Hierarchical core–shell structured Fe3O4@NiSiO3 magnetic microspheres: preparation, characterization, and adsorption of methylene blue from aqueous solution

Senwen Yuanab and Lang Zhao*a
aState Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, P. R. China. E-mail: Zhaolang@ciac.ac.cn; Fax: +86-431-85262878
bSchool of Chemistry and Life Science, Changchun University of Technology, Changchun 130012, P. R. China

Received 14th January 2016 , Accepted 9th May 2016

First published on 11th May 2016


Abstract

Hierarchical core–shell structured Fe3O4@NiSiO3 magnetic microspheres have been synthesized using a modified Stöber sol–gel process and solvothermal methods. The prepared composites were investigated by SEM-EDS, TEM, XRD, FTIR, SQUID magnetometry, and nitrogen adsorption/desorption measurements. The microspheres possess high porosity and magnetic properties, which allow the Fe3O4@NiSiO3 microspheres to exhibit efficient adsorption capacity and convenient separation. The adsorption isotherm and adsorption kinetics for the adsorption of methylene blue (MB) on the Fe3O4@NiSiO3 microspheres were analyzed. The adsorption isotherm is well fitted by the Langmuir isotherm model, which is valid for monolayer adsorption of MB on the surface of Fe3O4@NiSiO3 magnetic microspheres. The pseudo-second-order model accurately describes the adsorption kinetics process for the adsorption of MB on the Fe3O4@NiSiO3 microspheres, suggesting a chemical adsorption process. MB removal of the Fe3O4@NiSiO3 magnetic microspheres can reach 87% after the fifth adsorption, indicating good regeneration capacity and reusability.


1 Introduction

In recent years, water pollution has been a very big problem because of the inappropriate treatment of household and industrial wastes before they are drained into rivers.1 Dye producing and consuming industries, such as printing, food, paint, plastics, and pharmaceutical, trigger a large volume of toxic wastewater.2 MB is one of the basic dyes, which is widely employed in coloring paper, dyeing cottons, wools, coating paper and some medical uses. Massive exposure to MB causes increased heart rate, vomiting, shock, Heinz body formation, cyanosis, jaundice, quadriplegia and tissue necrosis in humans.3,4 Due to the molecular structures of dyes becoming more complicated and stable, it is difficult to treat the wastewater. Hence, removal of the dye from wastewater is of utmost importance.

A great number of physical and chemical methods such as precipitation, coagulation,5 reverse osmosis, ozonation,6,7 filtration, adsorption,8–11 and advanced oxidation processes,12,13 etc. have used for removal of dyes. Among these methods, adsorption takes advantages of its ease of operation, high efficiency, low energy consumption, low operating cost and low sensitivity to toxic environments, generally considered as the preferred method for removing organic dyes from aqueous solution.14,15 With the rapid advance of nanotechnology, there has brought various nanomaterials in the respect of treatment of dyeing aqueous solution. Different morphologies of micro/nanostructures have been used as adsorbents for the adsorption of organic dyes in aqueous solution, which exhibit remarkably enhanced sorption capacity. Among the various morphologies, the fabrication of three-dimensional (3D) hierarchical inorganic micro/nanomaterials by using nanoparticles (0D), nanorods (1D), and nanoplates (2D) as building blocks, has attracted considerable attention due to their large surface area, abundant surface functional groups, developed porous structure, good chemical stabilities, and their potential applications in catalysis, energy conversion and storage, environmental abatement, and sensing.16 To date, lots of micro/nanomaterials have been used as adsorbents for water decontamination.15,17–23 However, these adsorbents still suffer from issues involving separation inconvenience from a large volume of water, which limits the application of them in wastewater treatment. Thus, the development of new micro/nanomaterials as adsorbents with a facile separation property is great interest.24–28

It is worth mentioning that iron oxide nanoparticles, in particular magnetite (Fe3O4) and maghemite (γ-Fe2O3), have been demonstrated a rapid response to the external magnetic field which can be extensively applied in designing nanocomposite adsorbents. Recently, considerable efforts have been devoted to the functionalization magnetic nanoparticles by coating them with other materials, such as inorganic materials, polymers, etc.29–33 Surface functionalization has been found to enhance the stability and sorption capacity of magnetic nanocomposites. Thus, it is highly essential to explore new functional composites with efficiently magnetic separation and absorptive capacity while exhibit a low-cost and environmentally benign nature.

Nickel-based materials have an important function, and are employed in various fields, such as catalysis,34 biomolecule separation,35 and adsorption36 because of its high chemical and thermal stability. As a nickel-based material, nickel silicate with high surface area could enhance the accessibility of adsorbates to reactive sites. In the past years, researches have been conducted to functionalize the nickel silicate porous shell with magnetic nanoparticles core to apply for the affinity purification of his-tagged proteins.35,37 To the best of our knowledge, there have been few reports on the application of Fe3O4@NiSiO3 particles in adsorption of dyes from aqueous solution. Herein, we report a combination of Fe3O4 with NiSiO3 for synthesis of hierarchical core–shell structured Fe3O4@NiSiO3 magnetic microspheres (Scheme 1). In the synthetic process, the SiO2 as linker shell was coated on the Fe3O4 by modified Stöber sol–gel process and then the SiO2 shell was broken by hydroxide ions into silicate ions which reacted with nickel–ammonia complexes to form NiSiO3 shell on the Fe3O4. The synthetic Fe3O4@NiSiO3 magnetic microspheres with hierarchical shell and magnetic core possess high porosity and magnetic property, which allow the components to exhibit adsorption property and separation property in aqueous solution.


image file: c6ra01142j-s1.tif
Scheme 1 Schematic illustration of synthesis of hierarchical core–shell structured Fe3O4@NiSiO3 magnetic microspheres.

2 Experimental

Materials

FeCl3·6H2O, NiCl2·6H2O, NH4Cl, ethylene glycol (EG) and aqueous ammonia (25–28%) were purchased from Beijing Chemical Reagent Co. Na3Cit and anhydrous NaOAc were purchased from Beijing Yili Fine Chemicals Co., Ltd. Tetraethyl orthosilicate (TEOS) was purchased from Sinopharm Chemical Reagent Co., Ltd. All of this chemicals were of analytical grade reagents and used directly without further purification.

Synthesis of Fe3O4 microspheres

Fe3O4 microspheres were prepared by the solvothermal method. FeCl3·6H2O (0.54 g) was first mixed with EG (10 mL) under magnetic stirring to form a clear solution. Then, Na3Cit (0.20 g) was added, stirring for a period of time to form a homogeneous solution, followed by the addition of NaOAc (0.90 g). The homogeneous mixture was transferred into a Teflon-lined autoclave and sealed to heat at 230 °C for 10 h. The synthesized mixture was then allowed to cool naturally to room temperature. The black products were washed with ethanol and distilled water for three times respectively. At last, the Fe3O4 microspheres were redispersed in ethanol for farther use.

Synthesis of Fe3O4@SiO2 core–shell microspheres

In a classical procedure, above-mentioned Fe3O4 microspheres were dispersed in the mixture solution of H2O (20 mL) and ethanol (136 mL). The mixture solution was sonicated for 15 min. After that, concentrated NH4OH solution (25 wt%, 6.0 mL) mixed with H2O (20 mL) was added dropwise to the solution under sonication. After sonication for about 15 min, TEOS (1.0 mL) dispersed in ethanol (24 mL) was then added drop by drop to the solution under ultrasonication. The resulting solution was sonicated for 90 min at room temperature. The obtained particles were washed with ethanol and distilled water for three times respectively. At last, the Fe3O4@SiO2 microspheres were redispersed in ethanol for farther use.

Synthesis of Fe3O4@NiSiO3 core–shell microspheres

Fe3O4@NiSiO3 microspheres were prepared by the solvothermal method. The synthetic Fe3O4@SiO2 (40 mg) dispersed in ethanol (40 mL) was under ultrasonication for 30 min. NiCl2·6H2O (266.6 mg) and NH4Cl (553.0 mg) were dissolved in a mixture solution containing deionized water (20 mL), ethanol (20 mL) and ammonia solution (2.5 mL, 28%). These two solutions were mixed with each other and transferred into a Teflon-lined autoclave (100 mL) and sealed to heat at 160 °C for 12 h. After the reaction, the autoclave was cooled to the room temperature. The aurantius precipitate (Fe3O4@NiSiO3 microspheres) was collected by centrifugation and washed with ethanol and distilled water for three times respectively and then dried in a vacuum at 40 °C overnight. In addition, Ni element was replaced by Co, Cu and Zn element respectively (Fig. S2). The procedure was similar to the synthesis of Fe3O4@NiSiO3 except that MCl2 (M = Co, Cu and Zn) was used in place of NiCl2.

Dye adsorption and adsorbent regeneration experiment

The desired amount of the Fe3O4@NiSiO3 microspheres as adsorbent in the suspension was mixed with the aqueous solutions of MB. After ultrasonication for a certain time, the Fe3O4@NiSiO3 adsorbent was separated by a magnet and the supernatant solution was analyzed with UV-vis spectrometer at 664 nm to determine the concentrations of dyes in the solution. To study the change of the solution pH before and after adsorption, we have measured the solution pH of the addition of the adsorbents and the solution pH before and after the adsorption of MB by the pH test paper (Fig. S3). The solution pH kept at 7 (Fig. S3) during the adsorption of MB, which indicated that there was little effect on the solution pH. To estimate the adsorption capacity, the initial concentrations of MB were varied in the range of 1–200 mg L−1, and the dosage of Fe3O4@NiSiO3 adsorbent was kept at 5 mg. The mixture was sonicated at room temperature (300 K) for a certain time and the concentration of MB in the final solution was measured with UV-vis spectroscopy after the separation of adsorbent by a magnet. At the initial MB concentration of 18.11 mg L−1, the adsorption kinetic of Fe3O4@NiSiO3 adsorbent was obtained by monitoring the concentration of MB at various time intervals. In order to research the regeneration of Fe3O4@NiSiO3 adsorbent, 5 mg of Fe3O4@NiSiO3 adsorbent was added to 5 mL of dye solution with a concentration of 18.11 mg L−1, and the mixture was carefully sonicated at ambient temperature for 210 min. After magnetic separation, the concentration of the supernatant residual dye solution was monitored with UV-vis spectrometer. The MB-adsorbed Fe3O4@NiSiO3 adsorbent was washed using ethylene glycol and ethanol for several times, which could realize the regeneration of the adsorbent. The regeneration experiment was successively repeated five times.

Measurements and characterizations

A scanning electron microscope (SEM, S4800, Hitachi) equipped with an energy-dispersive X-ray spectrum was applied to determine the morphology and composition of the as-prepared samples. The SEM samples were prepared by depositing a dilute aqueous dispersion of the as-prepared samples onto a silicon wafer. Transmission electron microscope (TEM) measurements were carried out on a JEOL JEM-2010EX transmission electron microscope with a tungsten filament at an accelerating voltage of 200 kV. The samples were prepared by placing a drop of prepared solution on the surface of a copper grid and dried at room temperature. X-ray powder diffraction (XRD) measurements were performed on a D8 Focus diffractometer (Bruker) at a scanning rate of 2.5° min−1 in the 2θ range from 20° to 70°, with using of Cu-Kα radiation (λ = 0.15405 nm). FTIR spectra was measured using a Nicolet 6700 IR Fourier spectrometer equipped with smart iTR FTIR attachment in the range from 500 to 4000 cm−1. The magnetization curves of the products were recorded on a Quantum Design MPMS-XL7 superconducting quantum interference device (SQUID) magnetometer at 300 K. N2 adsorption–desorption isotherms were recorded on a Micromeritics ASAP 2020M automated sorption analyzer. UV-vis spectroscopy was carried out with a JASCO V-550 UV-vis spectrometer.

3 Results and discussion

The size and morphologies of the three as-synthesized samples were investigated by SEM and TEM (Fig. 1). Fig. 1a shows the SEM images of the highly water-dispersible Fe3O4 microspheres which synthesized by one-step solvothermal strategy. It can be seen that the Fe3O4 microspheres with rough surface have a relatively uniform diameter of about 350 nm. A typical TEM image (Fig. 1d) of the Fe3O4 microsphere shows that the microspheres are composed of a large quantity of small nanoparticles of about 10–15 nm, which could explain why the surface of Fe3O4 is rough. After coating the Fe3O4 microsphere with SiO2 via a typical modified Stöber sol–gel reaction, a SiO2-shell–Fe3O4-core structure was obtained, whose surface is much smoother than Fe3O4 (Fig. 1b). The Fe3O4 cores are black spheres, the silica shells of Fe3O4@SiO2 are a gray color with an average thickness of about 100 nm (Fig. 1e). The SEM image in Fig. 1c shows that the Fe3O4@NiSiO3 microspheres present a flowerlike hierarchical surface with a relatively average diameter of about 550 nm. The detailed morphologies of the as-synthesized Fe3O4@NiSiO3 microspheres are shown in Fig. 1f and S1, which reveal each NiSiO3 nanoflake perpendicularly grafted to the solid core so that the external surface of the microspheres is composed of the edges of the flakes. The dark regions represent the Fe3O4 microsphere cores and the bright regions represent the NiSiO3 nanoflake shells, which clearly displays that the Fe3O4 core is well wrapped by the coating NiSiO3 flowerlike nanoflake layers.
image file: c6ra01142j-f1.tif
Fig. 1 SEM image of (a) Fe3O4, (b) Fe3O4@SiO2, (c) Fe3O4@NiSiO3; TEM image of (d) Fe3O4, (e) Fe3O4@SiO2, (f) Fe3O4@NiSiO3.

A possible formation process of hierarchical NiSiO3 shell is due to the fact that the Si–O bonds were broken to form silicate ions under an alkaline condition, and the generated silicate ions continually reacted with nickel–ammonia complexes to form NiSiO3 nanoparticles under high temperature. Then these nanoparticles grow along the 2D direction, thereby resulting in the formation of nanoflakes. Third, the nanoflakes grow until all the nanoparticles are consumed, accompanied by their self-organization into the flowerlike hierarchical structure on the surfaces of Fe3O4 microspheres. Finally, the flowerlike Fe3O4@NiSiO3 hierarchical microspheres are obtained. In the experiments, the morphology of the as-synthesized NiSiO3 was a porous structure composed of flowerlike hierarchical structure. When Ni element was replaced by Co and Cu, there was only nanoparticles heterogeneous coating on the Fe3O4 microspheres, which was absent of porous flowerlike hierarchical structure (Fig. S2A and B). Furthermore, when Ni element was replaced by Zn, there was countable nanoflake perpendicularly grafted to the Fe3O4 microspheres, which was far from porous structure (Fig. S2C). The results indicate that the Ni element plays an important role in the formation of the flowerlike hierarchical structures.

In order to prevent the deposition of nickel hydrate during the process, NH3–NH4Cl buffer system was introduced to maintain the pH value throughout the experiments, which would be further confirmed by the XRD analysis later in the paper. The above results indicate that the synthetic composites are hierarchical core–shell structured microspheres.

The core–shell structure and the composition of the composites were father ascertained by using the energy dispersive X-ray (EDS) analyses. The EDS spectrum confirms the presence of Fe, Ni, Si, and O in the as-prepared Fe3O4@NiSiO3 microspheres (Fig. 2a). The sample was on the silicon wafer during the EDS analyses, which resulted in the strong peak of Si in the EDS spectrum. To further investigate their microstructure, elemental mapping is employed to investigate the elemental distributions in the core–shell structure. EDS mapping images (Fig. 2c–f) correspond to the elemental distribution of Fe, O, and Ni. The Fe element stays in the core region, and the Ni element is detected in the shell region, while the O element can be observed in both regions. The EDS line scanning data (Fig. 2b) is also characteristic of a core–shell type structure. Fe element as a component of the core part is dominant at the center of microsphere, and Ni element as a component of the microsphere shell is abundant at the edge of the microsphere. These results further confirm the core–shell structure of Fe3O4@NiSiO3 microspheres.


image file: c6ra01142j-f2.tif
Fig. 2 Energy-dispersive X-ray spectroscopy (EDS) spectrum (a), line scanning data (b) and mapping images(c–f): (d) Fe element, (e) Ni element, (f) O element of Fe3O4@NiSiO3 microsphere.

Fig. 3A shows the X-ray diffraction patterns of Fe3O4, Fe3O4@SiO2, and Fe3O4@ NiSiO3. The sharp diffraction peaks at 30.2°, 35.5°, 43.5°, 53.5°, 57.3°, and 62.8° are indexed as the (220), (311), (400), (422), (511), and (440) lattice planes of standard Fe3O4 (JCPDS no. 19-0629).38,39 No other peaks are observed in the pattern of Fe3O4, indicating that the sample is pure Fe3O4 crystalline phase. The reflection characteristic of amorphous SiO2 is appeared in the pattern of Fe3O4@SiO2. The presence of diffraction peaks at 24.6°, 34.6°, 40.6°, 53.7°, and 60.5° in the pattern of Fe3O4@NiSiO3 correspond to the (103), (110), (200), (210), (300) planes, respectively, of nickel silicate crystal (JCPDS 43-0664).40,41 Furthermore, no other peaks for other phases can be detected in the pattern of Fe3O4@NiSiO3, indicating that no other reaction occurred during the hierarchical NiSiO3 shell formation process.


image file: c6ra01142j-f3.tif
Fig. 3 (A) XRD patterns of (a) Fe3O4, (b) Fe3O4@SiO2, and (c) Fe3O4@NiSiO3; (B) FTIR spectra of (a) Fe3O4, (b) Fe3O4@SiO2, and (c) Fe3O4@NiSiO3.

Fig. 3B presents the FTIR spectra of Fe3O4, Fe3O4@SiO2, and Fe3O4@NiSiO3 nanocomposites. The absorption peak at 546 cm−1 corresponded to the Fe–O vibration related to the magnetite phase (Fig. 3Ba). The absorption bands at 1552 cm−1 and 1381 cm−1 are assigned to the stretching vibrations of carboxyl salt, which is because of that Na3Cit was added in the synthesis of Fe3O4. Compared with Fe3O4, the Fig. 3Bb presents an intense adsorption peak at 1066 cm−1 and two weak peaks at 953 cm−1 and 796 cm−1, which could be ascribed to the vibrations of Si–O–Si, Si–OH, and Si–O groups in the SiO2 shell. In Fig. 3Bc, two adsorption peaks at 953 cm−1 and 796 cm−1 disappear, and the peak at 1066 cm−1 shifts to 980 cm−1, which might be ascribed to the formation of Si–O–Ni bonds.41 Based on all above characterization results, the hierarchical core–shell microspheres with magnetic Fe3O4 cores and hierarchical NiSiO3 shells have been successfully synthesized via a facile and simple approach.

Fig. 4A shown that the magnetic hysteresis loops of Fe3O4, Fe3O4@SiO2, and Fe3O4@NiSiO3 at room temperature (T = 300 K) and the saturation magnetization values are 70.5 emu g−1, 38.4 emu g−1, and 21.0 emu g−1, respectively. All the samples have strong magnetism with negligible coercivity and remanence at room temperature. No hysteresis loop can be observed, which shows a superparamagnetic characteristic. The saturation magnetization of Fe3O4@NiSiO3 decreases evidently compared with Fe3O4. However, the value is strong enough to achieve a facile magnetic separation. Fe3O4@NiSiO3 could be fast aggregated by an external magnetic field from their homogeneous dispersion. After removing of the magnetic field, the redispersion of the nanocomposites occurred quickly with a slight shaking (inset of Fig. 4A). The magnetic property is essentially important for the separation and reuse of the synthetic Fe3O4@NiSiO3 microspheres as an adsorbent in aqueous solution.


image file: c6ra01142j-f4.tif
Fig. 4 (A) Room-temperature (300 K) magnetic hysteresis loops of (a) Fe3O4, (b) Fe3O4@SiO2, (c) Fe3O4@NiSiO3 and photograph of magnetic separation and redispersion process (inset) of Fe3O4@NiSiO3 in water solution; (B) N2 sorption isotherms and pore size distribution (inset) of Fe3O4@NiSiO3.

The N2 adsorption and desorption analysis was then introduced to investigate the specific surface area and porosity of the core–shell structured Fe3O4@NiSiO3 microspheres. The typical N2 adsorption–desorption isotherm for the Fe3O4@NiSiO3 microspheres and the corresponding pore size distribution are shown in Fig. 4B. This isotherm can be categorized as type IV with a hysteresis loop observed at a relative pressure of 0.01–1.0. At low relative pressure (P/P0 < 0.45), the adsorption and desorption curve coincide because of reversible monolayer adsorption. At a higher relative pressure region (0.45 < P/P0 < 1.0), the isotherm has significant hysteresis, which can be ascribed to the presence of a mesoporous structure in the interleaving nanoplates.42,43 The BET surface area and the total pore volume of the as-prepared Fe3O4@NiSiO3 microspheres are calculated to be 121.7 m2 g−1 and 0.23 cm3 g−1, respectively. The pore-size distribution plot (inset of Fig. 4B) confirms that the Fe3O4@NiSiO3 microspheres have well-developed mesopore with a diameter of 3.73 nm. The large specific surface area and high porosity of the core–shell structured Fe3O4@NiSiO3 microspheres make them very promising candidates for the adsorption of dyes in aqueous solution.

Nanomaterial adsorbents showed higher adsorption properties than bulk materials because of the nanoscale effects, which may provide an efficient way to adsorb of dyes from aqueous solution. Porous materials with high specific surface area are usually employed for adsorption of dyes from aqueous solution because of their capability to adsorb a large quantity of pollutants, and high efficiency in degrading the unwanted species. The as-prepared hierarchical core–shell structured Fe3O4@NiSiO3 microspheres with high surface area and large total pore volume can be used as adsorbent to remove organic dyes with a simple and rapid magnetic separation via a magnetic field. The adsorption of MB for Fe3O4@NiSiO3 microspheres was studied by UV-vis spectrophotometer at 664 nm. UV-vis absorption spectroscopy measurements were performed to determine the concentration of MB before and after adsorption experiments.

In the past years, there has brought various materials for adsorption of MB from aqueous solution.15,44

The adsorption isotherm is an effective method of investigating the adsorption ability of the adsorbent and understanding the interactions between the adsorbate and adsorbent. Fig. 5(a) shows the adsorption isotherm of MB on the Fe3O4@NiSiO3 microspheres. The amount of adsorbed MB dramatically increased at a lower final solution concentration, suggesting a high affinity between MB molecules and the Fe3O4@NiSiO3 microspheres surface. The adsorbed amount then reached a plateau at a higher equilibrium solution concentration (18.11 mg g−1), reflecting the saturated adsorption. Although the adsorption capacity of the Fe3O4@NiSiO3 is not extremely high compared with those materials obtained in the reported literature15,21,44,45 for the absorption of dye in aqueous solution, the magnetic absorber Fe3O4@NiSiO3 could be separated conveniently from a large volume of aqueous solution which could avoid the secondary pollution. In this study, Langmuir, Freundlich and Temkin isotherm models were used to describe the equilibrium adsorption. The fitting isotherms and corresponding parameters are shown in Fig. 5(b–d) and summarized in Table S1. The correlation coefficients (R2) indicate that the Langmuir isotherm model fits the experimental data best (RL2 = 0.9924), showing that MB adsorption on Fe3O4@NiSiO3 microspheres could be described by the Langmuir isotherm model, in accordance with monolayer adsorption. Also, the maximum adsorption capacity of Langmuir isotherm model is 19.14 mg g−1 (Table S1), which is close to the experimental value (18.11 mg g−1) than other two isotherm models. The Langmuir model is valid for monolayer adsorption on a surface containing a finite number of identical sites. There are no interactions among the adsorbate molecules and no further adsorption can take place at the sites that have been occupied by dye molecules.46 The result demonstrates the homogeneous nature of the sample surface and the formation of monolayer coverage of MB molecules on the adsorbent.


image file: c6ra01142j-f5.tif
Fig. 5 (a) Adsorption isotherm, (b) Langmuir, (c) Freundlich, (d) Temkin isotherm models for the adsorption of MB on the Fe3O4@NiSiO3 microspheres.

The adsorption kinetic behavior of the as-prepared Fe3O4@NiSiO3 microspheres for MB adsorption from aqueous solutions was investigated, which provided important data for understanding the dynamic of sorption reaction. As shown in Fig. 6, the composites exhibited a continuous adsorption process, with equilibrium time of approximately 210 min for MB adsorption. The pseudo-first-order and pseudo-second-order reaction rate equations are the most commonly applied models to examine the adsorption mechanism based on the empirical data.


image file: c6ra01142j-f6.tif
Fig. 6 Adsorption kinetic curve for the adsorption of MB on the Fe3O4@NiSiO3 microspheres.

The pseudo-first-order model:

image file: c6ra01142j-t1.tif

The pseudo-second-order model:

image file: c6ra01142j-t2.tif
where qe and qt represent the amount of dye adsorbed (mg g−1) at equilibrium and time t, respectively, and the k1 (min−1) or k2 (g mg−1 min−1) values are the kinetic rate constants.47,48

The Fig. 7(a) and (b) illustrate the straight line plot of ln(qeqt) against time for the pseudo-first order reaction, as well as the t/qt plot against time for the pseudo-second-order reaction in the adsorption of MB dye onto Fe3O4@NiSiO3 microspheres. The kinetic parameters (Table S2) indicate that the pseudo-second-order model with a high correlation coefficient (R22 = 0.9996) seemed better to accurately describe present adsorption process than the pseudo-first-order model. At the same time, the theoretical qe value, estimated from the pseudo-second order kinetic model, is very close to the experimental qe value. Therefore, the adsorption kinetic follows the pseudo-second-order model, suggesting a chemical adsorption process,49 which is as the rate-limiting process.50 The overall rate of adsorption can be described by the following three steps: (1) surface diffusion where the sorbate is transported from the bulk solution to the external surface of sorbent, (2) intraparticle or pore diffusion, where sorbate molecules move into the interior of sorbent particles, and (3) adsorption on the interior sites of the sorbent.


image file: c6ra01142j-f7.tif
Fig. 7 (a) pseudo-first-order kinetic plot and (b) pseudo-second-order kinetic plot for the adsorption of MB on the Fe3O4@NiSiO3 microspheres.

The regeneration ability of the adsorbent is crucial for its practical application and economic necessity. Fig. 8 shows the correlation between the adsorption efficiency of MB and the cycle number, confirming that the MB removal can reach 87% after the fifth adsorption, indicating good regeneration capacity and reusability. The decrease in the adsorption capacity was attributed to the incomplete desorption of dyes from the surface of Fe3O4@NiSiO3 microspheres.


image file: c6ra01142j-f8.tif
Fig. 8 The percentage MB removal by adsorption using the Fe3O4@NiSiO3 microspheres in recycle runs.

All in all, the prepared hierarchical core–shell structured Fe3O4@NiSiO3 magnetic microspheres as an adsorbent can be easily recycled and reused several times, which could make the adsorbent long-term use in adsorption of MB from aqueous solution.

4 Conclusions

In conclusion, hierarchical core–shell structured Fe3O4@NiSiO3 magnetic microspheres have been synthesized by modified Stöber sol–gel process and solvothermal method. The outer hierarchical NiSiO3 shell could support the Fe3O4@NiSiO3 microspheres exhibit excellent performance in the adsorption of MB, and the magnetic Fe3O4 core allows the composites to be conveniently separated by using a magnet. The adsorption isotherm demonstrates the monolayer adsorption of MB on the surface of Fe3O4@NiSiO3 magnetic microspheres and the adsorption kinetic shows a chemical adsorption process. In addition, the removal capacities of Fe3O4@NiSiO3 microspheres for MB were maintained well after several cycles, which could make the adsorbent long-term use in adsorption of MB from aqueous solution.

Acknowledgements

This work was financially supported the National Natural Science Foundation of China (Grants 21201160).

Notes and references

  1. Y.-F. Lin, H.-W. Chen, P.-S. Chien, C.-S. Chiou and C.-C. Liu, J. Hazard. Mater., 2011, 185, 1124–1130 CrossRef CAS PubMed.
  2. S. A. Saad, K. M. Isa and R. Bahari, Desalination, 2010, 264, 123–128 CrossRef CAS.
  3. V. Vadivelan and K. V. Kumar, J. Colloid Interface Sci., 2005, 286, 90–100 CrossRef CAS PubMed.
  4. K. V. Kumar, V. Ramamurthi and S. Sivanesan, J. Colloid Interface Sci., 2005, 284, 14–21 CrossRef CAS PubMed.
  5. J.-W. Lee, S.-P. Choi, R. Thiruvenkatachari, W.-G. Shim and H. Moon, Water Res., 2006, 40, 435–444 CrossRef CAS PubMed.
  6. M. Muthukumar, D. Sargunamani, N. Selvakumar and J. Venkata Rao, Dyes Pigm., 2004, 63, 127–134 CrossRef CAS.
  7. H. Selcuk, Dyes Pigm., 2005, 64, 217–222 CrossRef CAS.
  8. V. Janaki, K. Vijayaraghavan, B.-T. Oh, K.-J. Lee, K. Muthuchelian, A. K. Ramasamy and S. Kamala-Kannan, Carbohydr. Polym., 2012, 90, 1437–1444 CrossRef CAS PubMed.
  9. S. Chatterjee, M. W. Lee and S. H. Woo, Bioresour. Technol., 2010, 101, 1800–1806 CrossRef CAS PubMed.
  10. L. S. Silva, L. C. B. Lima, F. C. Silva, J. M. E. Matos, M. R. M. C. Santos, L. S. Santos Júnior, K. S. Sousa and E. C. da Silva Filho, Chem. Eng. J., 2013, 218, 89–98 CrossRef CAS.
  11. J. J. Lunhong Ai and J. Tang, Chin. J. Appl. Chem., 2010, 6, 710–715 Search PubMed.
  12. T.-C. Hsu, Fuel, 2008, 87, 3040–3045 CrossRef CAS.
  13. Q. F. Alsalhy, T. M. Albyati and M. A. Zablouk, Chem. Eng. Commun., 2013, 200, 1–19 CrossRef CAS.
  14. T. M. Al-Bayati, Part. Sci. Technol., 2014, 32, 616–623 CrossRef CAS.
  15. J. Gong, J. Liu, X. Chen, Z. Jiang, X. Wen, E. Mijowska and T. Tang, J. Mater. Chem. A, 2015, 3, 341–351 CAS.
  16. Q. Hao, S. Liu, X. Yin, Z. Du, M. Zhang, L. Li, Y. Wang, T. Wang and Q. Li, CrystEngComm, 2011, 13, 806–812 RSC.
  17. R. Rostamian, M. Najafi and A. A. Rafati, Chem. Eng. J., 2011, 171, 1004–1011 CrossRef CAS.
  18. Z. Gan, A. Zhao, Q. Gao, M. Zhang, D. Wang, H. Guo, W. Tao, D. Li, E. Liu and R. Mao, RSC Adv., 2012, 2, 8681–8688 RSC.
  19. S. Mahdavi, M. Jalali and A. Afkhami, Chem. Eng. Commun., 2013, 200, 448–470 CrossRef CAS.
  20. X. Wang, W. Cai, G. Wang, Z. Wu and H. Zhao, CrystEngComm, 2013, 15, 2956–2965 RSC.
  21. J. Gong, J. Liu, Z. Jiang, X. Wen, E. Mijowska, T. Tang and X. Chen, J. Colloid Interface Sci., 2015, 445, 195–204 CrossRef CAS PubMed.
  22. F. L. Yuli Yin, W. Rao, Z. h. Zhang and L. Yan, Chin. J. Appl. Chem., 2015, 32, 472–480 Search PubMed.
  23. W. Z. Zhongzheng Wu, D. Wang, L. Cun and Z. Wu, Chin. J. Appl. Chem., 2014, 9, 1089–1095 Search PubMed.
  24. M. N. Nadagouda, C. Bennett-Stamper, C. White and D. Lytle, RSC Adv., 2012, 2, 4198–4204 RSC.
  25. V. Yathindranath, M. Worden, Z. Sun, D. W. Miller and T. Hegmann, RSC Adv., 2013, 3, 23722–23729 RSC.
  26. J. C. Xu, P. H. Xin, Y. B. Han, P. F. Wang, H. X. Jin, D. F. Jin, X. L. Peng, B. Hong, J. Li, H. L. Ge, Z. W. Zhu and X. Q. Wang, J. Alloys Compd., 2014, 617, 622–626 CrossRef CAS.
  27. D. Guo, W. Ren, Z. Chen, M. Mao, Q. Li and T. Wang, RSC Adv., 2015, 5, 10681–10687 RSC.
  28. H. Y. Zhu, R. Jiang, S. H. Huang, J. Yao, F. Q. Fu and J. B. Li, Ceram. Int., 2015, 41, 11625–11631 CrossRef CAS.
  29. C. Chen, P. Gunawan and R. Xu, J. Mater. Chem., 2011, 21, 1218–1225 RSC.
  30. K. Mandel, F. Hutter, C. Gellermann and G. Sextl, ACS Appl. Mater. Interfaces, 2012, 4, 5633–5642 CAS.
  31. A. Farrukh, A. Akram, A. Ghaffar, S. Hanif, A. Hamid, H. Duran and B. Yameen, ACS Appl. Mater. Interfaces, 2013, 5, 3784–3793 CAS.
  32. X. Zhang, M. Lin, X. Lin, C. Zhang, H. Wei, H. Zhang and B. Yang, ACS Appl. Mater. Interfaces, 2014, 6, 450–458 CAS.
  33. Y. Xu, J. Jin, X. Li, Y. Han, H. Meng, C. Song and X. Zhang, Microchim. Acta, 2015, 182, 2313–2320 CrossRef CAS.
  34. X. Song and L. Gao, J. Phys. Chem. C, 2008, 112, 15299–15305 CAS.
  35. Z. Liu, M. Li, F. Pu, J. Ren, X. Yang and X. Qu, J. Mater. Chem., 2012, 22, 2935–2942 RSC.
  36. L. Tan, X. Zhang, Q. Liu, J. Wang, Y. Sun, X. Jing, J. Liu, D. Song and L. Liu, Dalton Trans., 2015, 44, 6909–6917 RSC.
  37. Y. Wang, G. Wang, Y. Xiao, Y. Yang and R. Tang, ACS Appl. Mater. Interfaces, 2014, 6, 19092–19099 CAS.
  38. Y. Wang, Y. Shen, A. Xie, S. Li, X. Wang and Y. Cai, J. Phys. Chem. C, 2010, 114, 4297–4301 CAS.
  39. X. Hou, W. Zhang, X. Wang, S. Hu and C. Li, Science Bulletin, 2015, 60, 884–891 CrossRef CAS.
  40. Z. Liu, M. Li, X. Yang, M. Yin, J. Ren and X. Qu, Biomaterials, 2011, 32, 4683–4690 CrossRef CAS PubMed.
  41. Y. Wu, G. Chang, Y. Zhao and Y. Zhang, Dalton Trans., 2014, 43, 779–783 RSC.
  42. Y. Meng, D. Chen and X. Jiao, J. Phys. Chem. B, 2006, 110, 15212–15217 CrossRef CAS PubMed.
  43. Y. Tao, H. Kanoh, L. Abrams and K. Kaneko, Chem. Rev., 2006, 106, 896–910 CrossRef CAS PubMed.
  44. Y. Wang, G. Wang, H. Wang, C. Liang, W. Cai and L. Zhang, Chem.–Eur. J., 2010, 16, 3497–3503 CrossRef CAS PubMed.
  45. Y. Lin, Z. Zeng, J. Zhu, Y. Wei, S. Chen, X. Yuan and L. Liu, Mater. Lett., 2015, 156, 169–172 CrossRef CAS.
  46. I. Langmuir, J. Am. Chem. Soc., 1918, 40, 1361–1403 CrossRef CAS.
  47. Q. Zhang, Z. Zhang, J. Teng, H. Huang, Q. Peng, T. Jiao, L. Hou and B. Li, Ind. Eng. Chem. Res., 2015, 54, 2940–2949 CrossRef CAS.
  48. R. Chen, J. Yu and W. Xiao, J. Mater. Chem. A, 2013, 1, 11682–11690 CAS.
  49. T. Lu, T. Xiang, X.-L. Huang, C. Li, W.-F. Zhao, Q. Zhang and C.-S. Zhao, Carbohydr. Polym., 2015, 133, 587–595 CrossRef CAS PubMed.
  50. H. K. Boparai, M. Joseph and D. M. O'Carroll, J. Hazard. Mater., 2011, 186, 458–465 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: TEM image of Fe3O4@NiSiO3 magnetic microspheres in Fig. S1. SEM images after replaced Ni element by Co (A), Cu (B) and Zn (C) element in Fig. S2. The solution pH of the addition of the adsorbents, (A) before (B) and after the adsorption of MB (C) in Fig. S3. Langmuir, Freundlich and Temkin model parameters for adsorption of MB onto the Fe3O4@NiSiO3 microspheres in Table S1. Kinetic adsorption parameters for MB adsorbed onto the Fe3O4@NiSiO3 microspheres obtained by using pseudo-first-order and pseudo-second-order models in Table S2. See DOI: 10.1039/c6ra01142j

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.