10,000 kg.
The magnetic properties were measured with the vibrating sample magnetometer (VSM 8600) Lakeshore (Lake Shore Cryotronics, Westerville, OH, USA).
Changes in the absorption spectra of the dye solutions were recorded with the UV/VIS circular dichroism spectrometer SKD-2MUF (OEP ISAN, Moscow, Russia). The quartz cells with optical path length of 5 mm were used.
For experiments with living cells, the aptamer-functionalized nanoparticles were incubated for 30 min with Ehrlich’s ascites carcinoma cells. The samples were placed in an alternating magnetic field Hmax = 100 Oe at 50 Hz for 20 min. Photographs were taken with the Levenhuk M1400 PLUS (Levenhuk LLC, Fremont, CA, USA) microscope camera at 10× magnification 2 h after exposure to a magnetic field.

3. Results and Discussion

3.1. Morphology, Structure, and Magnetic Properties

The morphology of Fe3O4 NPs is presented in Figure 3a. They are well-dispersed NPs of about 15 nm in diameter with a narrow size distribution. XRD patterns (Figure 3b) showed that the magnetic core of both Fe3O4 and Fe3O4@C NPs has the spinel ferrite crystal structure, with the parameters of the most intense peaks corresponding to the Fe3O4 phase (PDF Card # 04-005-4319). The XRD peaks’ narrowing in the case of Fe3O4@C NPs comparing with Fe3O4 NPs indicates ordering of the outer layers of particles and/or improving the quality of crystallites in the volume of particles during the formation of the carbon shell when the Fe3O4 NPs are kept in a glucose solution for a long time.
The FT-IR spectra are shown in Figure 4. In the spectrum of pure Fe3O4 NPs, an intense band at 584 cm1 was due to the Fe–O stretching vibrations (ν1-band) characteristic of magnetite [43,44,45,46]. This band was observed in the spectra of all samples. A splitted weaker band is seen for pure magnetite at 403 and 449 cm1, corresponding to the ν2-band of the Fe–O linkage in magnetite [44]. Weak C=O bands are seen in the pure Fe3O4 NPs at 1625 and 1715 cm1, which is explained by the NPs’ synthesis with hexane. At the transition from magnetite NPs to core–shell Fe3O4@C NPs and then to core–shell NPs after treatment in nitric acid Fe3O4@C–COOH, these bands become stronger, confirming the correct identification. In the carbon-coated particles, C-O bands appeared at 1050 and 812 cm1, the CH2-O-CH2 group was indicated by the appearance of a wide band in the region of 1150–1100 cm1 [46]. In the nanoparticles treated in acid, the carboxyl group was clearly identified by the 1384 cm1 band and CH2COOH by the 1475 cm1 band. A slightly noticeable band at 798 cm1 is characteristic of pendulum oscillations of the NH3 group and stretching vibrations –OH in the composition of the carboxyl group in the region of 3450 cm−1 [47]. Thus, FT-IR spectra confirmed, at first, the coating of nanoparticles with carbon and, second, its further carboxylation for bio-functionalization of the surface.
The high values of saturation magnetization Ms = 79 emu/g in the magnetite NPs and 64 emu/g in the carbon-coated NPs were close to the saturation magnetization of bulk magnetite 84 emu/g (Figure 5). The lower Ms value of Fe3O4@C NPs can be due to a certain fraction of carbon in the mass of NPs, as the total NPs mass is used to determine Ms from the measured magnetization. The magnetization curves’ shape with narrow hysteresis loop is typical for single-domain superparamagnetic particles. Each particle has a magnetic moment directed along its magnetic easy axis. In an assemble of noninteracting particles, their easy axes in the absence of magnetic field are oriented randomly, and the total magnetic moments of an assemble equals to zero (point at origin in Figure 5). If we imagine the magnetic moments of particles emitting from one point, then the picture will look like a hedgehog curled up into a ball. When particles are fixed in a space, the application of a magnetic field causes the magnetic moment of each particle to rotate in the direction of the field. To orient the magnetic moment of a particle along the direction of the magnetic field, the energy of the latter must exceed the energy barrier equal to ∆E = KV, where K is the anisotropy energy and V is the particle volume. Additionally, the magnetic field must be higher the larger the angle between the magnetic field and the easy axis directions. In this way, the magnetic moments of the particles will gradually turn towards the direction of the magnetic field until all are settled along the field. This process is described by Langevin’s equation.
L ( x ) = coth ( x ) 1 x
where x = g μ B J B k B T , g is g-factor, μ B   is the Bohr magneton, J is the total angle moment, B =   μ 0 H , μ 0 is the vacuum magnetic permeability, k B is the Boltzmann’s constant, and T is temperature. At x→∞, the magnetization saturates and the magnetic moments line up completely in the direction of the applied field. When the magnetic field decreases, the magnetic moments of the particles turn to their easy axes. However, in this case, in a zero field, they will all be oriented in a half-plane, resembling a fan, and thus some remnant magnetization MR of the entire ensemble will be preserved. Namely, such a situation occurs when measuring magnetization of the NPs powder densely packed in a special capsule. Being dispersed in a liquid, NPs participate in free motion—Brownian motion. The magnetic field greatly affects the type of this movement. In a DC gradient field, particles move along the gradient from a smaller field to a larger one and collect at the point of maximum field. This effect is of great interest for solving problems of water purification (and any liquids) from pollution, as it can be used to extract particles with the adsorbed pollutions from the water. In an AC magnetic field, nanoparticles behave like a compass needle, tending to settle in such a way that their magnetic moment is directed in the direction of the magnetic field and, at the same time, begins to rotate. This rotational motion can occur while maintaining a constant orientation of the magnetic moment inside the particle. To implement such a motion, the energy of the interaction of the magnetic field with the magnetization of the particle (Zeeman energy) must exceed the energy of the viscous deceleration of the particle in the fluid. Obviously, the Zeeman energy is greater the greater the magnetic moment of the particle. Such an action can be used in therapy to destroy cancer cells in the body. Rotational motion of free spherical magnetic NPs in a viscous medium under the action of a low-frequency magnetic field was considered, for example, in [35], where functionalized magnetic NPs with a size of 10–50 nm were shown to be capable to generate forces sufficient for cell destruction under an action of low-frequency magnetic fields (1 to 100 Hz) and magnetic induction of hundreds of mT.

3.2. Dyes Adsorption by Fe3O4 NPs

The optical absorption spectra of two anionic CR and EoY and two cationic MB and RhC organic dyes dissolved in distilled water (pH 5.5) in a given concentration were used as the reference spectra to find out the adsorption capacity of the investigated NPs for these dyes. Measurements were carried out at 25 °C. The dye concentration was tied to the absorption intensity at the wavelength corresponding to the maxima in the optical spectra of 495 nm for CR, 514 nm for EoY, 662 nm for MB, and 552 nm for RhC. The Figure 6 shows the optical absorption spectra of the listed dyes. The positions of the maxima are typical for these dye solutions and are close to those determined by other researchers for CR (498 nm) [48], MB (664 nm) [49], EoY (522 nm) [50], and RhC (553 nm) [51]. Thus, the change in dye concentration in the solution could be determined from the change in these lines’ intensities.
In the experiment, a certain concentration of the dye’s aqueous solution was mixed with NPs in an ultrasonic bath for 10 min, after that, NPs with adsorbed dyes were separated from the solution by applying a magnetic field, the intensities of the corresponding lines in the optical absorption spectra of the remaining solution were measured, and the remaining dye’s concentration was determined.
The values of the adsorption capacity qt (mg/g) and adsorption percentage (θ) of NPs were calculated as follows [52]:
q t = ( C 0 C t ) V m
θ = ( C 0 C t ) C 0 · 100 % ,
where C0 and Ct (mg/L) are the initial concentration and concentration of the dye at the contact time t (min), V (L) is the volume of the solution, and m (g) is the mass of the adsorbent (NPs).
The results of the effect of the contact time on the adsorption of Fe3O4 NPs for the various dyes are shown in Figure 7 at the initial dye concentration of C0 = 60 mg/L for CR and C0 = 30 mg/L for the other dyes.
The qt(t) dependencies are different for different dyes: Fe3O4 NPs adsorb the anionic dye CR very well, although the value of EoY adsorption is two and a half times less than for CR; both dependences are characterized by a short time to reach the equilibrium value. The kinetic of adsorption of the cationic dyes (MB and RhC) is very slow, and saturation is not reached for a very long time.
Usually, two main models were considered to describe kinetics of dye adsorption by NPs [52]—the pseudo-first
q t = q e ( 1 e k 1 t )
and pseudo-second order
q t = q e 2 k 2 t 1 + q e k 2 t
where k1 (1/mg) and k2 (g/(mg*min) are the rate constants of the sorption reaction in the pseudo-first and pseudo-second orders, respectively, and qe (mg/g) is the adsorption capacity at equilibrium. In Figure 7a,b, black solid and dotted lines are the results of the pseudo-first-order (PFO) and pseudo-second-order (PSO) nonlinear fitting of adsorption data. The advantages of nonlinear modeling were shown in ref. [53]. Nonlinear modeling was carried out using the OriginPro 2016 program (OriginLab Corporation, Northampton, MA, USA) utilizing the Orthogonal distance regression algorithm. The parameters extracted from the nonlinear fitting of the kinetic data are collected in Table 1. It can be seen that correlation coefficients (R2) describing the kinetic data for anionic dyes are close for both the pseudo-first and pseudo-second order, but it can be seen from Figure 7 and comparison qe with the experimental value that it is the pseudo-first-order description that makes it possible to describe the data well, and in this case, the greater thermodynamic benefit of the interaction of the sorbate–sorbent type prevails. In the case of cationic dyes, the kinetic curves are better described by the pseudo-second-order model.
Diffusion equation is often used to describe the adsorption kinetics based on the presentation of kinetic data in coordinates qt in dependence on t1/2, which makes it possible to estimate the contribution of internal and external diffusion processes [54]:
q t = k t 0.5 + C
where k (mg/g.min0.5) is the intraparticle diffusion rate constant, and C (mg/g) is the point of intersection of the linear curve with the y-axis; this parameter is related to the thickness of the boundary layer.
From Figure 8, it can be seen that such dependence is described by one straight line only passing through the origin in the case of MB. According to the Weber and Morris model [54], which is intensively used in the analysis of the mechanisms of adsorption of nanoparticles, if such dependence is described by a single straight line passing through the origin, then the adsorption process is controlled only by intra-particle diffusion. A more complex (polylinear) character of this dependence indicates the involvement of intra-particle diffusion processes as well. The parameters of the intra-particle diffusion model are given in Supplementary Material Table S1.
The rapid removal of the dyes in the initial stage, and the nearly horizontal curve at the later stages (Figure 8a), indicate the predominant role of external surface diffusion and the insignificant role of the dye molecules’ diffusion into the particle micropores in the case of anionic dyes.
Concentration dependencies of the adsorption efficiency are shown in Figure 9. It is seen from Figure 9a that the removal efficiency (right axis) of CR is very high at low dye concentration, up to 90% for C0 = 50 mg/L, and drops gradually up to 60 % with a further increase in the dye concentration. Instantaneous adsorption at low dye concentrations corresponds to the almost vertical initial part of the adsorption isotherm in Figure 9b. The maximum adsorption capacity at 25 °C in the concentration range studied is 58.9 mg/g for CR.
The adsorption isotherms for the several dyes and the experimental data fittings to models of adsorption from liquid phase are shown in Figure 9. It is seen that Langmuir curves (dotted lines in Figure 9b–d) for monomolecular adsorption well describes the isotherm for all investigated dyes in the low concentration range only, and the equilibrium saturation concentration postulated in this model is not reached at all. According to the monomolecular model, with an increase in the proportion of occupied adsorption sites, it is more difficult for the adsorptive molecules to find a vacant site, and the system comes to the equilibrium state. An increase in adsorption after filling the monolayer and the plateau transformation into an increasing curve may be due to polymolecular adsorption or reorientation of molecules, or an appearance of their associates relative to the NP surface. The polymolecular adsorption model was suggested by S. Brunauer, P. H. Emmett, and E. Teller [55]. Several cases were considered in this model, mentioned now as the BET model. The equation for polymolecular adsorption in Case 3 [56], when the number of adsorbing layers n = ∞, is as follows:
q = q m K S C e ( 1 K L C e ) ( 1 K L C e + K S C e ) ,
which for n = 1 coincides with the Langmuir equation
q = q m K S C e 1 + K S C e .
In these equations, KS (L/mg) and KL (L/mg) are equilibrium constants of adsorption for the first and upper layers, correspondingly, Ce (mg/L) is the residual dye concentration, and qm (mg/g) is the maximal adsorption capacity. For all dyes investigated here, experimental data are well described in the frames of the BET model (Figure 9b–d). Isotherm parameters calculated for CR according to Langmuir and BET models are shown in Table 2. The low value of the correlation coefficient R2 in the case of the Langmuir modeling and R2 value close to unity confirm the predominantly polymolecular nature of adsorption in the case under consideration. For other dyes, results are analogous.
Thus, adsorption by Fe3O4 NP can be used in solutions containing complex pollutants; they show preferential absorption of anionic dyes (especially CR), but they also adsorb cationic dyes, albeit at a lower capacity and rate. The adsorption processes on Fe3O4 NPs are well described on the bases of the theory of polymolecular (multilayer) adsorption.

3.3. Dyes Adsorption by Fe3O4@C NPs

Magnetite NPs coated with carbon shell showed selective adsorption of cationic dyes: they did not absorb anionic dyes CR and EoY at all, as well as a number of other dyes tested by us (methyl orange and aniline yellow). The magnitude and rate of adsorption of cationic dyes show larger values compared with Fe3O4 NPs without shell. Figure 10 shows the kinetic curves of the adsorption of MB and RhC. Results of the nonlinear curve fitting of the kinetic adsorption data are collected in Table 3.
It can be seen that the kinetic data for MB are described more closely by the pseudo-first-order model, while the adsorption of RhC is closer to being described by the pseudo-second order.
Interestingly, the adsorption isotherms of the two cationic dyes MB and RhC; are different (Figure 11). In the first case, it is closer to the Friendlich model [57] (also monomolecular as the Langmuir model) and, in the second case, to the BET model of polymolecular adsorption. The proximity to the Freundlich model shows that adsorbate molecules can interact with each other on the surface of the adsorbent, and the increase in adsorption does not stop after the formation of a monomolecular layer, in contrast to the Langmuir model. For the case of RhC, the adsorption isotherm has a polymolecular character (Figure 11b).

3.4. Comparison and Mechanism

Now, one can compare the adsorption capacity of the studied NPs with some of theliterature data. The large variation in the adsorption capacity values obtained by different authors (Table 4) is due to not only the variety of materials studied but also to the influence of various factors on the adsorption mechanism and their complex accounting in nanoscale systems. Note that the S-type sorption isotherms obtained here for CR, RhC, and EoY on Fe3O4 NPs and for RhC on Fe3O4@C NPs have not been observed by other authors who studied the adsorption of analogous dyes on magnetic nanoparticles. Additionally, in the majority of articles, kinetic data are well described by the pseudo-adsorption equation of the second order, and isotherms are close to the Langmuir model. Only in the work in [58], the authors propose a model of multilayer sorption with calculated thermodynamic parameters close to physical sorption, although the results are described by a second-order pseudo-equation and are in good agreement with the Langmuir model. As concerns the capacity value, both types of the studied NPs demonstrate low capacity for Methylene Blue. For the other three dyes, the absorption capacity is close to the results of other authors which, in combination with high magnetization, allows us to consider them as successful objects for studying promising adsorbents removed from water using a magnetic field.
The FT-IR spectra of Fe3O4 NPs and Fe3O4 NPs after adsorption of CR are shown in Figure 12. In the spectrum of nanoparticles after adsorption, changes in line intensities are noticeable, compared with the spectrum of particles before adsorption, and new features appeared. After adsorption of the dye, a line at 1046 cm−1 is clearly visible, characteristic of the formation of C=O bonds. In ref. [19], analogical peaks at 878 cm−1 and 1046 cm−1 appeared in the FT-IR Fe3O4@C spectra after MB adsorption, which were attributed to C-H in the benzene ring and the rocking vibration -CH3. Note that the CR molecule contains six benzene rings in its structure, and similar features are clearly distinguishable on the FT-IR spectrum of our Fe3O4 NPs after CR adsorption. Besides, slightly visible traces at 1178; at 1228 cm1, can be referred to the asymmetry stretching vibration of the S-O (SO3-H) group-characteristic of the CR dye [64]. In ref. [60], some shifts of the characteristic peaks and a decrease in their amplitude after the adsorption process were also noted, which the authors attribute to the electrostatic interaction of the adsorbent surface with the RC dye. Thus, the observed changes, i.e., the shift of the peaks and the appearance of new peaks, confirm the successful adsorption of the Congo Red dye on the surface of the Fe3O4 NPs. Similar results are obtained for other cases.
Differences in the adsorption capacity of the Fe3O4 and Fe3O4@C NPs with respect to cationic and anionic dyes indicate a significant difference between the adsorption mechanisms in these two cases. The main mechanisms of adsorption of the cationic dye MB on the carbon shell are considered, for example, in [61]. These are electrostatic interactions, hydrogen bonds, and chemical bonds between dye molecules and adsorbent. The presence of functional groups (such as -OH, COOH, and -CO) on the carbon surface provides a suitable platform for inducing electrostatic attraction between the cationic dyes and the adsorbent. Since, in our case, carbon-coated NPs exhibit selective adsorption of cationic dyes, in the case of Fe3O4@C NPs, electrostatic interaction predominates. A possible mechanism for the interaction of NPs with MB molecules is schematically shown in Figure 13a.
More complex situation takes place in the case of Fe3O4 NPs, they adsorb both cationic and anionic dyes, but a strong difference is observed in the NPs’ adsorption capacity with respect to different representatives of the same group. However, the adsorption isotherms have the same character for all studied dyes. The initial steep rise of the isotherm becomes gentler, then their steepness increases again. Qualitatively, it can be assumed that the decrease in steepness is associated with a decrease in the number of free places on the surface (which corresponds to the Langmuir model), and its subsequent increase is due to ever-increasing adsorption in the second and subsequent layers. Thus, it can be assumed that interaction occurs not only between the NPs surface and the arriving molecules but also between molecules arriving from the liquid phase and previously adsorbed molecules. The change in the electron density in arriving molecules is synchronized with the change in the electron density of already adsorbed molecules. Both orientational (in the case of polar molecules) and induced (in the case of charge induction in a molecule approaching the adsorbed layers) interactions arise. The contribution of the longitudinal interactions of molecules is negligibly small compared with perpendicular interactions. Simultaneously, adsorption proceeds on the surface of a solid body and on previously adsorbed molecules of the first, second, third or other layers, and only for the first layer the average residence time of molecules in it is longer than for molecules in other layers, where it is the same for any layer. On the surface, thus, vertical complexes are formed, consisting of one, two, three, etc., molecules and, in addition, there are free places. Such a process is shown schematically in Figure 13b.

3.5. Desorption and Reusability Studies

To establish the possibility of reusing the particles as adsorbents, desorption experiments were carried out. The regeneration of the Fe3O4 NPs after adsorption of all dyes was carried out by washing several times (4–5 times) in ethanol until the solution was transparent. The Fe3O4@C NPs after adsorption of MB were regenerated first by treatment in a 0.01 mol−1 HCl solution (5 min) and subsequent washing in ethanol several times (2–3 times) until the solution was transparent. Then, the NPs washed in deionized water were dried (6 h. at room temperature) and used in subsequent rounds.
The adsorption–desorption cycle was repeated for times using the same NPs, and the results for CR adsorption by Fe3O4 NPs and MB adsorption by Fe3O4@C NPs are shown in Figure 14. It was observed that the adsorption capacities in both cases decreases by no more than 10% after four cycles. The percentage of CR removal was 85.6, 82.5, 79.98, and 78, and for MB, 65.3, 64.9, 64.2, and 58.3% from cycle to cycle. The NPs thus showed a high regenerative capacity after four cycles of adsorption–desorption. The easy desorption is most likely due to the predominance of physical adsorption processes.

3.6. Magneto-Mechanical Destruction of Ehrlich Ascites Carcinoma Cells by Fe3O4@C NPs

For experiments with living cells, aptamer-functionalized NPs were incubated for 30 min with Ehrlich ascitic carcinoma cells and then exposed to a low-frequency alternating magnetic field. Destroyed cells are clearly visible in the microscope chamber as blue spots (Figure 15a).
To reveal the combined effect of immobilization of magnetic NPs by aptamers and a low-frequency magnetic field on the death of model cancer cells, we performed a statistical analysis of four cases: (1) control cells; (2) cells mixed with NP functionalized by aptamers without applying a magnetic field; (3) cells mixed with NP functionalized by aptamers after exposure to an alternating magnetic field; and (4) a mixture of cells with initial Fe3O4@C NPs after action of an alternating magnetic field. The percentage of disrupted ascites cells in the test samples was statistically evaluated on five independent probes of each sample, processing at least 10 different images of each so that the average number of particles in one image was 100–180 (Figure 15b).
Experimental data showed that the use of magnetic NPs increased the amount of disrupted cells. More than 28% of cells were dead in cells with aptamer-functionalized Fe3O4@C NPs after exposure to the alternating magnetic field compared with the ~6% control cell death. The percentage of destroyed cells in all cases is greater than in the control cells, which indicates the attachment of nanoparticles to cells in all cases. In our experiment, the frequency of the change in the magnetic field was determined by the frequency of the electric alternating current in the network and was 50 Hz. Experiments carried out by other authors, for example, in [65], showed that the most effective cell destruction frequency range is 10–40 Hz. However, the results of this work, when using a field of 90 Oe and a frequency of 50 Hz, are comparable with ours (28%), so the cell death of glioblastoma multiforme (aggressive form of brain cancer) was 8 and 23% when using magnetic particles coated with IrG and IL13, respectively. Based on the results presented in [65], and in our case, we can assume an increase in cell mortality of 80–90% with a decrease in the frequency of the applied magnetic field.

4. Conclusions

Thermal decomposition was used to synthesize Fe3O4 and Fe3O4@C nanoparticles (NPs) uniform in shape and size with an average size of 15 ± 2 nm, and their morphology, structure, and magnetic properties were studied. The formation of a carbon shell around the magnetite core and the modification of the NPs surface by carboxylated groups were confirmed by IR-Fourier spectra. The obtained high value of the saturation magnetization (more than 63 emu/g) of nanoparticles is their important advantage, since it facilitates the extraction of particles after the adsorption of pollutants from the treated water.
The NPs capability to adsorb the organic dyes Methylene Blue (MB), Rhodomine C (RhC), Congo Red (CR), and Eosin Y (EoY) was studied. It is shown that the adsorption kinetics depends not only on the type of dye but also on its composition. The highest capacity (58 mg/g) was observed upon adsorption of CR on Fe3O4 NPs. At dye concentrations up to 50 mg/L, more than 90 percent of the dye was removed instantly. At higher dye concentrations, the sorption reaction reached equilibrium in approximately 75 min. Nonlinear modeling showed that, in the case of Fe3O4 NPs, the kinetic data of the sorption reaction are better described by a pseudo-first-order equation for anionic dyes and a pseudo-second-order equation for cationic dyes. Adsorption isotherms in the region of low pollutant concentrations fit well into the Langmuir model for all the dyes studied, i.e., sorption can be due to a monomolecular model. With an increase in the concentration of the dye in the solution, the trend changes, and the experimental value of qe for all the dyes used continues to grow, that is, the adsorption process is more adequately described by the BET equation for the polymolecular adsorption model.
Fe3O4@C NPs do not adsorb anionic dyes, and thus have a selective adsorption capacity for cationic dyes. For MB, the kinetic data are more accurately described by a pseudo-first-order equation and, for RhC, by a pseudo-second-order equation. The MB adsorption isotherm is well described by a monomolecular process, and it is better in the Freundlich model than in the Langmuir model.
The results obtained for bio-functionalized Fe3O4@C particles can be used for targeted magneto-mechanical destruction of cancer cells. So, the percent of disrupted cells as results of interaction with Fe3O4@C NPs functionalized with aptamers after exposure to a low-frequency magnetic field was 27%, compared with the number of dead control cells of 6%. Thus, the investigated nanoparticles seem to be perspective for their use in water treatment technologies on one hand and interesting nano-objects for the fundamental study of the dye adsorption processes on the other hand. The first experiments on the use of these particles for the targeted magneto-mechanical destruction of cancer cells also gave encouraging results.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma16010023/s1, Table S1: Kinetic parameters of the intra-particle diffusion model parametres for the adsorption of the dyes (C0 = 60 mg/L for CR and C0 = 30 mg/L for other dyes) on Fe3O4 NPs at 25 °C.

Author Contributions

Conceptualization, I.S.E. and C.-R.L.; methodology, O.S.I., E.S.S. and K.A.L.; software, A.L.S.; validation, O.S.I., E.S.S. and K.A.L.; formal analysis, O.S.I.; investigation, E.S.S., N.P.S., A.L.S., K.A.L. and A.A.S.; resources, A.E.S. and Y.-Z.C.; data curation, O.S.I.; writing—original draft preparation I.S.E. and O.S.I.; writing—review and editing, I.S.E. and C.-R.L.; visualization, Y.-Z.C.; supervision, I.S.E. and C.-R.L.; project administration, I.S.E. and C.-R.L.; funding acquisition, I.S.E., K.A.L. and C.-R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded partly by the Ministry of Science and Higher Education of the Russian Federation, project FWES-2021-0035. C.-R.L., Y.-Z.C. and A.A.S. thank the National Science and Technology Council of Taiwan for the financial support, Grants NSTC № 108-2923-M-153-001-MY3 and № 110-2112-M-153-005-. Magnetic investigations were carried out in the Center for Collective Use of the Krasnoyarsk Regional Center of Research Equipment of Federal Research Center “Krasnoyarsk Science Center SB RAS”.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki and approved by the Institutional Review Board of Kirensky Institute of Physics (b/n 28.10.2022).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available since they are a part of ongoing research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khajeh, M.; Laurent, S.; Dastafkan, K. Nanoadsorbents: Classification, preparation, and applications (with emphasis on aqueous media). Chem. Rev. 2013, 113, 7728–7768. [Google Scholar] [CrossRef] [PubMed]
  2. Gehrke, I.; Geiser, A.; Somborn-Schulz, A. Innovations in nanotechnology for water treatment. Nanotechnol. Sci. Appl. 2015, 8, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Khan, A.; Malik, S.; Shah, S.; Ali, N.; Ali, F.; Ghotekar, S.; Dabhane, H.; Bilal, M. Nanoadsorbents for Environmental Remediation. In Nanotechnology for Environmental Remediation; Wiley: Hoboken, NJ, USA, 2022; pp. 103–120. [Google Scholar] [CrossRef]
  4. Nazir, M.A.; Najam, T.; Jabeen, S.; Wattoo, M.A.; Bashir, M.S.; Shah, S.S.A.; Rehman, A. Facile synthesis of Tri-metallic layered double hydroxides (NiZnAl-LDHs): Adsorption of Rhodamine-B and methyl orange from water. Inorg. Chem. Commun. 2022, 145, 110008. [Google Scholar] [CrossRef]
  5. Nazir, M.A.; Najam, T.; Bashir, M.S.; Javed, M.S.; Bashir, M.A.; Imran, M.; Azhar, U.; Shah, S.S.A.; Rehman, A. Kinetics, isothermal and mechanistic insight into the adsorption of eosin yellow and malachite green from water via tri-metallic layered double hydroxide nanosheets. Korean J. Chem. Eng. 2022, 39, 216–226. [Google Scholar] [CrossRef]
  6. Nazir, M.A.; Najam, T.; Shahzad, K.; Wattoo, M.A.; Hussain, T.; Tufail, M.K.; Shah, S.S.A.; Rehman, A. Heterointerface engineering of water stable ZIF-8@ZIF-67: Adsorption of rhodamine B from water. Surf. Interfaces 2022, 34, 102324. Available online: https://www.sciencedirect.com/science/article/pii/S246802302200585 (accessed on 2 September 2022). [CrossRef]
  7. Ouachtak, H.; Guerdaoui, A.E.; Haounati, R.; Akhouairi, S.; Haouti, R.E.; Hafid, N.; Addi, A.A.; Šljukić, B.; Santosa, D.M.F.; Taha, M.L. Highly efficient and fast batch adsorption of orange G dye from polluted water using superb organo-montmorillonite: Experimental study and molecular dynamics investigation. J. Mol. Liq. 2021, 335, 116560. [Google Scholar] [CrossRef]
  8. Alakhras, F.; Ouachtak, H.; Alhajri, E.; Rehman, R.; Al-Mazaideh, G.; Anastopoulos, I.; Lima, E.C. Adsorptive Removal of Cationic Rhodamine B Dye from Aqueous Solutions Using Chitosan-Derived Schiff Base. Sep. Sci. Technol. 2022, 57, 542–554. [Google Scholar] [CrossRef]
  9. Zhao, C.; Liu, J.; Deng, Y.; Tian, Y.; Zhang, G.; Liao, J.; Yang, J.; Yang, Y.; Liu, N.; Sun, Q. Uranium (Ⅵ) adsorption from aqueous solutions by microorganism-graphene oxide composites via an immobilization approach. J. Clean. Prod. 2019, 236, 117624. [Google Scholar] [CrossRef]
  10. Staroń, P.; Chwastowski, J. Raphia-Microorganism Composite Biosorbent for Lead Ion Removal from Aqueous Solutions. Materials 2021, 14, 7482. [Google Scholar] [CrossRef]
  11. Gómez-Pastora, J.; Bringas, E.; Ortiz, I. Recent progress and future challenges on the use of high performance magnetic nano-adsorbents in environmental applications. Chem. Eng. J. 2014, 256, 187–204. [Google Scholar] [CrossRef]
  12. Lima, M.M.; Macuvele, D.L.P.; Muller, L.; Nones, J.; Silva, L.L.; Fiori, M.A.; Soares, C.; Riella, H.G. Synthesis and Potential Adsorption of Fe3O4@C Core-Shell Nanoparticles for to Removal of Pollutants in Aqueous Solutions: A Brief Review. J. Adv. Chem. Eng. 2017, 7, 1000172. [Google Scholar]
  13. Hlongwane, G.N.; Sekoai, P.T.; Meyyappan, M.; Moothi, K. Simultaneous removal of pollutants from water using nanoparticles: A shift from single pollutant control to multiple pollutant control. Sci. Total Environ. 2019, 656, 808–833. [Google Scholar] [CrossRef] [PubMed]
  14. Abu-Dief, A.M.; Hamdan, S.K. Functionalization of Magnetic Nano Particles: Synthesis, Characterization and Their Application in Water Purification. Am. J. Nanosci. 2016, 2, 26–40. [Google Scholar] [CrossRef]
  15. 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] [Green Version]
  16. Ge, Y.; Li, Y.; Zu, B.; Zhou, C.; Dou, X. AM-DMC-AMPS Multi-Functionalized Magnetic Nanoparticles for Efficient Purification of Complex Multiphase Water System. Nanoscale Res. Lett. 2016, 11, 217. [Google Scholar] [CrossRef] [Green Version]
  17. Nehra, P.; Chauhan, R.P.; Garg, N.; Verma, K. Antibacterial and antifungal activity of chitosan coated iron oxide nanoparticles. Br. J. Biomed. Sci. 2018, 75, 13–18. [Google Scholar] [CrossRef]
  18. Vaytulevich, E.A.; Yurmazova, T.A.; Tuan, H.T. Sorbents Based on Magnetite Nanoparticles for Biomedical Applications. Nanotechnol. Russ. 2019, 14, 33–40. [Google Scholar] [CrossRef]
  19. Xiang, H.; Ren, G.; Zhong, Y.; Xu, D.; Zhang, Z.; Wang, X.; Yang, X. Fe3O4@C Nanoparticles synthesized by In situ solid-phase method for removal of methylene blue. Nanomaterials 2021, 11, 330. [Google Scholar] [CrossRef]
  20. Lin, C.-R.; Ivanova, O.S.; Edelman, I.S.; Knyazev, Y.V.; Zharkov, S.M.; Petrov, D.A.; Sokolov, A.E.; Svetlitsky, E.S.; Velikanov, D.A.; Solovyov, L.A.; et al. Carbon Double Coated Fe3O4@C@C Nanoparticles: Morphology Features, Magnetic Properties, Dye Adsorption. Nanomaterials 2022, 12, 376. [Google Scholar] [CrossRef]
  21. Qu, L.; Han, T.; Luo, Z.; Liu, C.; Mei, Y.; Zhu, T. One-step fabricated Fe3O4@C core-shell composites for dye removal: Kinetics, equilibrium and thermodynamics. J. Phys. Chem. Sol. 2015, 78, 20–27. [Google Scholar] [CrossRef]
  22. Ai, L.; Zhang, C.; Chen, Z. Removal of methylene blue from aqueous solution by a solvothermal-synthesized graphene/magnetite composite. J. Hazard. Mater. 2011, 192, 1515–1524. [Google Scholar] [CrossRef] [PubMed]
  23. Sun, Y.; Ni, P.; Zhu, M.; Yao, Y.; Fu, S. Preparation of Fe3O4@C submicron rods for adsorption of methylene blue and fast separation from water. Micro Nano Lett. 2019, 14, 962–966. [Google Scholar] [CrossRef]
  24. Zhang, X.; He, M.; Liu, J.H.; Liao, R.; Zhao, L.; Xie, J.; Wang, R.; Yang, S.T.; Wang, H.; Liu, Y. Fe3O4@C nanoparticles as high-performance Fenton-like catalyst for dye decoloration. Chin. Sci. Bull. 2014, 59, 3406–3412. [Google Scholar] [CrossRef]
  25. Zhang, Z.; Kong, J. Novel magnetic Fe3O4@C nanoparticles as adsorbents for removal of organic dyes from aqueous solution. J. Hazard. Mater. 2011, 193, 325–329. [Google Scholar] [CrossRef] [PubMed]
  26. Shen, Y.; Guan, C.; Huang, Y.; Zhang, P.; Lai, X.; Du, H.; Xu, X.; Li, J.; Zhang, W.; Qiu, K.; et al. Preparation of a Fe3O4@C magnetic materials with high adsorption capacity of methylene blue. Ferroelectrics 2020, 566, 94–103. [Google Scholar] [CrossRef]
  27. GOST 4453-74; Active Absorpting Powder Charcoal, Specifications. National Standard of the Russian Federation: Moscow, Russia, 1974.
  28. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; Rodriguez-Torres, M.P.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar] [CrossRef] [Green Version]
  29. Vangijzegem, T.; Stanicki, D.; Laurent, S. Magnetic iron oxide nanoparticles for drug delivery: Applications and characteristics. Expert Opin. Drug Deliv. 2019, 16, 69–78. [Google Scholar] [CrossRef]
  30. Nam, K.; Eom, K.; Yang, J.; Park, J.; Lee, G.; Jang, K.; Lee, H.; Lee, S.W.; Yoon, D.S.; Lee, C.Y.; et al. Aptamer-functionalized nano-pattern based on carbon nanotube for sensitive, selective protein detection. J. Mater. Chem. 2012, 22, 23348. [Google Scholar] [CrossRef] [Green Version]
  31. Arakha, M.; Pal, S.; Samantarrai, D.; Panigrahi, T.K.; Mallick, B.C.; Pramanik, K.; Mallick, B.; Jha, S. Antimicrobial activity of iron oxide nanoparticle upon modulation of nanoparticle-bacteria interface. Sci. Rep. 2015, 5, 14813. [Google Scholar] [CrossRef] [Green Version]
  32. Alphandéry, E. Biodistribution and targeting properties of iron oxide nanoparticles for treatments of cancer and iron anemia disease. Nanotoxicology 2019, 13, 573–596. [Google Scholar] [CrossRef]
  33. Naud, C.; Thebault, C.; Carrière, M.; Hou, Y.; Morel, R.; Berger, F.; Dieny, B.; Joisten, H. Cancer treatment by magneto-mechanical effect of particles, a review. Nanoscale Adv. 2020, 2, 3632. [Google Scholar] [CrossRef] [PubMed]
  34. Kolovskaya, O.S.; Zamay, T.N.; Zamay, G.S.; Babkin, V.A.; Medvedeva, E.N.; Neverova, N.A.; Kirichenko, A.K.; Zamay, S.S.; Lapin, I.N.; Morozov, E.V.; et al. Aptamer-conjugated superparamagnetic ferroarabinogalactan nanoparticles for targeted magnetodynamic therapy of cancer. Cancers 2020, 12, 216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Golovin, Y.I.; Klyachko, N.L.; Majouga, A.G.; Gribanovskii, S.L.; Golovin, D.Y.; Zhigachev, A.O.; Shuklinov, A.V.; Efremova, M.V.; Veselov, M.M.; Vlasova, K.Y.; et al. New Approaches to Nanotheranostics: Polyfunctional Magnetic Nanoparticles Activated by Non-Heating Low-Frequency Magnetic Field Control Biochemical System with Molecular Locality and Selectivity. Nanotechnol. Russ. 2018, 13, 215–239. [Google Scholar] [CrossRef]
  36. Sun, H.; Zhu, X.; Lu, P.Y.; Rosato, R.R.; Tan, W.; Zu, Y. Oligonucleotide aptamers: New tools for targeted cancer therapy. Mol. Ther. Nucleic Acids 2014, 3, e182. [Google Scholar] [CrossRef] [PubMed]
  37. Zhi, D.F.; Yang, T.; Yang, J.; Fu, S.; Zhang, S.B. Targeting strategies for superparamagnetic iron oxide nanoparticles in cancer therapy. Acta Biomater. 2019, 102, 13–34. [Google Scholar] [CrossRef]
  38. Zamay, G.S.; Zamay, T.N.; Lukyanenko, K.A.; Kichkailo, A.S. Aptamers increase biocompatibility and reduce the toxicity of magnetic nanoparticles used in biomedicine. Biomedicines 2020, 8, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Kudłak, B.; Wieczerzak, M. Aptamer based tools for environmental and therapeutic monitoring: A review of developments, applications, future perspectives. Crit. Rev. Environ. Sci. Technol. 2020, 50, 816–867. [Google Scholar] [CrossRef]
  40. Hori, S.I.; Herrera, A.; Rossi, J.J.; Zhou, J. Current Advances in Aptamers for Cancer Diagnosis and Therapy. Cancers 2018, 10, 9. [Google Scholar] [CrossRef] [Green Version]
  41. Belyanina, I.V.; Zamay, T.N.; Zamay, G.S.; Zamay, S.S.; Kolovskaya, O.S.; Ivanchenko, T.I.; Denisenko, V.V.; Kirichenko, A.K.; Glazyrin, Y.E.; Garanzha, I.V.; et al. In vivo cancer cells elimination guided by aptamer-functionalized gold-coated magnetic nanoparticles and controlled with low frequency alternating magnetic field. Theranostics 2017, 7, 3326–3337. [Google Scholar] [CrossRef]
  42. Oliveira, J.R.P.; Rodrigues, R.O.; Barros, L.; Ferreira, I.C.F.R.; Marchesi, L.F.; Koneracka, M.; Jurikova, A.; Zavisova, V.; Gomes, H.T. Carbon-Based Magnetic Nanocarrier for Controlled Drug Release: A Green Synthesis Approach. J. Carbon Res. 2019, 5, 1. [Google Scholar] [CrossRef]
  43. 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]
  44. Nunes, J.S.; de Vasconcelos, C.L.; Cabral, F.A.O.; de Araújo, J.H.; Pereira, M.R.; Fonseca, J.L.C. Synthesis and characterization of poly(ethyl methacrylate-co-methacrylic acid) magnetic particles via miniemulsion polymerization. Polymer 2006, 47, 7646–7652. [Google Scholar] [CrossRef]
  45. Mahdavi, M.; Namvar, F.; Ahmad, M.B.; Mohamad, R. Green Biosynthesis and Characterization of Magnetic Iron Oxide (Fe3O4) Nanoparticles Using Seaweed (Sargassum muticum) Aqueous Extract. Molecules 2013, 18, 5954–5964. [Google Scholar] [CrossRef] [PubMed]
  46. Bao, X.; Qiang, Z.; Chang, J.-H.; Ben, W.; Qu, J. Synthesis of carbon-coated magnetic nanocomposite (Fe3O4@C) and its application for sulfonamide antibiotics removal from water. J. Environ. Sci. 2014, 26, 962–969. [Google Scholar] [CrossRef] [PubMed]
  47. Bellamy, L.J. The Infrared Spectra of Complex Molecules; Springer: Dordrecht, The Netherlands, 1980; p. 290. [Google Scholar]
  48. Zare, K.; Sadegh, H.; Shahryari-Ghoshekandi, R.; Maazinejad, B.; Ali, V.; Tyagi, I.; Agarwal, S.; Kumar Gupta, V. Enhanced removal of toxic Congo red dye using multi walled carbon nanotubes: Kinetic, equilibrium studies and its comparison with other adsorbents. J. Mol. Liq. 2015, 212, 266–271. [Google Scholar] [CrossRef]
  49. Dinh, V.-P.; Huynh, T.-D.-T.; Le, H.M.; Nguyen, V.-D.; Dao, V.-A.N.; Hung, Q.; Tuyen, L.A.; Lee, S.; Yi, J.; Nguyen, T.D.; et al. Insight into the adsorption mechanisms of methylene blue and chromium (III) from aqueous solution onto pomelo fruit peel. RSC Adv. 2019, 9, 25847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Al-Kahlout, A.M.; El-Ghamri, H.S.; Al Dahoudi, N.; El-Agez, T.M.; Taya, S.A.; Abdel-Latif, M.S. A comparative study: Synthetic dyes as photosensitizers for dye-sensitized solar cells. Turk. J. Phys. 2015, 39, 272–279. [Google Scholar] [CrossRef]
  51. Nag, A.; Goswami, D. Solvent effect on two-photon absorption and fluorescence of rhodamine dyes. J. Photochem. Photobiol. A Chem. 2009, 206, 188–197. [Google Scholar] [CrossRef] [Green Version]
  52. Gopal, R.A.; Song, M.; Yang, D.; Lkhagvaa, T.; Chandrasekaran, S.; Choi, D. Synthesis of Hierarchically Structured γ-Fe2O3–PPy Nanocomposite as Effective Adsorbent for Cationic Dye Removal from Wastewater. Environ. Pollut. 2020, 267, 115498. [Google Scholar] [CrossRef]
  53. Revellame, E.D.; Fortela, D.L.; Sharp, W.; Hernandez, R.; Zappi, M.E. Adsorption kinetic modeling using pseudo-first order and pseudo-second order rate laws: A review. Clean. Eng. Technol. 2020, 1, 100032. [Google Scholar] [CrossRef]
  54. Weber, W.J.; Morris, J.C. Kinetics of adsorption on carbon from solutions. J. Sanit. Eng. Div. 1963, 89, 31–60. [Google Scholar] [CrossRef]
  55. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc. 1938, 2, 309–319. [Google Scholar] [CrossRef]
  56. Ebadi, A.; Mohammadzadeh, J.S.S.; Khudiev, A. What is the correct form of BET isotherm for modeling liquid phase adsorption? Adsorption 2009, 15, 65–73. [Google Scholar] [CrossRef]
  57. Freundlich, H. Über die Adsorption in Lösungen. Z. Phys. Chem. Stöchiometrie Verwandschaftslehre 1907, 57, 385–470. [Google Scholar] [CrossRef]
  58. Alvarez-Torrellas, S.; Boutahala, M.; Boukhalfa, N.; Munoz, M. Effective Adsorption of Methylene Blue dye onto Magnetic Nanocomposites. Modeling and Reuse Studies. Appl. Sci. 2019, 9, 4563. [Google Scholar] [CrossRef] [Green Version]
  59. Aksu, Z.; Tezer, S. Equilibrium and Kinetic Modelling of Biosorption of Remazol Black B by Rhizopus arrhizus in a Batch System: Effect of Temperature. Process Biochem. 2000, 36, 431–439. [Google Scholar] [CrossRef]
  60. Harja, M.; Buema, G.; Bucur, D. Recent advances in removal of Congo Red dye by adsorption using an industrial waste. Sci. Rep. 2022, 12, 6087. [Google Scholar] [CrossRef]
  61. Ghereghlou, M.; Esmaeili, A.A.; Darroudi, M. Adsorptive Removal of Methylene Blue from Aqueous Solutions Using Magnetic Fe3O4@C-dots: Removal and kinetic studies. Sep. Sci. Technol. 2022, 57, 2005–2023. [Google Scholar] [CrossRef]
  62. Peng, Z.; Fan, Z.; Chen, X.; Zhou, X.; Gao, Z.F.; Deng, S.; Wan, S.; Lv, X.; Shi, Y.; Han, W. Fabrication of Nano Iron Oxide–Modified Biochar from Co-Hydrothermal Carbonization of Microalgae and Fe(II) Salt for Efficient Removal of Rhodamine B. Nanomaterials 2022, 12, 2271. [Google Scholar] [CrossRef]
  63. Kazantsev, S.O.; Glazkova, E.A.; Lozhkomoev, A.S.; Bakina, O.V.; Khorobraya, E.G. Effect of the morphology of γ-Al2O3 nanosized particles on their adsorption properties. Russ. J. Phys. Chem. A 2020, 94, 806–809. [Google Scholar] [CrossRef]
  64. Bartošová, A.; Blinová, L.; Sirotiak, M.; Michalíková, A. Usage of ftir-atr as non-destructive analysis of selected toxic dyes, Research Papers Faculty of Materials Science and Technology Slovak University of Technology. Sciendo 2017, 25, 103–111. [Google Scholar] [CrossRef] [Green Version]
  65. Kim, D.-H.; Rozhkova, E.A.; Ulasov, I.V.; Bader, S.D.; Rajh, T.; Lesniak, M.S.; Novosad, V. Biofunctionalized magnetic-vortex microdiscs for targeted cancer-cell destruction. Nat. Mater. 2010, 9, 165–171. [Google Scholar] [CrossRef]
Figure 1. The scheme of NPs’ synthesis.
Figure 1. The scheme of NPs’ synthesis.
Materials 16 00023 g001
Figure 2. The scheme of the functionalization of carbon shell by aptamers. EDS and NHS—buffer solution with Ca2+ and Mg2+, pH = 7.4, containing: 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (1 mg) N-hydroxysuccinimide (1 mg) for 30 min.
Figure 2. The scheme of the functionalization of carbon shell by aptamers. EDS and NHS—buffer solution with Ca2+ and Mg2+, pH = 7.4, containing: 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (1 mg) N-hydroxysuccinimide (1 mg) for 30 min.
Materials 16 00023 g002
Figure 3. TEM image of the initial Fe3O4 NPs (a) and XRD patterns for Fe3O4 and Fe3O4@C NPs (b).
Figure 3. TEM image of the initial Fe3O4 NPs (a) and XRD patterns for Fe3O4 and Fe3O4@C NPs (b).
Materials 16 00023 g003
Figure 4. FT-IR spectra of synthesized Fe3O4, Fe3O4@C NPs, and Fe3O4@C NPs after nitric acid treatment.
Figure 4. FT-IR spectra of synthesized Fe3O4, Fe3O4@C NPs, and Fe3O4@C NPs after nitric acid treatment.
Materials 16 00023 g004
Figure 5. Room temperature magnetization curves for Fe3O4 and Fe3O4@C NPs. The remnant magnetization Mr = 12.7 emu/g and 9.7 emu/g for Fe3O4 and Fe3O4@C NPs, respectively.
Figure 5. Room temperature magnetization curves for Fe3O4 and Fe3O4@C NPs. The remnant magnetization Mr = 12.7 emu/g and 9.7 emu/g for Fe3O4 and Fe3O4@C NPs, respectively.
Materials 16 00023 g005
Figure 6. Optical absorption spectra of dye solutions (CR, EoY, RhC, and MB) with indication of the wavelength of the maxima that were used as reference to determine the concentration of dyes in solutions.
Figure 6. Optical absorption spectra of dye solutions (CR, EoY, RhC, and MB) with indication of the wavelength of the maxima that were used as reference to determine the concentration of dyes in solutions.
Materials 16 00023 g006
Figure 7. The qt dependence on the Fe3O4 contact time with CR and EoY (a), MB, and RhC (b). Black solid lines are nonlinear fitting of the results to the pseudo-first-order (PFO) model; dotted lines are the results nonlinear fitting to the pseudo-second-order (PSO) model. Experimental conditions: 25 °C, C0 = 60 mg/L for CR and C0 = 30 mg/L for other dyes, m (NPs) ≈ 3 mg in V = 1.5 mL.
Figure 7. The qt dependence on the Fe3O4 contact time with CR and EoY (a), MB, and RhC (b). Black solid lines are nonlinear fitting of the results to the pseudo-first-order (PFO) model; dotted lines are the results nonlinear fitting to the pseudo-second-order (PSO) model. Experimental conditions: 25 °C, C0 = 60 mg/L for CR and C0 = 30 mg/L for other dyes, m (NPs) ≈ 3 mg in V = 1.5 mL.
Materials 16 00023 g007
Figure 8. Representation of kinetic adsorption data of Fe3O4 NPs in coordinates qt vs. t1/2 for anionic (a) and cationic (b) dyes.
Figure 8. Representation of kinetic adsorption data of Fe3O4 NPs in coordinates qt vs. t1/2 for anionic (a) and cationic (b) dyes.
Materials 16 00023 g008
Figure 9. Adsorption isotherms of CR on Fe3O4 NPs at 25 °C in depending on initial C0 (a) and residual Ce (b) dye concentration. The right axis in (a) shows the percentage of dye removal. Experimental adsorption isotherms of EoY (c) and RhC (d) are shown by circles and squares, correspondingly. Adsorption isotherms according to the BET and Langmuir models are shown in panels (bd) by solid and dashed curves, correspondingly.
Figure 9. Adsorption isotherms of CR on Fe3O4 NPs at 25 °C in depending on initial C0 (a) and residual Ce (b) dye concentration. The right axis in (a) shows the percentage of dye removal. Experimental adsorption isotherms of EoY (c) and RhC (d) are shown by circles and squares, correspondingly. Adsorption isotherms according to the BET and Langmuir models are shown in panels (bd) by solid and dashed curves, correspondingly.
Materials 16 00023 g009
Figure 10. (a) qt dependence on the Fe3O4@C contact time with MB and RhC; solid line is fitting to the pseudo-first-order (PFO) and dotted line to pseudo-second-order (PSO) models. (b) Presentation of kinetic data in coordinates qt from t1/2.
Figure 10. (a) qt dependence on the Fe3O4@C contact time with MB and RhC; solid line is fitting to the pseudo-first-order (PFO) and dotted line to pseudo-second-order (PSO) models. (b) Presentation of kinetic data in coordinates qt from t1/2.
Materials 16 00023 g010
Figure 11. Adsorption isotherms of MB and RhC on Fe3O4@C NPs at at 25 °C—(a,b), correspondingly. Solid lines correspond to fitting to Freundlich (a) and BET (b) models, dotted lines to Langmuir model.
Figure 11. Adsorption isotherms of MB and RhC on Fe3O4@C NPs at at 25 °C—(a,b), correspondingly. Solid lines correspond to fitting to Freundlich (a) and BET (b) models, dotted lines to Langmuir model.
Materials 16 00023 g011
Figure 12. FT-IR spectra of Fe3O4 NPs before and after adsorption of CR. The black arrows indicate the position of the CR features, which correspond to data of Ref. [64]. The red lines indicate the positions of new lines in the spectrum of the Fe3O4@C nanoparticle after dye adsorption, correlating with data of Ref. [19].
Figure 12. FT-IR spectra of Fe3O4 NPs before and after adsorption of CR. The black arrows indicate the position of the CR features, which correspond to data of Ref. [64]. The red lines indicate the positions of new lines in the spectrum of the Fe3O4@C nanoparticle after dye adsorption, correlating with data of Ref. [19].
Materials 16 00023 g012
Figure 13. (a) Possible mechanism of electrostatic interaction of the MB molecule with the carbon shell, (b) mechanism of physical adsorption, and (c) molecule of the CR.
Figure 13. (a) Possible mechanism of electrostatic interaction of the MB molecule with the carbon shell, (b) mechanism of physical adsorption, and (c) molecule of the CR.
Materials 16 00023 g013
Figure 14. Reusability of Fe3O4 and Fe3O4@C NPs for adsorption Congo Red (CR) and Methylene Blue (MB), correspondingly.
Figure 14. Reusability of Fe3O4 and Fe3O4@C NPs for adsorption Congo Red (CR) and Methylene Blue (MB), correspondingly.
Materials 16 00023 g014
Figure 15. A typical picture in a microscope chamber; yellow objects with a blue border are living cells, blue spots are destroyed cells (a). Data of statistical processing of living and destroyed cells on five independent probes of each sample (b). (1) control cells (natural mortality), (2) a mixture of cells and Fe3O4@C NPs functionalized with aptamers without magnetic field exposure; (3) a mixture of cells and Fe3O4@C NPs with functionalized aptamers after action of a low-frequency magnetic field; and (4) a mixture of cells with initial NPs after action of a low-frequency magnetic field.
Figure 15. A typical picture in a microscope chamber; yellow objects with a blue border are living cells, blue spots are destroyed cells (a). Data of statistical processing of living and destroyed cells on five independent probes of each sample (b). (1) control cells (natural mortality), (2) a mixture of cells and Fe3O4@C NPs functionalized with aptamers without magnetic field exposure; (3) a mixture of cells and Fe3O4@C NPs with functionalized aptamers after action of a low-frequency magnetic field; and (4) a mixture of cells with initial NPs after action of a low-frequency magnetic field.
Materials 16 00023 g015
Table 1. Parameters of the best nonlinear fit to the pseudo-first- and pseudo-second-order kinetic models of adsorption data for Fe3O4 NPs, for experimental conditions of 25 °C, C0 = 60 mg/L for CR and C0 = 30 mg/L for other dyes, and m (NPs) ≈ 3 mg in V = 1.5 mL.
Table 1. Parameters of the best nonlinear fit to the pseudo-first- and pseudo-second-order kinetic models of adsorption data for Fe3O4 NPs, for experimental conditions of 25 °C, C0 = 60 mg/L for CR and C0 = 30 mg/L for other dyes, and m (NPs) ≈ 3 mg in V = 1.5 mL.
Parameters of the Nonlinear FitExperiment
DyesKinetic Modelqe (mg/g)k1 (1/min)
k2 (g/(mg*min)
R2qe
(mg/g)
CRPFO21.249 ± 0.3190.031 ± 0.00180.98921.6
PSO25.960 ± 0.700.00133 ± 0.000160.988
EoYPFO8.28 ± 0.1210.036 ± 0.00250.9878.3
PSO9.265 ± 0.2240.0054 ± 0.00080.984
MBPFO7.2 ± 0.850.002 ± 0.00030.9924.5 *
PSO12.58 ± 1.830.00009 ± 0.0000320.992
RhCPFO9.38 ± 0.2060.023 ± 0.00270.88910.2
PSO10.53 ± 0.2220.003 ± 0.00040.958
* The experimental value of qe for MB is significantly less than the value given by fitting to pseudo-models due to the fact that for the concentration and time of the experiment used, we did not reach the equilibrium value (see Figure 7b (data for MB)).
Table 2. Calculated for CR parameters of isotherm for different adsorption models.
Table 2. Calculated for CR parameters of isotherm for different adsorption models.
CR DyeBET-Case 3Langmuir
KL (L/mg)0.007 ± 0.0005
KS (L/mg)0.831 ± 0.00.7608 ± 1.136
qm (mg/g)23.73 ± 1.5627.49 ± 20.76
R20.96950.3188
Table 3. Parameters of the best nonlinear fitting of the experimental adsorption data for Fe3O4@C NPs to the pseudo-first- and pseudo-second-order kinetic models.
Table 3. Parameters of the best nonlinear fitting of the experimental adsorption data for Fe3O4@C NPs to the pseudo-first- and pseudo-second-order kinetic models.
Parameters of the Nonlinear FitExperiment
DyesKinetic Modelqe
(mg/g)
k1 (1/min)
k2 (g/(mg*min)
R2qe
(mg/g)
MBPFO10.59 ± 0.1590.013 ± 0.00060.99311
PSO13.33 ± 0.4870.0009 ± 0.000140.984
RhCPFO11.12 ± 0.1660.044 ± 0.00450.95012
PSO12.09 ± 0.1370.0059 ± 0.00050.987
Table 4. Comparison of several dyes’ adsorption capacities reported in the literature and obtained in the present work.
Table 4. Comparison of several dyes’ adsorption capacities reported in the literature and obtained in the present work.
AdsorbentDyesqe [mg/g]Refs.
Fe3O4 NPsCongo Red96.46[59]
Fly ash22.12[60]
Fe3O4 NPs58Present work
Fe3O4@C
NPs (30 nm with 2 nm carbon shell)
Methylene Blue18.52[19]
Fe3O4@C magnetic materials
(particle size, 1~100 μm)
270.51[26]
Fe3O4@C-dots124.9[61]
B-Fe3O4@C (~3 μm)42.11[21]
Fe3O4@C NPs15Present work
NiZnAl-LDH, ZnAl-LDH, NiAl-LDHRhB52–97[4]
Nano iron oxide–modified biochar286.4[62]
Fe3O4 NPsRhC14Present work
Fe3O4@C NPs35Present work
Nanoplates γ-Al2O3EoY6[63]
Fe3O4 NPs22Present work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ivanova, O.S.; Edelman, I.S.; Lin, C.-R.; Svetlitsky, E.S.; Sokolov, A.E.; Lukyanenko, K.A.; Sukhachev, A.L.; Shestakov, N.P.; Chen, Y.-Z.; Spivakov, A.A. Core–Shell Fe3O4@C Nanoparticles for the Organic Dye Adsorption and Targeted Magneto-Mechanical Destruction of Ehrlich Ascites Carcinoma Cells. Materials 2023, 16, 23. https://doi.org/10.3390/ma16010023

AMA Style

Ivanova OS, Edelman IS, Lin C-R, Svetlitsky ES, Sokolov AE, Lukyanenko KA, Sukhachev AL, Shestakov NP, Chen Y-Z, Spivakov AA. Core–Shell Fe3O4@C Nanoparticles for the Organic Dye Adsorption and Targeted Magneto-Mechanical Destruction of Ehrlich Ascites Carcinoma Cells. Materials. 2023; 16(1):23. https://doi.org/10.3390/ma16010023

Chicago/Turabian Style

Ivanova, Oxana S., Irina S. Edelman, Chun-Rong Lin, Evgeniy S. Svetlitsky, Alexey E. Sokolov, Kirill A. Lukyanenko, Alexander L. Sukhachev, Nikolay P. Shestakov, Ying-Zhen Chen, and Aleksandr A. Spivakov. 2023. "Core–Shell Fe3O4@C Nanoparticles for the Organic Dye Adsorption and Targeted Magneto-Mechanical Destruction of Ehrlich Ascites Carcinoma Cells" Materials 16, no. 1: 23. https://doi.org/10.3390/ma16010023

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