Characterization and mechanism analysis of AgBr mixed cuboid WO3 rods with enhanced photocatalytic activity

Shanshan Yao*a, Sikang Xuea, Junhao Zhangb and Xiangqian Shena
aInstitute for Advanced Materials, College of Materials Science and Engineering, Jiangsu University, Zhenjiang, Jiangsu 212313, P. R. China. E-mail: yaosshan@ujs.edu.cn
bSchool of Environmental and Engineering, Jiangsu University of Science and Technology, Zhenjiang, Jiangsu 212003, P. R. China

Received 25th June 2016 , Accepted 26th September 2016

First published on 26th September 2016


Abstract

Cuboid rods of monoclinic WO3 (m-WO3) were successfully prepared by employing preoxo-polytungstic acid as the precursor via a hydrothermal method. AgBr/m-WO3 photocatalysts with varying loadings of AgBr were synthesized and characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM). Photocatalytic degradation of rhodamine B (rhB), methyl orange (MO) and methyl blue (MB) were carried out to evaluate the photocatalytic activity of AgBr/m-WO3 under visible-light irradiation (λ ≥ 420 nm). Among such catalysts, the AgBr (30 wt%)/m-WO3 exhibits the highest visible-light-responsible photoactivity. However, these AgBr/m-WO3 materials gradually lost their photoactivity in the cycling photocatalytic tests. Possible mechanisms for both the enhance photocatalytic activity and deactivation of the AgBr/m-WO3 were proposed on basis of theoretical speculation and experimental observations.


Introduction

Metal–oxide semiconductors may be used in a wide variety of applications including photocatalytic degradation of organic compounds and gas monitoring and alarm systems.1,2 Materials based on tungsten trioxide (WO3) have been studied extensively because of their special electronic and optoelectronic properties, which have enormous potential applications in fields ranging from condensed-matter physics to solid-state chemistry,3 such as photo-electrochemical energy conversion,4 gas-sensors,5 photo-catalysts,6 lithium-ion batteries,7 solar cells,8 electron emitters,9 optical storage media.10 The band gap of WO3 is situated within the solar spectrum range, making it an important material for photocatalytic activity and photoconductivity.11 However, like other simple binary metal oxides, WO3 has a deep valence band which is mainly composed of O2p orbitals. The deep valence band combined with the small band gap results in a low conduction band level, which limits the photocatalyst to react with electron acceptors12–14 and then increases the recombination of photogenerated electron–hole pairs. This was one of the principles to improve the photocatalytic performance of WO3 for increasing the efficiency of electron–hole separation. Loading silver halides on the photocatalyst has been proved as effectual approach recently.16–20 Silver halides (AgX, X = Cl, Br, I) are important photosensitive semiconductors extensively employed in photography field. Under light irradiation AgX can absorb photons to generate electrons and holes. Therefore, it will be a promising way to utilize AgX as photosensitive components in photocatalytic field.21,22

For the above mentioned applications, the microstructures of photocatalyst also influence the photocatalytic activity significantly. WO3 has been produced by various synthetic routs such as electrospinning,23 chemical and physical vapor deposition,24 co-precipitation,25 sol–gel processing,26 thermal processing,27 microwave synthesis28 and solvothermal/hydrothermal methods.29 Herein, the cuboid rods of monocline WO3 (m-WO3) has been successful control using peroxo-polytungstic acid as the precursor under hydrothermal conditions. Furthermore, considering that AgX (X = Cl, Br) have highly photosensitive materials, we set out to investigate the photocatalytic performance of AgBr/m-WO3 and its working mechanism as visible-light-responsive photocatalysts.

Here, we reported the photocatalytic of AgBr/m-WO3 composite catalyst in the degradation reaction of rhodamine B (rhB), methyl orange (MO) and methyl blue (MB) under the simulated sunlight irradiation. The conduction band (CB) bottom and valence band (VB) tope of m-WO3 lie below the CB bottom and VB to of AgBr, respectively, which will result in the highly efficient separation of photoinduced electrons and holes. In addition, AgBr/m-WO3 can easily transform to be Ag/AgBr/m-WO3 system in the early stage of the photocatalytic reaction, which in the role of Ag nanoparticles. The stability of the photocatalyst was also investigated via the repetition tests. However, AgBr/m-WO3 gradually lost their photoactivity in the cycling photocatalytic tests. The possible mechanisms for the formation of photocatalysts, enhanced photocatalytic properties and deactivation of the AgBr/m-WO3 were investigated, respectively.

Experimental

Preparation and characterization

Tungsten powder, hydrogen peroxide, AgNO3, glycerol and cetyltrimethyl ammonium bromide (CTAB) are of analytical purity from Wako Pure Chemical Industries and used for the experiment without further purification. Deionized water was used throughout this study.

The cuboid rods of m-WO3 were synthesized using peroxo-polytungstic acid as precursor in hydrothermal reaction according to the previous report with minor changes.30 Typically, finely powdered tungsten (5 g, 12 μm) was slowly added to hydrogen peroxide (20 ml, 30 wt%) in a cold water bath (10 °C) to stabilize the reaction temperature below 30 °C. This produced a clear and colourless peroxo-polytungstic acid solution.

The as-prepared solution was then diluted in deionized water to obtain a tungsten concentration of 0.127 mol L−1. After that, the diluted solution was transferred into a Telfon-lined stainless steel autoclave (60 ml, filling ration 75%), and the autoclave was then sealed and maintained at 180 °C for 4 days. After cooled to room temperature, the as-prepared precipitates were isolated by centrifugation and repeatedly washed with absolute ethanol and deionized water for several times. Finally, the as-obtained precipitates were dried at 40 °C in a drying cabinet for 6 h. Then the cuboid rods of monoclinic WO3 (m-WO3) was obtained.

In a typical synthesis, the as-synthesized m-WO3 (0.5 g) was added to distilled water (30 ml), and the suspension was sonicated for 30 min. The glycerol (5 ml) was added to suspension, and the mixture was sonicated for 30 min. The solution was mixed various amounts of AgNO3 in order to obtain AgBr/m-WO3 composite with AgBr weight percentage of 5.0, 10.0, 20.0, 30.0, 40.0. The mixed solution was vigorously stirred for 30 min in the dark prior to adding to 25 ml deionized water containing stoichiometric amounts of CTAB. The resulting suspension was stirred at room temperature for 12 h. All the above processes were performed in the dark. The products were filtered, washed with absolute ethanol and deionized water. Finally, the precipitates were dried at 40 °C for 12 h in vacuum. Then, the AgBr/m-WO3 composite catalysts were obtained. Fig. 1 describes the detailed preparation process.


image file: c6ra16442k-f1.tif
Fig. 1 The schematic diagram of the preparation procedure of AgBr/m-WO3 photocatalyst.

The phase compositions of the AgBr/m-WO3 samples were determined by X-ray diffraction (XRD) analysis (RIGAKU Ultima IV, Japan) using Cu Kα radiation (λ = 1.5406 Å) with a scan speed of 10° min−1 ranging from 10° to 80°. The morphology of the sample was observed using a field-emission electron scanning microscope (FE-SEM, JEOL, JSM-6500F). Specific surface area was measured by nitrogen physisorption using the Brunauer–Emmett–Teller (BET) method (BELSORP-max, BEL Japan, Inc.). UV-vis diffuse reflectance spectroscopy (DRS) measurements were carried out using a Hitachi UV-3100 UV-vis spectrophotometer equipped with an integrating sphere attachment. The analysis range was from 300 to 650 nm. The band gap energy of the prepared catalysts can be calculated by the following formula:

 
αhν = A(Eg)n/2 (1)
where α, ν, Eg and A are absorption coefficient, light frequency, band gap energy, and a constant, respectively. Among them, n is determined by the type of optical transition of a semiconductor (n = 1 for direct transition and n = 4 for indirect transition). The values of n for AgBr and m-WO3 are 4 and 1, respectively. X-ray photoelectron spectroscopy (XPS) measurements were done on a KRATOA XSAM800 XPS system with Mg Kα source. All the binding energies were referenced to the C 1s peak at 284.8 eV of the surface adventitious carbon.

Evaluation of photocatalytic activities of AgBr/m-WO3

The photocatalytic properties of the as-obtained AgBr/m-WO3 samples were tested using a home-made system equipped with a 400 W metal–halogen lamp (IWASKI ELECTRIC Co., Ltd, Japan) and ultraviolet cut-off filter (UV-cut 420), providing a visible-light source (λ ≥ 420 nm). The aqueous solutions of rhB, MO and MB were used as the target pollutants to evaluate the visible-light-driven photocatalytic performance of these samples. The catalyst-free dye solution was analyzed with a HITACHI 200-10 spectrophotometer. All the experiments were conducted in ambient conditions.

Typically, 100 mg of photocatalysts were firstly dispersed in a 100 ml beaker, containing 100 ml of 10 mg L−1 rhB (or MO, MB) aqueous solutions under magnetic stirring in the dark. The distance between the bottom of the lamp and the top of the solution was 20 cm. The above suspensions with dye molecules were kept stirring in the dark for 60 min to reach an adsorption–desorption equilibrium of dye molecules on the photocatalysts. Then, the suspensions with photocatalysts and dye molecules were exposed to the visible light irradiation. At certain time intervals, 3 ml solution were sampled and centrifuged to remove the photocatalyst particles. The top transparent solutions obtained were then transferred to quartz cuvette to measure their absorption spectra in a wavelength range of 400–800 nm. The relative concentrations (C/C0) of the rhB, MO and MB solutions were determined by the absorbance (A/A0) at 554 nm, 464 nm and 664 nm, respectively, because of the relationship of C = kappA. Here, kapp is a constant, A is the absorbance of the dye aqueous solution at time t and A0 is the absorbance at the beginning of the visible-light irradiation. Similarly, C is the concentration of the dye aqueous solution at time t, and C0 is the concentration at the beginning of the visible-light irradiation.

For the stability measurements, 100 mg optimum photocatalyst (30 wt% AgBr/m-WO3) was dispersed in 100 ml of rhB solution (10 mg L−1) and the photoactivity tested as described above. After each photoactivity test, both the separated catalyst particles and the solution were returned to the reduction system. Photocatalysis stability tests were performed on the optimum composition in seven separated runs. Once these tests were completed the photocatalyst was reclaimed, washed, dried overnight at 40 °C, and then analyzed by XRD.

Results and discussion

The stable monoclinic WO3 (m-WO3) can have a ReO3-type structure (corner-sharing arrangement of octahedral). Each W ion is surrounded by six equidistant oxygen ions. An infinite array of corner-sharing WO6-octahedra is formed like in Fig. 2. The W ions occupy the corners of a primitive unit cell, and O ions bisect the unit cell edges.
image file: c6ra16442k-f2.tif
Fig. 2 Crystal structure of m-WO3.

In order to confirm the crystalline structure of prepared AgBr/m-WO3, powder XRD study was carried out. Fig. 3 shows the XRD patterns of fresh AgBr/m-WO3 with different silver halide contents. The patterns shows that WO3 substrates were monoclinic WO3 (JCPDS: 71-2141) and AgBr was of face-centered cubic structures. As evident form Fig. 3, with increasing silver halide content, the intensities of diffraction peaks of AgBr increased whereas those of substrate decreased simultaneously. The lattice parameters of AgBr/m-WO3, as listed in Table 1, are in good agreement with those in the literature. The average crystalline sizes of AgBr on the surface of m-WO3 were calculated by using Scherrer equation, respectively.31

 
image file: c6ra16442k-t1.tif(2)
where D is taken as crystalline size, K is a constant equals to 0.89, λ is 1.5406 Å, β is the FWHM measured in radians on the 2θ scale, θ is the Bragg angle for the diffraction peaks. The results were also showed in Table 1.


image file: c6ra16442k-f3.tif
Fig. 3 XRD patterns of AgBr/m-WO3 samples with different AgBr loadings.
Table 1 Lattice parameters values of AgBr/m-WO3 photocatalysts calculated from the XRD patters
Photocatalyst Lattice parameters Crystallite size of AgBra (nm)
a (Å) b (Å) c (Å) V3)
a Calculated using Scherrer equation on (2 0 0) diffraction of AgBr.
m-WO3 (JCPDS: 71-2141) 7.297 7.539 7.698 422.88
m-WO3 7.293 7.513 7.678 420.58
AgBr (5 wt%)/m-WO3 7.278 7.497 7.665 418.98 35.78
AgBr (10 wt%)/m-WO3 7.282 7.503 7.670 419.33 37.91
AgBr (20 wt%)/m-WO3 7.286 7.505 7.674 419.68 40.23
AgBr (30 wt%)/m-WO3 7.291 7.508 7.682 420.06 41.58
AgBr (40 wt%)/m-WO3 7.295 7.515 7.685 421.07 43.03


The SEM images and EDX result of AgBr (30 wt%)/m-WO3 photocatalyst were presented in Fig. 4. From Fig. 4a, the cuboid rods of monoclinic of WO3 substrates have been prepared by hydrothermal method. The loaded AgBr nanoparticles with average diameter of about 40 nm were uniformly dispersed on the surface of WO3. The EDX spectrum (Fig. 4b) shows that the composite catalysts are composed W, O, Ag and Br. It would be the direct evidence to prove the AgBr nanoparticles co-dispersed on the surface of WO3.


image file: c6ra16442k-f4.tif
Fig. 4 SEM images of (a) AgBr (30 wt%)/m-WO3 and (b) EDX spectrum of the sample.

Fig. 5 shows the UV-vis diffuse reflectance spectra of pure WO3, AgBr and AgBr/m-WO3. It is shown that m-WO3, AgBr and AgBr/m-WO3 all exhibited absorption in the visible light region, among which the absorption edge of AgBr/m-WO3 (471 nm) was lower than that of pure AgBr (479 nm) and m-WO3 (490 nm). This also suggests that the crystalline size of AgBr supported on the surface of m-WO3 is smaller than that of pure AgBr in our research.


image file: c6ra16442k-f5.tif
Fig. 5 DRS spectra of m-WO3, AgBr and AgBr/m-WO3 (AgBr 30 wt%).

According to eqn (1), Eg of AgBr was determined from the plot of (αhν)1/2 versus energy () (Fig. 6a). From the tangent line of the curve, extrapolated to the axis intercept, Eg of AgBr was found to be 2.60 eV. Similarly, Eg of m-WO3 was 2.69 eV (Fig. 6b).


image file: c6ra16442k-f6.tif
Fig. 6 Plots of (a) (αhν)1/2 versus energy () for AgBr and (b) (αhν)2 versus energy () for m-WO3.

BET surface area of as synthesized samples compared with that of commercial WO3 is shown in Table 2. It is revealed form the table that the composite photocatalysts have smaller BET surface area as compared to commercial WO3. Moreover, BET surface area decreased with increase in AgBr concentrations. Decrease in surface area might be due to the occupation of pores and interstitial lattice positions of m-WO3 by AgBr nanoparticles and it is the common characteristic of heterogeneous phase of the composite materials. Similar results have been reported for composite photocatalysts.15,32

Table 2 BET analysis and photodegradation rate constants of the as-synthesized AgBr/m-WO3 samples
Photocatalyst BET surface area (m2 g−1) kapp(rhB) (min−1) kapp(MO) (min−1) kapp(MB) (min−1)
Commercial WO3 6.179 0.00012 0.00013 0.00012
m-WO3 5.690 0.00036 0.00024 0.00036
AgBr (5 wt%)/m-WO3 5.981 0.00298 0.00220 0.00108
AgBr (10 wt%)/m-WO3 5.682 0.00591 0.00273 0.00188
AgBr (20 wt%)/m-WO3 5.611 0.01287 0.01056 0.00245
AgBr (30 wt%)/m-WO3 5.220 0.02267 0.01637 0.00588
AgBr (40 wt%)/m-WO3 4.704 0.01612 0.01199 0.00361


The photocatalystic activities of as-prepared samples were evaluated by the degradation of rhB, MO and MB under visible-light irradiation. Fig. 7a displays the degradation of rhB in the presence with different AgBr contents. Under visible light, WO3 had weaker photocatalytic activity for rhB though it can absorb visible light. The photocatalytic activity of AgBr/m-WO3 increased remarkably with increasing AgBr content, but at the higher AgBr level the catalyst photocatalytic activity decreased slightly, suggesting that the optimal AgBr content in AgBr/m-WO3 existed when the weight ratio 0.3. Moreover, the similar change tendency of degradation efficiency was also displayed in the process of MO (Fig. 7b) and MB (Fig. 7c) degradation as that of rhB degradation. The degradation efficiency of rhB was enhanced and remained unchanged after irradiation for 240 min with increasing AgBr content. According to Langmuir–Hinshelwood (L–H) kinetic model, the kapp of AgBr/m-WO3 were respectively calculated and displayed in Table 2. The results clearly demonstrate the optimum AgBr content was 0.3 with the maximal degradation rate. Loading of AgBr greater than 30 wt% leads to weaker catalytic activity because the majority of AgBr is not in close contact with the WO3 and may actively screen light irradiation. Also, the AgBr/m-WO3 exhibited excellent visible-light-driven photocatalytic efficiencies for the degradation of dyes, which was much higher the TiO2 (P25). In AgBr/m-WO3 system, only AgBr is the active component that absorbs visible light to initiate the photocatalytic reaction, so the AgBr content determines the number of electron–hole pairs that participate in the degradation of dyes.19,33


image file: c6ra16442k-f7.tif
Fig. 7 The degradation of rhB (a) by different photocatalysts with the same weigh of each visible-light-active component (image file: c6ra16442k-u1.tif) commercial WO3; (image file: c6ra16442k-u2.tif) prepared m-WO3; (image file: c6ra16442k-u3.tif) P25; (image file: c6ra16442k-u4.tif) 5% wt AgBr/m-WO3; (image file: c6ra16442k-u5.tif) 10% wt AgBr/m-WO3; (image file: c6ra16442k-u6.tif) 20% wt AgBr/m-WO3; (image file: c6ra16442k-u7.tif) 30% wt AgBr/m-WO3; (image file: c6ra16442k-u8.tif) 40% wt AgBr/m-WO3. The degradation of MO (b) and MB (c) were under the same conditions as rhB.

As a hybrid semiconductor, the photocatalytic activity of the composite relates directly to the electronic band structures of individual components, which determine the excitation, transportation, and ultimate fate of the photogenerated charge carriers. According to empirical Mulligen electronegativity principle, the conduction and valence band positions of AgBr and WO3 as the vital factors determined their photo-reactivity. Despite this methodology cannot give the discrete energy levels of individual semiconductors, it is plausible for understanding the transportation of photogenerated charge carriers and is helpful to roughly judge the thermodynamic trends of photoreaction. Because of the complexity of mechanistic energy principle, the charge carrier transportation mechanism based on energy bias law of heterojunctions has been widely accepted to simplify the depiction of photogenerated charge transportation.34

WO3 has more positive conduction band (CB) bottom and valence band (VB) top than AgBr.33,34 The possible pathway for the photocatalytic degradation of dyes with AgBr/m-WO3 photocatalyst was proposed as follows (Fig. 8): during the photocatalytic oxidation process, the AgBr/m-WO3 system was transformed to be Ag/AgBr/m-WO3 system. Ag, AgBr and WO3 can be simultaneously excited to form electron–hole pairs under visible-light irradiation. The photo-excited electrons (e) on CBAgBr would prefer to flow down to the CBWO3 crossing the interface, while the photogenerated positive carriers (holes, h+) on VBWO3 would transfer to the VBAgBr.35,36 Such transportation of the photogenerated carriers could either extend their transfer path or stabilize the photogenerated holes in the VBWO3, leading to prolonged lifetime of the charge carriers and successfully hindering the unfavorable recombination of electron–hole (e–h+) pairs. Meanwhile, the photogenerated electrons transfer form the CB bottom of AgBr to that of m-WO3 or are trapped by Ag NPs formed on AgBr particles. At the same time photogenerated holes also move in the opposite direction from the VB top of m-WO3 to that of AgBr. Besides the intrinsic photoexcitation of AgBr/m-WO3, the organic dyes may be excited by light irradiation (photo-activation) and the excited dye* will inject electron onto CB of either AgBr or m-WO3, and consequently dye* gets back to dye.


image file: c6ra16442k-f8.tif
Fig. 8 Schematic electronic band diagram of the AgBr/m-WO3 system.

From the photoelectrochemistry point of view, the photogenerated holes on VBAgBr and VBWO3 may oxidize organic dyes, while the electrons on the CBAgBr rather than CBWO3 are able to reduce soluble O2 to O2, an oxidative species that can break down dyes. AgBr are highly dispersed and closely contacted with WO3 that facilitated the transportation of charge carriers, meanwhile, the holes on VBWO3 are positive (more oxidative) than those on VBAgBr, hence such AgBr/m-WO3 showed higher activity. The VBWO3 and VBAgBr situate more positive than the standard redox potentials of OH˙/OH (1.99 eV) and H2O2 (1.77 eV), suggesting their photogenerated holes are far more oxidative than both OH˙ and H2O2.37,38 In addition, the photogenerated hole can not oxidize OH to form OH˙ because of the more negative (reductive) potential of W6+/W5+ (1.10 eV).39 The reduction–oxidation (redox) half reaction W6+/W5+ would not happen since the WO3 is much stable both in air and solution. Therefore, the photodegradation of organic dyes on the AgBr/m-WO3 can be associated with the photogenerated holes on the valance bands of AgBr and WO3.

Based on the discussions afore, the mechanism of photocatalytic degradation of organic dyes on the AgBr/m-WO3 may be proposed, as described in the eqn (3)–(7). We suppose electrons firstly generated, under visible-light irradiation, from the VB of the AgBr/m-WO3 and photoexcited organic dyes, then hop onto the CBs of the composite photocatalysts. The CB electrons would react with soluble oxygen to form reactive O2 radicals, which either attack organic dyes or pass to the CBWO3, correspondingly, the positively charged holes were left on VBs and excited dye (dye*). The photocatalytic degradation reaction would subsequently take place, in which the dye* would be transformed and decomposed into intermediate products and eventually converted to CO2 and H2O by both the VB holes and O2 radicals (eqn (8) and (9)). At the same time, electrons in the VB of AgBr could be excited up to a higher potential edge under visible-light illumination with energy less than 2.95 eV (λ > 420 nm), and these electrons will be trapped by Ag NPs and further react with O2 adsorbed on the surface of catalyst to generate reactive O2 that induced the degradation of dyes (eqn (10)).

(1) Photo-excitation and photosensitized process of organic dyes

 
WO3–AgBr + → WO3 (eCB + hVB+)–AgBr (eCB + hVB+) (3)
 
Organic dye + → dye* + e (4)
 
eCB + O2 → O2 (5)
 
AgBr/WO3 + dye* → dye+˙ + AgBr/WO3 (eCB) (6)
 
AgBr/WO3 (eCB) + O2 → O2 + AgBr/WO3 (7)

(2) Simultaneous photocatalytic reaction

 
Dye+˙ + O2 → intermediate products → CO2 + H2O (8)
 
Dye + h(AgBr/WO3)+ → intermediate products → CO2 + H2O (9)
 
˙O2 + dye → CO2 + H2O (10)

The stability of the AgBr/m-WO3 is a concern in photocatalysis due to the photosensitive AgBr involves. In order to investigate the stability of the photocatalysts, experiments of visible-light-driven photodegradation of rhB were repeated for certain times on the recycled AgBr/WO3 catalysts. After several cycling photocatalysis reaction, all the AgBr/m-WO3 samples partially lost their activity (Fig. 9), while WO3 remained its initial activity. AgBr (30% wt)/m-WO3 showed better initial activity and stability than other AgBr/m-WO3, though it deactivated and the deactivation was accelerated under prolonged visible-light irradiation (Fig. 10). For example, AgBr (30% wt)/m-WO3 completely decolorized rhB in 240 min in the first photocatalytic cycling test, but took nearly 480 min in the seventh cycling.


image file: c6ra16442k-f9.tif
Fig. 9 The degradation of rhB by AgBr/m-WO3 under visible-light irradiation.

image file: c6ra16442k-f10.tif
Fig. 10 Cycling tests of rhB photodegradation on AgBr (30 wt%)/m-WO3 samples under visible-light irradiation.

After each cycling test, the particulate AgBr/m-WO3 photocatalysts were recovered and examined to investigate the evolution of the samples during the photocatalysis stability tests. The crystalline structure of recycled AgBr (30 wt%)/WO3 samples were characterized by XRD. The diffraction patterns of the fresh and recycled AgBr (30 wt%)/m-WO3 are compared in Fig. 11. Both the fresh and recycled photocatalysts showed identical Bragg peaks for the m-WO3 and AgBr species, but new diffraction peaks centered at 38.20° were observed on the recycled catalyst. It can be well defined to cubic Ag. During the photocatalytic oxidation process, AgBr/m-WO3 system was transformed to be Ag/AgBr/m-WO3 system, which ascribed to convert some Ag+ ions on the surface of AgBr/m-WO3 to Ag0 species under visible light irradiation.40 However, after the fifth cycles, the other new diffraction peaks centered at 32.56° were clearly observed. The diffraction peak can be well defined to cubic Ag2O (Pn[3 with combining macron]ms, ICSD: 03-5540), a p-type oxide semiconductor with narrow bandgap of 1.30 eV.41 The diffraction intensity of the Ag2O became stronger with increased cycling times. For example, the XRD signal of Ag2O is stronger after seventh recycling than after the fifth cycling test. These changes of the recycled AgBr (30 wt%)/m-WO3 suggest that the AgBr was transformed into Ag2O along with the gradual deactivation of the photocatalysts under visible-light illumination.


image file: c6ra16442k-f11.tif
Fig. 11 XRD patterns of the fresh and recycled AgBr (30 wt%)/m-WO3 sample.

Thus, the AgBr/m-WO3 sample was examined by X-ray photoelectron spectroscopy, and the results are shown in Fig. 12. The XPS peak for C 1s is due to the adventitious hydrocarbon form the XPS instrument itself. In Fig. 12a, the Ag 3d spectra of AgBr/m-WO3 consist of two individual peaks at around 373 and 367 eV, which can be attributed to Ag 3d2/3 and Ag 3d5/2 binding energies, respectively. To further investigate the chemical state of Ag element, the XPS peaks of Ag element with the composite photocatalyst using for several consecutive photocatalysis experiment with solution of rhB. After the third cycling test, the Ag 3d5/2 peak is further divided into two different peaks at 367.6 and 368.8 eV, and Ag 3d3/2 peak is divided into two different peaks at 373.5 and 374.8 eV (Fig. 12b). The peaks at 368.8 and 374.8 eV are attributed to metal Ag0 and the peaks at 367.6 and 373.5 eV are attributed to Ag+ of AgBr, and those at 368.8 and 374.8 eV to metal Ag0 hence confirming the existence of Ag0.42 From the XPS peak areas, the surface Ag0 and Ag+ contents are calculated to be 0.33 and 7.8 mol%, respectively. According the XRD results, the phase of Ag2O was appeared with increasing cycling test. As for the Ag2O/Ag/AgBr/m-WO3 nanocomposite, the sample shows similar XPS peaks of Ag element with Ag/AgBr/m-WO3 phase (Fig. 12c). The calculated contents of the surface Ag+ of the corresponding samples are 11.3 mol%, which indicating the formation of Ag2O in the Ag/AgBr/m-WO3 composite. The result is good agreement with XRD analysis.


image file: c6ra16442k-f12.tif
Fig. 12 Ag 3d XPS spectras of (a) the as-prepared AgBr (30 wt%)/m-WO3; (b) the AgBr (30 wt%)/m-WO3 used for three consecutive photocatalysis experiments and (c) the AgBr (30 wt%)/m-WO3 used for seven consecutive photocatalysis experiments with solution of rhB.

The plausible formation process of Ag2O may be depicted thorough photochemical reaction eqn (10)–(13). During the photocatalysis process, trace amounts of Ag+ and Br would be generated via ionizing AgBr in photocatalysts (eqn (11)), under the visible light irradiation. The Ag+ will then react with OH to generate AgOH, which spontaneously convert to Ag2O (eqn (12)). The overall reaction is summarized as eqn (13), revealing that AgBr gradually transformed in to Ag2O and replaced the AgBr active site during photocatalytic reaction. These changes may be accelerated by light irradiation, in a consequence of the photocatalyst deactivation.

 
AgBr + H2O + → Ag+ + Br + Haq+ + OHaq (11)
 
2Ag+ + 2OH → 2AgOH → Ag2O + H2O (12)
 
In total: 2AgBr + H2O + → Ag2O + 2H+ + 2Br (13)

On the basis of the photo-degradation mechanism and stability study, the deactivation mechanism of our Ag/AgBr/WO3 samples can now be proposed. As schematically illustrated in Fig. 13, a small amount of AgBr was transformed into Ag2O on the surface of WO3 once visible-light stroke on the Ag/AgBr/WO3 samples for a certain period, meanwhile, the Ag2O concentration gradually increased after long-term irradiation. The XRD intensity of Ag2O was monotonically increased as observed in Ag/AgBr/WO3, revealing more Ag2O species were generated, in the multiple recycling photocatalytic tests. Eventually, the Ag2O would mask the active AgBr sites, resulting in the formation of Ag2O/Ag/AgBr/WO3, with reduced photoactivity.


image file: c6ra16442k-f13.tif
Fig. 13 Schematic evolution of AgBr/WO3 samples under long-term visible-light irradiation.

The Ag2O-induced deactivation can be well explained from the photoelectrochemistry point of view as illustrated in Fig. 14. The estimated VBAg2O top locates more negative than the potentials of H2O2, ˙OH and VBs of WO3 and AgBr, suggesting VBAg2O is a weaker photo-oxidant with limited oxidizing ability for dyes decomposition. On the other hand, the CBAg2O potential is more positive than that of AgBr and O2− but more negative than CBWO3, suggesting the electrons on CBAgBr would prefer to hop onto CBAg2O, which competed with the electron flowing onto CBWO3 and production of reactive O2−. The in situ-generated Ag2O would also deteriorate the intimate interfaces between AgBr and WO3. Ag2O may absorb more light due to its smaller bandgap energy and thus mask the active AgBr sites. Moreover, the narrow bandgap of Ag2O, implying short electron migration pathway, would inevitably lead to faster recombination of photogenerated charge carriers. Because the WO3 remains almost unchanged, the photogenerated Ag2O species should be the dominant factor responsible for the deactivation of the AgBr/WO3. In general, despite that the crystal of WO3 in the AgBr/WO3 remains stable, the transformation of AgBr to Ag2O would occur during photocatalytic reaction. Such Ag2O act as stronger recombination centers and would destroy the synergistic interaction between AgBr and WO3 in photocatalytic reaction, leading to reduce the surface reactivity arisen form AgBr.


image file: c6ra16442k-f14.tif
Fig. 14 Schematic band alignments of the deactivated Ag2O/Ag/AgBr/m-WO3 system.

At the same time, in the Ag2O/Ag/AgBr/WO3 system, Ag2O, Ag, AgBr and WO3 can be simultaneously excited to form electron–hole pairs under visible-light irradiation. Subsequently the photogenerated electrons transfer from the CB bottom of AgBr to that of WO3 or are trapped by AgNPs formed on AgBr particles. The photogenerated holes also move in the opposite direction from the VB top of WO3 to that of AgBr. Probably, electrons in the VB of AgBr could be excited up to a higher potential edge under visible-light illumination with energy less than 2.95 eV (λ ≥ 420 nm), and these electrons will be trapped by AgNPs and further react with O2 adsorbed on the surface of catalyst to generate reaction O2− that induced the degradation of rhB, MO and MB. At the same time, CBAg2O potential is more negative than CBWO3, suggesting the electrons on CBAgBr would prefer to hop onto CBAg2O, which competed with the electron flowing onto CBWO3 and production of reactive O2. It is worth noting that the results are much different from the AgBr/WO3 systems.

Meanwhile, on the one hand, reactive holes at the VB of AgBr can oxidize Br ions to Br0 atoms that are reactive radical species and degrade rhB, MO and MB;43 on the other hand, the holes generated on AgNPs can also oxidize the rhB, MO and MB directly. In summary, these organic dyes were decomposed by Ag2O/Ag/AgBr/WO3 under visible-light irradiation through O2−, Br0 or direct h+ oxidation pathway.

Conclusions

The cuboid rods of m-WO3 with different amounts of AgBr were successfully prepared and used for photodegradation of organic dyes (rhB, MO and MB) under visible-light irradiation. The initial activity of organic dyes photodegradation on the AgBr/m-WO3 is closely related to the AgBr loading. The AgBr (30 wt%)/m-WO3 showed the best initial activity and stability among these photocatalysts. Then enhanced photocatalytic activity is mainly attributed to the band structure. Loading of AgBr greater than 30 wt% leads to weaker catalytic activity because the majority of AgBr is not in close contact with the WO3 and may actively screen light irradiation.

All the AgBr/m-WO3 composite photocatalysts are prone to deactivation in the photocatalytic reactions since the AgBr is vulnerable and transformed to Ag2O under visible-light irradiation. The as-formed Ag2O reduces the intimate interface of AgBr/m-WO3 catalysts, masks the light excitation, and serves as recombination center of photogenerated charge carriers, resulting in the deactivation of the catalysts.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (Grant No. 51504101), the Natural Science Foundation of Jiangsu Province (Grant No. BK20150514), the Natural Science Foundation of Jiangsu Provincial Higher Education of China (Grant No. 15KJB430006), the Start-up Foundation of Jiangsu University for Senior Talents (Grant No. 15JDG014).

References

  1. H. Tong, S. X. Ouyang, Y. P. Bi, N. Umezawa, M. Oshikiri and J. H. Ye, Adv. Mater., 2012, 24, 229 CrossRef CAS PubMed.
  2. Y. H. Kim, J. S. Heo, T. H. Kim, S. Park, M. H. Yoon, J. Kim, M. S. Oh, G. R. Yi, Y. Y. Noh and S. K. Park, Nature, 2012, 489, 128 CrossRef CAS PubMed.
  3. S. Yamazaki, T. Yamate and K. Adachi, Appl. Catal., A, 2013, 454, 30 CrossRef CAS.
  4. S. J. Hong, H. Jun and L. S. Lee, Scr. Mater., 2010, 63, 757 CrossRef CAS.
  5. J. Zeng, M. Hu, W. D. Wang, H. Q. Chen and Y. X. Qin, Sens. Actuators, B, 2012, 161, 447 CrossRef CAS.
  6. G. C. Xi, J. H. Ye, Q. Ma, N. Su, H. Bai and C. Wang, J. Am. Chem. Soc., 2012, 134, 6508 CrossRef CAS PubMed.
  7. M. Sasidharan, N. Gunawardhana, M. Yoshi and K. Nakashima, Nano Energy, 2012, 1, 503 CrossRef CAS.
  8. S. K. Biswas and J. O. Baeg, Int. J. Hydrogen Energy, 2013, 38, 3177 CrossRef CAS.
  9. P. M. Kadam, N. L. Tarwal, S. S. Mali, H. P. Deshmukh and P. S. Patil, Electrochim. Acta, 2011, 58, 556 CrossRef CAS.
  10. A. K. Bhosale, N. L. Tarwal, P. S. Shinde, P. M. Kadam, R. S. Patil, S. R. Barman and P. S. Patil, Solid State Ionics, 2009, 180, 1324 CrossRef CAS.
  11. C. A. Bignozzi, S. Caramor, V. Cristino, R. Argazzi, L. Meda and A. Tacca, Chem. Soc. Rev., 2013, 42, 2228 RSC.
  12. A. Enesca and A. Duta, Appl. Phys. A: Mater. Sci. Process., 2013, 111, 639 CrossRef CAS.
  13. W. Erbs, J. Desilverstro, E. Borgarello and M. Gratzel, J. Phys. Chem., 1984, 88, 4001 CrossRef CAS.
  14. R. Abe, K. Sayama and H. Sugihara, J. Phys. Chem. B, 2005, 109, 16052 CrossRef CAS PubMed.
  15. S. M. Sun, W. Z. Wang, S. Z. Zeng, M. Shang and L. Zhang, J. Hazard. Mater., 2010, 178, 427 CrossRef CAS PubMed.
  16. J. Cao, B. D. Luo, H. L. Lin and S. F. Chen, J. Mol. Catal. A: Chem., 2011, 344, 138 CrossRef CAS.
  17. L. Kong, Z. Jiang, H. H. Lai, R. J. Nicholls, T. C. Xiao, M. O. Jones and P. P. Edwards, J. Catal., 2012, 293, 116 CrossRef CAS.
  18. Y. P. Bi, S. X. Ouyang, J. Y. Cao and J. H. Ye, Phys. Chem. Chem. Phys., 2011, 13, 10071 RSC.
  19. C. H. An, J. Z. Wang, W. Jiang, M. Y. Zhang, X. J. Ming, S. T. Wang and Q. H. Zhang, Nanoscale, 2012, 4, 5646 RSC.
  20. X. C. Ma, Y. Dai, M. Guo and B. B. Huang, ChemPhysChem, 2012, 13, 2304 CrossRef CAS PubMed.
  21. W. S. Wang, H. Du, R. X. Wang, T. Wen and A. W. Xu, Nanoscale, 2013, 5, 3315 RSC.
  22. C. C. Shen, Q. Zhu, Z. W. Zhao, T. Wen, X. K. Wang and A. W. Xu, J. Mater. Chem. A, 2015, 3, 14661 CAS.
  23. J. Y. Leng, X. J. Xu, N. Lv, H. T. Fan and T. Zhang, J. Colloid Interface Sci., 2011, 356, 54 CrossRef CAS PubMed.
  24. R. Q. Cabrera, E. R. Latimer, A. Kafizas, C. S. Blackman, C. J. Carmalt and I. P. Parkin, J. Photochem. Photobiol., A, 2012, 239, 60 CrossRef CAS.
  25. D. S. Martínez, A. M. Cruz and E. L. Cuéllar, Appl. Catal., A, 2011, 398, 179 CrossRef.
  26. P. K. Biswas, N. C. Pramanik, M. K. Mahapatra, D. Ganguli and J. Livage, Mater. Lett., 2003, 157, 4429 CrossRef.
  27. Y. Beak and K. Yong, J. Phys. Chem. C, 2007, 111, 1213 Search PubMed.
  28. K. H. Chang, C. C. Hu, C. M. Huang, Y. L. Liu and C. I. Chang, J. Power Sources, 2011, 196, 2387 CrossRef CAS.
  29. X. Q. Gao, X. T. Su, C. Yang, F. Xiao, J. D. Wang, X. D. Cao, S. J. Wang and L. Zhang, Sens. Actuators, B, 2013, 181, 537 CrossRef CAS.
  30. J. Y. Li, J. F. Huang, J. P. Wu, L. J. Cao and K. Yanagisawa, Ceram. Int., 2012, 38, 4495 CrossRef.
  31. M. Galceran, M. C. Pujol, C. Zaldo, F. Daz and M. Aguil, J. Phys. Chem. C, 2009, 113, 15497 CAS.
  32. R. Adhikari, G. Gyawali, T. Sekino and S. W. Lee, J. Solid State Chem., 2003, 197, 560 CrossRef.
  33. Y. G. Xu, H. Xu, J. Yan, H. M. Li, L. Y. Huang, Q. Zhang, C. J. Huang and H. L. Wan, Phys. Chem. Chem. Phys., 2013, 15, 5821 RSC.
  34. H. Ehrenreich and D. Turnbll, Harcourt Brace Hovanovich, Publishers, Academic Press, 1991 Search PubMed.
  35. R. H. Victora, Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 56, 4417 CrossRef CAS.
  36. L. Q. Ye, J. Y. Liu, C. Q. Gong, L. H. Tian, T. Y. Peng and L. Zan, ACS Catal., 2012, 2, 1677 CrossRef CAS.
  37. J. Tang and J. Ye, Chem. Phys. Lett., 2005, 410, 104 CrossRef CAS.
  38. Z. Jiang, F. Yang, G. Yang, L. Kong, M. O. Joes, T. Xiao and P. P. Edwards, J. Photochem. Photobiol., A, 2010, 212, 8 CrossRef CAS.
  39. L. Berggren, J. C. Jonsson and G. A. Niklasson, J. Appl. Phys., 2007, 102, 083538 CrossRef.
  40. Q. Zhu, W. S. Wang, L. Lin, G. Q. Gao, H. L. Guo, H. Du and A. W. Xu, J. Phys. Chem. C, 2013, 117, 5894 CAS.
  41. C. Sun, Q. Li, S. Gao, L. Gao and J. K. Shang, J. Am. Ceram. Soc., 2010, 93, 531 CrossRef CAS.
  42. P. Wang, B. B. Huang, X. Y. Qin, Z. Y. Zhang, Y. Dai and M. H. Whangbo, Inorg. Chem., 2009, 48, 10697 CrossRef CAS PubMed.
  43. L. Kuai, B. Geng, X. Chen, Y. Zhao and Y. Luo, Langmuir, 2010, 26, 18723 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2016