The crystal phase transformation of Ag2WO4 through loading onto g-C3N4 sheets with enhanced visible-light photocatalytic activity

Lei Shi bd, Jinling Goub, Lin Liangc, Fangxiao Wangb and Jianmin Sun*ab
aHarbin Institute of Technology (Shenzhen), Shenzhen Key Laboratory of Organic Pollution Prevention and Control, Shenzhen 518055, China. E-mail: sunjm@hit.edu.cn; Tel: +86 451 86403715
bState Key Laboratory of Urban Water Resource and Environment, MIIT Key Laboratory of Critical Materials Technology for New Energy Conversion and Storage, School of Chemistry and Chemical Engineering, Harbin Institute of Technology, Harbin 150080, China
cSchool of Life Science and Technology, Harbin Institute of Technology, Harbin 150080, China
dCollege of Chemistry, Chemical Engineering and Environmental Engineering, Liaoning Shihua University, Fushun 113001, China

Received 22nd July 2016 , Accepted 26th September 2016

First published on 29th September 2016


Abstract

A g-C3N4/Ag2WO4 composite photocatalyst was synthesized via a simple co-precipitation method at room temperature and was thoroughly characterized by X-ray diffraction, Fourier transform infrared spectroscopy, X-ray photoelectron spectroscopy, transmission electron microscopy, UV-visible diffuse reflectance spectroscopy and photoluminescence spectroscopy. The characterization results indicated that upon depositing Ag2WO4 onto the surface of g-C3N4 its morphology changed from primary rod-like particles to globular nanoparticles and its phase changed from α-type to β-type Ag2WO4. When the composite was used as the catalyst in the photodegradation of Rhodamine B, 40%-g-C3N4/Ag2WO4 exhibited the highest photocatalytic activity with its rate constant was about 10.8 times larger than that of pure g-C3N4 and 3.12 times larger than that of bare Ag2WO4. The enhanced photocatalytic activity can be attributed to the complementary potentials of the conduction bands and valence bands of g-C3N4 and Ag2WO4, which could not only realize the effective separation of photoinduced electron and hole pairs, but also retained the catalysts high stability and photocatalytic performance even after five recycles. Furthermore, a possible photocatalytic mechanism for the degradation of RhB over the g-C3N4/Ag2WO4 composite was proposed according to the results of active species quenching experiments.


1. Introduction

In recent years, semiconductor photocatalytic technology has revealed potential applications in converting solar energy to hydrogen and eliminating environmental contaminants.1,2 Among the numerous photocatalytic materials, TiO2 is one of the most widely studied photocatalysts due to its low price, non-toxicity, strong oxidation ability and good stability. However, the wide band gap of TiO2 (3.2 eV) can be excited by UV light, which only accounts for about 4% of the solar energy. Thus, it makes TiO2 difficult to be applied on large scale. Hence, it is very urgent to develop an efficient and stable photocatalyst with high activity under visible light.

Recently, various silver-based photocatalysts with unique physical and chemical properties, including AgX (X = Cl, Br, I),3–5 Ag2CO3,6,7 Ag2O,8 Ag3PO4,9,10 Ag2CrO4 (ref. 11) and AgVO3,12 have been proved to be efficient photocatalytic materials under visible light. As one of them, silver tungstate (Ag2WO4) also displays excellent photocatalytic capacity.13,14 Generally, Ag2WO4 has three different crystal forms: α, β, and γ phases. α-Ag2WO4 is the thermodynamically stable phase, β-Ag2WO4 and γ-Ag2WO4 are relatively unstable.15 In regard their photocatalytic properties, although pure β-Ag2WO4 exhibited improved photocatalytic activity when compared with pristine α-Ag2WO4 due to the different crystallinities and PL emissions, β-Ag2WO4 is easily inactivated due to photocorrosion during illumination.16 Hence, how to keep β-Ag2WO4 stable is also the important issue. Nowadays, graphite-like carbon nitride (g-C3N4) has been explored as a promising candidate for hydrogen evolution and environmental purification under visible-light irradiation.17–19 Although g-C3N4 possesses good chemical and thermal stability, the use of g-C3N4 in photocatalysis is limited by its high recombination of photogenerated electron–hole pairs. To solve this drawback, abundant strategies have been developed including doping,20 deposition21 and sensitization.22 Constructing a heterostructure is also an effective way for decreasing the recombination rate of the photogenerated carriers. Using the combination of g-C3N4 and other appropriate semiconductors, including AgX/g-C3N4 (X = Br, I),23 MoS2/g-C3N4,24,25 Ag2CrO4/g-C3N4 (ref. 26) and g-C3N4/Bi2O2CO3,27 the complementary potentials of the conduction band and valence band can realize the effective separation of photoinduced electron–hole pairs and thus, the photocatalytic activity would be anticipated to improve.28,29

Therefore, in this paper, g-C3N4/Ag2WO4 composites were successfully synthesized via a facile co-precipitation method. Amazingly, Ag2WO4 deposition on the surface of g-C3N4 lead to a change in its morphology from primary rod-like particles to globular nanoparticles and its phase also changed from the α-type to β-type phase. When the g-C3N4/Ag2WO4 composite was employed as a photocatalyst, it exhibited an enhanced photocatalytic activity when compared with pure g-C3N4 and Ag2WO4, and its photocatalytic property remained stable even after five recycles. Furthermore, a possible photocatalytic mechanism was also proposed based on the obtained experimental results.

2. Experimental

2.1 Preparation of the g-C3N4/Ag2WO4 composite

All reagents except for melamine were purchased from Sinopharm Chemical Reagent Co. Ltd, China. Melamine was supplied from Aladdin. All of the reagents were of analytical grade and used as received without any further purification.

g-C3N4 was prepared by heating melamine in a muffle furnace. 10 g of melamine was placed in a covered crucible and heated to 550 °C for 4 h in a muffle furnace at a heating rate of 2 °C min−1.

The process used to prepare g-C3N4/Ag2WO4 is described as follows: in a typical procedure, 0.16 g of g-C3N4 was dispersed in 20 mL of an aqueous solution containing 0.28 g of Na2WO4·2H2O by ultrasonication and then 20 mL of an aqueous solution containing 0.29 g of AgNO3 was slowly added dropwise into the above solution with stirring for 4 h at room temperature. The mixture was filtered, washed three times with deionized water and ethanol, respectively and dried for 24 h at room temperature to give 40%-g-C3N4/Ag2WO4 (the weight percentage of g-C3N4 in g-C3N4/Ag2WO4 was 40%). Similarly, a series of composites were fabricated by changing the mass of g-C3N4 added and are denoted as 10%-g-C3N4/Ag2WO4, 20%-g-C3N4/Ag2WO4 and 60%-g-C3N4/Ag2WO4. For comparison, pure Ag2WO4 was prepared without the addition of g-C3N4. For reference, a physical mixture of Ag2WO4 and g-C3N4 was prepared with 40% (weight percentage) of g-C3N4 and was denoted as g-C3N4/Ag2WO4 (mixed).

2.2 Characterization

X-ray diffraction (XRD) measurements were carried out on a Bruker D8 Advance X-ray powder diffractometer with Cu Kα radiation (40 kV, 40 mA) for phase identification. Fourier transform infrared spectroscopy (FTIR) was recorded on a Perkin Elmer spectrum 100 FTIR spectrometer using KBr discs. X-ray photoelectron spectroscopy (XPS) measurements were recorded using a Thermo Fisher Scientific Escalab 250. The morphology of the product was examined by Tecnai G2 T12 transmission electron microscopy. The BET surface area was collected at 77 K using a Micromeritics Tristar 3020 analyzer; the sample was outgassed at 120 °C for 12 h prior to the measurement. The UV-vis diffuse reflectance spectra (DRS) were measured using a Perkin Elmer Lambda 750 UV-vis spectrometer. The photoluminescence spectra (PL) were obtained on a Perkin Elmer LS55 spectrometer with an excitation wavelength of 325 nm.

2.3 Evaluation of the photocatalytic activity and stability

The photocatalytic performance of the g-C3N4/Ag2WO4 composites was evaluated by degrading Rhodamine B (RhB) under visible light. Visible light was provided by a 300 W Xe lamp with a UV cut off filter (λ > 400 nm). 50 mg of the g-C3N4/Ag2WO4 composite was dispersed into 50 mL of a RhB solution (5 mg L−1) under magnetic stirring at room temperature. Prior to the light irradiation, the dispersion was kept in the dark for 60 min under magnetic stirring to reach an adsorption–desorption equilibrium. Upon irradiation, the solutions were collected every 10 min, centrifuged to remove the catalyst and then analyzed according to the absorbance at 552 nm using a UV-vis spectrophotometer. The degradation efficiency of the solution was calculated according to the following formula:
 
Degradation efficiency = (C0C)/C0 × 100% (1)
C0 is the initial concentration of RhB and C is the RhB concentration after light irradiation at different times.

Recycling experiments of the g-C3N4/Ag2WO4 composite and the pure Ag2WO4 reference were also investigated under visible light irradiation. After the photodegradation of RhB, the separated photocatalyst was washed several times with water and ethanol, dried at room temperature for 24 h, then applied for the degradation of a fresh 5 mg L−1 aqueous solution of RhB under the same conditions for next cycle.

3. Results and discussion

3.1 The chemical structure and features of the g-C3N4/Ag2WO4 catalyst

To determine the crystal form of the as-prepared samples, the XRD patterns were investigated and shown in Fig. 1A. Pure Ag2WO4 displayed obvious diffraction peaks at 2θ = 16.7°, 30.2°, 31.4°, 33.0° and 45.4°, which were attributed to the (0 1 1), (0 0 2), (2 3 1), (4 0 0) and (4 0 2) diffraction planes of α-Ag2WO4 (JCPDS no. 34-0061).16 For the pure g-C3N4 sample, there were two diffraction peaks at 27.4° and 13.1°, corresponding to the graphite-like stacking and in-plane structural repeating motifs of the conjugated aromatic units of g-C3N4, which were indexed to the (0 0 2) and (1 0 0) planes of g-C3N4.30 Amazingly, for the resultant g-C3N4/Ag2WO4 composites, the crystal planes of α-Ag2WO4 were not observed, however, peaks appearing at 18.4°, 30.0°, 32.2°, 37.4° and 44.6° were indexed to the (0 2 0), (0 2 2), (2 2 0), (0 4 0) and (0 4 2) planes of a β-Ag2WO4 phase (JCPDS no. 33-1195),15 which implied that the α-Ag2WO4 phase was transformed into the β-Ag2WO4 phase in the g-C3N4/Ag2WO4 composites due to the addition of g-C3N4. To support the evidence of phase transformation, the XRD patterns were monitor with the preparation time as shown in Fig. 1B. It was clearly found that β-Ag2WO4 was produced in a short reaction time of 1 h and the XRD typical diffraction peaks of α-Ag2WO4 gradually decreased with increasing reaction time and correspondingly, those ascribed to β-Ag2WO4 were increased. Finally, the α-Ag2WO4 was completely transformed into the β-Ag2WO4 phase in 4 h.
image file: c6ra18648c-f1.tif
Fig. 1 (A) The XRD patterns obtained for (a) 10%-g-C3N4/Ag2WO4, (b) 20%-g-C3N4/Ag2WO4, (c) 40%-g-C3N4/Ag2WO4, (d) 60%-g-C3N4/Ag2WO4, (e) pure Ag2WO4 and (f) g-C3N4. (B) The XRD patterns obtained for 40%-g-C3N4/Ag2WO4 prepared with different reaction times.

Subsequently, the morphologies of pure Ag2WO4, g-C3N4 and 40%-g-C3N4/Ag2WO4 were investigated using TEM as shown in Fig. 2. Clearly, pure Ag2WO4 showed a rod-like morphology (Fig. 2A) in agreement with α-Ag2WO4.15 g-C3N4 exhibited a flat aggregated structure whose surface was smooth (Fig. 2B). As presented in Fig. 2C and D, after g-C3N4 and Ag2WO4 were composited, it was observed that small-sized nanoparticle-like Ag2WO4 were evenly loaded on the surface of the g-C3N4 and the possible mechanism for the formation of the nanoparticle-like Ag2WO4 was tentatively explained as follows: g-C3N4 exhibits a negative Z-potential,31 so Ag+ ions can be adsorbed on the surface of g-C3N4 through electrostatic interactions. After the WO42− ions reacted with Ag+, the as-formed Ag2WO4 was deposited on the g-C3N4. To reduce the interface energy, Ag2WO4 exists as an nanoparticle form in the composite. Hence, the morphology of Ag2WO4 was greatly changed after introduction of the g-C3N4 support, which may be the reason that the α-Ag2WO4 phase was transformed to β-Ag2WO4 in the g-C3N4/Ag2WO4 composites.


image file: c6ra18648c-f2.tif
Fig. 2 The TEM images of (A) pure Ag2WO4, (B) g-C3N4 and (C and D) the 40%-g-C3N4/Ag2WO4 composite.

Moreover, the surface areas of the as-prepared samples were determined. The surface areas of Ag2WO4, g-C3N4, 10%-g-C3N4/Ag2WO4, 20%-g-C3N4/Ag2WO4, 40%-g-C3N4/Ag2WO4 and 60%-g-C3N4/Ag2WO4 were 0.8, 4.7, 0.3, 0.5, 1.1 and 1.9 m2 g−1, respectively. Clearly, the small surface areas play a negligible role in the photocatalytic activity.

The FTIR spectra of pure Ag2WO4, g-C3N4 and the 40%-g-C3N4/Ag2WO4 composite are shown in Fig. 3. In the case of pristine g-C3N4, the peaks at 1247 cm−1, 1325 cm−1, 1408 cm−1 and 1640 cm−1 correspond to the stretching vibrations of C–N heterocyclic compounds, the peak at 808 cm−1 was ascribed to the breathing mode of the triazine unit32 and the peak in the range of 3000–3500 cm−1 was caused by N–H and O–H stretching vibrations.33 When compared with bare g-C3N4, a new peak at 860 cm−1 was observed in the 40%-g-C3N4/Ag2WO4 composite, attributed to the anti-symmetric stretching vibration of the O–W–O bond,34 which further revealed that g-C3N4 was successfully coupled with Ag2WO4 in the composite.


image file: c6ra18648c-f3.tif
Fig. 3 The FTIR patterns obtained for (a) Ag2WO4, (b) 40%-g-C3N4/Ag2WO4 and (c) g-C3N4.

In order to investigate the atomic species and the binding modes of the various atoms in the g-C3N4/Ag2WO4 composites, the XPS spectra of 40%-g-C3N4/Ag2WO4 was recorded and shown in Fig. 4. Obviously, the peaks of C, N, Ag, W and O in the survey spectra indicated that the sample incorporated g-C3N4 and Ag2WO4. The Ag 3d was located at 374.2 eV and 368.2 eV, which were attributed to the Ag(I) oxidation state (Fig. 4B).35 The W 4f5/2 and W 4f7/2 peaks were located at 36.1 and 34.1 eV (Fig. 4C), corresponding to W6+.36 The peaks at 531.1 and 532.3 eV were ascribed to the binding energies of O 1s (Fig. 4D). Besides, the C 1s spectrum exhibited two main peaks at 284.6 and 288.1 eV, the former was identified as carbon atoms in a pure carbonaceous environment, the latter was assigned to the sp2 C atom bonded in N–C[double bond, length as m-dash]N.37 Fig. 4F presents the N 1s XPS spectrum, which was deconvoluted into three peaks with binding energies at 398.8, 400.3 and 401.1 eV, which were attributed to the sp2-bonded N atom to two carbon atoms (C–N[double bond, length as m-dash]C), tertiary nitrogen (N–(C)3) and amino functional groups (N–H) respectively.38,39


image file: c6ra18648c-f4.tif
Fig. 4 The XPS spectra of the 40%-g-C3N4/Ag2WO4 composite: (A) survey spectra, (B) Ag 3d, (C) W 4f, (D) O 1s, (E) C 1s and (F) N 1s.

Furthermore, the optical properties of the 40%-g-C3N4/Ag2WO4 composite were measured by UV-vis DRS. As shown in Fig. 5, the absorption band edges of pure Ag2WO4 and g-C3N4 were estimated to be 425 and 466 nm. After g-C3N4 was coupled with Ag2WO4, there was red-shift in the band edge position to 502 nm, suggesting that the light harvested was shifted to the visible region, which is helpful in regard the utilization of the sunlight. Additionally, the band gap energy (Eg) was obtained according to the empirical formula Eg = 1240/λ (λ represents the absorption edge). Thus, the band gap energies of g-C3N4 and Ag2WO4 were estimated to be about 2.66 eV and 2.92 eV, respectively.


image file: c6ra18648c-f5.tif
Fig. 5 The UV-vis DRS patterns obtained for (a) Ag2WO4, (b) 40%-g-C3N4/Ag2WO4 and (c) g-C3N4.

PL spectroscopy has been widely used to investigate the mitigation, transfer and recombination process of photogenerated electron–hole pairs in a semiconductor. It is generally believed that a weaker PL intensity indicated the lower recombination probability of the photogenerated charge carriers, which results in higher photocatalytic activity. The PL spectra of pure g-C3N4 and the 40%-g-C3N4/Ag2WO4 composite at an excitation wavelength of 325 nm are shown in Fig. 6. The remarkable decrease in the PL intensity of the 40%-g-C3N4/Ag2WO4 implied that photogenerated electron–hole pairs were effectively separated in the g-C3N4/Ag2WO4 composite.


image file: c6ra18648c-f6.tif
Fig. 6 The photoluminescence spectra of (a) pure g-C3N4 and (b) the 40%-g-C3N4/Ag2WO4 composite.

3.2 The photocatalytic activity and stability of the g-C3N4/Ag2WO4 catalyst

The photocatalytic activities of pristine Ag2WO4, g-C3N4, g-C3N4/Ag2WO4 (mixed) and the as-prepared g-C3N4/Ag2WO4 composites were evaluated using the photodegradation of RhB dye under visible light irradiation. As shown in Fig. 7A, the blank experiment indicated that RhB was stable under visible light irradiation. When g-C3N4 and Ag2WO4 were used alone as the photocatalyst, about 32% and 76% of the RhB was decomposed within 40 min. The g-C3N4/Ag2WO4 (mixed) showed a 90% degradation rate for RhB. Attractively, the g-C3N4/Ag2WO4 composites showed significantly enhanced photocatalytic activities under the same conditions when compared with the reference g-C3N4 and Ag2WO4 catalysts, and the highest photocatalytic activity was obtained using 40%-g-C3N4/Ag2WO4, which contained an appropriate ratio of the two composites. To have a better understanding of the reaction kinetics of the RhB degradation catalyzed by the various samples, Fig. 7B shows the relationships between ln(C0/C) and irradiation time. From the linear relationships, the photocatalytic degradation curves obtained in this case all fit with first-order kinetics. Correspondingly, the 40%-g-C3N4/Ag2WO4 composite showed the highest reaction rate constant at 0.113 min−1, which was about 10.8 times larger than that found for pristine g-C3N4 (0.0104 min−1) and was 3.12 times larger than that found for bare Ag2WO4 (0.0362 min−1). For the g-C3N4/Ag2WO4 (mixed) reference, the highest reaction rate constant obtained was 0.0585 min−1.
image file: c6ra18648c-f7.tif
Fig. 7 (A) The degradation rates (B) first-order kinetic plots, (C) rate constants and (D) recycling runs for the photodegradation of RhB: (a) blank catalyst, (b) pristine g-C3N4, (c) Ag2WO4, (d) 10%-g-C3N4/Ag2WO4, (e) 20%-g-C3N4/Ag2WO4, (f) 40%-g-C3N4/Ag2WO4, (g) 60%-g-C3N4/Ag2WO4 and (h) g-C3N4/Ag2WO4 (mixed).

The stability of the as-prepared catalyst is a critical issue for its practical applications. To test the stability of the g-C3N4/Ag2WO4 composites, five cycles of the photocatalytic process were carried out as shown in Fig. 7D. The results showed that the 40%-g-C3N4/Ag2WO4 composite possessed excellent stability and the degradation rate could still reach 90% even after five cycles. While the activity of bare Ag2WO4 catalyst gradually decreased upon recycling; at the 5th recycle, the photodegradation rate was only 45%. Moreover, when the photocatalytic reaction was completed after 5 cycles, the primary light yellow color of the Ag2WO4 catalyst became black, suggesting that Ag nanoparticles were formed due to the reduction of Ag+ by the photoinduced electrons,40 which was further verified by the XRD characterization shown in Fig. 8A.


image file: c6ra18648c-f8.tif
Fig. 8 (A) The XRD patterns obtained for pure Ag2WO4. (B) The XRD patterns and (C) FTIR spectra obtained for 40%-g-C3N4/Ag2WO4 before and after the photocatalytic degradation of RhB.

To further determine the catalyst structure after five photocatalytic cycles, the XRD patterns of pristine Ag2WO4 and the 40%-g-C3N4/Ag2WO4 composite before and after the degradation of RhB were investigated as shown in Fig. 8A and B. In the case of the pristine Ag2WO4 catalyst after five catalytic runs, a new peak at 38.1° attributed to Ag appeared and the Ag intensity increased upon recycling, which was a result of the reduction of Ag+ in Ag2WO4 by the photogenerated electrons. For the resultant 40%-g-C3N4/Ag2WO4 composite, there were almost negligible peaks of Ag at 38.1° in the recycled samples when compared with the spent bare Ag2WO4 catalyst. This phenomenon may be a result of the photogenerated electrons rapidly transferring between g-C3N4 and Ag2WO4 in the g-C3N4/Ag2WO4 composite and thus, the photocorrosion of Ag2WO4 was effectively inhibited, retaining the g-C3N4/Ag2WO4 composites excellent structural stability and photocatalytic activity. In addition, the FTIR spectrum of the g-C3N4/Ag2WO4 composite after photodegradation was almost unaltered as shown in Fig. 8C, further proving the stability of the g-C3N4/Ag2WO4 photocatalyst.

3.3 Detection of the photocatalytic mechanism

In the photodegradation process, some active species such as holes (h+), hydroxyl radicals (˙OH) and superoxide radicals (˙O2) are formed upon light irradiation. The active species generated during the photodegradation process over the 40%-g-C3N4/Ag2WO4 composite were identified using active species scavenging experiments, where ammonium oxalate (AO) was used as a h+ scavenger,41 p-benzoquinone (p-BQ) used as a ˙O2 scavenger42 and t-butanol (t-BuOH) used as an ˙OH quencher.43 As shown in Fig. 9, after the addition of p-BQ into the reaction system, the degradation rate decreased moderately, suggesting that ˙O2 was an active species. For the addition of AO, the degradation efficiency of RhB exhibited a significant reduction, indicating that h+ were a key active species in this system. However, in the presence of t-BuOH, the effect was not obvious, showing that ˙OH was not the main active species. Hence, ˙O2 and h+ were the main active species during the degradation process over the g-C3N4/Ag2WO4 composite.
image file: c6ra18648c-f9.tif
Fig. 9 The influence of active species scavengers on the photodegradation of RhB under visible light irradiation upon the addition of: (a) AO (1 mM), (b) p-BQ (1 mM), (c) t-BuOH (5 mM) and (d) no scavenger.

To clearly understand the photocatalytic process over the g-C3N4/Ag2WO4 composite, the positions of the conduction band (CB) and valance band (VB) of Ag2WO4 and g-C3N4 were calculated using the following empirical equations:44,45

 
EVB = XEe + 0.5Eg (2)
 
ECB = EVBEg (3)
where EVB is the VB potential, ECB is the CB potential, X is the absolute electronegativity of the semiconductor, Ee is the energy of free electrons on hydrogen scale (4.5 eV) and Eg is the band gap. The X values of Ag2WO4 and g-C3N4 are 6 and 4.73, respectively.46,47 Thus, the calculated CB and VB positions of g-C3N4 were −1.1 eV and 1.56 eV, respectively. In the case of Ag2WO4, the CB and VB positions were located at +0.04 eV and 2.96 eV, respectively. Hence, the photoinduced electrons (e) on the CB of g-C3N4 can react with O2 to form ˙O2 radicals because the CB position of g-C3N4 was more negative than the O2/˙O2 potential (−0.33 eV vs. NHE).48 However, its VB potential was lower than the standard redox potential of ˙OH/H2O (2.68 eV vs. NHE),49,50 indicating that the photoinduced holes (h+) in the VB of g-C3N4 do not oxidize the adsorbed H2O to form ˙OH. On the contrary, for Ag2WO4, its CB potential was more positive than the of O2/˙O2 potential, implying it does not form ˙O2 by e with O2, whereas, its VB potential was higher than the standard redox potential of ˙OH/H2O, which means that the h+ of VB can oxidize the adsorbed H2O to form ˙OH. Finally, based on the band gap structures of the g-C3N4/Ag2WO4 composite and the effects observed from the scavenger experiments, a possible pathway for the photodegradation of RhB over the g-C3N4/Ag2WO4 composite was proposed and is shown in Fig. 10. Under the visible-light irradiation, g-C3N4 and Ag2WO4 were simultaneously excited to form electron–hole pairs. Because the CB and VB of Ag2WO4 lie below those of g-C3N4, the e in the CB of Ag2WO4 can transfer and recombine with the h+ in the VB of g-C3N4. As such, the some of the remaining h+ left in the VB of Ag2WO4 can directly oxidize the organic pollutant, the remaining h+ will reacted with water to form ˙OH radicals. At the same time, the photoinduced electrons on the CB of g-C3N4 will react with O2 to generate highly active ˙O2. Therefore, a Z-scheme mechanism was established, the photogenerated electron–hole pairs were effectively separated by the above pathway. The main active species, such as ˙O2 and h+, which were generated in the photocatalysis process are responsible for the degradation of RhB.


image file: c6ra18648c-f10.tif
Fig. 10 A possible pathway for the photodegradation of RhB over the g-C3N4/Ag2WO4 composite.

4. Conclusions

In summary, a stable and novel g-C3N4/Ag2WO4 composite photocatalyst has been synthesized successfully via a facile co-precipitation method at room temperature. Due to the introduction of g-C3N4, the morphology of the Ag2WO4 changed from primary rod-like particles to globular nanoparticles, which resulted in its phase change from the α-type to β-type. The g-C3N4/Ag2WO4 composite exhibited an enhanced photocatalytic activity for the degradation of RhB when compared with the reference samples of bare g-C3N4, pristine Ag2WO4 and its physical mixture g-C3N4/Ag2WO4 (mixed). The improved photocatalytic activity was mainly attributed to the suitable band edge positions of Ag2WO4 and g-C3N4, which contributed to accelerating the separation of the photogenerated electron–hole pairs. Besides, the transformed globular nanoparticles of β-Ag2WO4 on the g-C3N4 support also helped the g-C3N4/Ag2WO4 composite retain its photocatalytic and structural properties even after five recycles. Hence, the as-prepared g-C3N4/Ag2WO4 composite photocatalyst will have great potential due to its effective pollutant purification applications.

Acknowledgements

We sincerely acknowledge the financial supports from the National Natural Science Foundation of China (21373069, 21673060), Fund for Research and Development of Science and Technology in Shenzhen (No. JCYJ20160427184531017, No. ZDSYS201603301417588) and the State Key Lab of Urban Water Resource and Environment of Harbin Institute of Technology (HIT2015DX08).

References

  1. S. Cao, J. Low, J. Yu and M. Jaroniec, Adv. Mater., 2015, 27, 2150–2176 CrossRef CAS PubMed.
  2. X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959 CrossRef CAS PubMed.
  3. P. Wang, B. B. Huang, X. Y. Qin, X. Y. Zhang, Y. Dai and J. Y. Wei, Angew. Chem., Int. Ed., 2008, 47, 7931–7933 CrossRef CAS PubMed.
  4. D. Wang, Y. Duan, Q. Luo, X. Li and L. Bao, Desalination, 2011, 270, 174–180 CrossRef CAS.
  5. H. Yu, L. Liu, X. Wang, P. Wang, J. Yu and Y. Wang, Dalton Trans., 2012, 41, 10405–10411 RSC.
  6. H. J. Dong, G. Chen, J. X. Sun, C. M. Li, Y. G. Yu and D. H. Chen, Appl. Catal., B, 2013, 134, 46–54 CrossRef.
  7. C. Dong, K. L. Wu, X. W. Wei, X. Z. Li, L. Liu, T. H. Ding, J. Wang and Y. Ye, CrystEngComm, 2014, 16, 730–736 RSC.
  8. X. Wang, S. Li, H. Yu, J. Yu and S. Liu, Chem.–Eur. J., 2011, 17, 7777–7780 CrossRef CAS PubMed.
  9. Y. P. Bi, S. X. Ouyang, N. Umezawa, J. Y. Cao and J. H. Ye, J. Am. Chem. Soc., 2011, 133, 6490–6492 CrossRef CAS PubMed.
  10. G. Chen, M. Sun, Q. Wei, Y. Zhang, B. Zhu and B. Du, J. Hazard. Mater., 2013, 244–245, 86–93 CrossRef CAS PubMed.
  11. D. Xu, B. Cheng, S. Cao and J. Yu, Appl. Catal., B, 2015, 164, 380–388 CrossRef CAS.
  12. W. Zhao, Y. Guo, Y. Faiz, W. T. Yuan, C. Sun, S. M. Wang, Y. H. Deng, Y. Zhuang, Y. Li, X. M. Wang, H. He and S. G. Yang, Appl. Catal., B, 2015, 163, 288–297 CrossRef CAS.
  13. J. Li, C. Yu, C. Zheng, A. Etogo, Y. Xie, Y. Zhong and Y. Hu, Mater. Res. Bull., 2015, 61, 315–320 CrossRef CAS.
  14. R. X. Zhang, H. Cui, X. Yang, H. Tang, H. Liu and Y. Li, Micro Nano Lett., 2012, 7, 1285–1288 CAS.
  15. X. F. Wang, C. Fu, P. Wang, H. G. Yu and J. G. Yu, Nanotechnology, 2013, 24, 165602–165607 CrossRef PubMed.
  16. H. H. Chen and Y. M. Xu, Appl. Surf. Sci., 2014, 319, 319–323 CrossRef CAS.
  17. X. C. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domenet and M. Antonietti, Nat. Mater., 2009, 8, 76–80 CrossRef CAS PubMed.
  18. Y. Wang, X. Wang and M. Antonietti, Angew. Chem., Int. Ed., 2012, 51, 68–89 CrossRef CAS PubMed.
  19. S. Yin, J. Han, T. Zhou and R. Xu, Catal. Sci. Technol., 2015, 5, 5048–5061 CAS.
  20. S. C. Yan, Z. S. Li and Z. G. Zou, Langmuir, 2010, 26, 3894–3901 CrossRef CAS PubMed.
  21. L. Ge, C. Han, J. Liu and Y. Li, Appl. Catal., A, 2011, 409–410, 215–222 CrossRef CAS.
  22. S. Min and G. Lu, J. Phys. Chem. C, 2012, 116, 19644–19652 CAS.
  23. H. Xu, J. Yan, Y. Xu, Y. Song, H. Li, J. Xia, C. Huang and H. Wan, Appl. Catal., B, 2013, 129, 182–193 CrossRef CAS.
  24. J. Yan, Z. Chen, H. Ji, Z. Liu, X. Wang, Y. Xu, X. She, L. Huang, L. Xu, H. Xu and H. Li, Chem.–Eur. J., 2016, 22, 4764–4773 CrossRef CAS PubMed.
  25. J. Li, E. Liu, Y. Ma, X. Hu, J. Wan, L. Sun and J. Fan, Appl. Surf. Sci., 2016, 364, 694–702 CrossRef CAS.
  26. L. Shi, L. Liang, F. X. Wang, M. S. Liu and J. M. Sun, Dalton Trans., 2016, 45, 5815–5824 RSC.
  27. W. Shan, Y. Hu, Z. Bai, M. Zheng and C. Wei, Appl. Catal., B, 2016, 188, 1–12 CrossRef CAS.
  28. S. C. Yan, S. B. Lv, Z. S. Li and Z. G. Zou, Dalton Trans., 2010, 39, 1488–1491 RSC.
  29. J. C. Wang, H. C. Yao, Z. Y. Fan, L. Zhang, J. S. Wang, S. Q. Zang and Z. J. Li, ACS Appl. Mater. Interfaces, 2016, 8, 3765–3775 CAS.
  30. P. Niu, G. Liu and H. M. Cheng, J. Phys. Chem. C, 2012, 116, 11013–11018 CAS.
  31. Y. Xu, H. Xu, L. Wang, J. Yan, H. Li, Y. Song, L. Huang and G. Cai, Dalton Trans., 2013, 42, 7604–7613 RSC.
  32. L. Shi, L. Liang, J. Ma, F. X. Wang and J. M. Sun, Dalton Trans., 2014, 43, 7236–7244 RSC.
  33. Y. Zhao, Z. Liu, W. Chu, L. Song, Z. Zhang, D. Yu, Y. Tian, S. Xie and L. Sun, Adv. Mater., 2008, 20, 1777–1781 CrossRef CAS.
  34. L. S. Cavalcante, M. A. P. Almeida, W. Avansi Jr, R. L. Tranquilin, E. Longo, N. C. Batista, V. R. Mastelaro and M. S Li, Inorg. Chem., 2012, 51, 10675–10687 CrossRef CAS PubMed.
  35. Z. Lou, B. Huang, Z. Wang, X. Ma, R. Zhang, X. Zhang, X. Qin, Y. Dai and M. H. Whangbo, Chem. Mater., 2014, 26, 3873–3875 CrossRef CAS.
  36. J. Bao, S. Guo, J. Gao, T. Hu, L. Yang, C. Liu, J. Peng and C. Jiang, RSC Adv., 2015, 5, 97195–97204 RSC.
  37. L. Shi, L. Liang, F. X. Wang, M. S. Liu, S. F. Zhong and J. M. Sun, Catal. Commun., 2015, 59, 131–135 CrossRef CAS.
  38. Y. Yang, W. Guo, Y. Guo, Y. Zhao, X. Yuan and Y. Guo, J. Hazard. Mater., 2014, 271, 150–159 CrossRef CAS PubMed.
  39. L. Pi, R. Jiang, W. Zhou, H. Zhu, W. Xiao, D. Wang and X. Mao, Appl. Surf. Sci., 2015, 358, 231–239 CrossRef CAS.
  40. X. Y. Zhang, J. D. Wang, J. K. Liu, X. H. Yang and Y. Lu, CrystEngComm, 2015, 17, 1129–1138 RSC.
  41. Y. Tian, B. Chang, J. Lu, J. Fu, F. Xi and X. Dong, ACS Appl. Mater. Interfaces, 2013, 5, 7079–7085 CAS.
  42. L. Shi, L. Liang, J. Ma, Y. N. Meng, S. F. Zhong, F. X. Wang and J. M. Sun, Ceram. Int., 2014, 40, 3495–3502 CrossRef CAS.
  43. H. Lee and W. Choi, Environ. Sci. Technol., 2002, 36, 3872–3878 CrossRef CAS PubMed.
  44. L. Shi, L. Liang, F. X. Wang, M. S. Liu and J. M. Sun, RSC Adv., 2015, 5, 101843–101849 RSC.
  45. C. Xing, Z. Wu, D. Jiang and M. Chen, J. Colloid Interface Sci., 2014, 433, 9–15 CrossRef CAS PubMed.
  46. K. Vignesh and M. Kang, Mater. Sci. Eng., B, 2015, 199, 30–36 CrossRef CAS.
  47. W. Zhang, Y. Sun, F. Dong, W. Zhang, S. Duan and Q. Zhang, Dalton Trans., 2014, 43, 12026–12036 RSC.
  48. Y. Hong, Y. Jiang, C. Li, W. Fan, X. Yan, M. Yan and W. Shi, Appl. Catal., B, 2016, 180, 663–673 CrossRef CAS.
  49. S. Kumar, T. Surendar, A. Baruah and V. Shanker, J. Mater. Chem. A, 2013, 1, 5333–5340 CAS.
  50. L. Shi, L. Liang, J. Ma, F. X. Wang and J. M. Sun, Catal. Sci. Technol., 2014, 4, 758–765 CAS.

Footnote

The authors contributed equally to this study.

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