Xiaoqin Shena,
Xiaolei Chenb,
Dejun Sunb,
Tao Wu*b and
Yujiang Li*a
aShandong Provincial Research Center for Water Pollution Control, School of Environmental Science & Engineering, Shandong University, Jinan, 250100, PR China. E-mail: yujiang@sdu.edu.cn; Fax: +86-531-88363358; Tel: +86-531-88363358
bKey Laboratory of Colloid & Interface Science of Education Ministry, Shandong University, Jinan, 250100, PR China. E-mail: wutao@sdu.edu.cn; Fax: +86-531-88365437; Tel: +86-531-88365437
First published on 15th February 2018
A magnetic hybrid nanomaterial, which contains magnetite (Fe3O4) particles and diazonium functionalized-reduced graphene oxide (DF-RGO), was fabricated via a three-pot reaction. First, the reduced graphene oxide (RGO) was synthesized via a redox reaction. Second, diazonium functionalized-RGO was prepared via a feasible chemical reaction. Third, Fe3O4 particles were loaded onto the surface of DF-RGO by covalent bonding, fabricating the M-DF-RGO hybrid. The fabricated hybrid was characterized by SEM, TEM, AFM, XRD, XPS, FT-IR, TGA, Raman spectroscopy, and magnetometry. The resulting M-DF-RGO hybrid possessed unique magnetic properties and was applied to remove 4-chlorophenol (4-CP) and 2,4-dichlorophenol (2,4-DCP) from aqueous solution. The adsorption of 4-CP and 2,4-DCP on the M-DF-RGO hybrid was performed under various conditions, with respect to initial chlorophenol concentration, pH, and contact time. The results suggest that the adsorption of 4-CP and 2,4-DCP onto the M-DF-RGO hybrid is strongly dependent on pH and weakly dependent on contact time. In addition, the adsorption isotherm of 4-CP and 2,4-DCP on the M-DF-RGO hybrid fits the Freundlich model well and the adsorption capacities of 4-CP and 2,4-DCP on M-DF-RGO reached 55.09 and 127.33 mg g−1, respectively, at pH 6 and 25 °C. In this situation, intermolecular interactions including π–π interactions and hydrogen bonding are operative. The calculated results of density functional theory further demonstrate that 2,4-DCP molecules could be more easily absorbed than 4-CP molecules by the M-DF-RGO hybrid. Moreover, the M-DF-RGO hybrid could be easily separated by a magnetic separation process, and showed good recyclability of more than five cycles.
To date, various techniques such as biological degradation, chemical oxidation, photodegradation, UV/ozone oxidation, and adsorption have been researched for the treatment of phenolic wastewater at high concentration.6–8 Of all of the known technologies, adsorption is considered to be the most effective, simplest, lowest-costing, and most frequently used method for the removal of chlorophenols or phenolic compounds. Various adsorbent materials such as clay minerals, metal hydroxides, resin, activated carbon and polymeric adsorbents have been applied for the removal of chlorophenols or phenolic compounds from aqueous solution.9–12 However, these traditional adsorbent materials still have certain problems that limit their practical applications, such as low selectivity, low adsorption capacity, poor dispersion stability, and difficulty in regeneration of the adsorbent material. As a result, the design and fabrication of new adsorbent materials to solve these problems are of significant importance.
Graphene is an excellent two-dimensional (2D) material, which possesses structural units that are similar to benzene rings that connect to each other, forming a 2D planar structure.13 Its attractive properties include high surface area, high electron mobility at room temperature, tunable optical properties, good thermal conductivity, and excellent mechanical strength.13,14 As a result, graphene is applied in diverse fields, such as nanoelectronics, optical ultrasensitive sensors, field-effect transistors, and energy storage and conversion.14–17 However, graphene has poor dispersivity and process stability. Graphene sheets tend to aggregate in the liquid phase due to van der Waals forces between graphene sheets, and this aggregation restricts the applications. Hence, covalent modification of graphene has been used to accomplish improvements of the physicochemical properties of graphene. Previous studies have demonstrated that graphene can be covalently functionalized by various direct or indirect functionalization methods.16–20 Either directly or indirectly covalently functionalized graphene is generated from graphene oxide (GO) as the precursor. GO, one of the most important derivatives of graphene, has a variety of oxygen-containing functional groups, such as hydroxyl, carbonyl, epoxide, and carboxyl groups, on its basal planes and edges. Taking into account these abundant oxygen-containing functional groups, GO can not only be well dispersed in water to form a stable colloidal dispersion, but also can potentially be used as nano-scale 2D building blocks for the fabrication of graphene-based composites.21,22 Among these oxygen-containing functional groups, the carboxyl groups are mainly located at the edges of the graphene sheets, and are excellent precursors for functionalizing various graphene-based composites without distorting the graphene plane.23 Therefore, the carboxyl group plays an important role in fabricating various new composites. However, other oxygen-containing functional groups (e.g. epoxide and carbonyl groups) on the planar surface of GO may disturb the covalent modification process. In this case, GO can be selectively reduced by thiourea dioxide, leaving the desired carboxyl groups and removing the undesired epoxide and carbonyl groups. This type of selectively reduced graphene is called carboxyl graphene or reduced graphene oxide (RGO). RGO can be used as a precursor and reacts with the functional groups of various organic molecules to introduce new functional groups for designing and fabricating new adsorbent materials. To the best of our knowledge, a magnetic diazonium functionalized-reduced graphene oxide (M-DF-RGO) hybrid has not been synthesized and applied in the removal of 4-CP and 2,4-DCP from aqueous solution.
In this work, the M-DF-RGO hybrid was synthesized via a three-pot reaction and used as an adsorbent to remove 4-CP and 2,4-DCP from aqueous solution. The fabricated M-DF-RGO hybrid was characterized using scanning electron microscopy, transmission electron microscopy, atomic force microscopy, X-ray diffraction, X-ray photoelectron spectroscopy, Fourier transform infrared spectroscopy, thermogravimetric analysis, Raman spectroscopy, and magnetometry. 4-CP and 2,4-DCP were chosen as model pollutants to describe the adsorption behavior of chlorophenol molecules on the M-DF-RGO hybrid, and to elucidate the interactions between chlorophenol molecules and the M-DF-RGO hybrid. The effects of initial chlorophenol concentration, pH, and contact time on the adsorption capacity were investigated. At the same time, the interactions between chlorophenol molecules and the M-DF-RGO hybrid were also analyzed using density functional theory. Finally, the reusability of the M-DF-RGO hybrid for 4-CP and 2,4-DCP was explored.
Before the aryl diazonium salt solution was added into the RGO colloidal dispersion, the pH value of the RGO colloidal dispersion was adjusted to 6 by the addition of dilute HCl solution in an ice bath. During the adjustment of the pH value from basic to acidic, the RGO colloidal dispersion (1 g L−1) precipitated and produced brown flocculent precipitates. Under continuous stirring, the aryl diazonium salt solution was immediately added dropwise to the RGO flocculent precipitates. Then, the mixture was kept in the ice bath without stirring for 2 h in order to adsorb sufficient aryl diazonium ions onto the RGO sheets. After 2 h, the mixture was heated to 80 °C for 10 h, and then cooled to room temperature. A homogeneous black dispersion of diazonium salt functionalized-RGO sheets was obtained. The resulting product was filtered and washed with Milli-Q water and ethanol. Finally, the DF-RGO sample was obtained by freeze-drying for 24 h.
Recycling tests were carried out to assess the recyclability of the M-DF-RGO hybrid. After adsorption experiments, the adsorbent particles were collected using a magnet. The 4-CP- and 2,4-DCP-adsorbed adsorbent particles were washed five times with 0.01 mol L−1 NaOH solution and ethanol, respectively, and vacuum-dried at 60 °C for 4 h. The recovered adsorbent particles were again used for subsequent adsorption experiments. Five cycles were performed to evaluate the reusability of the M-DF-RGO hybrid.
Fig. S4† shows the molecular models of DF-RGO, and the added p-carboxyphenyl groups have been attached to the surface of the graphene sheets. It should be noted that in the conditions of our calculations, we intentionally neglect Fe3O4 particles, which did not work for the adsorption of chlorophenols (Fig. S5†). The adsorption energy (Ead) was calculated as Ead = EG + EX − EG–X, where EG corresponds to the total energies of DF-RGO; EX1 corresponds to the total energies of the 4-CP system in isolated form; EX2 corresponds to the total energies of the 2,4-DCP system in isolated form; EG–X1 is the total energy of the DF-RGO and 4-CP complex system; and EG–X2 is the total energy of the DF-RGO and 2,4-DCP complex system. A positive value of Ead indicates that the adsorption is exothermic, and a higher positive value of Ead corresponds to a stronger adsorption capacity of the adsorbate onto the DF-RGO hybrid.
Fig. 1 TEM images of (a) GO, (b) RGO, (c) DF-RGO, and (d) M-DF-RGO; and SEM images of (e) GO, (f) RGO, (g) DF-RGO, and (h) M-DF-RGO. |
AFM images confirmed that evaporated dispersions of GO, RGO, DF-RGO, and M-DF-RGO are composed of isolated graphitic sheets (Fig. S1a–d†). On average, the thickness of the GO sheets is ∼0.90 nm, while the thickness of RGO is ∼0.88 nm, which implies that the oxygenated groups are partially reduced to re-establish the conjugated graphite network. Considering the intrinsic monolayer ripples owing to thermal fluctuations, it can be concluded that the RGO sheets are mixtures of monolayers and bilayers. The thickness of single layer DF-RGO is measured to be 1.1–1.5 nm, which is slightly higher than that of RGO. The addition of thickness may be attributed to the perpendicular covalent attachment of the functional group moieties. The M-DF-RGO hybrid was redispersed to form a metastable dispersion and drop-dried on a mica substrate for the AFM study. We found that even after a long period of sonication during the preparation of the AFM specimen, the magnetic nanoparticles are still strongly anchored on the surface of the DF-RGO sheets, suggesting a strong interaction between the magnetic nanoparticles and the DF-RGO sheets. The thickness of the M-DF-RGO sheets was in the range of 10–12 nm. Since the size of the magnetic nanoparticles is about 10.12 nm (according to the XRD results) and the thickness of the DF-RGO is about 1.1–1.5 nm, it is reasonable to conclude that the hybrid consists of monolayer DF-RGO and magnetic nanoparticles.34
The powder X-ray diffraction (XRD) patterns are shown in Fig. 2a. Graphene oxide exhibited a diffraction peak at 2θ = 10.6°.35 As calculated from the Bragg equation, the interlayer spacing of GO (0.82 nm) was enhanced after graphite (0.34 nm) was oxidized, which may be due to the introduction of oxygen-containing functional groups and the adsorption of water molecules. For RGO, a weak and broad diffraction peak (002) was observed at 2θ = 22.8°, while the (001) diffraction peak of GO almost disappeared, which again proved that most of the oxygen-containing functional groups on the surface of GO had been removed. Concerning DF-RGO, a broad reflection peak (∼20°) exists that is associated with the (002) RGO planes. The interlayer distance of DF-RGO (0.90 nm) is larger than that of graphene oxide (0.82 nm) as a result of the presence of –C6H5–COOH groups on both sides of the RGO plane. The XRD pattern of M-DF-RGO exhibited weak reflections at 2θ = 35.84° (311), which is consistent with the existence of iron oxide either in Fe3O4 or γ-Fe2O3 phases.36 The adsorbed nanoparticles act as spacers that limit the restacking of graphene layers.37
Fig. 2 Characterization of samples. (a) XRD spectra; (b) FTIR spectra; (c) Raman spectra; (d) TGA curves. |
The FTIR spectra of the samples are shown in Fig. 2b. The characteristic absorption peaks of graphite oxide at 1723, 1620, 1400, 1223, and 1056 cm−1 are attributed to the stretching vibration of CO, CC, O–H, C–O–C, and C–O, respectively. An intense and broad peak at ∼3436 cm−1 could be assigned to the O–H stretching vibration.38 For RGO, the absorption peaks of the CO, O–H, C–O, and C–O–C groups decreased remarkably or even disappeared, while the peaks of the carbonyl groups were retained. The absorption peak at 794 cm−1 of DF-RGO is associated with stretching of the C–H vibration of the benzoic acid group.39 Moreover, the peak at 1636 cm−1 can be assigned to the CO in a carboxylic acid and the intensity is evidently higher than that for bare GO, indicating that RGO has been successfully functionalized. The FTIR spectrum of M-DF-RGO differs from that of DF-RGO, as evidenced by the weakening of the C–O peak at 3430 cm−1. This suggests that heat treatment in an alkaline medium removed oxygen from the surface of DF-RGO. Two additional vibrational peaks appeared at approximately 1440 and 1380 cm−1. These derived from the formation of either a monodentate complex or a bidentate complex between the carboxyl group and Fe of the magnetic particles, indicating that the iron oxide nanoparticles were covalently bonded to the surface of DF-RGO by Fe–O–C bonds.36
Raman spectroscopy is a tool utilized in order to analyze ordered and disordered crystal structures of graphene based materials. Fig. 2c shows the Raman spectra of GO, RGO, DF-RGO, and the M-DF-RGO hybrid. The ratio of intensities between the D and G bands (ID/IG) of the samples in Raman spectroscopy is regarded as a measure of defects in samples.40 The ratio of the D to G band intensities (ID/IG) increased from 1.19 to 1.44 after GO was reduced by TUD. Meanwhile, a general characteristic of the spectrum of RGO is the shift to a lower wavenumber, suggesting that the average size of the in-plane sp2 domain had decreased, while the number of sp2 carbon atoms increased for RGO.41 This means that the electronic conjugation of RGO was restored, which is consistent with the XRD results. The ID/IG ratio for DF-RGO is approximately 1.52, indicating that the aryl diazonium treatment led to a higher degree of disorder and defect bond formation, and some functional groups were grafted onto the surface of RGO.42 The D band of the M-DF-RGO composite is shifted downward by approximately 10 cm−1 compared to DF-RGO, while the ID/IG ratio was calculated to be approximately 1.46 due to the reduction of DF-RGO during the co-precipitation reaction process.43
The TGA curves of the samples are presented in Fig. 2d. For GO powders, significant weight loss at 100–150 °C and 150–300 °C of ∼19.03 wt% and ∼30.87 wt%, respectively, can be observed, which is attributed to the decomposition of oxygen functional groups and carbon skeleton oxidation, respectively.44 In contrast, RGO possesses better thermal stability than GO, which is shown by a mass loss of just ∼15.22 wt% at 130–262 °C that arose from the residual carboxyl groups on the plane of RGO. TGA was employed to gauge the degree of functionalization of DF-RGO, and it enables comparison of the weight loss of the functionalized sample and RGO. The defunctionalization of DF-RGO is determined to occur at 200–700 °C.45 In addition, M-DF-RGO exhibited a lower thermal stability than DF-RGO. There was also a slow weight loss (∼9.8 wt%) at a low temperature (<100 °C), which can be attributed to the loss of the residual or absorbed solvent. The obvious weight loss occurring at and above 595 °C may be ascribed to the breakdown of the –COO group coordinated with Fe3O4 nanoparticles in the M-DF-RGO hybrid.46
The magnetization curves of Fe3O4 and M-DF-RGO were measured using VSM at room temperature, as shown in Fig. 3. The saturation magnetization of the M-DF-RGO hybrid was smaller than that of pure bulk Fe3O4 (79.20 emu g−1) due to the smaller size of Fe3O4 nanoparticles and the existence of DF-RGO.33 Meanwhile, the specific saturation magnetization of M-DF-RGO is 28.46 emu g−1, which can be attributed to the relatively high amount of Fe3O4 nanoparticles loaded on DF-RGO.47
Fig. 3 The magnetization curves of the samples, with the inset showing the separation efficiencies of M-DF-RGO. |
X-ray photoelectron spectroscopy (XPS) is a sensitive analytical technique used to determine the surface elemental compositions and types of functional groups available on carbon materials.15,21,34 Table 1 lists the elementary compositions of the surfaces of the samples.
Sample | C 1s | O 1s | O 1s/C 1s | Fe 2p |
---|---|---|---|---|
GO | 63.88 | 36.12 | 0.56 | — |
RGO | 78.00 | 36.12 | 0.28 | — |
DF-RGO | 75.82 | 21.99 | 0.29 | — |
M-DF-RGO | 64.50 | 26.98 | 0.42 | 8.52 |
The high resolution XPS C 1s spectrum is shown in Fig. 4. For GO, the raw peaks can be divided into five individual peaks at 284.6 eV (CC), 285.6 eV (C–C), 286.6 eV (C–O), 287.4 eV (CO), and 288.6 eV (O–CO).23,43 The high resolution XPS C 1s spectrum of RGO showed a significant decrease of signals at 286–288 eV, which can be ascribed to the reduction of C–O/C–O–C and CO functionalities. The relative atomic percentage of oxygen obviously decreased from 36% for GO to 21% for RGO, while the relative atomic percentage of CC groups on RGO increased to some extent in comparison with that on GO. However, the actual amount of the carboxyl functional groups on RGO remained largely the same. Upon reaction with aryl diazonium salts, the ratio of the –COOH peak area increased rapidly to 4.2% for DF-RGO from 2.0% for GO. All of these data proved that the surface of RGO had been covalently modified by p-carboxyphenyl groups through aryl diazonium treatment. Compared with DF-RGO, the relative content of CC groups on M-DF-RGO is higher, and the relative intensities of C–O/C–O–C and CO functionalities are lower, indicating the likelihood of DF-RGO to be reduced partially under alkaline conditions during the preparation of the M-DF-RGO hybrid. In the Fe 2p spectrum (Fig. S2†), the peaks at 711.8 and 725.1 eV are attributed to Fe 2p3/2 and Fe 2p1/2. The spin-orbital splitting of the Fe 2p peak is broad due to a small chemical shift between Fe2+ and Fe3+,47 both of which are present in Fe3O4.
Fig. 4 High resolution C 1s core level XPS spectra of (a) GO, (b) RGO, (c) DF-RGO, and (d) M-DF-RGO. |
qe = KFCe1/n | (1) |
Some researchers have proved that graphene based materials with π–electron-rich properties had a good adsorption effect on aromatic compounds through their π–electron-coupling ability.55,56 The reduction of GO with TUD removed some of the functional groups on the plane of GO, and the introduction of the p-carboxyphenyl group can restrain the agglomeration of RGO in water. Efforts were made to increase the π–π interaction between RGO and chlorophenols. Moreover, some remaining oxygen-containing functional groups for DF-RGO could form hydrogen bonds with the hydroxyl groups of phenol.57,58
It should also be noted that the adsorption capacities of chlorophenols on M-DF-RGO were remarkably distinguished from each other. The adsorbed amounts of 2,4-DCP with two –Cl on phenol was higher than that of 4-CP with one, suggesting that the number of substituents would significantly influence donor/acceptor strength,55 and chlorophenol with more substituted groups had a higher adsorption capacity.
Fig. 8 Optimized structures of the adsorption of (a) 4-CP and (b) 2,4-DCP on DF-RGO (carbon atom: yellow; chlorine atom: green; oxygen atom: red; hydrogen atom: blue). |
System | Ead (kcal mol−1) |
---|---|
DF-RGO + 4-CP | 11.61 |
DF-RGO + 2,4-DCP | 14.56 |
As shown in Fig. 8a and b, chlorophenol molecules were parallel to the basal plane of graphene, indicating that the π–π interaction plays an important role in the removal of chlorophenols. As shown in Table 2, the Ead value for 2,4-DCP on DF-RGO is 14.56 kcal mol−1, which is slightly higher than the 11.61 kcal mol−1 for 4-CP on DF-RGO, indicating that the adsorption of 2,4-DCP on the DF-RGO model surface is more stable and the interaction between DF-RGO and 2,4-DCP is stronger. This might be because the 2,4-DCP molecule has an extra –Cl group that could strengthen the π–π interaction between the phenol ring and the adsorbent.
Therefore, based on the results of the theoretical calculations, the higher Ead value indicated that the DF-RGO had a stronger affinity to 2,4-DCP than 4-CP, which is in good agreement with the experimental results.
The regeneration and reuse of adsorbents is also important in water treatment based on stringent ecological and economic demands for sustainability. We found that M-DF-RGO can go through multiple rounds of reuse by thorough washing with 0.01 mol L−1 NaOH solution and ethanol, separately. Fig. 9 showed that the adsorption capacities of chlorophenols slightly decreased with increasing times of reuse, and were still over 80% after reuse in five cycles, suggesting that M-DF-RGO has good reusability.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c8ra00503f |
This journal is © The Royal Society of Chemistry 2018 |