Next Article in Journal
Fabrication and Characterization of Waste Wood Cellulose Fiber/Graphene Nanoplatelet Carbon Papers for Application as Electromagnetic Interference Shielding Materials
Next Article in Special Issue
Preparation of Dual-Layered Core–Shell Fe3O4@SiO2 Nanoparticles and Their Properties of Plasmid DNA Purification
Previous Article in Journal
Dual-Spectral Plasmon-Induced Transparent Terahertz Metamaterial with Independently Tunable Amplitude and Frequency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Multilayered Core–Shell Fe3O4-SnO2-C Nanoparticles via Polymeric/Silane–Amino Functionalization

1
Division of Materials Science and Engineering, Hanyang University, 222 Wangsimni-ro, Seongdong-gu, Seoul 04763, Korea
2
Department of Advanced Materials Engineering, Kyonggi University, 154-42 Gwanggyosan-ro, Yeongtong-gu, Suwon-si 16227, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(11), 2877; https://doi.org/10.3390/nano11112877
Submission received: 23 September 2021 / Revised: 25 October 2021 / Accepted: 26 October 2021 / Published: 28 October 2021
(This article belongs to the Special Issue Metal Oxide Core-Shell Nanoparticles)

Abstract

:
Multilayered core–shell Fe3O4-SnO2-C nanoparticles were prepared via surface treatment and carbonization at atmospheric pressure. Fe3O4-SnO2 nanoparticles were prepared by the carboxylation of the pivotal particles (Fe3O4) with an anionic surfactant to immobilize SnO2 nanoparticles. A method was proposed to externally surround hydrophilic carbon with amine-forming materials, polyethyleneimine (PEI), and (3-Aminopropyl) triethoxysilane (APTES). The synthesis strategy was based on the electrostatic bonding of the introduced amine group with the hydroxyl group on the carbon precursor and the carbonization of the coating layer by the catalytic reaction of sulfuric acid.

1. Introduction

Owing to their unique electrochemical and magnetic properties, magnetite (Fe3O4) nanoparticles have gained significant attention for application in various fields, including biomedical fields [1], catalysis [2], resistive switching memory [3], energy storage [4], and electromagnetic interference (EMI) shielding [5]. However, they suffer from low chemical stability and agglomeration owing to their relatively high surface energy [6,7]. Therefore, to overcome these limitations of Fe3O4, complementary and related maintenance functions are implemented through the formation of composites with various functional groups [8,9,10] and organic/inorganic materials [7,11,12,13,14,15].
Among the various materials that are used for forming composites with Fe3O4, tin dioxide (SnO2), an n-type semiconductor with a wide bandgap (Eg = 3.6 eV at 300 K), has been extensively investigated [16]. Owing to its unique characteristics, SnO2 can stabilize the electronic, thermal, and chemical properties of Fe3O4 through the proximity effect and equilibration of potentials [17,18,19], and various efforts have been made to realize the electrochemical applications of the composites of Fe3O4 and SnO2. However, Fe3O4-SnO2 composite materials show poor electrical conductivity, which limits their applications. Therefore, it is necessary to combine these composites with high-conductivity materials, such as precious metals and carbon materials, to enhance their conductivity [20,21].
In particular, numerous studies have been carried out on the preparation of composites of Fe3O4-SnO2 particles with carbon-based materials to improve their conductivity and performance for application in various fields. Wang et al. [22] prepared a composite obtained by combining core–shell Fe3O4@SnO2 with reduced graphene oxide (rGO), a carbon-based oxide, for application in EMI shielding. Iron oxide with high magnetic permeability and low complex permittivity improved the dielectric properties of the composite, thus improving its EM wave absorption performance. Furthermore, the combination of rGO (with high electrical conductivity) with Fe3O4@SnO2 resulted in electric dipolar polarization and interfacial polarization. Hence, the composite showed an enhanced electromagnetic wave absorption effect in the high-frequency region. Chen et al. [23] fabricated Fe3O4/SnO2/C composite particles for application as an anode for lithium-ion batteries. The complexation of Fe3O4 and SnO2 improved the reversibility of the LiO2 conversion reaction, thereby increasing the specific capacity and charge/discharge efficiency of the anode. In addition, the surface modification of the composite with carbon improved the conductivity of the active material and its affinity with the electrolyte. Thus, the Fe3O4/SnO2/C composite showed a more stable rate performance than the Fe3O4/SnO2 particles during cycling. Therefore, the preparation of Fe3O4-SnO2-C composites has become a topic of great scientific interest.
The fabrication of SnO2-decorated Fe3O4 composites through facile surface modification has been reported [24]. Fe3O4 nanoparticles were carboxylated using polyacrylic acid (PAA) or tri-sodium citrate dihydrate (tSCD), and the surface modification method using tSCD was more effective than that using PAA. The results demonstrated the feasibility of the preparation of spherical nanoparticles surrounded by ultrafine SnO2 nanoparticles. Motivated by these results, in this study, we prepared Fe3O4-SnO2-C composite particles with a three-layered core–shell structure through the dehydration of glucose under atmospheric pressure (Figure 1). Fe3O4-SnO2 nanoparticles were amino-functionalized using a polymeric precursor, PEI, and a silane precursor, APTES. Subsequently, the effects of the functionalizing amine groups on the structure of the multilayered particles were investigated. In addition, the mechanism underlying the formation of the core–shell structure by the reaction between the polymerized layer formed by combining glucose, a monosaccharide-based carbon precursor, and the aminated nanoparticles and sulfuric acid was investigated.

2. Materials and Methods

2.1. Materials

Ethylene glycol (EG, >99.5%, Samchun Pure Chemical, Pyeongtaek-si, Korea), ferric chloride hexahydrate (FeCl3·6H2O, >97%, Sigma Aldrich, USA), sodium acetate (NaOAc, >99.5%, Sigma Aldrich, St. Louis, MO, USA), ethanol (99.9%, Daejung Chemical, Siheung-si, Korea), tri-sodium citrate dihydrate (tSCD, HOC(COONa)(CH2COONa)2·2H2O, ≥99.0%, Sigma Aldrich, St. Louis, MO, USA), sodium hexahydroxostannate (Na2SnO3·3H2O, 95%, Junsei, Tokyo, Japan), (3-Aminopropyl)triethoxysilane (APTES, 98%, Sigma Aldrich, St. Louis, MO, USA), polyethyleneimine (PEI, branched, Mw ~25,000, Sigma Aldrich, St. Louis, MO, USA), D-(+)-glucose (>99.5%, Sigma Aldrich, St. Louis, MO, USA) and sulfuric acid (H2SO4, >98%, Daejung Chemical, Siheung-si, Korea) were used without further treatment.

2.2. Preparation of Core–Shell Structured Fe3O4-SnO2 Nanoparticles

Fe3O4-SnO2 nanoparticles were prepared as described in our previous report [24]. To prepare the Fe3O4 nanoparticles, 45 g of FeCl3·6H2O was added to distilled water under stirring at room temperature, and 0.5 M NaOAc was dissolved in 1000 mL of EG. These two solutions were then transferred to a three-necked round-bottom flask, and the resulting solution was heated to boiling with mechanical stirring and maintained at this temperature for 18 h. After the completion of the reaction, the solution was naturally cooled to room temperature, and the reactants were separated from the solution using a magnet. Subsequently, the reactants were rinsed with distilled water and ethanol several times to eliminate the organic and inorganic by-products.
The prepared Fe3O4 nanoparticles (1 g) were dispersed in 100 mL of distilled water. The resulting suspension was then transferred to a three-neck round-bottom flask, followed by the addition of 0.5 M tSCD. The mixed solution was mechanically stirred at 300 rpm for 24 h at room temperature. After the completion of the reaction, the surface-modified particles were separated from the solution and subsequently rinsed several times with distilled water.
The surface-modified Fe3O4 nanoparticles (0.5 g) were ultrasonically redispersed in 300 mL of distilled water for 30 min. Subsequently, 1 g of Na2SnO3·3H2O was added to the suspension under stirring at 300 rpm for 24 h at 80 °C. The reaction-terminated suspension was naturally cooled to room temperature, and the particles were separated using a magnet. The resultant particles were washed several times with distilled water.

2.3. Amino Functionalization of the Core–Shell Fe3O4-SnO2 Nanoparticles

The surface amino-functionalization process was carried out using two different methods. In the case of the inorganic-precursor-based amino functionalization, the Fe3O4-SnO2 nanoparticles (0.5 g) were ultrasonically dispersed for 15 min in a 100 mL solution of distilled water and ethanol (1:1 vol%). The dispersed Fe3O4-SnO2 suspension was then transferred to a three-neck round-bottom flask, followed by the injection of 0.09 M of APTES. The suspension was mechanically stirred at 350 rpm and 70 °C for 48 h. After the reaction, the solution was naturally cooled to room temperature and the reactants were separated from the solution using a magnet. The resultant particles were washed several times with ethanol and distilled water.
In the case of the polymeric-precursor-based amino functionalization, the Fe3O4-SnO2 nanoparticles (0.5 g) were ultrasonically redispersed for 20 min in 200 mL of distilled water with 5 wt% of PEI. Subsequently, the dispersed suspension was mechanically stirred at 300 rpm in a round-bottom flask and heat-treated at 80 °C for 18 h. The solution was cooled to room temperature, and the resulting particles were washed with distilled water several times to eliminate the by-products.

2.4. Carbonization of the Core–Shell Fe3O4-SnO2 Nanoparticles

The amino-functionalized Fe3O4–SnO2 nanoparticles (0.5 g) were redispersed in 400 mL of EG for 30 min via ultrasonication. The mixture solution was transferred to a three-neck round-bottom flask equipped with a reflux condenser, followed by the addition of 30 g of D-(+)-glucose. Subsequently, the solution was mechanically stirred at 300 rpm for 20 h at 180 °C under reflux conditions. Subsequently, the heated solution was slowly cooled to room temperature and immediately reheated to 60 °C. Then, 2 mL of 0.1 M H2SO4 was added to the solution, and the resulting solution was heated to 180 °C and maintained for 24 h. After the reaction, the solution was naturally cooled to room temperature, and the particles were separated from the solution using a magnet. The synthesized particles were washed several times with ethanol and distilled water.

2.5. Characterization

The surface characteristics of the surface-modified particles were investigated using Fourier transform infrared (FTIR) spectroscopy (Nicolet 5700, Thermo Electron, Waltham, MA, USA). The surface charges and dispersion properties of the particles were evaluated by carrying out zeta potential and particle size distribution analyses (Zetasizer Nano ZS, Malvern, UK). The crystal structures of the synthesized nanoparticles were analyzed using X-ray diffraction (XRD, UltimaIV, Rigaku, Japan) with Cu Kα radiation (λ = 1.5418 Å). The morphologies of the nanoparticles were investigated using high-resolution transmission electron microscopy (HRTEM, Tecnai G2 F30 S-Twin, FEI, Hillsboro, OR, USA). The degree of graphitization of the carbon layer on the particle surface was investigated using Raman spectroscopy (NRS-3100, Jasco, Easton, PN, USA), with an excitation wavelength of 532 nm. Moreover, the magnetic properties of the particles were evaluated using a vibrating sample magnetometer (Lake Shore 7400, Cryotronics Inc., Westerville, OH, USA) at the applied field of −10–10 kOe at room temperature.

3. Results and Discussion

To identify the surface functional groups of the as-prepared nanoparticles, their FTIR spectra were analyzed, as shown in Figure 2. The Fe3O4 nanoparticles exhibited a peak at 582 cm−1, corresponding to the stretching of the Fe-O bond in the tetrahedral sites [8]. In contrast, the SnO2 nanoparticles synthesized under the same experimental conditions exhibited peaks corresponding to the stretching of the Sn–OH, O–Sn–O, and Sn-O bonds at 550, 618, and 941 cm−1, respectively [25]. The peaks at 1639 and 3415 cm−1 can be ascribed to the –OH stretching vibration of the particle surface of the hydroxyl group, which was generated by the absorption of H2O from the ambient atmosphere [25]. The tSCD-treated Fe3O4 nanoparticles exhibited –CH2 and –CH3 vibration peaks at 2857 and 2920 cm−1, respectively. In addition, peaks corresponding to the vibration of the C–O and COO–Fe bonds were observed at 1393 and 1619 cm−1, respectively [26]. This confirms the presence of carboxyl groups on the particle surface.
The Fe3O4-SnO2 nanoparticles exhibited peaks at 582 and 550 cm−1, corresponding to the Fe–O and O–Sn–O bonds, respectively. Additionally, the peak at 941 cm−1 can be ascribed to the stretching vibration of the Sn–O bond, which originated from the SnO2 formed on the Fe3O4 surface. In the case of the APTES-treated Fe3O4-SnO2 nanoparticles, an absorption peak was observed at approximately 3300 cm−1, corresponding to the symmetric and asymmetric stretching modes of NH/NH2 [27,28]. The peak at 1048 cm−1 can be ascribed to the Si–O–Si bond, which originated from the silane group of APTES [27]. On the other hand, the PEI-treated Fe3O4-SnO2 nanoparticles exhibited –CH2 and –CH3 vibration peaks at 2843 and 2919 cm−1, respectively [29]. In addition, symmetric and asymmetric NH/NH2 stretching modes were observed at 1639 and approximately 3300 cm−1, respectively [30]. Therefore, it can be stated that various organic functional groups were present on the surface of the PEI- and APTES-modified Fe3O4-SnO2 nanoparticles with different structures.
The zeta potentials of the nanoparticles were measured to examine the changes in their surface properties after their surface modification under distilled water conditions (Table 1). In the table, each sample displacement is the result of 10 measurements under the same conditions. The zeta potential value of the Fe3O4 nanoparticles was −5.4 mV, which indicates that these particles showed a small amount of surface charge. The Fe3O4-SnO2 nanoparticles showed a zeta potential of −42.3 mV. In contrast, the zeta potentials of both the APTES- and PEI-treated particles were positive (31.8 and 32.9, respectively) because of the formation of amine groups on their surfaces. Although these modified nanoparticles showed similar zeta potential values, the average value and standard deviation of the PEI-treated particles were slightly larger. Therefore, the initial glucose conjugation amount could be set identically by minimizing the effect of electrostatic attraction that might have been caused by the surface charge on each particle during the carbon coating process.
Figure 3 shows the particle size distributions of the as-prepared Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles in distilled water. The distribution curve of the Fe3O4 nanoparticles showed a bimodal shape with a wide size distribution range of 200–2650 nm and a mean size of 821 nm. This can be attributed to the presence of particles aggregated because of the high surface energy generated at the nanoscale and the low surface charge of the Fe3O4 nanoparticles [31]. In contrast, the surface-decorated/functionalized nanoparticles exhibited monomodal curves. The measured mean size distribution values of the Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, PEI-treated Fe3O4-SnO2 nanoparticles were 476, 551, and 578 nm, respectively. In the case of the Fe3O4-SnO2 nanoparticles, the hydroxyl groups present on the SnO2 surface generated a high surface charge, improving the dispersibility of the particles in the solvent. On the other hand, the APTES- and PEI-treated Fe3O4-SnO2 nanoparticles showed relatively low surface charge, and thus showed slightly lower dispersibility than the unfunctionalized Fe3O4-SnO2 nanoparticles.
The crystal structures of the Fe3O4, Fe3O4-SnO2, PEI-treated Fe3O4-SnO2, and APTES-treated Fe3O4-SnO2 nanoparticles were investigated using XRD (Figure 4). The Fe3O4 nanoparticles exhibited significant diffraction peaks at 2θ = 30.2, 35.5, 43.2, 53.5, 57.1, 62.7, and 74.3°, corresponding to the (220), (311), (400), (422), (511), (440), and (533) planes of the inverse-spinel structure (JCPDS No. 19-0629) [32]. The Fe3O4-SnO2 nanoparticles exhibited peaks corresponding to Fe3O4 in addition to those at 2θ = 26.1, 33.5, 37.3, 51.7, 65.4, and 71.7° corresponding to the (110), (101), (200), (211), (220), and (202) planes of the tetragonal rutile structure (JCPDS card no. 41-1445), respectively [33]. The diffraction peaks corresponding to SnO2 were broad because of its crystalline size of less than 5 nm [34]. This was further confirmed by calculating the average crystallite sizes of the prepared nanoparticles using the Scherrer’s formula, D = Kλ/βcosθ, where D is the average crystallite size, K is the shape factor (K = approximately 0.94 for spherical crystallites), λ is the X-ray wavelength (λ = 1.5418 Å for Cu Kα radiation), β is the full width at half maximum of the high-intensity diffraction peak (in radians), and θ is the Bragg’s angle (in radians). The average crystallite sizes of Fe3O4 and SnO2 were calculated to be 18.09 and 4.36 nm, respectively. Meanwhile, the XRD patterns of the APTES and PEI-treated Fe3O4-SnO2 nanoparticles were the same as that of the Fe3O4-SnO2 nanoparticles. This indicates that the Fe3O4-SnO2 nanoparticles maintained their crystallinity even after the amino-functionalization treatment.
Figure 5 shows the TEM and HR-TEM images showing the morphologies and microstructures of the synthesized particles before carbon coating. As shown in the figure, all the synthesized particles were spherical, and the Fe3O4 particles, which acted as the core, had a diameter of approximately 300 nm (Figure 5a). The magnified TEM (Figure 5(b-1)) and HRTEM (Figure 5(b-2)) images revealed that SnO2 particles with a diameter of approximately 4.5 nm were formed on the surface of the Fe3O4 particles to a thickness of approximately 20 nm. This is almost consistent with the average crystallite sizes of the Fe3O4 and SnO2 nanoparticles, as calculated from the XRD data (Figure 4) according to Scherrer’s formula. In addition, lattice patterns with the interplanar spacings of 0.268, 0.334, and 0.233 nm corresponding to the (101), (110), and (200) planes, respectively, were observed on the particle surface. In Figure 5(c-1), the outer layer of the PEI-treated Fe3O4-SnO2 nanoparticles can be clearly distinguished from that of the untreated nanoparticles. This outer layer was confirmed to be grafted onto the particle surface with a thickness of approximately 6.5 nm because of the polymerization of the polymeric precursor. In contrast, no significant difference was observed in the images of the Fe3O4-SnO2 nanoparticles before and after the APTES treatment (Figure 5(d-1)). Figure 5(c-2,d-2) show the HRTEM images of the PEI- and APTES-treated Fe3O4-SnO2 nanoparticles. Both the amine-treated nanoparticles showed lattice patterns with the interplanar spacings of 0.334 and 0.233 nm corresponding to the (110) and (200) planes of SnO2, respectively. This is consistent with the XRD results and confirms the successful amino functionalization of the Fe3O4-SnO2 nanoparticles through the polymerization of PEI or the silane bonding of APTES. However, the surface microstructures of the amino-functionalized nanoparticles prepared using these functionalization methods were significantly different.
Figure 6 shows the carbon coating procedure of the amino-functionalized Fe3O4-SnO2 nanoparticles using glucose. EG, which was used as the solvent, played an important role in forming a stable dispersion of the reaction medium and functionalized magnetic nanoparticles. First, owing to its negative charge, the hydroxyl group of glucose bonded strongly to the amine group (with a positive charge) on the surface of the Fe3O4-SnO2 particles by electrostatic attraction. Glucose transformed into oligosaccharide chains by the condensation reaction and surrounded the particle surface. In addition, the rotational energy generated by stirring facilitated the deposition of a uniform layer of the oligosaccharide chains on the surface of the particles. Then, the added sulfuric acid dehydrated the oligosaccharide chains bonded to the surface, and a hydrophilic carbon layer with a hydroxyl group at the end was finally formed.
The TEM images of the amino-functionalized nanoparticles subjected to the carbon coating process are shown in Figure 7. In the case of the PEI-treated nanoparticles, only naked nanoparticles with the polymer layer removed were observed after the carbon coating process (Figure 7a). The polymer layer disappeared because of dissolution by high temperature or sulfuric acid during the reaction. Since PEI, as a binding polymeric material, has a melting point of approximately 75 °C, it is sensitive to heat and acidic environments [35]. However, it seems that the SnO2 particles formed on the surface of the Fe3O4 particles by electrostatic attraction were not removed by sulfuric acid (Figure 7b). To analyze the structure of the nanoparticles in detail, their selected area electron diffraction (SAED) patterns were recorded, which confirmed the presence of the (101), (110), and (200) planes of SnO2 in the composite nanoparticles (Figure 7c) [36].
In contrast, the particles treated with APTES formed a new coating layer on the surface with a thickness of approximately 17 nm (Figure 7d,e). In addition, the lattice pattern corresponding to the (110) plane was observed with an interplanar spacing of 0.334 nm in the HRTEM image. Ring patterns corresponding to the (333) and (311) planes of Fe3O4 and the (101) plane of SnO2 were observed, as shown in Figure 7f [29]. This can be attributed to the carbonization of the nanoparticles after their electrostatic bonding with glucose because the amino-functional group formed at the end of the particles by the silane bond of APTES was resistant to high temperatures and acidic environments. Therefore, it can be stated that the formation of the carbon layer was directly related to the thermal/chemical stability of the amine retention layer, and the surface treatment using APTES was more effective in carbon layer formation than that using the polymeric precursor, PEI. Finally, multilayered core–shell Fe3O4-SnO2-C nanoparticles were synthesized by the amino functionalization of the Fe3O4-SnO2 particles through APTES.
Figure 8a shows the XRD pattern of the Fe3O4-SnO2-C composite nanoparticles. The nanoparticles showed peaks of Fe3O4 and SnO2 corresponding to the (110), (101), (200), (211), (220), (202), (220), (311), (400), (422), (511), and (440) planes. In addition, the low-intensity peak at 2θ = 22.1° corresponds to the (002) plane of carbon, which originated from the carbon layer formed by the carbonization of glucose [37].
As shown in Figure 8b, the Raman spectrum of the Fe3O4-SnO2-C composite nanoparticles confirmed the formation of carbon on their surface. Two strong characteristic carbon peaks were observed over the wavenumber range of 1000–2800 cm−1. In general, the D band, which appears only when disordered or finite-sized graphite crystals are present, is not identified in well-crystallized graphite [38]. The D band at 1371 cm−1 indicates the presence of sp3 defects in the carbon network, whereas the G band at 1584 cm−1 corresponds to the E2g stretching vibration of graphite with sp2 electronic configuration.
The intensity ratio of the D- and G-bands (ID/IG) provides useful information about the crystallinity of carbon material. The closer the ID/IG value to zero, the higher is the degree of structural alignment of the formed carbon material [39]. The ID/IG ratio of the Fe3O4-SnO2-C composite nanoparticles was approximately 0.5706, as determined using Lorentz fitting. This indicates that the formed carbon layer had a fairly ordered structure. However, this carbon layer can be considered amorphous carbon because of its high ID value. Thus, the layer formed on the surface of the Fe3O4-SnO2 nanoparticles was confirmed to be amorphous graphitized carbon, which was formed when the sulfuric acid used in the coating process promoted the dehydration reaction of the oligosaccharide obtained from glucose. Thus, the TEM and Raman analysis results indicate that the Fe3O4-SnO2-C nanoparticles were synthesized by the formation of a structurally aligned carbon layer in the reflux process under atmospheric pressure.
Figure 9a shows the FTIR spectra of the Fe3O4–SnO2 and Fe3O4–SnO2–C nanocomposites. Unlike Fe3O4-SnO2, the Fe3O4-SnO2-C nanocomposites exhibited a –OH stretching vibration peak over a broad spectral region of 3630–3200 cm−1. The peaks at 2917 and 2871 cm−1 correspond to the vibration of the –CH2 and –CH3 groups, respectively. The strong peak at 1636 cm−1 can be ascribed to the C–C stretching vibration. In addition, the characteristic peaks at approximately 1387–1301 and 1042–1083 cm−1 correspond to the stretching vibration of the C–O and C–OH bonds, respectively, which were probably formed because of the hydroxyl group formed on the carbon layer [40]. As a result, the Fe3O4-SnO2-C nanoparticles exhibited high hydrophilicity because of the large number of hydroxyl groups present in the carbon coating layer.
Figure 9b shows the size distribution curves of the Fe3O4-SnO2-C nanoparticles. As can be observed, the estimated mean size of the particles was 669 nm, and their size distribution was in the range of 225–1700 nm. As compared with the Fe3O4-SnO2 nanoparticles (mean size of 476 nm with a zeta potential of −42.3 mV), the Fe3O4-SnO2-C nanoparticles (with a higher zeta potential of −51 mV) showed aggregation. This can be attributed to the carbonization of the Fe3O4-SnO2 nanoparticles in a bonded state between the oligosaccharide coating layers during the carbon coating process. Nevertheless, the Fe3O4-SnO2-C nanoparticles showed high dispersibility in aqueous solvents because of the hydroxyl groups present on the surface of the carbon layer.
As can be observed from the lower right inset of Figure 10, the coercivity (Hc) and residual magnetization (Mr) of the Fe3O4, Fe3O4-SnO2, and Fe3O4-SnO2-C nanoparticles were almost zero. This suggests that all the prepared nanoparticles were superparamagnetic. As the prepared Fe3O4 particles gradually evolved into the Fe3O4-SnO2-C particles, their saturation magnetization (MS) value tended to decrease. The measured MS values of the Fe3O4, Fe3O4-SnO2, and Fe3O4-SnO2-C nanoparticles were 121, 95, and 81 emu/g, respectively (Figure 10). This decrease in the MS of the Fe3O4-SnO2-C nanoparticles can be attributed to the increase in the weight of the single particles because of the formation of non-magnetic SnO2 and carbon in the Fe3O4 nanoparticles. As can be seen from the optical image inserted, the Fe3O4-SnO2-C nanoparticles reacted strongly to the external magnetic fields despite their relatively lower magnetization than that of the Fe3O4-SnO2 nanoparticles.

4. Conclusions

Multilayered core–shell-structured Fe3O4-SnO2-C nanoparticles were fabricated via surface treatment and carbonization at atmospheric pressure. The carboxylation of Fe3O4 nanoparticles using tSCD fixed SnO2 nanoparticles electrostatically on them, with a mean size of 4.5 nm while maintaining a layer thickness of 20 nm. Thereafter, the surface amino functionalization of the Fe3O4-SnO2 nanoparticles using PEI and APTES was carried out at the hydroxyl group at the end of the glucose molecule by electrostatic attraction. Unfortunately, when PEI was used, the delicate polymer layer was destroyed by high temperatures and sulfuric acid during the carbon coating process, and the final product was Fe3O4-SnO2. Therefore, the polymer-based amino functionalization was not suitable for the carbonization process. When APTES was used, the amine group formed on the surface of the nanoparticles by silane bonding facilitated the formation of a carbon layer even under severe coating conditions. The carbonized layer formed by the dehydration of glucose with sulfuric acid was hydrophilic amorphous graphitized carbon. The Fe3O4-SnO2-C nanoparticles showed an MS of approximately 81 emu/g and showed excellent separation performance, which is typical of superparamagnetic nanoparticles.

Author Contributions

Conceptualization, methodology, validation, formal analysis, investigation, data curation, writing—original draft preparation, and visualization, J.U.H. and G.S.A.; writing—review and editing, resources, and supervision, G.S.A. and S.-C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Korea Institute for Advancement of Technology (KIAT) grant, funded by the Korean Government (MOTIE) (P0017012, Human Resource Development Program for Industrial Innovation).

Data Availability Statement

All data used to support the findings of this study are included in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yew, Y.P.; Shameli, K.; Miyake, M.; Ahmad Khairudin, N.B.B.; Mohamad, S.E.B.; Naiki, T.; Lee, K.X. Green biosynthesis of superparamagnetic magnetite Fe3O4 nanoparticles and biomedical applications in targeted anticancer drug delivery system: A review. Arab. J. Chem. 2020, 13, 2287–2308. [Google Scholar] [CrossRef]
  2. Chen, X.; Zhang, Y.; Zhao, Y.; Wang, S.; Liu, L.; Xu, W.; Guo, Z.; Wang, S.; Liu, Y.; Zhang, J. Encapsulating Pt Nanoparticles through Transforming Fe3O4 into MIL-100(Fe) for Well-Defined Fe3O4@Pt@MIL-100(Fe) Core-Shell Heterostructures with Promoting Catalytic Activity. Inorg. Chem. 2019, 58, 12433–12440. [Google Scholar] [CrossRef] [PubMed]
  3. Ishibe, T.; Maeda, Y.; Terada, T.; Naruse, N.; Mera, Y.; Kobayashi, E.; Nakamura, Y. Resistive switching memory performance in oxide hetero-nanocrystals with well-controlled interfaces. Sci. Technol. Adv. Mater. 2020, 21, 195–204. [Google Scholar] [CrossRef] [Green Version]
  4. Xu, J.S.; Zhu, Y.J. Monodisperse Fe3O4 and γ-Fe2O3 magnetic mesoporous microspheres as anode materials for lithium-ion batteries. ACS Appl. Mater. Interfaces 2012, 4, 4752–4757. [Google Scholar] [CrossRef]
  5. Wu, N.; Liu, C.; Xu, D.; Liu, J.; Liu, W.; Shao, Q.; Guo, Z. Enhanced Electromagnetic Wave Absorption of Three-Dimensional Porous Fe3O4/C Composite Flowers. ACS Sustain. Chem. Eng. 2018, 6, 12471–12480. [Google Scholar] [CrossRef]
  6. Di Liu, X.; Chen, H.; Liu, S.S.; Ye, L.Q.; Li, Y.P. Hydrothermal synthesis of superparamagnetic Fe3O4 nanoparticles with ionic liquids as stabilizer. Mater. Res. Bull. 2015, 62, 217–221. [Google Scholar] [CrossRef]
  7. Liu, S.; Yu, B.; Wang, S.; Shen, Y.; Cong, H. Preparation, surface functionalization and application of Fe3O4 magnetic nanoparticles. Adv. Colloid Interface Sci. 2020, 281, 102165. [Google Scholar] [CrossRef]
  8. Singh, S.; Barick, K.C.; Bahadur, D. Surface engineered magnetic nanoparticles for removal of toxic metal ions and bacterial pathogens. J. Hazard. Mater. 2011, 192, 1539–1547. [Google Scholar] [CrossRef]
  9. Hur, J.U.; Choi, J.S.; Choi, S.C.; An, G.S. Highly dispersible Fe3O4 nanoparticles via anionic surface modification. J. Korean Ceram. Soc. 2020, 57, 80–84. [Google Scholar] [CrossRef]
  10. Zhao, Y.G.; Shen, H.Y.; Pan, S.D.; Hu, M.Q. Synthesis, characterization and properties of ethylenediamine-functionalized Fe3O4 magnetic polymers for removal of Cr(VI) in wastewater. J. Hazard. Mater. 2010, 182, 295–302. [Google Scholar] [CrossRef] [PubMed]
  11. 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]
  12. Ge, R.; Li, X.; Lin, M.; Wang, D.; Li, S.; Liu, S.; Tang, Q.; Liu, Y.; Jiang, J.; Liu, L.; et al. Fe3O4@polydopamine composite theranostic superparticles employing preassembled Fe3O4 nanoparticles as the core. ACS Appl. Mater. Interfaces 2016, 8, 22942–22952. [Google Scholar] [CrossRef] [PubMed]
  13. Cai, W.; Guo, M.; Weng, X.; Zhang, W.; Owens, G.; Chen, Z. Modified green synthesis of Fe3O4@SiO2 nanoparticles for pH responsive drug release. Mater. Sci. Eng. C 2020, 112, 110900. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, K.; Qin, Y.; Muhammad, Y.; Zhu, Y.; Tang, R.; Chen, N.; Shi, H.; Zhang, H.; Tong, Z.; Yu, B. Effect of Fe3O4 content and microwave reaction time on the properties of Fe3O4/ZnO magnetic nanoparticles. J. Alloys Compd. 2019, 781, 790–799. [Google Scholar] [CrossRef]
  15. Reddy, N.N.K.; Godavarthi, S.; Kumar, K.M.; Kummara, V.K.; Vattikuti, S.P.; Akkera, H.S.; Bitla, Y.; Jilani, S.A.K.; Manjunath, V. Evaluation of temperature dependent electrical transport parameters in Fe3O4/SiO2/n-Si metal–insulator-semiconductor (MIS) type Schottky barrier heterojunction in a wide temperature range. J. Mater. Sci. Mater. Electron. 2019, 30, 8955–8966. [Google Scholar]
  16. Chiodini, N.; Paleari, A.; Dimartino, D.; Spinolo, G. SnO2 nanocrystals in SiO2: A wide-band-gap quantum-dot system. Appl. Phys. Lett. 2002, 81, 1702–1704. [Google Scholar] [CrossRef]
  17. Leostean, C.; Pana, O.; Stefan, M.; Popa, A.; Toloman, D.; Senila, M.; Gutoiu, S.; Macavei, S. New properties of Fe3O4@SnO2 core shell nanoparticles following interface charge/spin transfer. Appl. Surf. Sci. 2018, 427, 192–201. [Google Scholar] [CrossRef]
  18. Mallikarjuna, K.; Bari, G.A.R.; Vattikuti, S.P.; Kim, H. Synthesis of carbon-doped SnO2 nanostructures for visible-light-driven photocatalytic hydrogen production from water splitting. Int. J. Hydrog. 2020, 45, 32789–32796. [Google Scholar] [CrossRef]
  19. Paramarta, V.; Taufik, A.; Saleh, R. Comparison of photocatalytic performance of different types of grapheme in Fe3O4/SnO2 composites. J. Phys. Conf. Ser. 2017, 820, 012027. [Google Scholar] [CrossRef] [Green Version]
  20. Hong, L.R.; Zhao, J.; Lei, Y.M.; Yuan, R.; Zhuo, Y. Efficient Electrochemiluminescence from Ru(bpy)32+ Enhanced by Three-Layer Porous Fe3O4@SnO2@Au Nanoparticles for Label-Free and Sensitive Bioanalysis. Electrochim. Acta 2017, 241, 291–298. [Google Scholar] [CrossRef]
  21. Luo, G.; Lu, Y.; Zeng, S.; Zhong, S.; Yu, X.; Fang, Y.; Sun, L. Synthesis of rGO-Fe3O4-SnO2-C Quaternary Hybrid Mesoporous Nanosheets as a High-performance Anode Material for Lithium Ion Batteries. Electrochim. Acta 2015, 182, 715–722. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, Y.; Peng, Z.; Jiang, W. Controlled synthesis of Fe3O4@SnO2/RGO nanocomposite for microwave absorption enhancement. Ceram. Int. 2016, 42, 10682–10689. [Google Scholar] [CrossRef]
  23. Chen, Y.; Song, B.H.; Lu, L.; Xue, J.M. Synthesis of carbon coated Fe3O4/SnO2 composite beads and their application as anodes for lithium ion batteries. Mater. Technol. 2013, 8, 254–259. [Google Scholar] [CrossRef]
  24. Hur, J.U.; Han, J.S.; Shin, J.R.; Park, H.-Y.; Choi, S.-C.; Jung, Y.-G.; An, G.S. Fabrication of SnO2-decorated Fe3O4 nanoparticles with anionic surface modification. Ceram. Int. 2019, 45, 21395–21400. [Google Scholar] [CrossRef]
  25. Elci, A.; Demirtas, O.; Ozturk, I.M.; Bek, A.; Esenturk, E.N. Synthesis of tin oxide-coated gold nanostars and evaluation of their surface-enhanced Raman scattering activities. J. Mater. Sci. 2018, 53, 16345–16356. [Google Scholar] [CrossRef]
  26. Yang, D.; Hu, J.; Fu, S. Controlled Synthesis of Magnetite-Silica Nanocomposites via a Seeded Sol-Gel Approach. J. Phys. Chem. C 2009, 113, 7646–7651. [Google Scholar] [CrossRef]
  27. Rao, X.; Abou Hassan, A.; Guyon, C.; Zhang, M.; Ognier, S.; Tatoulian, M. Plasma Polymer Layers with Primary Amino Groups for Immobilization of Nano- and Microparticles. Plasma Chem. Plasma Process. 2020, 40, 589–606. [Google Scholar] [CrossRef]
  28. Majoul, N.; Aouida, S.; Bessaïs, B. Progress of porous silicon APTES-functionalization by FTIR investigations. Appl. Surf. Sci. 2015, 331, 388–391. [Google Scholar] [CrossRef]
  29. An, G.S.; Shin, J.R.; Hur, J.U.; Oh, A.H.; Kim, B.G.; Jung, Y.-G.; Choi, S.-C. Fabrication of core-shell structured Fe3O4@Au nanoparticle via self-assembly method based on positively charged surface silylation/polymerization. J. Alloys Compd. 2019, 798, 360–366. [Google Scholar] [CrossRef]
  30. Zhang, R.; Liu, Y.; Yang, Z.; Li, Y.; Rong, X.; Wang, L.; Guo, C.; Li, S.; Liu, J.; Li, M.; et al. Construction of nanoparticles based on amphiphilic PEI-PA polymers for bortezomib and paclitaxel co-delivery. RSC Adv. 2015, 5, 15453–15460. [Google Scholar] [CrossRef]
  31. An, G.S.; Han, J.S.; Shin, J.R.; Chae, D.H.; Hur, J.U.; Park, H.Y.; Jung, Y.G.; Choi, S.C. In situ synthesis of Fe3O4@SiO2 core–shell nanoparticles via surface treatment. Ceram. Int. 2018, 44, 12233–12237. [Google Scholar] [CrossRef]
  32. Oh, A.H.; Park, H.Y.; Jung, Y.-G.; Choi, S.-C.; An, G.S. Synthesis of Fe3O4 nanoparticles of various size via the polyol method. Ceram. Int. 2020, 46, 10723–10728. [Google Scholar] [CrossRef]
  33. Zhang, Q.; Liu, P.; Miao, C.; Chen, Z.; Lawrence Wu, C.M.; Shek, C.H. Formation of orthorhombic SnO2 originated from lattice distortion by Mn-doped tetragonal SnO2. RSC Adv. 2015, 5, 39285–39290. [Google Scholar] [CrossRef]
  34. Dey, T.; Khuntia, P.; Mahajan, A.V.; Kumar, N.; Sundaresan, A. 119Sn NMR probe of magnetic fluctuations in SnO2 nanoparticles. Europhys. Lett. 2011, 96, 67008. [Google Scholar] [CrossRef] [Green Version]
  35. Lungu, C.N.; Diudea, M.V.; Putz, M.V.; Grudziński, I.P. Linear and branched PEIs (Polyethylenimines) and their property space. Int. J. Mol. Sci. 2016, 17, 555. [Google Scholar] [CrossRef] [PubMed]
  36. Wongsaprom, K.; Bornphotsawatkun, R.A.; Swatsitang, E. Synthesis and characterization of tin oxide (SnO2) nanocrystalline powders by a simple modified sol-gel route. Appl. Phys. A Mater. Sci. Process. 2014, 114, 373–379. [Google Scholar] [CrossRef]
  37. Lokman, I.M.; Rashid, U.; Taufiq-Yap, Y.H.; Yunus, R. Methyl ester production from palm fatty acid distillate using sulfonated glucose-derived acid catalyst. Renew. Energy 2015, 81, 347–354. [Google Scholar] [CrossRef]
  38. Deng, Y.; Zheng, Y.; Zhang, D.; Han, C.; Cheng, A.; Shen, J.; Zeng, G.; Zhang, H. A novel and facile-to-synthesize three-dimensional honeycomb-like nano-Fe3O4@C composite: Electromagnetic wave absorption with wide bandwidth. Carbon 2020, 169, 118–128. [Google Scholar] [CrossRef]
  39. Wang, B.; Zhang, X.; Liu, X.; Wang, G.; Wang, H.; Bai, J. Rational design of Fe3O4@C yolk-shell nanorods constituting a stable anode for high-performance Li/Na-ion batteries. J. Colloid Interface Sci. 2018, 528, 225–236. [Google Scholar] [CrossRef]
  40. Li, Q.; Tang, G.; Xiong, X.; Cao, Y.; Chen, L.; Xu, F.; Tan, H. Carbon coated magnetite nanoparticles with improved water-dispersion and peroxidase-like activity for colorimetric sensing of glucose. Sens. Actuators B Chem. 2015, 215, 86–92. [Google Scholar] [CrossRef]
Figure 1. Schematic for the preparation of the Fe3O4-SnO2-C nanoparticles.
Figure 1. Schematic for the preparation of the Fe3O4-SnO2-C nanoparticles.
Nanomaterials 11 02877 g001
Figure 2. FTIR spectra of the as-prepared Fe3O4, SnO2, tSCD-treated Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles.
Figure 2. FTIR spectra of the as-prepared Fe3O4, SnO2, tSCD-treated Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles.
Nanomaterials 11 02877 g002
Figure 3. Particle size distribution curves of the as-prepared Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles.
Figure 3. Particle size distribution curves of the as-prepared Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles.
Nanomaterials 11 02877 g003
Figure 4. XRD patterns of the as-prepared Fe3O4, Fe3O4-SnO2, PEI-treated Fe3O4-SnO2, and APTES-treated Fe3O4-SnO2 nanoparticles.
Figure 4. XRD patterns of the as-prepared Fe3O4, Fe3O4-SnO2, PEI-treated Fe3O4-SnO2, and APTES-treated Fe3O4-SnO2 nanoparticles.
Nanomaterials 11 02877 g004
Figure 5. TEM and HRTEM images of the (a) as-prepared Fe3O4, (b,b-1,b-2) Fe3O4-SnO2, (c,c-1,c-2) PEI-treated Fe3O4-SnO2, and (d,d-1,d-2) APTES-treated Fe3O4-SnO2 nanoparticles.
Figure 5. TEM and HRTEM images of the (a) as-prepared Fe3O4, (b,b-1,b-2) Fe3O4-SnO2, (c,c-1,c-2) PEI-treated Fe3O4-SnO2, and (d,d-1,d-2) APTES-treated Fe3O4-SnO2 nanoparticles.
Nanomaterials 11 02877 g005
Figure 6. Schematic for the carbon-coating mechanism of amino-functionalized Fe3O4-SnO2 nanoparticles through glucose and sulfuric acid.
Figure 6. Schematic for the carbon-coating mechanism of amino-functionalized Fe3O4-SnO2 nanoparticles through glucose and sulfuric acid.
Nanomaterials 11 02877 g006
Figure 7. HRTEM images and SAED patterns of (ac) the PEI-treated Fe3O4-SnO2 and (df) APTES-treated Fe3O4-SnO2 nanoparticles after the carbon coating process.
Figure 7. HRTEM images and SAED patterns of (ac) the PEI-treated Fe3O4-SnO2 and (df) APTES-treated Fe3O4-SnO2 nanoparticles after the carbon coating process.
Nanomaterials 11 02877 g007
Figure 8. XRD patterns (a) and Raman spectra (b) of the Fe3O4-SnO2-C composite nanoparticles.
Figure 8. XRD patterns (a) and Raman spectra (b) of the Fe3O4-SnO2-C composite nanoparticles.
Nanomaterials 11 02877 g008
Figure 9. FTIR spectra (a) and particle size distribution curves (b) of the Fe3O4-SnO2 and Fe3O4-SnO2-C nanoparticles.
Figure 9. FTIR spectra (a) and particle size distribution curves (b) of the Fe3O4-SnO2 and Fe3O4-SnO2-C nanoparticles.
Nanomaterials 11 02877 g009
Figure 10. Magnetization curves of the Fe3O4, Fe3O4-SnO2, and Fe3O4-SnO2-C nanoparticles at room temperature.
Figure 10. Magnetization curves of the Fe3O4, Fe3O4-SnO2, and Fe3O4-SnO2-C nanoparticles at room temperature.
Nanomaterials 11 02877 g010
Table 1. Zeta potentials of the as-prepared Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles in the presence of distilled water.
Table 1. Zeta potentials of the as-prepared Fe3O4, Fe3O4-SnO2, APTES-treated Fe3O4-SnO2, and PEI-treated Fe3O4-SnO2 nanoparticles in the presence of distilled water.
SampleZeta Potential (mV)Standard Deviation
Prepared Fe3O4−5.40.08
Fe3O4-SnO2−42.31.17
APTES-treated Fe3O4-SnO231.81.05
PEI-treated Fe3O4-SnO232.91.68
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hur, J.U.; An, G.S.; Choi, S.-C. Preparation of Multilayered Core–Shell Fe3O4-SnO2-C Nanoparticles via Polymeric/Silane–Amino Functionalization. Nanomaterials 2021, 11, 2877. https://doi.org/10.3390/nano11112877

AMA Style

Hur JU, An GS, Choi S-C. Preparation of Multilayered Core–Shell Fe3O4-SnO2-C Nanoparticles via Polymeric/Silane–Amino Functionalization. Nanomaterials. 2021; 11(11):2877. https://doi.org/10.3390/nano11112877

Chicago/Turabian Style

Hur, Jae Uk, Gye Seok An, and Sung-Churl Choi. 2021. "Preparation of Multilayered Core–Shell Fe3O4-SnO2-C Nanoparticles via Polymeric/Silane–Amino Functionalization" Nanomaterials 11, no. 11: 2877. https://doi.org/10.3390/nano11112877

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