Ag-Doped g-C3N4 film electrode: fabrication, characterization and photoelectrocatalysis property

Fanjing Qiab, Yibing Lib, Yanbin Wanga, Yan Wanga, Shanshan Liua and Xu Zhao*a
aKey Laboratory of Drinking Water Science and Technology, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, P. O. Box 2871, 18 Shuangqing Road, Haidian District, Beijing 100085, China. E-mail: zhaoxu@rcees.ac.cn; Tel: +86-010-62849667
bSchool of Civil Engineering and Transportation, Hebei University of Technology, Tianjin 300401, China

Received 10th July 2016 , Accepted 28th July 2016

First published on 28th July 2016


Abstract

Ag-Doped graphitic carbon nitride films (Ag/g-C3N4) were synthesized easily onto ITO substrates by a liquid-based reaction process. Ag/g-C3N4 films were comprehensively characterized by SEM, HRTEM, XRD, UV/vis DRS, and XPS. The results indicated that Ag and Ag2O disperse homogeneously in the matrix of the g-C3N4 film. The photocurrent response of the Ag/g-C3N4 films increased remarkably with increasing Ag content and the best performance was observed with the sample of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10). The Ag/g-C3N4 films exhibited a high photoelectrocatalytic activity for the degradation of methylene blue. The enhancement of photoelectrocatalysis owing to more visible light could be harvested and photogenerated electron and interfacial electron could transfer more easily after modifying Ag in the g-C3N4 film. Thus, a possible photoelectrocatalysis mechanism was proposed. Beside g-C3N4, electron–hole pairs could be generated by Ag under visible light irradiation, and the photogenerated electron was captured by O2 or ˙O2 and then forms ˙OH radicals.


1. Introduction

Polymeric graphitic carbon nitride (g-C3N4) has attracted considerable attention as a potential material recently.1,2 It has two-dimensional planes of tri-s-triazine and π-conjugated phase between the layers and shows photocatalytic, electrocatalytic, and heterogeneous catalytic activities.3–5 g-C3N4 can be simply synthesized from a certain amount of nitrogen- as well as carbon-rich precursors such as urea,4 dicyandiamide5 or melamine.6,7 With an optical band gap of 2.70 eV, g-C3N4 showed a response to visible-light up to 460 nm.8–10

Most of the researches focus on the g-C3N4 powders. For the electrochemical or photoelectrochemical applications, including water electrosplitting, fuel cells, and solar cells, a direct and continuous g-C3N4 layer is preferred. Recently, numerous studies have been devoted to the synthesis of a g-C3N4 thin film, including bubble template method,11,12 sol–gel techniques,13 chemical exfoliation method,14 deposition method,15,16 as well as spin-coating method.17

It was recognized that the photocatalytic activity of g-C3N4 is limited by its high recombination probability of photogenerated electron–hole pairs. Therefore, a great deal of effort has been made, which include coupling with other semiconductors (e.g., C3N4/TiO2,18 C3N4/BiPO4,19 metal or nonmetal doping (e.g., Fe,20 Ag,21 Bi22 and P,23 S,24 F25), and other chemical doping. Metal doping is an effective strategy to tune the electronic and optical properties, as well as the conduction band edge of the semiconductors and their surface properties. Ag is an ideal dopant to improve the photoelectrocatalytic abilities followed by the enhancement of the visible light response and the improvement of the charge separation and transfer. Ag/g-C3N4 powder materials were synthesized with a metallic silver content in the 1–10 wt% range through a microemulsion method; the introduction of silver in the 1–10 wt% range enhanced the g-C3N4 photocatalytic activity.26

Herein, we report a Ag doped g-C3N4 film electrode with an remarkable enhancement of photoelectrochemical properties afforded by a mixture of AgNO3, cyanuric acid, benzoguanamine and grown thin films using the supramolecular approach on the ITO substrate by a liquid-based reaction. The Ag species on the g-C3N4 film were analyzed and a proposed mechanism of the photoelectrochemical response enhancement was proposed.

2. Experimental

2.1 Materials

Cyanuric acid (>98%) and benzoguanamine (>98%) were purchased from Tokyo Kasei Kogyo Co., Ltd. AgNO3 (AR) was provided by Sinopharm Chemical Reagent Co., Ltd. (Beijing, China). All the reagents were used without further purification. ITO (Indium-Tin Oxide) conductive film glasses with a size of 10 × 2.5 × 0.2 cm were provided by Shenzhen Jingweite Technology Co., Ltd. (Shenzhen, China).

2.2 Preparation of the film electrode

The precursor powder was prepared by a concussion-hybrid method and the film electrodes were afforded by liquid-based growth reaction, as described by Xu et al.15 In a typical process, 4 g of a mixture of cyanuric acid and benzoguanamine (molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1) and the required amount of AgNO3 were dispersed in 40 mL deionized water and stirred continuously for 6 hours. The white suspension was then centrifuged at about 7500 rpm and the sediment was freeze-dried for 24 h.

One piece of ITO glass was placed in a quartz crucible with the conducting side up, and then an adequate amount of precursor powder was transferred to the crucible, totally covering the substrate placed at the bottom. The crucible was then capped and heated at 500 °C for 2 h with a ramp rate of about 3 °C min−1 for both the heating and cooling process in a nitrogen atmosphere. In all cases, continuous g-C3N4 and Ag/g-C3N4 film electrodes were obtained, demonstrating complete coverage across the substrate. Ag/C3N4 film electrode with different Ag contents were denoted as Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10), Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]25), and Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50), where the number denotes the AgNO3 to g-C3N4 molar ratios.

2.3 Characterization

The surface morphology was characterized on a Hitachi SU-8020 scanning electron microscope (SEM, Hitachi Ltd., Japan). Powder and film X-ray diffraction (XRD) patterns were recorded on an X'pert PRO MPD PC system, utilizing Cu Kα radiation at a scan rate (2θ) of 0.5° s−1. The accelerating voltage and applied current were 40 kV and 80 mA, respectively. Fourier transform infrared (FT-IR) spectra were obtained on a Nicolet 8700 spectrometer (Thermo Fisher Scientific) in the frequency range of 4000–800 cm−1. X-ray photoelectron spectroscopy (XPS) was measured using a PHI Quantera SXM instrument (ULVAC-PHI, Japan). The binding energies were calibrated with C 1s = 284.8 eV.

2.4 Photoelectrochemical measurements

Photoelectrochemical measurements were performed in a three-electrode experimental system using the CHI660D Electrochemical Workstation (Shanghai Chenhua Instrument Co., Ltd., Shanghai, China). The prepared photoelectrodes (10 × 2.5 × 0.2 mm), saturated calomel electrode, and Pt electrode acted as the working, reference, and counter electrodes, respectively. The electrolyte was 0.05 mol L−1 Na2SO4. The photocurrent density with time (it curve) was performed at 1 V bias potential under visible light irradiation. The photons generated by a 300 W Xe arc lamp (PLS-SXE500, BoPhilae technology co., LTD, Beijing, China) passed through a quartz reactor, equipped on the side of the three-electrode cell, and irradiated the backside of the photoelectrode. Visible light was obtained by placing a 400 nm cutoff filter to remove the UV irradiation. Electrochemical impedance spectroscopy (EIS) was performed at the open circuit potential over the frequency range between 106 and 10−2 Hz, with an AC voltage magnitude of 5 mV, using 12 points per decade.

In the case of the photoelectrocatalytic degradation experiment, 10 mg L−1 MB was added to the quartz reactor with a three-electrode experimental system. The prepared photoelectrodes (10 × 2.5 × 0.2 mm), saturated calomel electrode, and Pt electrode acted as the working, reference, and counter electrodes, respectively. The electrolyte was 0.05 mol L−1 Na2SO4. The absorbance of the reaction solution was measured by a UV-vis spectrophotometer (U-3010, Hitachi Ltd., Japan).

3. Results and discussion

3.1 SEM-EDS analysis of g-C3N4 and Ag/g-C3N4 composites

Fig. 1 shows SEM images of g-C3N4 film (a) and Ag/g-C3N4 films (b–d). As observed in Fig. 1(a), pure g-C3N4 shows a typical flat layer structure with a size of several micrometers. For Ag/g-C3N4 films (Fig. 1(b–d)), Ag particles with a polyhedral structure could be seen on the surface of the Ag/g-C3N4 films and linked with g-C3N4, indicating that Ag and Ag2O disperse homogeneously in the matrix of the g-C3N4 film. To further confirm the existence of Ag nanoparticles, EDS was performed and the result is shown in Fig. 2(b). 47.93 wt% of Ag was observed in the partial area of the Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) film electrode, suggesting Ag with a polyhedral structure was introduced to the g-C3N4 film (Table 1).
image file: c6ra17613e-f1.tif
Fig. 1 SEM images of g-C3N4 and Ag/g-C3N4 with different doping ratios: (a) g-C3N4; (b) Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50); (c) Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]25); (d) Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10).

image file: c6ra17613e-f2.tif
Fig. 2 EDS image of the Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) sample.
Table 1 Atomic ratio of the Ag/g-C3N4 sample (1[thin space (1/6-em)]:[thin space (1/6-em)]10) sample
Element Intensity (c/s) Atomic% Conc. Units
C 59.36 31.769 16.63 wt%
N 55.90 58.037 35.43 wt%
Ag 177.51 10.194 47.93 wt%
Total 100.000 100.00 wt%


3.2 HRTEM-STEM analysis of g-C3N4 and Ag/g-C3N4

Fig. 3 shows the HRTEM images of Ag/g-C3N4 and the corresponding selected area electron diffraction (SAED). As seen in Fig. 3(a–c), the morphology of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) exhibits a disordered state and a lamellar structure with no lattice fringes. However, the diffraction ring was observed in Fig. 3(d), which corresponds to (002) of the g-C3N4 layered structure formed by accumulation. The STEM-EDS of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) are shown in Fig. S1. Since a heavier element is indicated as a brighter image in the STEM result, the brighter parts are supposed to be Ag nanoparticles. Furthermore, as shown in Fig. S1(c), the percentage of Ag is 39.5 wt% for the brighter parts, further confirming that the brighter parts are Ag particles. In addition, the STEM and corresponding element mappings in Fig. S2 detect the presence of Ag in addition to C, N, and O and the size of Ag is about 1 micrometer, proving the existence of Ag in Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) and that this method can successfully fabricate a Ag/g-C3N4 film.
image file: c6ra17613e-f3.tif
Fig. 3 HRTEM images of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) and the corresponding selected area electron diffraction pattern.

3.3 XRD patterns of Ag/g-C3N4

The X-ray diffraction pattern for the pure g-C3N4 film has two distinct peaks at 13.2° and 27.3°, corresponding to the (100) and (002) planes of g-C3N4, respectively. The peak at 27.3° is due to the stacking of the conjugated aromatic system, which was consistent with the reported literature.27,28 As shown in Fig. 4, the relative intensity of the peak at 27.3° decreased with increasing Ag contents, suggesting that the diffracted intensity of graphite with tri-s-triazine-based connection sheet became weaker after doping with Ag. For the Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) sample, five peaks at 38.1°, 44.3°, 64.4°, 77.4°, and 81.5° were observed, corresponding to the (111), (200), (220), (311), and (222) planes of the face-centered cube structure of Ag,29,30 respectively. However, no significant diffraction peaks of Ag were observed for Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]25) and Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50) owing to the low Ag content of these samples.
image file: c6ra17613e-f4.tif
Fig. 4 XRD patterns of g-C3N4 and Ag/g-C3N4 with different Ag content.

3.4 FT-IR analysis of Ag/g-C3N4

Fig. 5 represents the FT-IR spectra of g-C3N4 and Ag/g-C3N4 with different Ag ratios. Three main absorption regions are observed for the pure g-C3N4. The broad peak at 3000–3500 cm−1 was ascribed to the stretching vibration of N–H and O–H in physically adsorbed water.31 The intensity of this broad peak decreases with increasing Ag content, which can be attributed to the destruction part of N–H and O–H by inserting Ag atoms. The bands at 1407, 1320 and 1240 cm−1 are assigned to the typical stretching vibration of aromatic C–N, whereas the bands at 1640 and 1567 cm−1 are ascribed to the C[double bond, length as m-dash]N vibration.30,32 The intensity of these peaks does not change obviously, indicating that Ag/g-C3N4 has a reservation of aromatic CN heterocycles. In addition, the intensity of the peak at 808 cm−1, which was assigned to the breathing mode of the triazine units, decreased with the decreasing of g-C3N4 content. However, the triazine rings still remain even at the highest doping level, indicating that the lattice changes are only partial. In particular, the peaks at 2160 and 2340 cm−1 can be assigned to C[triple bond, length as m-dash]C and C[triple bond, length as m-dash]N triple bonds, respectively. Their intensity increases with increasing Ag content, indicating that new C[triple bond, length as m-dash]C and C[triple bond, length as m-dash]N triple bonds are formed in place of the typical stretching vibration of aromatic C–N stretching. These results imply that some triazine rings of g-C3N4 framework were broken and sp2 C–N bonds transformed into C[triple bond, length as m-dash]C and C[triple bond, length as m-dash]N triple bonds upon the introduction of Ag.
image file: c6ra17613e-f5.tif
Fig. 5 FT-IR spectra of the prepared g-C3N4 and Ag/g-C3N4 samples.

3.5 UV/vis diffuse densities of Ag/g-C3N4 composite

Fig. 6 shows the UV/vis diffuse reflectance spectra of pure g-C3N4 and Ag/g-C3N4 with different Ag contents. The maximum absorption wavelength can be obtained by drawing tangent along the critical fall part of the UV/vis diffuse reflectance spectrum and extending to the horizontal axis. The maximum absorption wavelength of g-C3N4 was estimated to be about 462 nm, which indicates that g-C3N4 responds to visible light. In contrast, with increasing Ag doping ratio, the maximum absorption wavelengths of Ag/g-C3N4 are approximately 521, 595 and 602 nm, respectively. The band gap of g-C3N4 calculated based on the Oregan and Gratzel method is 2.68 eV, which accorded well with previous reports.33,34 The band gap of Ag/g-C3N4 with a Ag[thin space (1/6-em)]:[thin space (1/6-em)]g-C3N4 ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]50, 1[thin space (1/6-em)]:[thin space (1/6-em)]25, and 1[thin space (1/6-em)]:[thin space (1/6-em)]10 are 2.38, 2.08, and 2.05 eV, respectively. These results demonstrate the existence of strong interactions between Ag and nitrogen, which alter the electronic structure of g-C3N4. This phenomenon suggests that more visible light can be harvested with increasing Ag doping content.
image file: c6ra17613e-f6.tif
Fig. 6 UV/vis diffuse reflectance spectra of the g-C3N4 and Ag/g-C3N4 samples.

3.6 XPS analysis of Ag/g-C3N4 composite samples

The surface elemental composition and valence state of g-C3N4 and Ag/g-C3N4 were investigated by XPS. Fig. S3 shows the survey spectrum and high resolution XPS spectra of C 1s, N 1s and Ag 3d. As shown in Fig. 7(a), two peaks centered at 284.6 and 288.0 eV were observed, which can be ascribed to the C–C bond of carbon species adsorbed on the surface of g-C3N4[thin space (1/6-em)]12,34 and N–C[double bond, length as m-dash]N species,35,36 respectively. The ratio of the intensity between C–C and N–C[double bond, length as m-dash]N decreases with the increasing Ag doping. Moreover, the signal of N–C[double bond, length as m-dash]N slightly shifts to a higher binding energy, indicating that the chemical environment changed due to the interaction between Ag and g-C3N4. The N 1s XPS spectrum of g-C3N4 is shown in Fig. 7(b), which can be deconvoluted into four peaks that correspond to nitrogen atoms in different functional groups: sp2 hybridized aromatic N bonded to carbon atoms of C–N[double bond, length as m-dash]C at 398.5 eV,37 tertiary N bonded to carbon atoms in the form of N–C3 at 399.7 eV, amino functional groups with hydrogen (C–N–H) at 401.2 eV,38 and π-excitations at 404.3 eV.3,35 The N 1s XPS spectrum of Ag/g-C3N4 is shown in Fig. 7(c); remarkable shifts to lower binding energy are observed, suggesting that the electronic structure of g-C3N4 is changed by Ag modification. In the high-resolution Ag 3d XPS spectrum (Fig. 7(d), two major peaks with binding energies of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10 and 1[thin space (1/6-em)]:[thin space (1/6-em)]25) at 367.9 and 374.0 eV are fitted to Ag 3d5/2 and Ag 3d3/2, respectively, suggesting that the major phase are Ag0. The two peaks at 368.2 and 273.9 eV of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50) indicate Ag2O to be the major phase.
image file: c6ra17613e-f7.tif
Fig. 7 High-resolution XPS spectra of C 1s region (a), N 1s region of pure g-C3N4 (b), N 1s region of Ag/g-C3N4 (c), and Ag 3d region (d).

3.7 Photoelectrochemical response of Ag/g-C3N4

EIS was used to investigate the migration ability of the electrons, which was related to the photocatalytic and photoelectrochemical properties of semiconductor materials. The Nyquist plots of g-C3N4 and Ag/g-C3N4 photoanodes are shown in Fig. 8. The diameter of the semicircular Nyquist plot was significantly smaller compared to pure g-C3N4 with the introduction of Ag, suggesting a faster charge transfer rate and more effective separation of the photogenerated electron–hole pairs.39
image file: c6ra17613e-f8.tif
Fig. 8 EIS spectra of g-C3N4 and Ag/g-C3N4 photoelectrodes (λ > 400 nm, [Na2SO4] = 50 mM, applied bias = 1.0 V).

The photocurrent response of g-C3N4 and Ag/g-C3N4 photoelectrodes under visible light irradiation are shown in Fig. 9. The photogenerated current density is 0.93 μA mm−2 for g-C3N4 under visible light irradiation and it increases with increasing Ag content. For example, the photogenerated current density was 1.93, 3.86, 6.40 μA mm−2 for the Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50), Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]25) and Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) samples, respectively. The photocurrent results confirm the more efficient separation of photogenerated electron–hole pairs for the Ag/g-C3N4 samples than that of pure g-C3N4. It indicates that Ag has been combined effectively with Ag/g-C3N4, which leads to efficient separation of the photogenerated electron–hole pairs in the g-C3N4.


image file: c6ra17613e-f9.tif
Fig. 9 Photocurrent response of g-C3N4 and Ag/g-C3N4 photoelectrodes under visible light (λ > 400 nm, [Na2SO4] = 50 mM, U = 1.0 V) (a), g-C3N4; (b), Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50); (c), Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]25); (d), Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10).

To evaluate the photoelectrocatalytic activities of the Ag/g-C3N4 samples, photoelectrocatalytic degradation of MB using the Ag/g-C3N4 film as a photoanode under visible light irradiation (λ > 400 nm) with a bias potential of 1.0 V was performed. As shown in Fig. 10(a), the MB degradation efficiency was 38.96%, 45.66%, 64.23%, and 72.90% for g-C3N4, Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]50), Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]25), and Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10), respectively. Compared with the g-C3N4 case, it is clearly seen that the degradation efficiency for MB was improved using Ag/g-C3N4 and increased with increasing Ag content. Among these photoanodes, Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) shows the best photoelectrocatalytic activity. This shows that doping with Ag can increase the photo response ability of the g-C3N4 film electrode and enhance the photoelectrocatalytic activity. As shown in Fig. 10(b), the rate of MB degradation with Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) as the photoanode increases with increasing voltage in the range of 0–1.0 V. The degradation efficiency of MB was the highest at a bias potential of 1.0 V and reached up to 72.90% in 2.5 h. A possible reason is that the higher applied bias, the greater level of electron transfer.


image file: c6ra17613e-f10.tif
Fig. 10 (a) MB degradation curves of g-C3N4 and Ag/g-C3N4 under visible light irradiation at 1.0 V; (b) effect of the applied voltages on the degradation rate of MB with Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) (λ > 400 nm, initial concentration of MB = 10 mg L−1, [Na2SO4] = 50 mM).

3.8 Mechanism considerations

To investigate the main reactive oxidative species involved in the photoelectrocatalytic process, reactive oxidative species trapping experiments were performed. Ethanol, EDTA, and N2 were used as hydroxyl radical (˙OH), hole (h+) and superoxide radical (˙O2) scavenger, respectively. Fig. 11(a) presents the effects of various scavengers on the degradation of MB in the g-C3N4 photoelectrocatalysis system. The removal of MB decreased from 48.21% to 28.54% and 38.96% with N2 and ethanol added, respectively, implying that ˙O2 and ˙OH are involved in the photoelectrocatalysis process and ˙O2 is the major reactive oxidative species for the degradation of MB. However, the degradation of MB improved with the addition of EDTA, indicating that the photogenerated hole was captured by EDTA and the recombination of electron–hole pairs was inhibited. Therefore, more e can react with O2 to form ˙O2 and ˙OH (eqn (1)–(3)), which enhances the photoelectrocatalytic performance. As for Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) system (Fig. 11(b)), the inhibition performance followed the order: ethanol > N2 > EDTA. The results suggest that ˙OH radicals are not only formed from ˙O2 (eqn (1)–(3)), but are also formed from the reaction between h+ and OH (eqn (4)). However, when enriched electrons gathered on the Ag connect with other semiconductors, a multiple-electron reduction reaction of oxygen will occur.37
 
e + O2 → ˙O2 (1)
 
˙O2 + 2H+ + e → H2O2 (2)
 
H2O2 + e → ˙OH + OH (3)
 
OH + h+ → ˙OH (4)

image file: c6ra17613e-f11.tif
Fig. 11 Influence of various scavengers on the PEC processes of (a) pure g-C3N4 and (b) Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) toward the degradation of MB (λ > 400 nm, initial concentration of MB = 10 mg L−1, [Na2SO4] = 50 mM).

Based on the abovementioned results and analysis, a possible mechanism for the enhanced photoelectrocatalytic performance of the Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) film electrode was proposed, and the scheme is shown in Fig. 12. The band gap of Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) was smaller than that of the pure g-C3N4 film photoelectrode and more light could be harvested. In addition, a faster charge transfer rate and more effective separation of the photogenerated electron–hole pairs could be achieved according to EIS. Therefore, the photogenerated electron could transfer easily to the Pt electrode and be captured by O2, leading to the generation of ˙O2 and ˙OH radicals. Finally, MB can be degraded by ˙OH radicals. In addition, the Ag metal also can generate electron–hole pairs under visible light irradiation, which also can capture O2 molecules to form ˙O2 radicals, leaving behind holes (Ag+) that can be filled by electrons from Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10).30


image file: c6ra17613e-f12.tif
Fig. 12 Proposed mechanism of the photoelectrocatalytic degradation of MB with the Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) photoanode.

4. Conclusions

Ag-Modified g-C3N4 film electrodes were prepared by a liquid-based reaction onto an ITO substrate during the calcination treatment. The characterized results indicated that metallic silver and Ag2O exist in the Ag/g-C3N4 film electrode. The photoelectrochemical response of the g-C3N4 film electrode increased after the modification of Ag, because of the enhanced migration rate of the photogenerated electrons and interfacial electron. The Ag/g-C3N4 (1[thin space (1/6-em)]:[thin space (1/6-em)]10) film electrode showed the best performance and its photocurrent density was approximately 6.8 times greater than that of the pristine g-C3N4 film electrode. The involved active radicals including ˙O2 and ˙OH were confirmed for the degradation of MB.

Acknowledgements

This study was supported by projects of National Key Research and Development Program of China (No. 2016YFA0203101) and National Natural Science Foundation of China (No. 21377148, 51438011).

References

  1. Y. Gong, M. Li, H. Li and Y. Wang, Green Chem., 2015, 17, 715–736 RSC.
  2. S. Cao, J. Low, J. Yu and M. Jaroniec, Adv. Mater., 2015, 27, 2150–2176 CrossRef CAS PubMed.
  3. K. Wang, Q. Li, B. Liu, B. Cheng, W. Ho and J. Yu, Appl. Catal., B, 2015, 176–177, 44–52 CrossRef CAS.
  4. Q. Li, N. Zhang, Y. Yang, G. Z. Wang and D. H. L. Ng, Langmuir, 2014, 30, 8965–8972 CrossRef CAS PubMed.
  5. Y. Cui, Z. Ding, P. Liu, M. Antonietti, X. Fu and X. Wang, Phys. Chem. Chem. Phys., 2012, 14, 1455–1462 RSC.
  6. M. Xu, L. Han and S. Dong, ACS Appl. Mater. Interfaces, 2013, 5, 12533–12540 CAS.
  7. Y. Tao, Q. Ni, M. Wei, D. Xia, X. Li and A. Xu, RSC Adv., 2015, 5, 44128–44136 RSC.
  8. S. Zuluaga, L.-H. Liu, N. Shafiq, S. M. Rupich, J.-F. Veyan, Y. J. Chabal and T. Thonhauser, Phys. Chem. Chem. Phys., 2015, 17, 957–962 RSC.
  9. W. Shan, Y. Hu, Z. Bai, M. Zheng and C. Wei, Appl. Catal., B, 2016, 188, 1–12 CrossRef CAS.
  10. J. Liu, J. Alloys Compd., 2016, 672, 271–276 CrossRef CAS.
  11. M. Zhang, J. Xu, R. Zong and Y. Zhu, Appl. Catal., B, 2014, 147, 229–235 CrossRef CAS.
  12. J. Liu, H. Wang, Z. P. Chen, H. Moehwald, S. Fiechter, R. van de Krol, L. Wen, L. Jiang and M. Antonietti, Adv. Mater., 2015, 27, 712–718 CrossRef CAS PubMed.
  13. J. Zhang, M. Zhang, L. Lin and X. Wang, Angew. Chem., Int. Ed., 2015, 54, 6297–6301 CrossRef CAS PubMed.
  14. F. Liang and Y. Zhu, Appl. Catal., B, 2016, 180, 324–329 CrossRef CAS.
  15. J. Xu, T. J. K. Brenner, L. Chabanne, D. Neher, M. Antonietti and M. Shalom, J. Am. Chem. Soc., 2014, 136, 13486–13489 CrossRef CAS PubMed.
  16. M. Shalom, S. Gimenez, F. Schipper, I. Herraiz-Cardona, J. Bisquert and M. Antonietti, Angew. Chem., Int. Ed., 2014, 53, 3654–3658 CrossRef CAS PubMed.
  17. Y. Bu, Z. Chen, J. Yu and W. Li, Electrochim. Acta, 2013, 88, 294–300 CrossRef CAS.
  18. Y. Wu, L. Tao, J. Zhao, X. Yue, W. Deng, Y. Li and C. Wang, Res. Chem. Intermed., 2016, 42, 3609–3624 CrossRef CAS.
  19. X. Zou, C. Ran, Y. Dong, Z. Chen, D. Dong, D. Hu, X. Li and Y. Cui, RSC Adv., 2016, 6, 20664–20670 RSC.
  20. R. B. N. Baig, S. Verma, R. S. Varma and M. N. Nadagouda, ACS Sustainable Chem. Eng., 2016, 4, 1661–1664 CrossRef CAS.
  21. S. Ma, S. Zhan, Y. Jia, Q. Shi and Q. Zhou, Appl. Catal., B, 2016, 186, 77–87 CrossRef CAS.
  22. F. Dong, Z. Zhao, Y. Sun, Y. Zhang, S. Yan and Z. Wu, Environ. Sci. Technol., 2015, 49, 12432–12440 CrossRef CAS PubMed.
  23. W. Tian, N. Li and J. Zhou, Appl. Surf. Sci., 2016, 361, 251–258 CrossRef CAS.
  24. L. Cao, R. Wang and D. Wang, Mater. Lett., 2015, 149, 50–53 CrossRef CAS.
  25. D. Gao, Y. Liu, M. Song, S. Shi, M. Si and D. Xue, J. Mater. Chem. C, 2015, 3, 12230–12235 RSC.
  26. O. Fontelles-Carceller, M. J. Muñoz-Batista, M. Fernández-García and A. Kubacka, ACS Appl. Mater. Interfaces, 2016, 8, 2617–2627 CAS.
  27. S. Li, G. Dong, R. Hailili, L. Yang, Y. Li, F. Wang, Y. Zeng and C. Wang, Appl. Catal., B, 2016, 190, 26–35 CrossRef CAS.
  28. D. B. Hernández-Uresti, A. Vázquez, D. Sanchez-Martinez and S. Obregón, J. Photochem. Photobiol., A, 2016, 324, 47–52 CrossRef.
  29. K. Tian, W. J. Liu and H. Jiang, ACS Sustainable Chem. Eng., 2015, 3, 269–276 CrossRef CAS.
  30. X. Lu, J. Shen, J. Wang, Z. Cui and J. Xie, RSC Adv., 2015, 5, 15993–15999 RSC.
  31. S. C. Yan, Z. S. Li and Z. G. Zou, Langmuir, 2009, 25, 10397–10401 CrossRef CAS PubMed.
  32. W. Bing, Z. Chen, H. Sun, P. Shi, N. Gao, J. Ren and X. Qu, Nano Res., 2015, 8, 1648–1658 CrossRef CAS.
  33. M. Tahir, C. Cao, F. K. Butt, F. Idrees, N. Mahmood, Z. Ali, I. Aslam, M. Tanveer, M. Rizwan and T. Mahmood, J. Mater. Chem. A, 2013, 1, 13949–13955 CAS.
  34. Y. Fu, T. Huang, L. Zhang, J. Zhu and X. Wang, Nanoscale, 2015, 7, 13723–13733 RSC.
  35. S. W. Hu, L. W. Yang, Y. Tian, X. L. Wei, J. W. Ding, J. X. Zhong and P. K. Chu, Appl. Catal., B, 2015, 163, 611–622 CrossRef CAS.
  36. J. Wang and W.-D. Zhang, Electrochim. Acta, 2012, 71, 10–16 CrossRef CAS.
  37. Y. Yang, W. Guo, Y. Guo, Y. Zhao, X. Yuan and Y. Guo, J. Hazard. Mater., 2014, 271, 150–159 CrossRef CAS PubMed.
  38. Y. Yang, Y. Guo, F. Liu, X. Yuan, Y. Guo, S. Zhang, W. Guo and M. Huo, Appl. Catal., B, 2013, 142–143, 828–837 CrossRef CAS.
  39. M. Yang, S. Hu, F. Li, Z. Fan, F. Wang, D. Liu and J. Gui, Ceram. Int., 2014, 40, 11963–11969 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra17613e

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