Next Article in Journal
O-Band Emitting InAs Quantum Dots Grown by MOCVD on a 300 mm Ge-Buffered Si (001) Substrate
Previous Article in Journal
Enhanced Terahertz Amplification Based on Photo-Excited Graphene-Dielectric Hybrid Metasurface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Photocatalytic Hydrogen Production of the Polyoxoniobate Modified with RGO and PPy

1
Henan Key Laboratory of Polyoxometalate Chemistry, College of Chemistry and Chemical Engineering, Henan University, Kaifeng 475004, China
2
College of Chemistry and Chemical Engineering, Anyang Normal University, Anyang 455002, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(12), 2449; https://doi.org/10.3390/nano10122449
Submission received: 27 October 2020 / Revised: 20 November 2020 / Accepted: 2 December 2020 / Published: 7 December 2020

Abstract

:
The development of high-efficiency, recyclable, and inexpensive photocatalysts for water splitting for hydrogen production is of great significance to the application of solar energy. Herein, a series of graphene-decorated polyoxoniobate photocatalysts Nb6/PPy-RGO (Nb6 = K7HNb6O19, RGO = reduced graphene oxide, PPy = polypyrrole), with the bridging effect of polypyrrole were prepared through a simple one-step solvothermal method, which is the first example of polyoxoniobate-graphene-based nanocomposites. The as-fabricated photocatalyst showed a photocatalytic H2 evolution activity without any co-catalyst. The rate of 1038 µmol g−1 in 5 h under optimal condition is almost 43 times higher than that of pure K7HNb6O19·13H2O. The influencing factors for photocatalysts in photocatalytic hydrogen production under simulated sunlight were studied in detail and the feasible mechanism is presented in this paper. These results demonstrate that Nb6O19 acts as the main catalyst and electron donor, RGO provides active sites, and PPy acted as an electronic bridge to extend the lifetime of photo-generated carriers, which are crucial factors for photocatalytic H2 production.

Graphical Abstract

1. Introduction

With the widespread energy consumption and environmental problems, it is necessary and urgent to search for sources of green and clean energy to replace traditional fossil fuels [1,2,3,4]. Since the first discovery of photoelectrochemical water splitting into H2 via the semiconductor TiO2 by Fujishima and Honda in 1972, photocatalytic water splitting to hydrogen production has long been viewed as a promising and attractive strategy [5]. In recent years, many semiconductor materials have been reported for photocatalytic hydrogen evolution, in which the band gap, the position of the conduction band, the separation of electrons and holes and the lifetime of carriers are the key factors affecting photocatalysis [6,7]. Nonetheless, the exploitation of highly effective semiconductor materials for photocatalytic hydrogen evolution is still in the infant stage [8,9,10].
Polyoxometalates (POMs) have been applied in photocatalytic H2 or O2 evolution and the photodegradation of organic dyes [11,12,13,14,15,16], due to their controllable and stable structures, superior redox and semiconductor properties [17,18,19,20,21,22]. However, the lower H2 production rates and the introduction of the scarce noble metal have largely limited the scale of their potential applications. It is a feasible strategy to design and prepare photocatalyst composites based on POMs, such as [PW12O40]3−, [PMo12O40]3−, [SiW12O40]4− and [PMo10V2O40]5− [23,24,25,26,27,28,29,30]. In 2018, Zhang reported an interesting investigation on the preparation of a novel Ag3PO4/POM/graphene oxide (GO) heterostructure photocatalyst based on molybdophosphoric acid, which has excellent photocatalytic activity for photodegraded rhodamine B and water splitting for H2 evolution, because of its increased surface area, electronegativity and structural stability [28]. This is a pioneering work in the study of photocatalyst composites from well conductive graphene and POMs for H2 production.
Graphene is an ideal support material and as a solid electron mediator in photocatalysis, owing to the special 2D sp2-hybridied honeycomb structure [2,31], high charge carrier mobility (ca. 10,000 cm2 V−1 s−1) and large theoretical surface area (2600 m2 g−1) [32,33,34]. Therefore, the combination of graphene with semiconductors can not only inhibit the recombination of the photogenerated electron-hole pairs, but also adjust the band gap of photocatalysts and provide more photoactive sites [35,36,37]. More importantly, functionalization of N-doped graphene is a good way to improve the photocatalytic behavior. In particular, pyrrole-N-doped graphene is used as an oxygen-reduction active site to accelerate the interface-catalytic process [38]. Therefore, it is a promising work to introduce pyrrole into POM and graphene-based composites for photocatalytic decomposition of water into H2O2 or hydrogen and oxygen.
As an important kind of POMs, polyoxoniobates have been widely used in photocatalytic H2 evolution due to higher negative charges and stronger basicity of [Nb6O19]8−, which can easily accept electrons and keep an intact structure. Recently, several studies on photocatalytic H2 evolution by polyoxoniobates have been reported, in which precious metals are essential components as co-catalysts [39,40,41,42,43]. Taking these studies into account, we hope to design a new type of POM-based composite photocatalysts with good photocatalytic H2 production performance without any co-catalyst. Therefore, graphene was introduced into [Nb6O19]8− for the first time to produce composite photocatalysts, where it could act as the active sites to produce H2, avoiding the use of the scarce noble metal, while using economical and non-toxic pyrrole as a bridging material to prolong the lifetime of photo-generated carriers [44,45], decreasing the recombination of electron-hole pairs. Herein, Nb6/PPy-RGO (Nb6 = K7HNb6O19, RGO = reduced graphene oxide, PPy = polypyrrole), composites with the appearance of peony flowers were prepared using a simple one-pot solvothermal process. By adjusting the molar rate of Nb6, the volume of pyrrole and solvothermal temperature, a series of Nb6/PPy-RGO nanocomposites were obtained. Among the achieved Nb6/PPy-RGO composites, Nb6/PPy-RGO-0.25 shows an outstanding photocatalytic activity with H2 production rate of 1038 µmol g−1 in 5 h under the simulated sunlight without co-catalyst. Moreover, as-obtained Nb6/PPy-RGO composites demonstrate high stability and superb recyclability, which can be attributed to the collaborations among Nb6, RGO and PPy.

2. Materials and Methods

2.1. Materials

All the reagents and solvents used in this experiment were reagent grade and were used without further treatment. Graphite powder, sulfuric acid (H2SO4, 98%), sodium sulfate (Na2SO4, 99%), sodium nitrate (NaNO3, 99%), potassium permanganate (KMnO4, 99.5%) and pyrrole (C4H5N, 98%) were purchased from Aladdin Biochemical Technology Co. (Shanghai, China). The K7HNb6O19·13H2O was synthesized according to the method in the literature [46], and its purity was determined by infrared spectra (IR). Graphene oxide (GO) was manufactured by a modified Hummers’ method and the purity of GO was confirmed by X-ray diffraction (XRD) [47,48].

2.2. Synthesis of Nb6/PPy-RGO

As shown in Scheme 1, 4 mg graphene oxide (GO) was added to 2 mL distilled water in a 25 mL Teflon lining and ultrasonic for 1 h. Then, a solution containing 110 μL pyrrole (denoted as Py) and 1 mL ethanol was added to the above solution., After stirring for 30 min, Nb6 (345 mg, 0.25 mmol) was added to the mixed solution and heated at 180 °C for 12 h. After cooling to room temperature, the achieved product, Nb6/PPy-RGO-0.25, was respectively washed with deionized water four times and ethanol two times to remove the excess reactants, and finally dried at 60 °C under vacuum for 12 h. According to the same procedure, except for the amount of Nb6 (173 mg, 0.125 mmol) and (523 mg, 0.375 mmol), the samples Nb6/PPy-RGO-0.125 and Nb6/PPy-RGO-0.375 were respectively obtained. Similarly, to gain further insight into the effect of the amount of Nb6 and reaction temperature on the photocatalytic activity of the catalysts, controlling the amount of Nb6 (0.25 mmol), GO and temperature, we achieved Nb6/PPy-RGO-90 and Nb6/PPy-RGO-130 by adjusting the amount of pyrrole (90 and 130 μL). Simultaneously, controlling the amount of Nb6 (0.25 mmol), GO and pyrrole (110 μL), the samples Nb6/PPy-RGO-160 °C and Nb6/PPy-RGO-190 °C were fabricated. As a comparison, the PPy-RGO was also obtained without the addition of Nb6.

2.3. Characterization

The morphology and microstructure of the synthetic samples were characterized using a field emission scanning electron microscopy (FESEM, JSM-7610F, Electronics Co., LTD., Tokyo, Japan) at 10 kV and a transmission electron microscopy (TEM, JEM-2100F, Electronics Co., LTD., Tokyo, Japan) at 200 kV). Powder X-ray diffraction (XRD, Bruker optics Instruments company, Karlsruhe, Germany) was measured from 5° to 80° at room temperature on a Bruker D8 Advance diffractometer with Cu-Kα radiation and the Fourier transform-infrared spectra (FT-IR, KBr pellets, Bruker optics Instruments company, Karlsruhe, Germany) were recorded on a Nicolet 170 SXFT-IR spectrophotometer ranging from 4000 to 500 cm−1. X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, Waltham, MA, USA) was recorded on an ESCALAB 250Xi X-ray photoelectron spectrometer with an Al-Kα (hv = 1486.6 eV) monochromatic radiation source. The binding energy peak position of each sample was calibrated by the C 1s peak at 284.8 eV. UV-visible spectra were achieved on a Shimadzu UV–2600 UV/Vis spectrophotometer (Shimadzu Instrument Co., LTD, Kyoto-fu, Japan) using BaSO4 as a reference. Raman spectroscopy (Renishaw company, London, UK) was prepared using a Renishaw in Via Raman spectrometer with excitation by 325 nm Photoluminescence (PL) spectra were obtained at excitation wavelength 385 nm with TU-1900 (PuXi Company, Beijing, China).

2.4. Photocatalytic Hydrogen Production

Photocatalytic hydrogen producing experiments were performed using a method similar to the literature [6]. Photocatalyst (50 mg) was added to a solution of deionized water (40.0 mL) and CH3OH (10.0 mL, pH = 7) in a quartz vessel. The reaction system was irradiated by a 300 W Xe lamp (Perfect light PLS-SXE300, Perfectlight Technology Co., Ltd, Beijing, China), and the temperature was maintained at about 5 °C. The product H2 was measured by gas chromatograph (FL-9790) on line. After each photocatalytic test, photocatalysts were collected by centrifugation without reactivated before each cycle, washed with H2O and C2H5OH to remove impurities, then used after drying.

2.5. Photocatalytic Measurement

The Mott–Schottky curves were tested on an AMETEK Princeton Applied Research (Versa STAT 4, Princeton, NJ, USA) electrochemical workstation (FTO = fluorine-doped tin oxide substrate, 1 cm × 1.5 cm). The electrochemical impedance spectroscopy (EIS) and photocurrent response were recorded on an electrochemical workstation (CHI660, Chenhua, Shanghai, China) equipped with a three-electrode system with a complex/FTO as the working electrode, platinum foil as the counter electrode and Ag/AgCl as the reference electrode in 0.2 M sodium sulfate solution (Na2SO4). The working electrode complex/FTO was prepared by dropwise adding 50 μL of sample suspensions containing Nb6/PPy-RGO composites (3.0 mg), ethanol (1.0 mL), and nafion (20 μL) onto a FTO substrate (1 cm × 1.5 cm).

3. Results and Discussion

3.1. Structural Characterizations

The XRD patterns of GO, Nb6, and Nb6/PPy-RGO composites were investigated. As shown in Figure 1 and Figure S1, pristine GO and RGO show a single peak near 12° and 26°, respectively, while pristine K7HNb6O19 has sharp diffraction peaks at 9.7°, 26° and 48°, respectively, which is consistent with the literature reports [46,47,48]. The broad peaks centered on 27.5° of Nb6/PPy-RGO composites can be attributed to the combined action of PPy and RGO, while the characteristic peaks at 46° are attributed to the superposition of Nb6 and PPy [49,50,51]. In Nb6/PPy-RGO composite materials, although it is not obvious in Nb6/PPy-RGO-0.325 and Nb6/PPy-RGO-0.25, the characteristic peaks of raw materials can still be found. No additional diffraction peaks were observed, confirming the formation of the Nb6/PPy-RGO composites. The crystallite sizes of the catalysts calculated from the XRD patterns were shown in Table 1.
The structure of Nb6/PPy-RGO-0.25 composite has also been evidenced by Fourier transform-infrared (FT-IR) spectroscopy (Figure 2). The peak at 1398 cm−1 is assigned to the C-O stretching vibration due to π–π interaction between PPy and RGO. The peak at 1637 cm−1 is assigned to H-O-H stretching vibration and the peak at 1258 cm−1 was assigned to C-H and C-N in-plane deformation vibration, further implying the combination of GO with PPy [52,53]. The characteristic peaks at 840 and 887 cm−1 could be attributed to the terminal Nb-O stretching vibration, and the peaks appearing at 532 and 597 cm−1 correspond to Nb-O-Nb bonds [54]. The result shows that Nb6 was successfully inserted into the final composite.
From the Raman spectra shown in Figure 3, the higher wavenumber bands at 1334 and 1584 cm−1 are ascribed to D band and G band of GO. It is worth noting that the peak intensity ratio (ID/IG) is a popular method to evaluate the disorder and reduction of graphene materials. The ID/IG ratio of the pure GO is about 0.76, while that of the Nb6/PPy-RGO-0.25 composite is about 0.99. The numeric addition indicated that the oxygen-containing functional groups were partially reduced, meaning the transition of GO to RGO in the process of solvothermal reaction [54]. Compared with the characteristic peaks of Nb6 (541, 827 and 875 cm−1) [55,56], the redshift could be found for Nb6/PPy-RGO-0.25 (878 and 620 cm−1), which may be assigned to the strong interaction among the Nb6, PPy, and RGO of the composites.
In order to further identify the structure of composites, the XPS spectra were tested for Nb6/PPy-RGO-0.25 composites and GO, as well as Nb6 (Figure 4). As shown in Figure 4A, the full spectrum clearly shows the presence of C, O, N, K, and Nb elements in the composite sample, which is consistent with the chemical composition of the photocatalyst. As shown in Figure 4B, the binding energies of C 1s, including 284.05, 285.60, 286.47 and 288.37 eV, could be attributed to carbon of the non-oxygenated ring, C-OH, C-O and C=O in Nb6/PPy-RGO-0.25, respectively, which are lower and weaker than that of pure GO (284.74, 285.97, 287.07 and 289.08 eV). These results demonstrate that the GO has been successfully reduced in composites, which is in agreement with Raman spectra [57,58,59,60]. For pure GO, the XPS of O 1s contains two peaks at 532.20 and 533.08 eV assigned to C=O/C-O and O-C=O, respectively. The peaks at 531.96, 529.05, and 530.63 eV in the pure Nb6 sample belong to Nb-O-Nb, Nb-O-H, and crystal water, respectively. There are two main peaks at 529.73 and 531.74 eV in Nb6/PPy-RGO-0.25, which can be attributed to binding energies of O 1s in Nb-O-Nb and Nb-O-H (Figure 4C [61]. Figure 4D illustrates a comprehensive Nb 3d XPS analysis of the Nb6/PPy-RGO-0.25 and Nb6. The binding energies of 206.49 and 209.26 eV could be assigned to Nb 3d5/2 and Nb 3d3/2 of Nb5+ (highest oxidation state) in Nb6/PPy-RGO-0.25. Compared with Nb6, the peaks of Nb 3d shifted to higher binding energy by ~1 eV, which may be caused by electron transfer from Nb6 to RGO in composites [54]. As shown in Figure 4E, the N 1s peaks centered at 398.01 and 399.55 eV are vested to the pyridinic and pyrrolic of N atom in polymer. In addition, there is a weak peak at the center of 394.11 eV, corresponding to the Nb-N bond, indicating the formation of chemical bonds between PPy and Nb6 [62].

3.2. Nanosphere Morphologies

As shown in Figure 5, the results show that the morphology of Nb6 is rodlike. However, when fabricated with PPy and RGO, Nb6/PPy-RGO-X (X = 0.125, 0.25, 0.375) display the flower-like methodology. To deeply understand the formation of flower-like methodology, the controlled experiments including solvothermal reaction temperatures and the amount of Py have been respectively conducted using Nb6/PPy-RGO-0.25 as a representative. As shown in Figures S2 and S3, the samples probably agglomerated when changing the conditions of 160 °C or Py (130 μL). However, the lower amount of Py, with 90 μL, or higher temperature of 190 °C may reduce the surface area and could not form morphology of flowers for Nb6/PPy-RGO-0.25. These results demonstrate that the flower-like morphology for Nb6/PPy-RGO-0.25 largely depends on the temperature and the amounts of Py in the reaction. The corresponding EDS elemental mapping of Nb6/PPy-RGO-0.25 indicated the elements of C, N, O, K and Nb were uniformly distributed throughout the whole composite (Figure 5E).

3.3. Photochemistry and Electrochemistry

The UV-Vis diffuse reflectance spectra of Nb6/PPy-RGO composites were recorded at room temperature to discuss their optical properties, together with Nb6 for comparison. As shown in Figure 6, Nb6/PPy-RGO composites exhibit broad absorptions in the range of 200–340 nm. Compared with the absorption of Nb6 (200–315 nm), the light absorption of Nb6/PPy-RGO samples was obviously widened, which is related to the introduction of PPy and RGO. Due to the zero band gap for RGO, with the increasing amounts of Nb6, the absorbance intensity for Nb6/PPy-RGO composites become weaker (Figure 6a). Simultaneously, the color becomes lighter and the ability to capture light is decreased (Figure S4), hinting that the enhanced absorption is mainly attributed to the introduction of RGO. The band-gap values of Nb6 and Nb6/PPy-RGO composites (from 0.125 to 0.375) obtained by the Kubelka–Munk function via the UV-Vis diffuse reflectance are 4.14, 3.66, 3.71, and 3.76 eV, respectively, indicating the semiconductor characters of Nb6/PPy-RGO composites (Figure 6b). Compared to the band gap value of Nb6, those of Nb6/PPy-RGO composites are obviously narrowed through the incorporation of RGO and PPy.
Furthermore, the potentials of conduction band (CB) of Nb6/PPy-RGO composites could be obtained from electrochemical Mott–Schottky plots (Figure S5). The plots of C−2 vs. potential exhibited positive slopes, indicating that we should consider the composites as typical n-type semiconductors [57]. The flat band position of Nb6/PPy-RGO composites determined from the intersection of the Mott–Schottky plot and X axis are approximately −1.01 V (Nb6/PPy-RGO-0.125), −1.34 V (Nb6/PPy-RGO-0.25) and −1.23 V (Nb6/PPy-RGO-0.375) (vs. NHE, NHE = Normal hydrogen electrode). These results imply that the potentials of CB of Nb6/PPy-RGO composites are more negative than the redox potential of H+/H2 (−0.41 V vs. NHE, pH = 7) [63], indicating the title composites should be used as photocatalysts for hydrogen evolution via water splitting.
As shown in Figure S6, compared with other composites, the Nb6/PPy-RGO-0.25 composite displays the weakest PL intensity at 378 nm, and the amount of pyrrole and the solvothermal temperature had little effect on the separation of photo-excited charge and holes, indicative of the lowest combination of electron-hole pairs. Because K7HNb6O19 has good water solubility, electrochemical measurement cannot be performed under the same conditions. As shown in Figure 7a, the highest photocurrent response of Nb6/PPy-RGO-0.25 shows better separation ability and lower recombination of photogenerated electron-hole pairs [42]. As depicted in Figure 7b, the smallest radius of Nb6/PPy-RGO-0.25 indicates lower charge transfer resistance. This result confirms that more photo-generated charge carriers can be separated by constructing Nb6/PPy-RGO-0.25, which is consistent with the result of PL analysis and photocurrent response.

3.4. Photocatalytic Hydrogen Evolution

Photocatalytic activity of Nb6/PPy-RGO composites for hydrogen evolution via water splitting was evaluated in aqueous solution with methanol (CH3OH) as a sacrificial electron donor. As shown in Figure 8, the H2 production for Nb6/PPy-RGO composites was 351.5 (Nb6/PPy-RGO-0.125), 1038 (Nb6/PPy-RGO-0.25) and 690.5 μmol g−1 (Nb6/PPy-RGO-0.375) in 5 h, respectively. Under the same conditions, the H2 production was 25 μmol g−1 for the Nb6 catalyst. Apparently, the composites exhibit higher photocatalytic activity by introduction of PPy and RGO. Compared with the other two composites, Nb6/PPy-RGO-0.125 and Nb6/PPy-RGO-0.375, Nb6/PPy-RGO-0.25 exhibits better photocatalytic efficiency and the turnover frequency (TOF) value reaches 830.4 × 10−3 h−1 (Table 2). These results agree well with EIS and photo-current measurements.
Turnover frequency (TOF) was based on Nb6, which has been widely used to evaluate the property of catalysts, especially for POM-based catalysts. The details are as follows:
TOF   =   n ( H 2 ) / h n ( N b 6 )
Taking Nb6/PPy-RGO-0.25 as an example:
TOF ( Nb 6 / PPy - RGO - 0.25 ) = n ( H 2 ) / h n ( N b 6 ) = 1038 × 10 6 mol / 5 h 0.25 × 10 3 mol = 830.4 × 10 3 h 1
From the above table, the sample Nb6/PPy-RGO-0.25 showed the better photocatalytic activity.
The photocatalytic system has also been optimized by changing the amount or species of sacrificial electron donors by using the optimal as Nb6/PPy-RGO-0.25 photocatalyst. As shown in Figure S7a, when the concentration of CH3OH increased from 0 to 20% (v:v), the hydrogen evolution increased from 0 to 1038 μmol g−1, but decreased to 274 and 94 μmol g−1 with higher concentrations of 25% (v:v) and 30% (v:v), respectively. These results imply that the amount of CH3OH played an important role in photocatalytic hydrogen evolution. The more sacrificial electron donors may react with holes of photocatalyst to release more photo-excited electrons to produce more hydrogen. However, with an excess of CH3OH, it should also decrease the concentrations of substrate, leading to lower photocatalytic efficiency. Furthermore, the species of sacrificial electron donor was optimized by controlling the ratio of sacrificial agent/H2O (v:v = 10:40) (Figure S7b). The title photocatalyst exhibited the optimal H2 production of 1038 μmol g−1 with CH3OH as the sacrificial agent, but varied in the range of 249.5–834 μmol g−1 with other sacrificial agents. As shown in Figure S8, the best sample, Nb6/PPy-RGO-0.25, exhibits the highest photocatalytic H2 production, indicating the optimal reaction conditions with pyrrole dosage of 110 μL and reaction temperature of 180 °C.
In order to prove the reusability of Nb6/PPy-RGO-0.25, the recycling experiment was carried out (Figure 9). There was no noticeable deactivation observed for the catalyst and the yields of H2 evolution for the four recycling experiments all remained at 1000 μmol g−1 in 5 h. In addition, the XRD pattern (Figure S9), Raman spectra (Figure S10) and SEM image (Figure S11) of the recovered catalysts were alike to those of the as-prepared sample, further demonstrating the recyclability and stability of Nb6/PPy-RGO-0.25 in the photocatalysis reaction.

3.5. Proposed Photocatalytic Mechanism

According to the Mott–Schottky plot shown in Figure S5, the conduction band minimum (CBM) of the composites could be estimated to explain the enhanced photocatalytic activity mechanism, which is about −1.34 eV for the sample Nb6/PPy-RGO-0.25 possessing the best H2 evolution performance. Meanwhile, valence band maximum (VBM) potential of Nb6/PPy-RGO-0.25 can be determined based on the formula EVBM = ECBM + Eg, which is about 2.37 eV [64]. As displayed in Figure 10, the pure Nb6 exhibits negligible photocatalytic activity, which may be due to the fast recombination of electrons and holes. Higher photocatalytic efficiency can be achieved for Nb6/PPy-RGO composites by the modification with PPy and RGO. It has been reported that conductive polymers can be used as photocatalysts for H2 production from water splitting [65]. However, in such Nb6/PPy-RGO-0.25 composites, pure PPy and RGO cannot exhibit photocatalytic activity due to their combination of electrons and holes. On the other hand, they can prolong the lifetime of charge carriers to promote the separation efficiency of electron-hole pairs in Nb6/PPy-RGO-0.25 composites due to their excellent π conjugated chain structure and electrical conductivity. In addition, RGO can also serve as active sites to produce H2, avoiding the use of noble metals, while PPy acted as conductive polymers, promoting the transfer of the electrons [52,53,66,67,68,69]. In particular, the π–π interaction between RGO and PPy further strengthens the electrical conductivity. Based on the above results, a probable photocatalytic mechanism for composites could be proposed as follows: under light irradiation, the electrons in the valence band (VB) of Nb6/PPy-RGO-0.25 can be excited to the CB, causing holes in the VB. On one hand, the holes in the VB could be quenched by accepting electrons from CH3OH. On the other hand, the photoelectron in the CB of Nb6/PPy-RGO-0.25 could be transferred to RGO through the bridge of PPy. Acting as photoactive sites, the RGO accepting electrons would react with H2O to produce H2.

4. Conclusions

In summary, a series of unprecedented polyoxoniobate-graphene nanocomposites, Nb6/PPy-RGO, have been successfully synthesized through a simple one-step solvothermal method. This is the first example of polyoxoniobate-graphene nanocomposites for photocatalytic H2 production. This work presents a new idea for the design and synthesis of high-performance polyoxometalates-graphene photocatalysts in photocatalytic H2 production. Nb6/PPy-RGO was optimized by adjusting the amount of Nb6, volume of pyrrole and temperature, displaying superb photocatalytic H2 production of 1038 μmol g−1 in 5 h without a co-catalyst, almost 43 times more than pure Nb6, in which the collaborations among Nb6, PPy and RGO play vital roles in such a photocatalytic system. In further work, we will focus on studying more fascinating polyoxoniobate-graphene nanocomposites with new interesting properties.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/12/2449/s1, Figure S1: X-ray diffraction (XRD) patterns for series of concentration of pyrrole and temperature and the corresponding starting materials. Figure S2: SEM images for (a) Nb6-RGO; (b) Nb6/PPy-RGO-90 μL; (c) Nb6/PPy-RGO-110 μL; (d) Nb6/PPy-RGO-130 μL (SEM = scanning electron microscopy, Nb6 = K7HNb6O19, RGO = reduced graphene oxide, PPy = polypyrrole). Figure S3: SEM images for (a) Nb6/PPy-RGO-160 °C; (b) Nb6/PPy-RGO-180°C; (c) Nb6/PPy-RGO-190 °C. Figure S4: The color change of samples. Figure S5: Mott-Schottky plots of (a) Nb6/PPy-RGO-0.125; (b) Nb6/PPy-RGO-0.25; (c) Nb6/PPy-RGO-0.375, in 0.2 M Na2SO4 aqueous solution with pH = 7. Figure S6: Photoluminescence (PL) spectra recorded at room temperature in the range of 325–450 nm with an excitation wavelength of 378 nm for adjust (a) molar ratio of Nb6; (b) volume of pyrrole; (c) temperature. Figure S7: (a) Rate of H2 evolution as a function of CH3OH concentration, (b) Rate of H2 evolution as a function of type of sacrificial agents. Figure S8: Photocatalytic hydrogen production in concentration of pyrrole (a) and temperature (b) in aqueous solution with MeOH 20%. Figure S9: XRD patterns of Nb6/PPy-RGO-0.25 before and after photocatalytic H2 evolution reaction. Figure S10: Raman spectra of Nb6/PPy-RGO-0.25 before and after photocatalytic H2 evolution reaction. Figure S11: SEM images for (a) Nb6/PPy-RGO-0.25 and (b) that after the recycle of photocatalytic hydrogen evaluation.

Author Contributions

Conceptualization, S.H. and D.D.; investigation, S.H., W.L., H.L. and M.Z.; writing—original draft preparation, S.H. and L.L.; writing—review and editing, Y.B., D.D. and L.L.; supervision, Y.B. and D.D.; funding acquisition, Y.B. and J.P. All authors have read and agreed to the published version of the manuscript.

Funding

The author thanks the support of the National Natural Science Foundation of China (No. 21971055), the Foundation of the Education Department of Henan Province of China (No. 21A150011 and No. 19B150001), the Postdoctoral Science Foundation of China (No. 2019M652518), and the Postdoctoral Scientific Research Foundation of Henan Province.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, H.; Sun, Y.; Yuan, Z.Y.; Zhu, Y.P.; Ma, T.Y. Titanium Phosphonate Based Metal-Organic Frameworks with Hierarchical Porosity for Enhanced Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2018, 57, 3222–3227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Li, X.; Yu, J.G.; Wageh, S.; Al-Ghamdi, A.A.; Xie, J. Graphene in Photocatalysis: A Review. Small 2016, 12, 6640–6696. [Google Scholar] [CrossRef]
  3. Zhang, G.; Wang, G.; Liu, Y.; Liu, H.; Qu, J.; Li, J. Highly Active and Stable Catalysts of Phytic Acid-Derivative Transition Metal Phosphides for Full Water Splitting. J. Am. Chem. Soc. 2016, 138, 14686–14693. [Google Scholar] [CrossRef] [PubMed]
  4. Gao, M.M.; Zhu, L.L.; Ong, W.L.; Wang, L.; Ho, G.W. Structural design of TiO2-based photocatalyst for H2 production and degradation applications. Catal. Sci. Technol. 2015, 5, 4703–4726. [Google Scholar] [CrossRef]
  5. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  6. Li, L.; Zhu, S.; Hao, R.; Wang, J.J.; Yang, E.C.; Zhao, X.J. Amino group promoted photocatalytic hydrogen evolution activity observed in two copper(ii)–based layered complexes. Dalton Trans. 2018, 47, 12726–12733. [Google Scholar] [CrossRef]
  7. Zhao, X.X.; Feng, J.R.; Liu, J.W.; Lu, J.; Shi, W.; Yang, G.M.; Wang, G.C.; Feng, P.Y.; Cheng, P. Metal-Organic Framework-Derived ZnO/ZnS Heteronanostructures for Efficient Visible-Light-Driven Photocatalytic Hydrogen Production. Adv. Sci. 2018, 5, 1700590–1700598. [Google Scholar] [CrossRef]
  8. Afroz, K.; Moniruddin, M.; Bakranov, N.; Kudaibergenov, S.; Nuraje, N. A heterojunction strategy to improve the visible light sensitive water splitting performance of photocatalytic materials. J. Mater. Chem. A 2018, 6, 21696–21718. [Google Scholar] [CrossRef]
  9. Stroyuk, O.; Raevskaya, A.; Gaponik, N. Solar light harvesting with multinary metal chalcogenide nanocrystals. Chem. Soc. Rev. 2018, 47, 5354–5422. [Google Scholar] [CrossRef]
  10. Kumar, P.; Boukherroub, R.; Shankar, K. Sunlight-driven water-splitting using two-dimensional carbon based semiconductors. J. Mater. Chem. A 2018, 6, 12876–12931. [Google Scholar] [CrossRef]
  11. Li, S.J.; Liu, S.M.; Liu, S.X.; Liu, Y.W.; Tang, Q.; Shi, Z.; Ou yang, S.X.; Ye, J.H. {Ta12}/{Ta16} Cluster-Containing Polytantalotungstates with Remarkable Photocatalytic H2 Evolution Activity. J. Am. Chem. Soc. 2012, 138, 19716–19721. [Google Scholar] [CrossRef] [PubMed]
  12. Qin, J.S.; Du, D.Y.; Guan, W.; Bo, X.J.; Li, Y.F.; Guo, L.P.; Su, Z.M.; Wang, Y.Y.; Lan, Y.Q.; Zhou, H.C. Ultrastable Polymolybdate-Based Metal-Organic Frameworks as Highly Active Electrocatalysts for Hydrogen Generation from Water. J. Am. Chem. Soc. 2015, 137, 7169–7177. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, L.; Ding, Y.; Zheng, M. Polyoxometalate-based manganese clusters as catalysts for efficient photocatalytic and electrochemical water oxidation. Appl. Catal. B Environ. 2017, 209, 45–52. [Google Scholar] [CrossRef]
  14. Geletii, Y.V.; Botar, B.; Kögerler, P.; Hillesheim, D.A.; Musaev, D.G.; Hill, C.L. An All-Inorganic, Stable, and Highly Active Tetraruthenium Homogeneous Catalyst for Water Oxidation. Angew. Chem. Int. Ed. 2008, 47, 3896–3899. [Google Scholar] [CrossRef]
  15. Du, X.Q.; Zhao, J.L.; Mi, J.Q.; Ding, Y.; Zhou, P.P.; Ma, B.C.; Zhao, J.W.; Song, J. Efficient photocatalytic H2 evolution catalyzed by an unprecedented robust molecular semiconductor {Fe11} nanocluster without cocatalysts at neutral conditions. Nano Energy 2015, 16, 247–255. [Google Scholar] [CrossRef]
  16. Chen, L.; Chen, W.L.; Wang, X.L.; Li, Y.G.; Su, Z.M.; Wang, E.B. Polyoxometalates in dye-sensitized solar cells. Chem. Soc. Rev. 2019, 48, 260–284. [Google Scholar] [CrossRef]
  17. Troupis, A.; Hiskia, A.; Papaconstantinou, E. Photocatalytic Reduction and Recovery of Copper by Polyoxometalates. Environ. Sci. Technol. 2002, 36, 5355–5362. [Google Scholar] [CrossRef]
  18. Kornarakis, I.; Lykakis, I.N.; Vordos, N.; Armatas, G.S. Efficient visible-light photocatalytic activity by band alignment in mesoporous ternary polyoxometalate–Ag2S–CdS semiconductors. Nanoscale 2014, 6, 8694–8703. [Google Scholar] [CrossRef]
  19. Hu, B.; Wang, P.F.; Hou, J.; Wang, C.; Qian, J.; Zhang, N.N.; Yuan, Q.S. Effects of titanium dioxide (TiO2) nanoparticles on the photodissolution of particulate organic matter: Insights from fluorescence spectroscopy and environmental implications. Environ. Pollut. 2017, 229, 19–28. [Google Scholar] [CrossRef]
  20. Park, H.; Choi, W. Photocatalytic conversion of benzene to phenol using modified TiO2 and polyoxometalates. Catal. Today 2005, 101, 291–297. [Google Scholar] [CrossRef]
  21. Zhao, S.; Zhao, X. Polyoxometalates-derived metal oxides incorporated into graphitic carbon nitride framework for photocatalytic hydrogen peroxide production under visible light. J. Catal. 2018, 366, 98–106. [Google Scholar] [CrossRef]
  22. Li, K.X.; Yan, L.S.; Zeng, Z.X.; Luo, S.L.; Luo, X.B.; Liu, X.M.; Guo, H.Q.; Guo, Y.H. Fabrication of H3PW12O40-doped carbon nitride nanotubes by one–step hydrothermal treatment strategy and their efficient visible-light photocatalytic activity toward representative aqueous persistent organic pollutants degradation. Appl. Catal. B Environ. 2014, 156–157, 141–152. [Google Scholar] [CrossRef]
  23. Gamelas, J.A.F.; Evtuguin, D.V.; Esculcas, A.P. Transition metal substituted polyoxometalates supported on amine-functionalized silica. Transit. Met. Chem. 2007, 32, 1061–1067. [Google Scholar] [CrossRef]
  24. Sajjad, S.; Leghari, S.A.K.; Iqbal, A. Study of Graphene Oxide Structural Features for Catalytic, Antibacterial, Gas Sensing, and Metals Decontamination Environmental Applications. ACS Appl. Mater. Inter. 2017, 9, 43393–43414. [Google Scholar] [CrossRef] [PubMed]
  25. Kamat, P.V. Graphene-Based Nanoassemblies for Energy Conversion. J. Phys. Chem. Lett. 2011, 2, 242–251. [Google Scholar] [CrossRef]
  26. Zhang, N.; Zhang, Y.H.; Xu, Y.J. Recent progress on graphene-based photocatalysts: Current status and future perspectives. Nanoscale 2012, 4, 5792–5813. [Google Scholar] [CrossRef]
  27. Liu, J.; Zhang, H.C.; Tang, D.; Zhang, X.; Yan, L.K.; Han, Y.Z.; Huang, H.; Liu, Y.; Kang, Z.H. Carbon Quantum Dot/Silver Nanoparticle/Polyoxometalate Composites as Photocatalysts for Overall Water Splitting in Visible Light. ChemCatChem 2014, 6, 2634–2641. [Google Scholar] [CrossRef]
  28. Liu, G.D.; Zhao, X.F.; Zhang, J.; Liu, S.J.; Sha, J.Q. Z-scheme Ag3PO4/POM/GO heterojunction with enhanced photocatalytic performance for degradation and water splitting. Dalton Trans. 2018, 47, 6225–6232. [Google Scholar] [CrossRef]
  29. Tang, Y.J.; Wang, Y.; Wang, X.L.; Li, S.L.; Huang, W.; Dong, L.Z.; Liu, C.H.; Li, Y.F.; Lan, Y.Q. Molybdenum Disulfide/Nitrogen-Doped Reduced Graphene Oxide Nanocomposite with Enlarged Interlayer Spacing for Electrocatalytic Hydrogen Evolution. Adv. Energy Mater. 2016, 6, 1600116. [Google Scholar] [CrossRef]
  30. Tang, Y.J.; Liu, C.H.; Huang, W.; Wang, X.L.; Dong, L.Z.; Li, S.L.; Lan, Y.Q. Bimetallic Carbides-Based Nanocomposite as Superior Electrocatalyst for Oxygen Evolution Reaction. ACS Appl. Mater. Inter. 2017, 9, 16977–16985. [Google Scholar] [CrossRef]
  31. Iwase, A.; Ng, Y.H.; Yoshimi, I.; Kudo, A.; Amal, R. Reduced Graphene Oxide as a Solid-State Electron Mediator in Z-Scheme Photocatalytic Water Splitting under Visible Light. J. Am. Chem. Soc. 2011, 133, 11054–11057. [Google Scholar] [CrossRef] [PubMed]
  32. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Xiang, Q.J.; Yu, J.G.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2012, 41, 782–796. [Google Scholar] [CrossRef]
  34. Geim, A.K. Graphene: Status and Prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Wei, T.; Zhang, M.; Wu, P.; Tang, Y.J.; Li, S.L.; Shen, F.C.; Wang, X.L.; Zhou, X.P.; Lan, Y.Q. POM-based metal-organic framework/reduced graphene oxide nanocomposites with hybrid behavior of battery-supercapacitor for superior lithium storage. Nano Energy 2017, 34, 205–214. [Google Scholar] [CrossRef]
  36. Li, H.L.; Gupta, J.; Wang, S.; Zhang, N.; Bubeck, C. Photoreduction of graphene oxide with polyoxometalate clusters and its enhanced saturable absorption. J. Colloid Interface Sci. 2014, 427, 25–28. [Google Scholar] [CrossRef] [PubMed]
  37. Huang, Y.Y.; Xiao, A.X.; Hou, G.H.; Li, H.T.; Guo, T.; Guan, B.O. Photocatalysis in an evanescent field: An in situ approach to studying photocatalytic performance by tracing interfacial refractive index changes and kinetics. J. Mater. Chem. A 2018, 6, 20513–20522. [Google Scholar] [CrossRef]
  38. Zhang, N.; Yang, M.Q.; Liu, S.Q.; Sun, Y.G.; Xu, Y.J. Waltzing with the Versatile Platform of Graphene to Synthesize Composite Photocatalysts. Chem. Rev. 2015, 115, 10307–10377. [Google Scholar] [CrossRef]
  39. Wang, Z.L.; Tan, H.Q.; Chen, W.L.; Li, Y.G.; Wang, E.B. A copper(ii)-ethylenediamine modified polyoxoniobate with photocatalytic H2 evolution activity under visible light irradiation. Dalton Trans. 2012, 41, 9882–9884. [Google Scholar] [CrossRef] [Green Version]
  40. Shen, J.Q.; Zhang, Y.; Zhang, Z.M.; Li, Y.G.; Gao, Y.Q.; Wang, E.B. Polyoxoniobate-based 3D framework materials with photocatalytic hydrogen evolution activity. Chem. Commun. 2014, 50, 6017–6019. [Google Scholar] [CrossRef]
  41. Geng, Q.H.; Liu, Q.S.; Ma, P.T.; Wang, J.P.; Niu, J.Y. Synthesis crystal structure and photocatalytic properties of an unprecedented arsenic-disubstituted Lindqvist-type peroxopolyoxoniobate ion: {As2Nb4(O2)4O14H1.5}4.5−. Dalton Trans. 2014, 43, 9843–9846. [Google Scholar] [CrossRef] [PubMed]
  42. Hu, J.F.; Wang, Y.; Zhang, X.N.; Chi, Y.N.; Yang, S.; Li, J.K.; Hu, C.W. Controllable Assembly of Vanadium-Containing Polyoxoniobate-Based Three-Dimensional Organic-Inorganic Hybrid Compounds and Their Photocatalytic Properties. Inorg. Chem. 2016, 55, 7501–7507. [Google Scholar] [CrossRef] [PubMed]
  43. Niu, J.Y.; Li, F.; Zhao, J.W.; Ma, P.T.; Zhang, D.D.; Bassil, B.; Kortz, U.; Wang, J.P. Tetradecacobalt(II)-Containing 36-Niobate [Co14(OH)16(H2O)8Nb36O106]20− and Its Photocatalytic H2 Evolution Activity. Chem. Eur. J. 2014, 20, 9852–9857. [Google Scholar] [CrossRef] [PubMed]
  44. Wu, F.; Xie, A.M.; Sun, M.X.; Wang, Y.; Wang, M.Y. Reduced graphene oxide (RGO) modified spongelike polypyrrole (PPy) aerogel for excellent electromagnetic absorption. J. Mater. Chem. A 2015, 3, 14358–14369. [Google Scholar] [CrossRef]
  45. Abaci, U.; Guney, H.Y.; Kadiroglu, U. Morphological and electrochemical properties of PPy, PAni bilayer films and enhanced stability of their electrochromic devices (PPy/PAni–PEDOT, PAni/PPy–PEDOT). Electrochim. Acta 2013, 96, 214–224. [Google Scholar] [CrossRef]
  46. Filowitz, M.; Ho, R.K.C.; Klemperer, W.G.; Shum, W. 17O nuclear magnetic resonance spectroscopy of polyoxometalates. 1. Sensitivity and resolution. Inorg. Chem. 1979, 18, 93–103. [Google Scholar] [CrossRef]
  47. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  48. Bao, C.L.; Song, L.; Xing, W.Y.; Yuan, B.H.; Wilkie, C.A.; Huang, J.L.; Guo, Y.Q.; Hu, Y. Preparation of graphene by pressurized oxidation and multiplex reduction and its polymer nanocomposites by masterbatch-based melt blending. J. Mater. Chem. 2012, 22, 6088–6096. [Google Scholar] [CrossRef]
  49. Suvina, V.; Krishna, S.M.; Nagaraju, D.H.; Melo, J.S.; Balakrishna, R.G. Polypyrrole-reduced graphene oxide nanocomposite hydrogels: A promising electrode material for the simultaneous detection of multiple heavy metal ions. Mater. Lett. 2018, 232, 209–212. [Google Scholar] [CrossRef]
  50. Yin, Y.; Zhang, H.T.; Huang, P.R.; Xiang, C.L.; Zou, Y.J.; Xu, F.; Sun, L.X. Inducement of nanoscale Cu–BTC on nanocomposite of PPy–rGO and its performance in ammonia sensing. Mater. Res. Bull. 2018, 99, 152–160. [Google Scholar] [CrossRef]
  51. Liu, R.J.; Li, S.W.; Yu, X.L.; Zhang, G.J.; Zhang, S.J.; Yao, J.N.; Keita, B.; Nadjo, L.; Zhi, L.J. Facile Synthesis of Au-Nanoparticle/Polyoxometalate/Graphene Tricomponent Nanohybrids: An Enzyme-Free Electrochemical Biosensor for Hydrogen Peroxide. Small 2012, 8, 1398–1406. [Google Scholar] [CrossRef] [PubMed]
  52. Han, Y.Q.; Zhang, Z.L.; Yang, M.; Li, T.X.; Wang, Y.M.; Cao, A.P.; Chen, Z.X. Facile preparation of reduced graphene oxide/polypyrrole nanocomposites with urchin-like microstructure for wide-potential-window supercapacitors. Electrochim. Acta 2018, 289, 238–247. [Google Scholar] [CrossRef]
  53. Jiang, Y.Z.; Gu, Y.; Nie, G.D.; Chi, M.Q.; Yang, Z.Z.; Wang, C.; Wei, Y.; Lu, X.F. Synthesis of RGO/Cu8S5/PPy Composite Nanosheets with Enhanced Peroxidase-Like Activity for Sensitive Colorimetric Detection of H2O2 and Phenol. Part. Part. Syst. Charact. 2017, 3, 1600233–1600240. [Google Scholar] [CrossRef]
  54. Gan, Q.Y.; Shi, W.L.; Xing, Y.J.; Hou, Y.A. Polyoxoniobate/g-C3N4 Nanoporous Material with High Adsorption Capacity of Methylene Blue from Aqueous Solution. Front. Chem. 2018, 6, 7–16. [Google Scholar] [CrossRef]
  55. Aureliano, M.; Ohlin, C.A.; Vieira, M.O.; Marques, M.P.M.; Casey, W.H.; Batista de Carvalho, L.A.E. Characterization of decavanadate and decaniobate solutions by Raman spectroscopy. Dalton Trans. 2016, 45, 7391–7399. [Google Scholar] [CrossRef] [Green Version]
  56. Fullmer, L.B.; Malmberg, C.E.; Fast, D.B.; Wills, L.A.; Cheong, P.H.Y.; Dolgos, M.R.; Nyman, M. Aqueous tantalum polyoxometalate reactivity with peroxide. Dalton Trans. 2017, 46, 8486–8493. [Google Scholar] [CrossRef]
  57. Akhavan, O.; Ghaderi, E. Photocatalytic Reduction of Graphene Oxide Nanosheets on TiO2 Thin Film for Photoinactivation of Bacteria in Solar Light Irradiation. J. Phys. Chem. C 2009, 113, 20214–20220. [Google Scholar] [CrossRef]
  58. Ma, J.; He, Y.; Shi, H.; Fan, Y.; Yu, H.; Li, Y.B. Stable graphene oxide-based composite membranes intercalated with montmorillonite nanoplatelets for water purification. J. Mater. Sci. 2019, 54, 2241–2255. [Google Scholar] [CrossRef]
  59. Zhang, L.B.; Li, X.T.; Wang, X.H.; Han, L.; Liu, R.Y.; Zheng, B.B.; Li, W.T.; Liu, X.L.; Liu, Y.C. Photo-assisted preparation and patterning of large-area reduced graphene oxide-TiO2 conductive thin film. Chem. Commun. 2010, 46, 3499–3501. [Google Scholar]
  60. Liu, Q.Q.; Shen, J.Y.; Yang, X.F.; Zhang, T.R.; Tang, H. 3D reduced graphene oxide aerogel-mediated Z-scheme photocatalytic system for highly efficient solar-driven water oxidation and removal of antibiotics. Appl. Catal. B Environ. 2018, 232, 562–573. [Google Scholar] [CrossRef]
  61. Bai, X.; Du, Y.Y.; Hu, X.Y.; He, Y.D.; He, C.L.; Liu, E.Z.; Fan, J. Synergy removal of Cr (VI) and organic pollutants over RP-MoS2/rGO photocatalyst. Appl. Catal. B Environ. 2018, 239, 204–213. [Google Scholar] [CrossRef]
  62. Yang, J.; Huang, J.H.; Fan, D.Y.; Chen, S.H. First-principles investigation on the electronic property and bonding configuration of NbC (111)/NbN (111) interface. J. Alloys Compd. 2016, 689, 874–884. [Google Scholar] [CrossRef]
  63. Kulkarni, A.K.; Praveen, C.S.; Sethi, Y.A.; Panman, R.P.; Arbuj, S.S.; Naik, S.D.; Ghule, A.V.; Kale, B.B. Nanostructured N-doped orthorhombic Nb2O5 as an efficient stable photocatalyst for hydrogen generation under visible light. Dalton Trans. 2017, 46, 14859–14868. [Google Scholar] [CrossRef] [PubMed]
  64. Ma, L.F.; Li, F.Y.; Sun, Z.X.; Liu, M.M.; Wang, Y.H.; Xu, L. Synergetic effect of polyoxoniobate and NiS as cocatalysts for enhanced photocatalytic H2 evolution on Cd0.65Zn0.35S. RSC Adv. 2014, 4, 21369–21372. [Google Scholar] [CrossRef]
  65. Li, T.; Wang, K.; Fang, Q.T.; Zhang, Y.; Wang, B.; Li, R.; Lin, Y.Z.; Liu, K.C.; Xie, H.Q.; Li, K. Conductive polymer supported and confined iron phosphide nanocrystals for boosting the photocatalytic hydrogen production of graphitic carbon nitride. J. Mater. Chem. C 2020, 8, 14540–14547. [Google Scholar] [CrossRef]
  66. Zhang, W.Y.; Shen, C.; Lu, G.X.; Ni, Y.R.; Lu, C.H.; Xu, Z.Z. Synthesis of PPy/RGO-based hierarchical material with super-paramagnetic behavior and understanding its robust photo current driven by visible light. Synth. Met. 2018, 241, 17–25. [Google Scholar] [CrossRef]
  67. Yuan, X.J.; Dragoe, D.; Beaunier, P.; Uribe, D.B.; Ramos, L.; Méndez-Medrano, M.G.; Remita, H. Polypyrrole nanostructures modified with monoand bimetallic nanoparticles for photocatalytic H2 generation. J. Mater. Chem. A 2020, 8, 268–277. [Google Scholar] [CrossRef]
  68. Wang, K.L.; Endo-Kimura, M.; Belchi, R.; Zhang, D.; Habert, A.; Bouclé, J.; Ohtani, B.; Kowalska, E.; Herlin-Boime, N. Carbon/Graphene-Modified Titania with Enhanced Photocatalytic Activity under UV and Vis Irradiation. Materials 2019, 12, 4158. [Google Scholar] [CrossRef] [Green Version]
  69. Yuan, X.J.; Floresyona, D.; Aubert, P.H.; Bui, T.T.; Remita, S.; Ghosh, S.; Brisset, F.; Fabrice, G.; Remita, H. Photocatalytic degradation of organic pollutant with polypyrrole nanostructures under UV and visible light. Appl. Catal. B Environ. 2019, 242, 284–292. [Google Scholar] [CrossRef]
Scheme 1. Illustration of the synthesis process of Nb6/PPy-RGO composites (K7HNb6O19·13H2O = Nb6, RGO = reduced graphene oxide, PPy = polypyrrole).
Scheme 1. Illustration of the synthesis process of Nb6/PPy-RGO composites (K7HNb6O19·13H2O = Nb6, RGO = reduced graphene oxide, PPy = polypyrrole).
Nanomaterials 10 02449 sch001
Figure 1. X-ray diffraction (XRD) patterns of graphene oxide (GO), RGO (insert figure), Nb6 and Nb6/PPy-RGO composites.
Figure 1. X-ray diffraction (XRD) patterns of graphene oxide (GO), RGO (insert figure), Nb6 and Nb6/PPy-RGO composites.
Nanomaterials 10 02449 g001
Figure 2. Fourier transform-infrared (FT-IR) spectra of Nb6, GO and Nb6/PPy-RGO-0.25.
Figure 2. Fourier transform-infrared (FT-IR) spectra of Nb6, GO and Nb6/PPy-RGO-0.25.
Nanomaterials 10 02449 g002
Figure 3. Raman spectra of Nb6, GO, RGO and Nb6/PPy-RGO-0.25.
Figure 3. Raman spectra of Nb6, GO, RGO and Nb6/PPy-RGO-0.25.
Nanomaterials 10 02449 g003
Figure 4. X-ray photoelectron spectroscopy (XPS) of GO, Nb6, and Nb6/PPy-RGO-0.25, (A) survey; (B) C 1s; (C) O 1s; (D) Nb 3d; (E) N 1s.
Figure 4. X-ray photoelectron spectroscopy (XPS) of GO, Nb6, and Nb6/PPy-RGO-0.25, (A) survey; (B) C 1s; (C) O 1s; (D) Nb 3d; (E) N 1s.
Nanomaterials 10 02449 g004
Figure 5. Scanning electron microscopy (SEM) images for: (A) Nb6; (B) Nb6/PPy-RGO-0.125; (C) Nb6/PPy-RGO-0.25; (D) Nb6/PPy-RGO-0.375. (E) Transmission electron microscopy (TEM) images and the element maps of Nb6/PPy-RGO-0.25.
Figure 5. Scanning electron microscopy (SEM) images for: (A) Nb6; (B) Nb6/PPy-RGO-0.125; (C) Nb6/PPy-RGO-0.25; (D) Nb6/PPy-RGO-0.375. (E) Transmission electron microscopy (TEM) images and the element maps of Nb6/PPy-RGO-0.25.
Nanomaterials 10 02449 g005
Figure 6. (a) UV-vis diffuses reflectance spectra of Nb6 and Nb6/PPy-RGO composites; (b) Kubelka–Munk function-calculated band gap of Nb6 and Nb6/PPy-RGO composites.
Figure 6. (a) UV-vis diffuses reflectance spectra of Nb6 and Nb6/PPy-RGO composites; (b) Kubelka–Munk function-calculated band gap of Nb6 and Nb6/PPy-RGO composites.
Nanomaterials 10 02449 g006
Figure 7. (a) Transient photocurrent response and (b) electrochemical impedance spectroscopy (EIS) plots of Nb6/PPy-RGO-0.125, Nb6/PPy-RGO-0.25 and Nb6/PPy-RGO-0.375.
Figure 7. (a) Transient photocurrent response and (b) electrochemical impedance spectroscopy (EIS) plots of Nb6/PPy-RGO-0.125, Nb6/PPy-RGO-0.25 and Nb6/PPy-RGO-0.375.
Nanomaterials 10 02449 g007
Figure 8. Hydrogen evolution of Nb6 and Nb6/PPy-RGO composites in aqueous solution with MeOH 20%.
Figure 8. Hydrogen evolution of Nb6 and Nb6/PPy-RGO composites in aqueous solution with MeOH 20%.
Nanomaterials 10 02449 g008
Figure 9. Changes in photocatalytic performance of Nb6/PPy-RGO-0.25 during the recycled uses.
Figure 9. Changes in photocatalytic performance of Nb6/PPy-RGO-0.25 during the recycled uses.
Nanomaterials 10 02449 g009
Figure 10. The proposed photocatalytic hydrogen production mechanism of Nb6/PPy-RGO-0.25 (NHE = normal hydrogen electrode).
Figure 10. The proposed photocatalytic hydrogen production mechanism of Nb6/PPy-RGO-0.25 (NHE = normal hydrogen electrode).
Nanomaterials 10 02449 g010
Table 1. The crystallite sizes of Nb6 and Nb6/PPy-RGO composites (Nb6 = K7HNb6O19, RGO = reduced graphene oxide, PPy = polypyrrole).
Table 1. The crystallite sizes of Nb6 and Nb6/PPy-RGO composites (Nb6 = K7HNb6O19, RGO = reduced graphene oxide, PPy = polypyrrole).
SamplesNb6Nb6/PPy-RGO-0.125Nb6/PPy-RGO-0.25Nb6/PPy-RGO-0.375
Crystalite sizes (nm)2661214
Table 2. The turnover frequency (TOF) values of Nb6 and Nb6/PPy-RGO composites.
Table 2. The turnover frequency (TOF) values of Nb6 and Nb6/PPy-RGO composites.
SamplesNb6Nb6/PPy-RGO-0.125Nb6/PPy-RGO-0.25Nb6/PPy-RGO-0.375
TOF (h−1)137 × 10−3562.6 × 10−3830.4 × 10−3368 × 10−3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Heng, S.; Li, L.; Li, W.; Li, H.; Pang, J.; Zhang, M.; Bai, Y.; Dang, D. Enhanced Photocatalytic Hydrogen Production of the Polyoxoniobate Modified with RGO and PPy. Nanomaterials 2020, 10, 2449. https://doi.org/10.3390/nano10122449

AMA Style

Heng S, Li L, Li W, Li H, Pang J, Zhang M, Bai Y, Dang D. Enhanced Photocatalytic Hydrogen Production of the Polyoxoniobate Modified with RGO and PPy. Nanomaterials. 2020; 10(12):2449. https://doi.org/10.3390/nano10122449

Chicago/Turabian Style

Heng, Shiliang, Lei Li, Weiwei Li, Haiyan Li, Jingyu Pang, Mengzhen Zhang, Yan Bai, and Dongbin Dang. 2020. "Enhanced Photocatalytic Hydrogen Production of the Polyoxoniobate Modified with RGO and PPy" Nanomaterials 10, no. 12: 2449. https://doi.org/10.3390/nano10122449

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop