Facile preparation of novel quaternary g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites: magnetically separable visible-light-driven photocatalysts with significantly enhanced activity

Anise Akhundi and Aziz Habibi-Yangjeh*
Department of Chemistry, Faculty of Science, University of Mohaghegh Ardabili, P.O. Box 179, Ardabil, Iran. E-mail: ahabibi@uma.ac.ir; Fax: +98 045 33514701; Tel: +98 045 33514702

Received 12th May 2016 , Accepted 2nd November 2016

First published on 2nd November 2016


Abstract

In this paper, we successfully fabricated quaternary g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites with various weight percents of Bi2S3, as novel magnetically separable visible-light-driven photocatalysts, using a facile refluxing method at 96 °C. The prepared samples were characterized by XRD, EDX, SEM, TEM, UV-vis DRS, FT-IR, TGA, BET, PL, and VSM techniques. Photocatalytic activity of the samples was investigated by degradation of rhodamine B, methylene blue, and fuchsine under visible-light illumination. It was found that photocatalytic activity of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite, as the optimal photocatalyst, in the degradation of RhB is nearly 56, 44, 6.5, and 16-folds higher than those of the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/Bi2S3 (30%) samples, respectively. Furthermore, the effects of refluxing time, calcination temperature, scavengers of the reactive species, and number of recycling runs on the photocatalytic activity were investigated. Based on the results, the improved photocatalytic activity was attributed to the enhanced visible-light absorption ability and matching energy levels of the counterparts that facilitate generation and separation of electron–hole pairs, respectively.


Introduction

In recent decades, a major cause of water pollution has been organic compounds from different industrial sources. These persistent organic pollutants, introduced into natural water resources or wastewaters, are highly toxic and hazardous to living organisms.1 Hence, removal of these pollutants from water prior to discharge into the environment is highly desirable. Among different treatment strategies such as physical, chemical, and biological, heterogeneous photocatalytic processes are promising technology for decontamination of water containing stable organic compounds.2,3 In these processes, after irradiating a photocatalyst with proper photons (their energy equal to or greater than band gap of the photocatalyst), electrons are moved from the valence band (VB) to the conduction band (CB) of photocatalyst to create some electron–hole pairs. After that, the photogenerated charge carriers start up different oxidation and reduction reactions over the surface of the photocatalyst to produce different reactive species.4 Many semiconductors including TiO2, ZnO, and SnO2 are commonly used in photocatalytic processes.2,5,6 However, these photocatalysts can be activated only by the light with a wavelength less than about 390 nm.3 It is well known that this radiation accounts for about 4% of the solar energy, resulting in restricted large-scale utilization of this technology in different disciplines. Therefore, in recent years, extensive efforts have been devoted to looking for novel visible-light-driven photocatalysts with considerable activity.6–16

Graphitic carbon nitride (g-C3N4), a new kind of conjugated polymeric semiconductor, has been discovered as a fascinating metal-free organic photocatalyst with moderate activity under visible-light irradiation.17 This semiconductor has found broad applications, such as splitting of water to produce hydrogen gas, degradation of different pollutants, reduction of carbon dioxide to different fuels, catalysis of different reactions, solar cells, and sensing owing to its high stability, appealing electronic structure, and medium band gap.18–22 However, pure g-C3N4 suffers from some drawbacks, due to recombination of the charge carriers with high rate, low visible-light harvesting ability, and difficulties in recycling from the treated solutions.23 Therefore, it is very important to design and prepare more efficient g-C3N4-based photocatalysts by expanding the light absorption ability further into the visible range. Combining g-C3N4 with narrow band gap semiconductors, having matched energy levels, is the most effective strategy to overcome the first and second shortcomings.24–29 Furthermore, deposition of magnetic materials over g-C3N4 sheets is fascinating method to tackle the third drawback by application of an external magnetic field.30–32

On the other hand, considerable attention has been recently paid to the bismuth-based semiconductors.33–35 Bismuth sulfide (Bi2S3) is a promising photocatalyst and it has attracted more attention due to its narrow band gap of 1.4 eV.35,36 Hence, Bi2S3 absorbs a wide range of visible-light irradiation. In this regard, different Bi2S3-containing photocatalysts have been reported.37–39 Very recently, we prepared g-C3N4/Fe3O4/AgI (20%) nanocomposite as visible-light-driven photocatalysts with high photocatalytic activity.40 However, because of its weak ability for absorption of visible light, this nanocomposite has moderate activity under visible-light illumination. Hence, it seems that by combining g-C3N4/Fe3O4/AgI (20%) nanocomposite with Bi2S3 particles, we could prepare novel magnetically separable quaternary nanocomposites with significantly enhanced activity. For the best of our knowledge, there is not any report about magnetically separable photocatalysts containing g-C3N4 and Bi2S3. Furthermore, preparation and investigation photocatalytic activity of the quaternary g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites have not reported.

In the present study, we successfully prepared g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites with different weight percents of Bi2S3, as novel visible-light-driven photocatalysts, using a facile large-scale method and they were characterized using different techniques. Photocatalytic activity of the prepared samples was investigated by degradation of rhodamine B (RhB), methylene blue (MB), and fuchsine under visible-light irradiation. The results indicated that the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite showed the highest photocatalytic activity. We also illustrated that different parameters such as refluxing time, calcination temperature, and scavengers of the reactive species played particularly important roles on the photocatalytic performance of the nanocomposites. Finally, the recycling runs for the degradation reaction of RhB were performed to evaluate stability of the optimal nanocomposite.

Experimental

Materials

Ferric chloride (FeCl3·6H2O, 99.5%), ferrous chloride (FeCl2·4H2O, 98.0%), ammonia solution (30%), melamine (C3H6N6, 99.2%), silver nitrate (99.9%), sodium hydroxide (98%), and benzoquinone were purchased from Loba Chemie and used as received. Bismuth nitrate (Bi(NO3)3·4H2O, 99%), potassium iodide (99.89%), thioacetamide (C2H5NS, 99%), 2-propanol, ammonium oxalate, RhB, MB, fuchsine, and absolute ethanol with high quality were obtained from Merck and used as received. Deionized water was used throughout this work.

Instruments

The X-ray diffraction (XRD) patterns were recorded by a Philips Xpert X-ray diffractometer with Cu Kα radiation (λ = 0.15406 nm). Surface morphology and distribution of particles were studied by LEO1430VP scanning electron microscopy (SEM) of model LEO1430VP, using an accelerating voltage of 15 kV. The purity and elemental analysis of the products were obtained by energy dispersive analysis of X-rays (EDX) on the same SEM instrument. The transmission electron microscopy (TEM) investigations were performed by a Zeiss-EM10C instrument with an acceleration voltage of 80 kV. The UV-vis diffuse reflectance spectroscopy (DRS) data were recorded by a Scinco 4100 apparatus. The Fourier transform-infrared (FT-IR) spectra were obtained using a Perkin-Elmer Spectrum RXI apparatus. The thermogravimetric analysis (TGA) of the samples was performed on Linseis STAPT1000 by heating under air atmosphere from room temperature to 700 °C at 10 °C min−1. The specific surface area and pore properties of the samples were calculated using the Brunauer–Emmett–Teller (BET) and Barret–Joyner–Halenda (BJH) models with nitrogen adsorption–desorption isotherms collected by Belsorp apparatus at −196 °C. Prior to the experiments, the samples were degassed at 120 °C for 15 h. The photoluminescence (PL) spectra of the samples were studied using a Perkin-Elmer (LS55) fluorescence spectrophotometer with an excitation wavelength of 300 nm. The conditions such as amount of the sample, volume of solvent, ultrasonic irradiation time and its power (applied for dispersion of the sample) were fixed in order to compare the PL intensities. UV-vis spectra of the degradation reaction were studied using a Cecile 9000 spectrophotometer. Magnetic properties of the samples were obtained using vibrating sample magnetometer (VSM, Meghnatics Kavir Kashan Co., Iran). The ultrasound radiation was performed using a Bandeline ultrasound processor HD3100 (12 mm diameter Ti horn, 75 W, 20 kHz).

Preparation of the samples

Pyrolysis of melamine, as the starting material, was employed to prepare the g-C3N4 sample in a semi closed system.18 The g-C3N4/Fe3O4 (2[thin space (1/6-em)]:[thin space (1/6-em)]1), where 2[thin space (1/6-em)]:[thin space (1/6-em)]1 is weight ratio of g-C3N4 to Fe3O4, and g-C3N4/Fe3O4/AgI (20%), where 20% is weight percent of AgI, were prepared using the reported methods by our research group.40 The g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites were simply prepared by refluxing method at 96 °C. Typically for preparation of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite, where 30% is weight percent of Bi2S3, 0.3 g of the g-C3N4/Fe3O4/AgI (20%) nanocomposite was dispersed in 150 mL of water via ultrasonication for 10 min and then 0.241 g of bismuth nitrate were added to the suspension. After stirring for 120 min, an aqueous solution of thioacetamide (0.056 g in 20 mL of water) was dropwise added to the suspension with subsequent stirring for 30 min and then refluxed at 96 °C for 60 min. After that, the product was collected by centrifugation, washed with water and absolute ethanol, and dried in an oven at 60 °C for 24 h. According to this method, different weight percents of Bi2S3 were loaded and they were labeled as g-C3N4/Fe3O4/AgI/Bi2S3 (10%), g-C3N4/Fe3O4/AgI/Bi2S3 (20%), g-C3N4/Fe3O4/AgI/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 (40%) nanocomposites. For comparison purposes, the g-C3N4/Fe3O4/Bi2S3 (30%) sample was synthesized by the same method. Typically, 0.241 g of bismuth nitrate was added to 0.3 g of the ultrasonicated g-C3N4/Fe3O4 (2[thin space (1/6-em)]:[thin space (1/6-em)]1) nanocomposite. Next, the suspension was stirred for 120 min and thioacetamide solution (0.056 g in 20 mL of water) was dropwise added under stirring for 30 min. After refluxing for 60 min, the resultant suspension washed with water and ethanol for two times and dried in an oven at 60 °C for 24 h (Scheme 1).
image file: c6ra12414c-s1.tif
Scheme 1 The procedure for preparation of the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites.

Photocatalysis experiments

To investigate photocatalytic activity of the prepared samples, several photocatalytic experiments were comparatively carried out by degradation 250 mL of RhB (2.50 × 10−5 M), MB (2.50 × 10−5 M), and fuchsine (9.2 × 10−6 M) solutions under visible-light irradiation. Due to high absorption coefficient of fuchsine, concentration of fuchsine was very lower than those of RhB and MB dyes. For this purpose, 0.1 g of the powdered photocatalyst was dispersed under ultrasonication for 6 min. For the irradiation system, an LED lamp with 50 W was used as visible-light source at the distance of 20 cm from the solution in a dark box. The emission spectrum of the source has high intensity in visible range and its intensity rapidly decreases in wavelengths near to UV and IR ranges.41 During the degradation reactions, different samples were withdrawn regularly from the reactor and the catalyst was removed by a magnet to analyze the transparent solutions spectrophotometrically. The absorbances of RhB, MB, and fuchsine solutions were determined as a function of the irradiation time at wavelengths of 553, 664, and 540 nm, respectively.

Results and discussion

Fig. 1a shows XRD patterns of the g-C3N4, g-C3N4/Fe3O4, and quaternary g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites with different amounts of Bi2S3 along with the patterns for the g-C3N4/Fe3O4/AgI (20%) and g-C3N4/Fe3O4/Bi2S3 (30%) samples. Two distinct diffraction peaks at 2θ of 13.1 and 27.3° can be found for pure g-C3N4, which correspond to the (100) and (002) planes of hexagonal g-C3N4 (JPCDS no. 87-1526), respectively.18 All diffraction peaks of Bi2S3 can be well indexed to orthorhombic phase (JCPDS no. 84-0279).42 Due to presenting XRD patterns of many samples in Fig. 1a, some of the peaks are not clearly seen. For this reason, the XRD pattern of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite was shown in Fig. 1b. As can be seen, diffraction peaks of g-C3N4, Fe3O4, AgI, and Bi2S3 counterparts are clearly seen, suggesting the successful deposition of Fe3O4, AgI, and Bi2S3 particles on the surface of g-C3N4.
image file: c6ra12414c-f1.tif
Fig. 1 (a) XRD patterns for the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites with different weight percents of Bi2S3. (b) XRD pattern for the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite.

The EDX spectra of g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples were provided to further illustrate purity of the prepared samples. Fig. 2a shows the EDX spectra of the samples, which reveals that the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite possesses C, N, Fe, O, Ag, I, Bi, and S elements. Moreover, distributions of g-C3N4, Fe3O4, AgI, and Bi2S3 counterparts in the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite were analyzed through EDX mapping of C, N, Fe, O, Ag, I, Bi, and S elements, as seen in Fig. 2b–i. It can be observed that all of the elements are detectable and distributed evenly, confirming formation of heterojunction between counterparts of the nanocomposite. In addition, weight percents of the elements in the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite are 15.2, 21.8, 12.2, 7.20, 6.48, 6.83, 24.4, and 5.92% for C, N, Fe, O, Ag, I, Bi, and S elements, respectively. The theoretical weight percents of C, N, Fe, O, Ag, I, Bi, and S elements are 15.3, 21.8, 12.1, 7.24, 6.51, 6.82, 24.3, and 5.92%, respectively. Hence, as can be observed, the experimental and theoretical values of the weight percents are comparable. Based on the experimental values, weight percents of g-C3N4, Fe3O4, AgI, and Bi2S3 counterparts in the nanocomposite were calculated to be 36.7, 18.7, 14.2, and 30.4%, respectively.


image file: c6ra12414c-f2.tif
Fig. 2 (a) EDX spectra for the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples. (b–i) EDX mapping of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite.

Fig. 3 shows the SEM and TEM images of the quaternary g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite. In this figure, aggregates of deposited particles are observed over the g-C3N4 sheet.


image file: c6ra12414c-f3.tif
Fig. 3 (a) SEM and (b) TEM images of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite.

Fig. 4a displays UV-vis DRS spectra of the as-prepared samples. The pristine g-C3N4 has an absorption edge at about 470 nm. Due to low band gap of about 1.4 eV for Bi2S3, absorption edges of the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites gradually shift toward longer wavelengths and the absorption intensities obviously increase in the visible region with the increase of Bi2S3 amount. The band gap energies (Eg) of the prepared samples were estimated using equation αhν = B(Eg)n/2, in which α, ν, and B are absorption coefficient, the light frequency, and proportionality constant, respectively.43 The value of n depends on the characteristics of the transition in the semiconductor. It is well known that g-C3N4, AgI, and Bi2S3 are direct band gap semiconductors.18,35,44 The Eg values for the prepared samples were estimated by extrapolation of the linear part of the curves obtained by plotting (αhν)2 versus hν (Fig. 4b). It can be observed that the Eg values of samples are between 1.56 and 2.73 eV and band gap of the quaternary nanocomposites decreases with increasing weight percent of Bi2S3.


image file: c6ra12414c-f4.tif
Fig. 4 (a) UV-vis DRS for the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites with different weight percents of Bi2S3. (b) Plots for calculation of band gap of the prepared samples. (c) Magnetization curves for the Fe3O4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples.

Magnetization versus magnetic field curves (MH) for the Fe3O4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples were measured at room temperature using a VSM instrument with an applied field of 8500 Oe and the results are shown in Fig. 4c. The saturation magnetizations of these samples were measured to be 55.5, 20.6, 16.9, and 7.13 emu g−1 for the Fe3O4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples, respectively. In particular, it is evident from Fig. 4b that saturation magnetizations of these samples increased with an increase in Fe3O4 contents in the samples. The magnetic separability of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite was also examined by placing a magnet near the glass bottle. The grey particles were tightly attracted toward the magnet (inset of Fig. 4b), revealing that the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite possesses excellent magnetic separability. These results clearly state that the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites may be an excellent candidate as novel type of magnetically separable photocatalysts for potential environmental applications.

To further confirm composition of the as-prepared samples, FT-IR spectra were provided and the results are shown in Fig. 1S. The strong IR absorption bands in the pure g-C3N4 sample revealed a typical molecular structure of g-C3N4.18 Several bands in the 1240–1650 cm−1 region are corresponding to the typical stretching modes of CN heterocycles. Moreover, a peak at 806 cm−1 is related to the out-of-plane breathing vibration of triazine units.18 For the Fe3O4 containing samples, the intense signal at 630 cm−1 is assigned to vibration of the Fe–O bond.40 Also, similar to the literature, there is not any peak for Ag–I bond.40 For the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite, a new peak was appeared at 462 cm−1, due to the presence of Bi2S3.45

The amounts of g-C3N4 in the nanocomposites were estimated by TGA analysis. Fig. 2S shows TGA curves of the g-C3N4, g-C3N4/Fe3O4, and quaternary g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites along with those of the AgI and Bi2S3 samples from room temperature to 700 °C in the air condition. For the pure g-C3N4 sample, a 97% of weight loss is observed. The deviation from 100% may be related to the baseline changes during the heating process. This weight loss is due to burning of g-C3N4 starts from nearly 530 °C to produce carbon dioxide and different oxides of nitrogen.18,19 This implies that the g-C3N4 can be almost completely burned in air up to 700 °C. As can be seen, similar to many reports, pure g-C3N4 is stable up to 500 °C.18 It is evident that the AgI and Bi2S3 samples have considerable stability in the heating process. In the case of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) sample, the total weight loss is about 35% in the temperature range of 25–700 °C, indicating that this sample contains about 35% of g-C3N4. The actual loading amounts of g-C3N4 in the g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI/Bi2S3 (10%), g-C3N4/Fe3O4/AgI/Bi2S3 (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (40%) samples were estimated to be about 67, 45, 40, and 30 wt%, respectively. In addition, it is evident that all of the prepared samples are thermally stable up to 430 °C. Similar to many reports, stability of g-C3N4 decreases with loading different materials over that.31,32

The textural properties of the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples were characterized using nitrogen adsorption–desorption experiments and the results are displayed in Fig. 5. It is evident that the samples have typical IV isotherms with H1 hysteresis, representing that the samples are mesoporous. The specific surface area and pore properties of the samples were calculated by BET and BJH models, respectively and the results are shown in Table 1. As can be seen, the specific surface area of the g-C3N4/Fe3O4 sample has increased compared with the g-C3N4, due to formation of the hierarchical structure. However, after loading of AgI and Bi2S3 particles on the g-C3N4/Fe3O4 nanocomposite, the specific surface area was decreased. This decrease may be ascribed to covering of the g-C3N4/Fe3O4 surface by AgI and Bi2S3 particles, resulting in blocking some parts of the active sites. As can be seen, there are not major differences between textural properties of the g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposites. Hence, textural properties cannot describe the highly enhanced photocatalytic activity of the quaternary nanocomposite relative to the ternary nanocomposites.


image file: c6ra12414c-f5.tif
Fig. 5 Nitrogen adsorption–desorption data for the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples.
Table 1 The textural properties of g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples
Photocatalyst Surface area (m2 g−1) Mean pore diameter (nm) Total pore volume (cm3 g−1)
g-C3N4 14.6 14.2 0.0500
g-C3N4/Fe3O4 29.4 21.6 0.159
g-C3N4/Fe3O4/AgI (20%) 16.0 14.5 0.0581
g-C3N4/Fe3O4/Bi2S3 (30%) 22.3 17.9 0.0998
g-C3N4/Fe3O4/AgI/Bi2S3 (30%) 14.9 19.6 0.0728


Photocatalytic activity of the prepared samples was investigated by degradation of RhB under visible-light irradiation and the results are depicted in Fig. 6a. As it can be observed, the pure g-C3N4 presents poor photocatalytic activity. For comparison, photocatalytic activities of the g-C3N4/Fe3O4/AgI (20%) and g-C3N4/Fe3O4/Bi2S3 (30%) nanocomposites were also tested under the same conditions, which can degrade 64 and 41% of RhB within 60 min, respectively. It is evident that all of the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites exhibit higher photocatalytic activity than those of the g-C3N4/Fe3O4/AgI (20%) and g-C3N4/Fe3O4/Bi2S3 (30%) samples, indicating that a synergistic effect exists between AgI and Bi2S3 particles in enhancing photocatalytic activity. As can be seen, with the increase of Bi2S3 content, photocatalytic activity of the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites exhibits a rise followed by a decline and the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) sample demonstrates the highest photocatalytic activity, which can degrade 99% of RhB after 60 min. By loading more amount of Bi2S3, they could destruct the heterojunction between the counterparts due to aggregation of its particles, leading to the decrease of the photocatalytic activity.


image file: c6ra12414c-f6.tif
Fig. 6 (a) Photodegradation of RhB over the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites with different weight percents of Bi2S3. The UV-vis spectra for degradation of RhB under visible-light irradiation over the (b) g-C3N4, (c) g-C3N4/Fe3O4, (d) g-C3N4/Fe3O4/AgI (20%), (e) g-C3N4/Fe3O4/Bi2S3 (30%), and (f) g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples.

Fig. 6b–f show spectral changes of RhB during the degradation reaction over the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), g-C3N4/Fe3O4/Bi2S3 (30%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples under visible-light irradiation. It is clear that intensity of the main absorption peak of RhB, located at 553 nm, decreases rapidly with increasing irradiation time and stable intermediates are not formed during the degradation reaction.46,47 Moreover, the quaternary nanocomposite completely degrades RhB after 60 min and compared with the other samples, this photocatalyst exhibits superior activity.

As compared in Fig. 7a, the rate constant for degradation of RhB over the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite is 731.8 × 10−4 min−1, which is significantly higher than those of the g-C3N4 (16.0 × 10−4 min−1), g-C3N4/Fe3O4 (20.2 × 10−4 min−1), g-C3N4/Fe3O4/AgI (20%) (139 × 10−4 min−1), and g-C3N4/Fe3O4/Bi2S3 (30%) (54.9 × 10−4 min−1) samples. Hence, activity of the quaternary nanocomposite is about 56, 44, 6.5, and 16-folds higher than those of the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/Bi2S3 (30%) samples, respectively. It is well known that PL spectra can be applied to display recombination rate of photogenerated charge carriers. Herein, we conducted PL measurements to study the effect of Bi2S3 on recombination rate of the electron–hole pairs in the g-C3N4/Fe3O4/AgI (20%) nanocomposite. It can be seen that the g-C3N4 sample exhibits the strongest PL spectrum with an emission peak at about 435 nm (Fig. 7b). This strong peak is attributed to the band–band PL phenomenon with the energy of light approximately equal to the band gap energy of g-C3N4.18 Compared with the pristine g-C3N4, the PL intensities of the g-C3N4/Fe3O4 and g-C3N4/Fe3O4/AgI (20%) samples are weaker, reflecting their slow recombination rates of the charge carriers. The weakest intensity of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite suggests that the presence of Bi2S3 can sharply reduce recombination rate of the photogenerated charge carriers in the quaternary nanocomposite.


image file: c6ra12414c-f7.tif
Fig. 7 (a) The degradation rate constants of RhB over different samples. (b) PL spectra for the g-C3N4, g-C3N4/Fe3O4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples.

Preparation time of a photocatalyst is a very important parameter, affecting crystallinity, size, aggregation, and morphology of particles. Thus a series of experiments were conducted by preparation of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite at different refluxing times. Fig. 3S demonstrates the preparation time dependence of the degradation rate constant of RhB over the quaternary nanocomposite under visible-light irradiation. As illustrated, the nanocomposite prepared by refluxing for 2 h has the highest photocatalytic activity and with more increasing of the preparation time, the degradation rate constant starts to decrease. In addition, Fig. 3S shows SEM image of the nanocomposite prepared by refluxing for 4 h. As can be seen, the deposited particles over g-C3N4 considerably aggregated relative to the sample prepared by refluxing for 2 h (see Fig. 3a). Hence, by more aggregation of the deposited particles over g-C3N4, heterojunction formed between the counterparts were destructed, leading to decrease of the photogenerated electron–hole pair separations. Therefore, the nanocomposite prepared by refluxing for 2 h was selected for further investigations.

It is notable that calcination temperature of photocatalysts could remarkably affect on crystallization and size of particles. In this step, the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite prepared by refluxing for 120 min was calcined at 200, 300, 400, and 500 °C for 2 h. As can be seen in Fig. 3S, the degradation rate constant of RhB over the nanocomposite decreases with an increase in the calcination temperature. Moreover, Fig. 3S shows SEM image of the nanocomposite calcined at 500 °C. Compared with the uncalcined sample (Fig. 3a), it is evident that the deposited particles have considerably aggregated and formed bigger sizes.48 The decrease of the rate constant is attributed to increase the aggregation of AgI and Bi2S3 particles deposited on the g-C3N4 sheet, resulting in destruction of heterojunction between counterparts of the nanocomposite. Hence, photogenerated electrons could not easily transfer from g-C3N4 to AgI and Bi2S3 particles, leading to decrease of the photocatalytic activity.

Depending on the reaction mechanism, a photocatalytic degradation reaction involves formation of various active species that can be different from those of the other reactions.2,3 To understand the reaction mechanism, it is important to identify the active species participating in the degradation reaction. Using trapping experiments, the roles of holes (h+), hydroxyl radicals (˙OH), and superoxide ion radicals (˙O2) were identified using ammonium oxalate, 2-propanol, and benzoquinone, respectively. It can be seen in Fig. 4S that the degradation rate constant significantly decreased upon addition of benzoquinone. On the contrary, the photocatalytic activity was slightly decreased when ammonium oxalate and 2-propanol were added to the reaction system. The above results indicate that superoxide ion radicals are the main reactive species in the degradation reaction of RhB over the quaternary nanocomposite.

The band structures and matching of the band energies are responsible for the efficient generation and separation of the charge carriers in photocatalysts. Hence, VB potentials of the counterparts were estimated by the following empirical equation: EVB = XEe + 0.5Eg, where EVB is the VB potential, X is the electronegativity of the semiconductor (the geometric mean of the electronegativity of the constituent atoms), Ee is the energy of free electrons on the hydrogen scale (∼4.5 eV), Eg is the band gap energy of the photocatalyst.49 Moreover, the ECB potentials were calculated by ECB = EVBEg formula and the results were tabulated (Table 2). Based on the experimental and calculation results, the plausible mechanism for the enhanced photocatalytic activity over the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites is presented in Fig. 8. It is well known that g-C3N4, AgI, and Bi2S3 are n-type semiconductors and their Fermi levels are close to their CB levels. Thus, after contacting these semiconductors to each other, n–n heterojunctions will be formed in the contacting regions.50–54 Band gaps of g-C3N4, AgI, and Bi2S3 are 2.7, 2.8, and 1.4 eV, respectively. As a result, they are simply activated under visible-light irradiation to produce electron–hole pairs. The photogenerated electrons transfer from the CB of AgI and Bi2S3 to that of g-C3N4 under the driving force of the produced electrostatic fields.11,13 As can be seen in Table 1, the CB potential of g-C3N4 is more negative than the potential of O2/˙O2 (E0(O2/˙O2) = −0.33 eV/NHE),55 hence, photogenerated electron on g-C3N4 react with adsorbed O2 to form ˙O2 species, as a major reactive species participating in the degradation reaction. Furthermore, the VB energies for AgI and Bi2S3 are less positive than that of g-C3N4. As a consequence, the photogenerated holes transfer from the VB of g-C3N4 to those of the AgI and Bi2S3. Hence, the photogenerated electrons and holes are gathered on the CB of g-C3N4 and VB of AgI and Bi2S3, respectively. Therefore, due to formation of tandem n–n heterojunctions between these semiconductors, the charge carriers are effectively separated from each other, resulting in enhanced photocatalytic activity in the quaternary nanocomposites.50–54

Table 2 The VB, CB, and Eg values for counterparts of the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites
Semiconductor VB energy (eV) CB energy (eV) Eg (eV)
g-C3N4 +1.58 −1.12 2.70
AgI +2.38 −0.42 2.80
Bi2S3 +1.50 +0.10 1.40



image file: c6ra12414c-f8.tif
Fig. 8 The mechanism for the enhanced photocatalytic activity of the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites.

Additionally, another two dyes (MB and fuchsine) were chosen as target pollutants to further evaluate photocatalytic activity of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite relative to its counterparts. As can be seen in Fig. 9a and b, similar to RhB degradation, the quaternary nanocomposite exhibited the highest photocatalytic activity in degradations of MB and fuchsine. The degradation rate constants of MB over the g-C3N4, g-C3N4/Fe3O4/AgI, and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples are 19.2 × 10−4, 44.5 × 10−4, and 83.5 × 10−4 min−1, respectively. Also, the rate constants for degradation of fuchsine over the g-C3N4, g-C3N4/Fe3O4/AgI, and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples are 20.8 × 10−4, 27.6 × 10−4, and 172.4 × 10−4 min−1, respectively. Therefore, it is evident that activities of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite in degradation of MB and fuchsine are about 4.3 and 8.2-folds greater than those of the g-C3N4.


image file: c6ra12414c-f9.tif
Fig. 9 The degradation rate constants of (a) MB and (b) fuchsine over the g-C3N4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/AgI/Bi2S3 (30%) samples.

In addition to photocatalytic activity, the stability of a photocatalyst is also very important parameter in widespread practical applications. The results shown in Fig. 10 revealed that the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite did not exhibit significant loss of its activity after four runs in degradation of RhB. Only a slight decrease was observed, which can be attributed to adsorption of degradation intermediates over the surface of the photocatalyst during the degradation process. Hence, this work showed that the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites could be an excellent candidate as novel type of magnetically separable photocatalysts for potential environmental applications.


image file: c6ra12414c-f10.tif
Fig. 10 Reusability of the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite for four successive runs.

In order to investigate the effect of solution pH on the degradation reaction of RhB over the g-C3N4/Fe3O4/AgI/Bi2S3 (30%) nanocomposite, a series of experiments were carried out by changing initial pH of the solution between 2 and 10 by adding HCl and NaOH and the results are shown in Fig. 5S. As can be seen, the degradation rate constant enhances with increasing pH of solution. Very recently, it was found that isoelectric point of g-C3N4 is nearly in solutions with pH = 5.55 Consequently, in solutions with pH < 5, surface of g-C3N4 is positively charged due to protonation of its amine groups.56 On the other hand, RhB is a cationic pollutant. Hence, in solutions with low pH, electrostatic repulsion between the pollutant and the catalyst restricts adsorption of RhB. As a result, the degradation rate constant decreases in acidic solutions. However, in alkaline solutions, surface of g-C3N4 is negatively charged through reaction with hydroxide ions.55 Therefore, the positively charged pollutant is easily adsorbed over the catalyst in solutions with high pH, resulting in enhanced activity in these solutions.

Conclusions

Quaternary g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites, as novel magnetically separable visible-light-driven photocatalysts, were synthesized via a simple method for the first time and their physicochemical properties were systematically characterized using different techniques. The optimal content of Bi2S3 in the g-C3N4/Fe3O4/AgI/Bi2S3 nanocomposites was found to be 30 wt% and the optimal refluxing time was 120 min. This nanocomposite showed 56, 6.5, and 16-folds enhanced activity in degradation of RhB compared to the g-C3N4, g-C3N4/Fe3O4/AgI (20%), and g-C3N4/Fe3O4/Bi2S3 (30%) samples, respectively. Moreover, it was found that superoxide ions played a major role, whereas hydroxyl radicals and holes had minor roles. The recycling of the photocatalyst was also investigated and the activity had little decrease after four cycles. Due to the presence of narrow band gap semiconductor and formation of tandem n–n heterojunctions, the results demonstrated that the introduction of Bi2S3 could modify the visible-light absorption ability and separation of the charge carriers in the ternary g-C3N4/Fe3O4/AgI photocatalyst.

Acknowledgements

The authors wish to acknowledge University of Mohaghegh Ardabili-Iran, for financial support of this work.

References

  1. T. Reemtsma and M. Jekel, Organic Pollutants in the Water Cycle: Properties, Occurrence, Analysis and Environmental Relevance of Polar Compounds, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, 2006 Search PubMed.
  2. V. Etacheri, C. D. Valentin, J. Schneider, D. Bahnemann and S. C. Pillai, J. Photochem. Photobiol., C, 2015, 25, 1 CrossRef CAS.
  3. S. Dong, J. Feng, M. Fan, Y. Pi, L. Hu, X. Han, M. Liu, J. Sun and J. Sun, RSC Adv., 2015, 5, 14610 RSC.
  4. M. Pelaez, N. T. Nolan, S. C. Pillai, M. K. Seery, P. Falaras, A. G. Kontos, P. S. M. Dunlop, J. W. J. Hamilton, J. A. Byrne, K. O'Shea, M. H. Entezari and D. D. Dionysiou, Appl. Catal., B, 2012, 125, 331 CrossRef CAS.
  5. R. Fagan, D. E. McCormack, D. D. Dionysiou and S. C. Pillai, Mater. Sci. Semicond. Process., 2016, 42, 2 CrossRef CAS.
  6. K. M. Lee, C. W. Lai, K. S. Ngai and J. C. Juan, Water Res., 2016, 88, 428 CrossRef CAS PubMed.
  7. M. Pirhashemi and A. Habibi-Yangjeh, Appl. Surf. Sci., 2013, 283, 1080 CrossRef CAS.
  8. S. Khanchandani, P. K. Srivastava, S. Kumar, S. Ghosh and A. K. Ganguli, Inorg. Chem., 2014, 53, 8902 CrossRef CAS PubMed.
  9. S. Pal, S. Maiti, U. N. Maiti and K. K. Chattopadhyay, CrystEngComm, 2015, 17, 1464 RSC.
  10. S. Shaker-Agjekandy and A. Habibi-Yangjeh, Mater. Sci. Semicond. Process., 2015, 34, 74 CrossRef CAS.
  11. F. Kiantazh and A. Habibi-Yangjeh, Mater. Sci. Semicond. Process., 2015, 39, 671 CrossRef CAS.
  12. C. Dong, X. Xiao, G. Chen, H. Guan and Y. Wang, Mater. Chem. Phys., 2015, 155, 1 CrossRef CAS.
  13. M. Pirhashemi and A. Habibi-Yangjeh, J. Mater. Sci.: Mater. Electron., 2016, 27, 4098 CrossRef CAS.
  14. H.-P. Lin, W. W. Lee, S.-T. Huang, L.-W. Chen, T.-W. Yeh, J.-Y. Fu and C.-C. Chen, J. Mol. Catal. A: Chem., 2016, 417, 168 CrossRef CAS.
  15. X. Lv, J. Wang, Z. Yan, D. Jiang and J. Liu, J. Mol. Catal. A: Chem., 2016, 418, 146 CrossRef.
  16. W. Zhao, Y. Liu, Z. Wei, S. Yang, H. He and C. Sun, Appl. Catal., B, 2016, 185, 242 CrossRef CAS.
  17. S. Cao, J. Low, J. Yu and M. Jaroniec, Adv. Mater., 2015, 27, 2150 CrossRef CAS PubMed.
  18. Y. Zheng, J. Liu, J. Liang, M. Jaroniec and S. Z. Qiao, Energy Environ. Sci., 2012, 5, 6717 CAS.
  19. G. Dong, Y. Zhang, Q. Pan and J. Qiu, J. Photochem. Photobiol., C, 2014, 20, 33 CrossRef CAS.
  20. J. Zhu, P. Xiao, H. Li and S. A. C. Carabineiro, ACS Appl. Mater. Interfaces, 2014, 6, 16449 CAS.
  21. Y. Gong, M. Li, H. Li and Y. Wang, Green Chem., 2015, 17, 715 RSC.
  22. S. Ye, R. Wang, M.-Z. Wu and Y.-P. Yuan, Appl. Surf. Sci., 2015, 358, 15 CrossRef CAS.
  23. Z. Zhao, Y. Sun and F. Dong, Nanoscale, 2015, 7, 15 RSC.
  24. C. Xing, Z. Wu, D. Jiang and M. Chen, J. Colloid Interface Sci., 2014, 433, 9 CrossRef CAS PubMed.
  25. H. Xu, Y. Song, Y. Song, J. Zhu, T. Zhu, C. Liu, D. Zhao, Q. Zhang and H. Li, RSC Adv., 2014, 4, 34539 RSC.
  26. X. Rong, F. Qiu, J. Yan, H. Zhao, X. Zhu and D. Yang, RSC Adv., 2015, 5, 24944 RSC.
  27. H. Xu, H. Zhao, Y. Song, W. Yan, Y. Xu, H. Li, L. Huang, S. Yin, Y. Li, Q. Zhang and H. Li, Mater. Sci. Semicond. Process., 2015, 39, 726 CrossRef CAS.
  28. N. Wang, Y. Zhou, C. Chen, L. Cheng and H. Ding, Catal. Commun., 2016, 73, 74 CrossRef CAS.
  29. D. Ma, J. Wu, M. Gao, Y. Xin, T. Ma and Y. Sun, Chem. Eng. J., 2016, 290, 136 CrossRef CAS.
  30. K. Vignesh, A. Suganthi, B.-K. Min and M. Kang, J. Mol. Catal. A: Chem., 2014, 395, 373 CrossRef CAS.
  31. M. Mousavi and A. Habibi-Yangjeh, Mater. Chem. Phys., 2015, 163, 421 CrossRef CAS.
  32. A. Habibi-Yangjeh and A. Akhundi, J. Mol. Catal. A: Chem., 2016, 415, 122 CrossRef CAS.
  33. M. Gao, D. Zhang, X. Pu, H. Li, D. Lv, B. Zhang and X. Shao, Sep. Purif. Technol., 2015, 154, 211 CrossRef CAS.
  34. C. Li, G. Chen, J. Sun, J. Rao, Z. Han, Y. Hu, W. Xing and C. Zhang, Appl. Catal., B, 2016, 188, 39 CrossRef CAS.
  35. R. P. Panmand, Y. A. Sethi, R. S. Deokar, D. J. Late, H. M. Gholap, J.-Q. Baeg and B. B. Kale, RSC Adv., 2016, 6, 23508 RSC.
  36. J. Zhang, L. Zhang, N. Yu, K. Xu, S. Li, H. Wang and J. Liu, RSC Adv., 2015, 5, 75081 RSC.
  37. W. Xu, J. Fang, Y. Chen, S. Lu, G. Zhou, X. Zhu and Z. Fang, Mater. Chem. Phys., 2015, 154, 30 CrossRef CAS.
  38. T. J. Entradas, J. F. Cabrita, B. Barrocas, M. R. Nunes, A. J. Silvestre and O. C. Monteiro, Mater. Res. Bull., 2015, 72, 20 CrossRef CAS.
  39. N. Qin, Y. Liu, W. Wu, L. Shen, X. Chen, Z. Li and L. Wu, Langmuir, 2015, 31, 1203 CrossRef CAS PubMed.
  40. A. Akhundi and A. Habibi-Yangjeh, Mater. Chem. Phys., 2016, 174, 59 CrossRef CAS.
  41. M. Shekofteh-Gohari and A. Habibi-Yangjeh, Ceram. Int., 2015, 41, 1467 CrossRef CAS.
  42. Z. Xuan, Y. Shiyue, L. Yumei, W. Zuoshan and L. Weifeng, Mater. Lett., 2015, 145, 23 CrossRef CAS.
  43. X. Li and J. Ye, J. Phys. Chem. C, 2007, 111, 13109 CAS.
  44. R. H. Victora, Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 56, 4417 CrossRef CAS.
  45. T. K. Patil and M. I. Talele, Adv. Appl. Sci. Res., 2012, 3, 1702 CAS.
  46. X. Li and J. Ye, J. Phys. Chem. C, 2007, 111, 13109 CAS.
  47. A. Martinez-dela Cruz and U. M. Garcia Perez, Mater. Res. Bull., 2010, 45, 135 CrossRef CAS.
  48. M. Shekofteh-Gohari and A. Habibi-Yangjeh, RSC Adv., 2016, 6, 2402 RSC.
  49. S. R. Morrison, Electrochemistry at semiconductor and oxidized metal electrode, Plenum, New York, 1980 Search PubMed.
  50. Y. Yan, H. Guan, S. Liu and R. Jiang, Ceram. Int., 2014, 40, 9095 CrossRef CAS.
  51. M. Xu, T. Ye, F. Dai, J. Yang, J. Shen, Q. He, W. Chen, N. Liang, J. Zai and X. Qian, ChemSusChem, 2015, 8, 1218 CrossRef CAS PubMed.
  52. Y.-K. Hsu, Y.-C. Chen and Y.-G. Lin, ACS Appl. Mater. Interfaces, 2015, 7, 14157 CAS.
  53. M. Pirhashemi and A. Habibi-Yangjeh, Ceram. Int., 2015, 41, 14383 CrossRef CAS.
  54. M. Shekofteh-Gohari and A. Habibi-Yangjeh, J. Ind. Eng. Chem., 2016, 44, 174 CrossRef CAS.
  55. B. Zhu, P. Xia, W. Ho and J. Yu, Appl. Surf. Sci., 2015, 344, 188 CrossRef CAS.
  56. Y.-P. Sun, W. Ha, J. Chen, H.-Y. Qi and Y.-P. Shi, TrAC, Trends Anal. Chem., 2016, 84, 12 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra12414c

This journal is © The Royal Society of Chemistry 2016