DOI:
10.1039/C6RA01408A
(Paper)
RSC Adv., 2016,
6, 35228-35238
One-step hydrothermal synthesis of hydrophilic Fe3O4/carbon composites and their application in removing toxic chemicals†
Received
17th January 2016
, Accepted 4th April 2016
First published on 6th April 2016
Abstract
Hydrophilic Fe3O4/C nanocomposites are prepared by a simple one-step hydrothermal reaction route using FeCl3 and glucose as raw materials, and sodium acetate as an alkali source. The phase structure, morphology, and composition are characterized by X-ray diffraction, Raman spectroscopy, transmission electron microscopy, mass spectrometry, Fourier transform infrared spectroscopy, and X-ray photoelectron spectroscopy, revealing that the glucose plays dual roles in the formation of final products: (I) the reduction of partial Fe3+ to Fe2+ ions and the formation of size-controlled Fe3O4 particles; (II) the stabilization and further functionalization of Fe3O4 nanoparticles by coating carbonous layers. The influence of crucial reaction factors such as the glucose concentration, pH value, reaction temperature and time on the aimed products is also investigated. The prepared Fe3O4/C samples demonstrate typical ferromagnetic behaviors and high removal capacities in removing the toxic Cr(VI) ions and organic pollutant Rhodamine B from wastewater, together with facile magnetic separability and good recyclability.
1. Introduction
The synthesis of nanomaterials has attracted extensive research interest over the past few decades due to their superior physical, chemical, and biological properties to bulk materials.1–3 Magnetic iron oxide Fe3O4 nanoparticles are a kind of novel functional nanomaterial, which have been widely used in catalysis, environmental protection, sensors, magnetic storage media, and clinical diagnosis and treatment, based on their advantages such as magnetic properties, chemical stability, biocompatibility, and low toxicity.4–8 Currently, various strategies have been developed to prepare magnetite nanoparticles with the desired physical and chemical properties, including the coprecipitation of ferric and ferrous ions with a 1
:
2 molar ratio in a basic system,9 thermal decomposition of the metal organic precursor (iron(III)acetylacetonate) at high temperature,10,11 hydrothermal synthesis,12,13 nonaqueous route,14,15 sonochemical synthesis16 and microemulsion.17 However, in these previous studies, the process usually involves in the nitrogen or argon protection to avoid the oxidation of Fe2+ and ensure the formation of Fe3O4 nanoparticles, inevitably undergoing the complicated and harsh synthetic condition. In addition, the used metal organic precursors in the thermal decomposition technique are high-cost and the conventionally used organic solutions and reducing agents, such as ethylene glycol and diethylene glycol, are inferior to the aqueous media in view of the cost and the potential environmental or biological risks. In order to keep free from the trivial, expensive, and toxic fabrication process, a simple and green synthesis route to prepare magnetic nanoparticles at low cost is in demand.
Size and surface functionality have been proved to be two crucial factors in determining the applicable performance of magnetite materials.18,19 However, the aggregation of magnetite nanoparticles, caused by the crystal-face attraction, electrostatic and dipolar fields, usually results in the large-sized particles.5 Moreover, the naked magnetite nanoparticles are easily oxidized in air due to the high chemical activity, leading to the loss of magnetism. The dispersibility and stability of magnetic nanoparticles could be resolved by encapsulating the nanoparticles in polymeric or inorganic matrixes.20–22 Especially, the water-dispersible magnetic particles coated with the hydrophilic layer are highly in demand with respect to the biomedical and adsorbent application, since the magnetic particles are usually hydrophobic. For example, Zhang et al.9 have prepared the β-cyclodextrin coated Fe3O4 using a co-precipitation of FeCl3·6H2O and FeCl2·4H2O in NH3·H2O solution containing the hydrophilic small organic molecule of β-cyclodextrin under the protection of nitrogen. Sun et al.23 have firstly synthesized the oleic acid coated iron oxide nanoparticles by thermal decomposition of Fe(acac)3 (acac = acetylacetonate) in the presence of oleic acid and oleylamine, and then used the ligand-exchange strategies to offer them hydrophilic surface and aqueous dispersibility. Ge and co-workers15 have synthesized highly water-dispersible magnetite by using the surfactant of poly(acrylic acid) as the coated layer. However, some non-ignorable problems confined their practical application, such as the toxic and non-biodegradable coated agents, or high cost, and multiple, complicated procedures. Direct synthesis of water-soluble Fe3O4 nanoparticles coated with the low-cost and eco-friendly agent is very desirable for potential applications.
Heavy metal ions and organic pollutants in water system are great threats to the environment and human health. Among various water treatment technologies, such as the chemical precipitation, ion-exchange, membrane filtration, adsorption, and coagulation–flocculation, adsorption has been a preferred one because of its simplicity, high efficiency, and availability of a large number of adsorbents.24,25 As one of the potential adsorbents, the magnetic iron oxide nanoparticles and nanocomposites have received much attention due to the high removal capacity for heavy metal ions (As(V), Cr(VI), etc.) and organic pollutants, as well as the convenient separation caused by their unique magnetic properties.25–28 Moreover, the water-soluble magnetic nanocomposites with the functional groups and porous coating structure could improve dispersion and adsorption behaviors, providing a solid basis for the wastewater treatment.
In this work, our efforts were focused on the design of simple, green, and low-cost synthesis route for the preparation of carbon-coated magnetite nanoparticles. The one-step synthesis of carbon-coated magnetite nanoparticles using the glucose assisted hydrothermal reaction method was performed. The reaction parameters including the glucose concentration, pH value, reaction temperature and time were systematically studied. Based on the optimized synthesis strategy, the particle sizes of magnetite nanoparticles were well controlled in the range of 120–140 nm due to the confinement of carbon layer. Furthermore, the gluconic acid and glucose were anchored on the surface of magnetite/carbon composite particles, offering a large number of surface hydroxyl groups, and thus the composites showed good water solubility and an excellent removal capacity for the heavy metal ion Cr(VI) and organic pollutant Rhodamine B (RhB).
2. Experimental
2.1 Sample preparation
Ferric chloride (FeCl3), glucose and sodium acetate, obtained from Tianjin Guangfu Chemical Co., were used as chemical reagents for the preparation of magnetite/carbon nanoparticles. All the chemicals were of analytical grade and used as received without further purification.
In a typical synthesis, 0.3, 0.6 or 1.1 g of glucose was firstly dissolved in 80 mL of deionized water under vigorous stirring, followed by the addition of 1.6 g of FeCl3. Subsequently, 4.0 g of NaOAc was introduced to adjust the pH value of the solution to ca. 5. After stirring for 30 min, the final mixture was transferred to a Teflon-lined autoclave (150 mL capacity) and heated at 200 °C for 10 h. The resulting product was filtrated, washed with water, and dried at 60 °C overnight, which was labeled as Fe3O4/C-a, -b, or -c for the carbon-coated Fe3O4 composite sample obtained with 0.3, 0.6 or 1.1 g of glucose, respectively. For comparison, a control experiment without the addition of glucose was also carried out. Moreover, various factors in affecting the aimed products were also explored in the same synthesis procedure, including the glucose concentration, pH value, reaction temperature and time.
2.2 Sample purification
The prepared sample was stirred for 1 h, with the pure carbon spheres floating on the surface of the aqueous solution due to the lower density than those of Fe3O4 and Fe3O4/C. Then, the Fe3O4 and Fe3O4/C were collected by a magnet from the aqueous solution and put into the deionized water again, which continually stirred for 30 min and remained still for 10 min. Since the bigger particle size and larger density of the uncovered Fe3O4 compared to those of Fe3O4/C, the sediments in the bottom were abandoned and the upper suspension were obtained. In the end, the upper suspension was centrifuged at 8000 rpm for 10 min to obtain the final purified Fe3O4/C.
2.3 Characterization
Transmission electron microscopy (TEM) was performed on a JEOL JEM 2010F at 200 kV. All the samples subjected to TEM measurements were ultrasonically dispersed in ethanol and drop-casted onto copper grids covered with carbon film. X-ray diffraction (XRD) patterns were recorded on a Bruker D8 Focus diffractometer with Cu Kα radiation (λ = 1.5406 Å) operated at 40 kV and 40 mA. The diffraction data were collected over the 2θ angle of 10° to 80° at a scan rate of 6° min−1. Raman spectroscopy was performed on a Renishaw-1000 Raman spectrometer. Thermogravimetric analysis (TGA) of the samples was conducted on a TA SDT Q600 analyzer in nitrogen atmosphere with a heating rate of 10 °C min−1. N2 adsorption–desorption isotherms were recorded on a Quantachrome NOVA 2000e sorption analyzer at liquid nitrogen temperature (77 K). The specific surface area was obtained by the Brunauer–Emmett–Teller (BET) method, and the pore size distribution was derived from the adsorption branches of isotherms using the DFT method. Mass spectrometry (MS) measurement was carried out on a Shimadzu LCMS-2020 mass spectrometer under the following conditions: interface voltage, 4.5 kV; DL temperature, 250 °C; heat block temperature, 200 °C. Fourier transform infrared (FT-IR) spectrum was recorded on a Bruker Vector 22 FT-IR spectrophotometer using a KBr pellet. X-ray photoelectron spectroscopy (XPS) measurement was performed on a Kratos Axis Ultra DLD (delay line detector) spectrometer equipped with a monochromatic Al Kα X-ray source (1486.6 eV). The XPS survey spectra were recorded with a pass energy of 160 eV, and high-resolution spectra with a pass energy of 40 eV. Binding energies were calibrated by using the containment carbon (C 1s 284.6 eV). Magnetization studies were performed at room temperature using a LDJ 9600 VSM magnetometer.
2.4 Adsorption measurements
For Cr(VI) adsorption: the aqueous solutions with different Cr(VI) concentrations (20, 40, 60, 80 and 100 mg L−1) were prepared by dissolving different amounts of K2Cr2O7 in deionized water. The pH values of the solution were adjusted to 4 by hydrochloric acid. 20 mg of Fe3O4/C-b was added into 100 mL of Cr(VI) solution with different concentrations. The mixture was continually stirred for 2.5 h in dark, followed by sampling the solution at defined time intervals. For analysis of Cr(VI), 5 mL of obtained clear solution was diluted with different amount of deionized water to ensure that the total Cr(VI) mass in the 50 mL colorimetric cylinder was less than 50 μg. Then 5 mL of diluted Cr(VI) solution and 2 mL of diphenylcarbazide were successively added into the colorimetric cylinder, diluted to 50 mL with deionized water, and adjusted pH to 1–2 (by adding H2SO4 (1
:
5, v/v) under the test of the pH-meter). The adsorbed capacity of Cr(VI) was monitored by measuring the UV absorption at λmax = 540 nm of the initial and final solutions based on the Cr(VI) standard curve formula of C = 8.25092A − 0.00214, where C represents the Cr(VI) concentration, and A is the adsorption intensity.
In order to investigate the recycling potential, the Cr(VI)-loaded Fe3O4/C-b was separated from the solution by a conventional magnet, and ultrasonically cleaned in deionized water for 30 min, and subsequently soaked in 0.25 mol L−1 NaOH for 3 h. The resulted Fe3O4/C-b sample was rinsed with 0.05 mol L−1 NaOH and distilled water in succession, and such washing procedures were repeated until the rinsed water becomes colorless. Finally, the regenerated product was dried at 60 °C overnight before reuse as adsorbent.
For RhB adsorption: 20 mg of the Fe3O4/C-b sample was placed into a tubular quartz reactor filled with 100 mL of RhB aqueous solution (10−5 mol L−1). The mixture was stirred for 2.5 h in dark at room temperature (30 °C). A magnet of about 0.002 T was used for collecting the magnetite nanoparticles in the solution. Finally, the content of RhB in the solution was measured by a SP-752 spectrophotometer at λmax = 554 nm.
3. Results and discussion
3.1 Glucose assisted hydrothermal synthesis of Fe3O4/C nanocomposites
Fig. 1 shows the XRD patterns of the prepared samples obtained from the hydrothermal method with and without the use of glucose. The diffraction peaks of the Fe3O4/C-b prepared with the assistance of glucose are well consistent with the characteristic of the cubic Fe3O4 (a = b = c = 8.396 Å, Fd3m (227), JCPDS 19-0629) or γ-Fe2O3 (maghemite, JCPDS 39-1346), while the sample prepared in the absence of glucose exhibits the typical feature of the α-Fe2O3. This indicates the crucial role of glucose in the formation of Fe3O4 or γ-Fe2O3. However, the specific structural phase of the Fe3O4/C-b sample can still not be identified between Fe3O4 and γ-Fe2O3 only from the XRD data since they have similar XRD patterns. Raman spectroscopy is thus performed to further distinguish the structural phase of Fe3O4/C-b, as the two main bands of Fe3O4 centered at 668 (A1g) and 535 (T2g) cm−1, and the ones of γ-Fe2O3 are located at 350, 500, and 720 cm−1.12 As shown in Fig. 2, the two main peaks at 663 and 535 cm−1 are observed, assignable to the typical feature of Fe3O4, suggesting the presence of iron oxide in the form of Fe3O4 nanoparticles in Fe3O4/C-b. In addition, two broad peaks in the range of 1500–1650 cm−1 and 1350–1450 cm−1 are also observed in the Raman spectrum (Fig. 2 (inset)), which can be assigned to the crystalline graphitized carbon (G-band) and disordered amorphous carbon (D-band), respectively.29,30 The much stronger peak intensity of G-band than that of D-band implies that glucose, as the carbon source, is partially pyrolyzed to form graphitized carbon layer after the hydrothermal treatment.
 |
| | Fig. 1 XRD patterns of as-prepared Fe3O4/C-b and α-Fe2O3. | |
 |
| | Fig. 2 Raman spectra of Fe3O4/C-b nanocomposites. | |
Fig. 3 shows TEM images and SAED pattern of the prepared Fe3O4/C-b. As shown in Fig. 3A and B, the dark Fe3O4 particles were encapsulated into the carbon spheres and the mean diameters of composite particles are in the range of 120–140 nm. The carbon layer is ca. 20 nm in thickness and the coated Fe3O4 particles exhibits quasi-spheric morphology with the size of 100–120 nm. The ordered lattice fringes were clearly observed in the high-resolution TEM image of Fe3O4/C-b (Fig. 3C), with the fringe separation of 0.48 nm corresponding to (111) plane of Fe3O4. The selected-area electron diffraction (SAED) in Fig. 3D also confirms the formation of Fe3O4.
 |
| | Fig. 3 TEM images (A–C) and SAED pattern (D) of Fe3O4/C-b. | |
Fig. 4 shows the FT-IR spectra of the synthesized Fe3O4/C-b. An obvious absorption peak at 571 cm−1 is regarded as the vibration of Fe–O bond, representing the magnetite phase of product.31 The broad absorption band in the range of 3160–3660 cm−1 is assigned to O–H stretching vibration, arising from hydroxyl groups on nanoparticles, the adsorbed glucose and gluconic acid.30 The two obvious peaks at the 1427 and 1593 cm−1 correspond to the asymmetric and symmetric stretching vibrations of COO−, further confirming the existence of the gluconic acid on the surface of Fe3O4.32 While the band at 1050 cm−1 of C–O stretching almost comes from the residue glucose, and the peaks around 2849 and 2927 cm−1, assignable to asymmetric and symmetric vibrations of C–H in –CH2–, is related to the carbon layer due to the carbonization of glucose.33 The FT-IR analysis confirms that Fe3O4 nanoparticles not only be synthesized by the reduction of partial Fe3+ ions to Fe2+ ions with the glucose, but further stabilized and modified by the carbon layer, along with the residual glucose and gluconic acid. Moreover, the large amount of OH and COO− groups covalently bonded to the carbon frameworks offers the hydrophilic surface of Fe3O4 nanoparticles and the Fe3O4 particles show the good water dispersibility, superior to the sample obtained by the traditional method (Fig. 4 (inset)).34
 |
| | Fig. 4 FT-IR spectrum of as-synthesized Fe3O4/C-b, and the photograph of aqueous solution of Fe3O4/C-b and the one synthesized according to the literature32 (inset). | |
XPS analysis was performed to reveal the compositions and chemical structures of the prepared Fe3O4/C-b (Fig. 5). The Fe 2p and O 1s peaks arise from the contribution of Fe3O4 nanoparticles, and the C 1s and O 1s peaks point to the carbon layer, glucose and gluconic acid which are coated on the surface of Fe3O4. The high-resolution Fe 2p spectrum shows that the Fe 2p3/2 and Fe 2p1/2 are located at the binding energy of 711.6 and 725.2 eV, respectively. The broadening of these two peaks and the disappearance of charge transferring satellites of Fe 2p3/2 around 719 eV indicate the appearance of dual iron oxidation states (Fe2+ and Fe3+), well consistent with the previous reports.35–37 This also provides another piece of proof that the structural phase of the prepared sample is Fe3O4.
 |
| | Fig. 5 XPS survey scans (a) and high-resolution Fe 2p spectrum (b) of the synthesized Fe3O4/C-b. | |
As discussed above, the Fe3O4 particles have been successfully prepared from the FeCl3 and glucose in the sodium acetate solution. In this hydrothermal synthesis route, the introduction of glucose is a crucial factor for the preparation of Fe3O4, since the structural phase of the product is pure α-Fe2O3 without the addition of glucose. It is suggested that the partial Fe3+ ions have been reduced to Fe2+ ions for the formation of Fe3O4 in the glucose-assisted synthesis route. As no external reductant was presented in the solution, the glucose is assumed to be responsible for the reduction reaction. The mass spectrometry in the ESI (Fig. S1†) further provides a solid evidence for the hydrothermal reaction process, by analyzing the composition of the initial and final reaction solution, respectively. Only the pristine glucose is present in the solution at the initial stage (Fig. S1a†), while the gluconic acid is detected in the reaction solution after hydrothermal treatment (Fig. S1b†), suggesting that the glucose has been oxidized to the gluconic acid due to the reduction of partial Fe3+ ions. Herein, the glucose plays the reduction role in reducing the partial Fe3+ ions to Fe2+ ions for the composition of Fe3O4.
Another observation is that the carbon layer derived from the carbonization of glucose is coated on the surface of Fe3O4 particles, as evidenced by the TEM images and Raman spectroscopy. A possible explanation is that the glucose would firstly reduce partial Fe3+ ions to Fe2+ ions to form the Fe3O4 crystalline nucleus and grow up, and the gluconic acid arising from the oxidation of glucose would be anchored on the surface of magnetite particles.38 Meanwhile, the glucose as the carbon source also involves other chemical reaction in the hydrothermal process (Fig. 6): glucose is gradually polymerized into polysaccharides by the intermolecular dehydration and then carbonized into the carbon nanosphere, and such process would prefer to be conducted around the magnetite particles due to their catalytic carbonization role.32,33 As a result, the carbon layer along with residual gluconic acid and glucose are completely or incompletely coated on the surface of Fe3O4 particles, thus stabilizing and modifying the surface properties of Fe3O4 particles.
 |
| | Fig. 6 Schematic illustration of the glucose-assisted hydrothermal synthesis of the Fe3O4/C nanocomposites. | |
3.2 Factors influencing the formation of Fe3O4/C nanoparticles
Based on above powerful tools, such as XRD, TEM, mass spectrometry, Raman and FT-IR spectroscopy, the hydrothermal reaction route has been confirmed as an effective strategy to prepare carbon-coated magnetic nanoparticles. Moreover, the synthesis route is nontoxic and low-cost, without involving the organic solvents, initiators, or surfactants which are commonly used for the preparation of functionalized Fe3O4 particles. For the novel synthesis approach, systematic studies have revealed that many factors influence the successful formation of magnetite particles, such as the glucose concentration, pH value, reaction temperature and time, which should be discussed in detail.
3.2.1 Effect of the glucose concentration. The glucose plays the decisive role in the preparation of Fe3O4 nanoparticles in consideration of its reduction of partial Fe3+ ions to Fe2+ ions and further functionality of the as-prepared product. Fig. 7 shows the XRD patterns of magnetite nanoparticles obtained by changing the glucose amount from 0.3 to 1.1 g while keeping the other reaction conditions at the same level. The XRD patterns of all the samples agree well with the characteristic peaks of Fe3O4. A significant decrease in the relative intensity of diffraction peaks was observed with the increased glucose amount from 0.3 g of sample Fe3O4/C-a to 0.6 g of sample Fe3O4/C-b, accompanied with the broadening of Fe3O4 diffractions peaks, indicating a dramatic reduction of the crystal particle size. However, when the amount of added glucose increased to 1.1 g for sample Fe3O4/C-c, the diffraction peaks become sharper and higher again, suggesting that the Fe3O4 particles grew into larger crystallite size in the higher glucose concentration. The smallest particles were achieved when the appropriate amount of glucose was added (0.6 g), with the average size of 126 nm calculated from the Scherrer's equation (D = 0.89k/β
cos
θ).10
 |
| | Fig. 7 XRD patterns of magnetite nanoparticles obtained by using different glucose concentration. | |
Fig. 8 displays TEM images of magnetite nanoparticles prepared by using different glucose concentration, which provide more detailed morphology information and size distribution of the carbon-coated Fe3O4. TEM analysis shows that the morphology and size of the prepared Fe3O4 particles are concentration-dependent. In a low glucose concentration, the Fe3O4 have a cubic shape with the average particle size of approximately 1.25 μm (Fig. 8A and B), and the coating layer of the carbon, glucose and gluconic acid with the thickness of ca. 10 nm are coated on the surface of Fe3O4 particles (Fig. 8B). When the glucose concentration increased, the Fe3O4 particles show the quasi-spheric structure and the reduced particle size (ca. 130 nm) (Fig. 8C and D). However, the particle size of Fe3O4/C-c centered on ca. 350 nm, bigger than those of Fe3O4/C-b, implying particles grew larger with further increment of glucose concentration. Consequently, the carbon layer may be too thin to restrict the particle size of Fe3O4 in low glucose concentration, while the glucose tends to form the large-sized carbon sphere in high concentration, which all leads to the undesired particle size and morphology of carbon-coated Fe3O4. The optimum particle size was achieved for the Fe3O4/C-b by controlling the glucose concentration, which also agrees well with the analysis of XRD data. The results demonstrate that the glucose concentration played an important role in controlling the morphology of the Fe3O4/C. The mechanism may be explained as follows: in the low glucose concentration, the thin carbon layer (shell) would tend to coat on the surface of crystallized Fe3O4 particles (core), forming the typical core–shell structure. The unique structural feature could restrict the growth of Fe3O4 particles and control their cubic morphology. However, once the glucose concentration increased to a certain degree, the glucose would be inclined to polymerize into carbon microspheres and then serve as the nucleating agent. Subsequently, the Fe3+ diffused into the carbon microspheres and crystallized to form the spherical morphology of Fe3O4/C particles.
 |
| | Fig. 8 TEM images of Fe3O4/C nanoparticles prepared by using different glucose concentration: (A and B) Fe3O4/C-a, (C and D) Fe3O4/C-b, and (E and F) Fe3O4/C-c. | |
The coating agents in the synthesized Fe3O4/C samples were analyzed by the TGA measurements. All the samples were heated to 800 °C in nitrogen atmosphere to avoid the oxidation of Fe2+ ions, during which the weight loss of the samples is only attributed to the thermal decomposition of glucose, gluconic acid and glucose polymers. As shown in Fig. S2,† the TGA curves of synthesized magnetite nanoparticles present three weight loss processes. The first weight loss between 25 °C and 200 °C should be ascribed to the dehydration of the samples due to the loss of surface adsorbed water and crystal water. The second weight loss in the temperature range of 200–500 °C corresponds to the desorption and subsequent evaporation of the coating agents, including glucose, gluconic acid and glucose polymers, and the last one at 500–800 °C is due to the decomposition and carbonization of the glucose, gluconic acid and glucose polymers. In our experiment, the higher initial glucose concentration resulted in the higher weight loss, implying the larger amount of residual glucose and gluconic acid.
Fig. 9 shows the nitrogen adsorption–desorption isotherms and the corresponding pore size distribution curves of samples Fe3O4/C-b and Fe3O4/C-c. The nitrogen isotherms of Fe3O4/C-b and Fe3O4/C-c are correspondingly close to type-II and IV curves, with an increased nitrogen adsorbed volume at relative pressure P/P0 of 0.8–1.0.39 The BET surface areas of Fe3O4/C-b and Fe3O4/C-c are 75 and 84 m2 g−1, respectively. The pore diameters of Fe3O4/C-b and Fe3O4/C-c are both distributed in the range of 1–14 nm, accompanied with the pore volume of 0.125 and 0.171 cm3 g−1, respectively. The sample Fe3O4/C-a obtained by low glucose concentration solution has almost no surface area because of the limited coating agents, which is not discussed here in detail.
 |
| | Fig. 9 N2 adsorption–desorption isotherms and the corresponding DFT pore size distribution curves (inset) of samples Fe3O4/C-b and Fe3O4/C-c. | |
3.2.2 Effect of the sodium acetate (NaOAc) amount. The role of NaOAc was also found to be very critical in this synthesis strategy. In a control experiment, different amounts of NaOAc (1.0–8.0 g) were added in the solution. As shown in Fig. S3 (ESI†), when the amount of NaOAc was more than 6.0 g or less than 3.0 g in this synthesis route, the XRD patterns of the prepared samples can only show the characteristics of α-Fe2O3. It is noteworthy that the NaOAc amount of 4.0 g resulted in the high-crystalline Fe3O4 particles. Generally, NaOAc is used to provide a steady OH− ion supply through its hydrolysis. In our experiment, FeCl3 hydrolyzes to form ferric hydroxide and releases H+ ions, leading to the byproducts of HCl. However, the accumulation of HCl would inhibit the further reduction of partial Fe3+ and thus affect the composition of the iron oxide samples.5 The appropriate addition of NaOAc would provide OH− and neutralize the HCl, thus allowing the completion of reduction reaction and the formation of Fe3O4 particles.
3.2.3 Effect of the hydrothermal reaction temperature. Fig. S4† displays the XRD patterns of iron oxide samples obtained by different hydrothermal temperature. Notably, when the hydrothermal temperature is fixed on the 160 °C, no apparent Fe3O4 particles can be achieved. While when the temperature increased to 180 °C, the typical diffraction peaks of Fe3O4 could be observed. Moreover, the degree of crystallinity of Fe3O4 was higher along with the further increased temperature to 200 °C, which was taken as the optimum hydrothermal reaction temperature.
3.2.4 Effect of the hydrothermal treatment time. Since the high-quality Fe3O4 particles can be prepared when the glucose amount is 0.6 g, the NaOAc amount is 4.0 g, and hydrothermal temperature is 200 °C, the investigation of the effect of hydrothermal treatment time on the structural phase of the product was carried out based on the optimum composition and reaction temperature. Fig. S5† shows the XRD patterns of iron oxide samples prepared with different hydrothermal treatment time. It can be observed that the products are presented in the form of α-Fe2O3 when the hydrothermal treatment time was less than 4 hours, whereas the α-Fe2O3 gradually transformed into the Fe3O4 along with the reaction time exceeding 10 hours, as evidenced by the analysis of Fe3O4/C-b. No obvious change had been observed for the products as the reaction time ranges from 10 to 70 hours.
3.3 The magnetic properties of the prepared Fe3O4/C particles
Fig. 10 shows the magnetic hysteresis curves of the Fe3O4/C samples. All the samples with the small hysteresis loops exhibited diamagnetic behaviors, which could be reflected by Br and Hc values in Table 1.35 The phenomenon is attributed to the large particle size (>130 nm) of magnetite particles in this work, since the Fe3O4 nanoparticles with a particle size <30 nm can have superparamagnetic properties.40 Nevertheless, when the magnetic field was 6000 Oe, the samples Fe3O4/C-a, Fe3O4/C-b and Fe3O4/C-c show the Bs values of 26.26, 63.48 and 57.05 emu g−1, respectively, suggesting the prepared samples had strong magnetic response to the external magnetic field, especially the sample Fe3O4/C-b. The higher purity of Fe3O4/C-b than that of Fe3O4/C-a is mainly responsible for the larger magnetism. However, the excessive glucose concentration for the Fe3O4/C-c resulted in the relatively low mass percentage of Fe3O4 in the sample, which decreased the magnetism of Fe3O4/C-c. The highest saturation magnetization would render the Fe3O4/C-b a convenient magnetic separation process in the practical application.
 |
| | Fig. 10 Magnetic hysteresis curves taken at room temperature for the prepared Fe3O4/C samples (inset: the enlarged view). | |
Table 1 Magnetic parameters of the prepared Fe3O4/C samples
| Sample |
Bsa/emu g−1 |
Hcb/Oe |
Brc/emu g−1 |
| Saturation magnetization. Coercive force. Remanent magnetization. |
| Fe3O4/C-a |
26.26 |
40.37 |
1.096 |
| Fe3O4/C-b |
63.48 |
46.91 |
10.03 |
| Fe3O4/C-c |
57.05 |
62.67 |
3.729 |
3.4 The adsorption performance of Fe3O4/C particles in removing toxic chemicals
The heavy metal ions and organic chemicals are highly toxic environmental pollutants existing in industrial wastewater, which are very deleterious to humans, even in trace amount. The adsorption of these toxic environmental pollutants is one of the effective strategies to resolve the environmental problems involving wastewater treatment. Technologically, the constructed Fe3O4/C-b particles with high specific surface areas, strong magnetic properties, and excellent water dispersibility, could be used as an effective adsorbent for the removal of some toxic chemicals, such as Cr(VI) and Rhodamine B (RhB). In our experiment, the prepared Fe3O4/C-b particles were used to remove the Cr(VI) and RhB from wastewater. The effects of adsorption time, the reclaimed sample's performance and its equilibrium adsorption isotherms, and adsorption kinetics were investigated.
3.4.1 The adsorption performance of Fe3O4/C-b particles in removing Cr(VI).
(I) Adsorption isotherms of Fe3O4/C-b particles. The adsorption isotherm plays an important role in understanding sorption systems since the maximum capacity of adsorption could be determined for the adsorbent.41 Fig. 11 shows the adsorption performance of the prepared Fe3O4/C-b adsorbent for Cr(VI). As shown in Fig. 11a, the adsorption of Cr(VI) on Fe3O4/C-b adsorbent at different Cr(VI) concentration can all achieve the absorption equilibrium within 360 min. As a result, the equilibrium concentration of Cr(VI) solution was obtained. The amount of adsorbed Cr(VI) at equilibrium qe was calculated from eqn (1):42| |
 | (1) |
where C0 (mg L−1) is the initial Cr(VI) concentration, Ce (mg L−1) is the equilibrium concentration of Cr(VI) solution, V (L) is the volume of the solution, and m (g) is the mass of adsorbent added to the solution.
 |
| | Fig. 11 The removed percentage of Cr(VI) on as-prepared Fe3O4/C-b particles in different Cr(VI) concentration solution (a), the plot of qe versus Ce for the non-linear adsorption isotherm of Cr(VI) onto Fe3O4/C-b (b), and adsorption isotherms of the Langmuir (c) and Freundlich (d) model. | |
The Langmuir and Freundlich models are two commonly used to investigate the adsorption process and mechanism of the adsorbent. The Langmuir model is primarily applied to the monolayer adsorption on a homogeneous surface without any interaction between the adsorbed molecules, and the Langmuir equation is commonly presented as eqn (2):21,40
| | |
Ce/qe = Ce/qmax + 1/KLqmax
| (2) |
where
qmax (mg g
−1) and
Ce (mg L
−1) represent the maximum capacity and the equilibrium concentration of adsorbate, respectively.
KL (L mg
−1) is the Langmuir constant related to adsorption capacity.
The Freundlich model is widely used for the adsorption in a multilayer manner on an energetically heterogeneous surface, which can be expressed as eqn (3):43
| |
log qe = log KF + 1/n log Ce
| (3) |
where
KF and
n are the Freundlich constants that are related to adsorption capacity and intensity, respectively.
Fig. 11b shows the variation of qe with respect to Ce for the adsorption of Cr(VI). The adsorption isotherms of Fe3O4/C–Cr(VI) adsorption system were built according to the two models (Fig. 11c and d), and the corresponding coefficients and non-linear R2 values were shown in Table 2. It is shown that the R2 value of the Langmuir isotherm model is higher than that of the Freundlich model, suggesting that a monolayer adsorption of Cr(VI) on the surface of Fe3O4/C-b occurs, which may originate from the inverse electrostatic adsorption and irreversible redox reaction on the Fe3O4 particles.44 The maximum adsorption capacity (qmax) of Fe3O4/C-b for Cr(VI) is 61.69 ± 0.5 mg g−1, much higher than previously reported values of other Cr(VI)-removal adsorbents, such as pristine Fe3O4 nanospheres (2.95 mg g−1)45 and Fe3O4 particles (4.38 mg g−1),5 and comparable to the ethylenediamine-functionalized Fe3O4 magnetic polymers (61.35 mg g−1).27 The suitable coating agents (proper functional groups) and high specific surface area of the Fe3O4/C-b particles may result in such good removal performance to heavy metal ion Cr(VI).
Table 2 The kinetics parameters for the adsorption of Cr(VI) and RhB on the Fe3O4/C-b adsorbent
| |
C0 (mg L−1) |
qe,exp (mg g−1) |
Pseudo-first order |
Pseudo-second order |
| k1 (min−1) |
R2 |
qe (mg g−1) |
k2 (g mg−1 min−1) |
R2 |
qe (mg g−1) |
| Cr(VI) |
20 |
26.50 ± 0.2 |
0.004 |
0.959 |
21.17 ± 0.2 |
0.000689 |
0.995 |
29.68 ± 0.2 |
| 40 |
23.27 ± 0.2 |
0.003 |
0.914 |
18.01 ± 0.2 |
0.000723 |
0.982 |
25.62 ± 0.2 |
| 60 |
18.25 ± 0.2 |
0.004 |
0.905 |
12.38 ± 0.2 |
0.001322 |
0.992 |
19.94 ± 0.2 |
| 80 |
14.88 ± 0.2 |
0.005 |
0.948 |
9.06 ± 0.25 |
0.002683 |
0.999 |
15.70 ± 0.2 |
| 100 |
12.00 ± 0.3 |
0.006 |
0.846 |
7.46 ± 0.3 |
0.002948 |
0.998 |
12.89 ± 0.25 |
| RhB |
4.79 |
27.39 ± 0.2 |
0.011 |
0.785 |
25.83 ± 0.2 |
0.000596 |
0.956 |
27.55 ± 0.2 |
(II) Kinetic adsorption rate equation of Fe3O4/C-b particles. The pseudo-first order equation and pseudo-second order equation are used to understand the adsorption kinetics of Fe3O4/C-b in the adsorption of Cr(VI), which could provide useful information about the adsorption type and the crucial factors involved in the adsorption. The two kinetic models can be expressed by eqn (4) and (5), respectively:27,40| |
log(qe − qt) = log qe − k1t
| (4) |
| | |
t/qt = 1/(k2qe2) + t/qe
| (5) |
where qe (mg g−1) and qt (mg g−1) are the adsorbed amounts of Cr(VI) on Fe3O4/C-b at equilibrium and time t, respectively, and k1 (min−1) and k2 (g mg−1 min−1) are the rate constants of the pseudo-first-order adsorption and pseudo-second-order adsorption, respectively.The kinetic adsorption data for the Fe3O4/C–Cr(VI) adsorption system were simulated using the pseudo-first order rate equation and the pseudo-second order rate equation, as shown in Fig. 12. The fitted kinetic parameters are summarized in Table 2. The correlation coefficients (>0.982) for the pseudo-second order equation are higher than those (<0.959) for the pseudo-first-order rate equation, and the calculated qe from the pseudo-second order equation was closer to the experimental value. Therefore, it can be concluded that the adsorption of Cr(VI) on the Fe3O4/C-b particles is a second-order reaction rather than a first-order one, which may involve the surface complexation between the Fe3O4/C surface and the toxic Cr(VI) ions in the aqueous solution.5,43
 |
| | Fig. 12 The pseudo-first order adsorption rate equation (a) and pseudo-second order adsorption rate equation (b). | |
(III) Cycling performance of Fe3O4/C-b particles. In order to investigate the recyclability of Fe3O4/C-b, recycling experiments were carried out. Firstly, the Cr(VI)-loaded Fe3O4/C adsorbent is conveniently separated from the solution by the conventional magnet, and further regenerated by dilute basic solution. The results of recycling experiments are shown in Fig. 13. It is found that, compared to the removal efficiency (26.5 ± 0.5%) of pristine Fe3O4/C-b within 6 h, the regenerated Fe3O4/C-b with a high saturation magnetization of 52.54 emu g−1 shows the slightly inferior adsorption performance for the Cr(VI), but the 22 ± 0.5% of Cr-removal percentage can still be achieved after 3 cycles. The easy separation from the solution, the good recycling adsorption performance for Cr(VI) and the high remained magnetism suggest that Fe3O4/C particles can be utilized as an effective adsorbent for the wastewater treatment.
 |
| | Fig. 13 The Cr(VI) removal percentage by using the Fe3O4/C-b composite in recycle runs (a) and the magnetism variation with the recycle runs (b). | |
3.4.2 The adsorption performance of Fe3O4/C particles in removing RhB. Fig. S6† shows the adsorption performance and the theoretical analysis of the RhB adsorption process for the Fe3O4/C-b absorbent. As shown in Fig. S6a,† the removal percentage of RhB can be up to 83.64 ± 0.5% at the initial concentration of 10 μmol L−1, displaying a high adsorption capacity for the dye molecule. In order to investigate the maximum adsorption capacity, similarly, the Langmuir and Freundlich equations are applied to fit the adsorption data derived from the Fig. S6b.† In comparison of the values of the correlation coefficients (R2) in Fig. S6c and d,† the adsorption process is closer to the Langmuir model. This indicates that the adsorption of RhB is a monolayer adsorption process, and the calculated qmax is 43.86 ± 0.5 mg g−1, higher than those of previously reported adsorbents, such as polystyrene/Fe3O4/graphene oxide (13.8 mg g−1),46 activated carbon (16.1 mg g−1),47 and sodium montmorillonite (42.2 mg g−1).48The kinetic adsorption process for the Fe3O4/C-b absorbent was also investigated by the pseudo-first order and pseudo-second order models, as shown in Fig. S6e–g.† The fitting of adsorption data suggests that the pseudo-second order model with the larger correlation coefficient fits better with the experiment data than the pseudo-first order model. Thus, the adsorption of RhB on the Fe3O4/C-b is likely to be controlled by chemisorption,49 which may be attributed to the hydrogen bonding interaction of RhB and functional groups of glucose and gluconic acid on the surface of Fe3O4/C-b.46
4. Conclusions
The Fe3O4/C nanoparticles have been synthesized via a one-step glucose-assisted hydrothermal reaction approach, in which the Fe3+ ions were partially reduced to the Fe2+ by the glucose, and moreover, the carbon layer with the retention of functional groups, such as glucose and gluconic acid, were also coated on the surface of Fe3O4 nanoparticles. For this novel synthesis route, some crucial factors, such as the glucose concentration, pH value, reaction temperature and time, influence the formation of coated Fe3O4 nanoparticles. When applied as the adsorbent, the coated Fe3O4 nanoparticles with carbon layer, glucose and gluconic acid exhibited good hydrophilicity, superior magnetic properties, and excellent adsorption capacity of toxic chemicals of Cr(VI) and RhB. The Fe3O4 nanoparticles prepared from the facile and low-cost synthesis route may also be extended to the other fields, such as high-density magnetic recording media and catalysis.
Acknowledgements
This research was financially supported by the Foundation of Henan Educational Committee (No. 16A150023), Nanhu Scholars Program for Young Scholars of XYNU, the Doctoral Start-up Research Fund of Xinyang Normal University (15006) and the College of Chemistry and Chemical Engineering of Xinyang Normal University.
Notes and references
- M. E. F. Brollo, R. López-Ruiz, D. Muraca, S. J. A. Figueroa, K. R. Pirota and M. Knobel, Sci. Rep., 2014, 4, 6839–6844 CrossRef CAS PubMed.
- N. B. Saleh, N. Aich, J. Plazas-Tuttle, J. R. Lead and G. V. Lowry, Environ. Sci.: Nano, 2015, 2, 11–18 RSC.
- N. Saleh, K. Sirk, Y. Liu, T. Phenrat, B. Dufour, K. Matyjaszewski, R. D. Tilton and G. V. Lowry, Environ. Eng. Sci., 2007, 24, 45–57 CrossRef CAS.
- K. Naka, A. Narita, H. Tanaka, Y. Chujo, M. Morita, T. Inubushi, I. Nishimura, J. Hiruta, H. Shibayama, M. Koga, S. Ishibashi, J. Seki, S. Kizaka-Kondoh and M. Hiraoka, Polym. Adv. Technol., 2008, 19, 1421–1429 CrossRef CAS.
- L. S. Zhong, J. S. Hu, H. P. Liang, A. M. Cao, W. G. Song and L. J. Wan, Adv. Mater., 2006, 18, 2426–2431 CrossRef CAS.
- Y. H. Deng, Y. Cai, Z. K. Sun, J. Liu, C. Liu, J. Wei, W. Li, C. Liu, Y. Wang and D. Y. Zhao, J. Am. Chem. Soc., 2010, 132, 8466–8473 CrossRef CAS PubMed.
- G. Shi, B. Sun, Z. Jin, J. Liu and M. Li, Sens. Actuators, B, 2012, 171–172, 699–704 CrossRef CAS.
- T. P. Almeida, T. Kasama, A. R. Muxworthy, W. Williams, L. Nagy, T. W. Hansen, P. D. Brown and R. E. Dunin-Borkowski, Nat. Commun., 2014, 5, 5154–5159 CrossRef CAS PubMed.
- X. Zhang, Y. Wang and S. Yang, Carbohydr. Polym., 2014, 114, 521–529 CrossRef CAS PubMed.
- S. Sun and H. Zeng, J. Am. Chem. Soc., 2002, 124, 8204–8205 CrossRef CAS PubMed.
- S. Sun, H. Zeng, D. B. Robinson, S. Raoux, P. M. Rice, S. X. Wang and G. Li, J. Am. Chem. Soc., 2004, 126, 273–279 CrossRef CAS PubMed.
- T. J. Daou, G. Pourroy, S. Bégin-Colin, J. M. Grenèche, C. Ulhaq-Bouillet, P. Legarè, P. Bernhardt, C. Leuvrey and G. Rogez, Chem. Mater., 2006, 18, 4399–4404 CrossRef CAS.
- J. Wang, J. Sun, Q. Sun and Q. Chen, Mater. Res. Bull., 2003, 38, 1113–1118 CrossRef CAS.
- J. Liu, Z. Sun, Y. Deng, Y. Zou, C. Li, X. Guo, L. Xiong, Y. Gao, F. Li and D. Zhao, Angew. Chem., Int. Ed., 2009, 48, 5875–5879 CrossRef CAS PubMed.
- J. Ge, Y. Hu, M. Biasini, W. P. Beyermann and Y. Yin, Angew. Chem., Int. Ed., 2007, 46, 4342–4345 CrossRef CAS PubMed.
- A.-L. Morel, S. I. Nikitenko, K. Gionnet, A. Wattiaux, J. Lai-Kee-Him, C. Labrugere, B. Chevalier, G. Deleris, C. Petibois, A. Brisson and M. Simonoff, ACS Nano, 2008, 2, 847–856 CrossRef CAS PubMed.
- T. Lu, J. Wang, J. Yin, A. Wang, X. Wang and T. Zhang, Colloids Surf., A, 2013, 436, 675–683 CrossRef CAS.
- W. Wu, Q. He and C. Jiang, Nanoscale Res. Lett., 2008, 3, 397–415 CrossRef CAS PubMed.
- D. Wang and D. Astruc, Chem. Rev., 2014, 114, 6949–6985 CrossRef CAS PubMed.
- J. Deng, C. He, Y. Peng, J. Wang, X. Long, P. Li and A. S. C. Chan, Synth. Met., 2003, 139, 295–301 CrossRef CAS.
- M. Bhaumik, A. Maity, V. V. Srinivasu and M. S. Onyango, J. Hazard. Mater., 2011, 190, 381–390 CrossRef CAS PubMed.
- N. R. Jana, C. Earhart and J. Y. Ying, Chem. Mater., 2007, 19, 5074–5082 CrossRef CAS.
- S. I. C. J. Palma, M. Marciello, A. Carvalho, S. Veintemillas-Verdaguer, M. d. P. Morales and A. C. A. Roque, J. Colloid Interface Sci., 2015, 437, 147–155 CrossRef CAS PubMed.
- F. Fu and Q. Wang, J. Environ. Manage., 2011, 92, 407–418 CrossRef CAS PubMed.
- J.-S. Hu, L.-S. Zhong, W.-G. Song and L.-J. Wan, Adv. Mater., 2008, 20, 2977–2982 CrossRef CAS.
- H. Mittal and S. B. Mishra, Carbohydr. Polym., 2014, 101, 1255–1264 CrossRef CAS PubMed.
- Y.-G. Zhao, H.-Y. Shen, S.-D. Pan and M.-Q. Hu, J. Hazard. Mater., 2010, 182, 295–302 CrossRef CAS PubMed.
- C. Chen, P. Gunawan and R. Xu, J. Mater. Chem., 2011, 21, 1218–1225 RSC.
- Y. Liu, Z. Ren, Y. Wei, B. Jiang, S. Feng, L. Zhang, W. Zhang and H. Fu, J. Mater. Chem., 2010, 20, 4802–4808 RSC.
- S. Xuan, L. Hao, W. Jiang, X. Gong, Y. Hu and Z. Chen, Nanotechnology, 2007, 18, 035602 CrossRef PubMed.
- Y. Wang, S. Wang, H. Niu, Y. Ma, T. Zeng, Y. Cai and Z. Meng, J. Chromatogr. A, 2013, 1283, 20–26 CrossRef CAS PubMed.
- X. Sun and Y. Li, Angew. Chem., Int. Ed., 2004, 43, 597–601 CrossRef PubMed.
- Z. Wang, H. Guo, Y. Yu and N. He, J. Magn. Magn. Mater., 2006, 302, 397–404 CrossRef CAS.
- I. Martínez-Mera, M. E. Espinosa-Pesqueira, R. Pérez-Hernández and J. Arenas-Alatorre, Mater. Lett., 2007, 61, 4447–4451 CrossRef.
- S. Gao, Y. Shi, S. Zhang, K. Jiang, S. Yang, Z. Li and E. Takayama-Muromachi, J. Phys. Chem. C, 2008, 112, 10398–10401 CAS.
- Y. Lai, W. Yin, J. Liu, R. Xi and J. Zhan, Nanoscale Res. Lett., 2009, 5, 302–307 CrossRef PubMed.
- L. Zhou, H. Deng, J. Wan, J. Shi and T. Su, Appl. Surf. Sci., 2013, 283, 1024–1031 CrossRef CAS.
- H. Hosseini-Monfared, F. Parchegani and S. Alavi, J. Colloid Interface Sci., 2015, 437, 1–9 CrossRef CAS PubMed.
- Y.-P. Zhu, J. Li, T.-Y. Ma, Y.-P. Liu, G. Du and Z.-Y. Yuan, J. Mater. Chem. A, 2014, 2, 1093–1101 CAS.
- F.-H. Wang, W. Jiang, Y. Fang and C.-W. Cheng, Chem. Eng. J., 2015, 259, 827–836 CrossRef CAS.
- T. Hao, X. Rao, Z. Li, C. Niu, J. Wang and X. Su, J. Alloys Compd., 2014, 617, 76–80 CrossRef CAS.
- Y. Gao, D. Zhong, D. Zhang, X. Pu, X. Shao, C. Su, X. Yao and S. Li, J. Chem. Technol. Biotechnol., 2014, 89, 1859–1865 CrossRef CAS.
- Y. Li, S. Zhu, Q. Liu, Z. Chen, J. Gu, C. Zhu, T. Lu, D. Zhang and J. Ma, Water Res., 2013, 47, 4188–4197 CrossRef CAS PubMed.
- J. Hu, I. M. Lo and G. Chen, Water Sci. Technol., 2004, 50, 139–146 CAS.
- J. Wang, B. Tang, T. Tsuzuki, Q. Liu, X. Hou and L. Sun, Chem. Eng. J., 2012, 204–206, 258–263 CrossRef CAS.
- P. Yuan, D. Liu, M. Fan, D. Yang, R. Zhu, F. Ge, J. Zhu and H. He, J. Hazard. Mater., 2010, 173, 614–621 CrossRef CAS PubMed.
- K. Kadirvelu, C. Karthika, N. Vennilamani and S. Pattabhi, Chemosphere, 2005, 60, 1009–1017 CrossRef CAS PubMed.
- P. P. Selvam, S. Preethi, P. Basakaralingam, N. Thinakaran, A. Sivasamy and S. Sivanesan, J. Hazard. Mater., 2008, 155, 39–44 CrossRef CAS PubMed.
- Y. S. Ho and G. McKay, Process Biochem., 1999, 34, 451–465 CrossRef CAS.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra01408a |
|
| This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.