Next Article in Journal
Insights in the Rhodium-Catalyzed Tandem Isomerization-Hydroformylation of 10-Undecenitrile: Evidence for a Fast Isomerization Regime
Previous Article in Journal
Porous Organic Polymers-Supported Metallocene Catalysts for Ethylene/1-Hexene Copolymerization
Previous Article in Special Issue
Sustainable Carbon Dioxide Photoreduction by a Cooperative Effect of Reactor Design and Titania Metal Promotion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing Light-Driven Production of Hydrogen Peroxide by Anchoring Au onto C3N4 Catalysts

1
State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, China
2
School of Chemical Engineering, Zhengzhou University, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(4), 147; https://doi.org/10.3390/catal8040147
Submission received: 27 February 2018 / Revised: 30 March 2018 / Accepted: 31 March 2018 / Published: 4 April 2018

Abstract

:
Light-driven production of hydrogen peroxide (H2O2) is a green and sustainable way to achieve solar-to-chemical energy conversion. During such a conversion, both the high activity and the stability of catalysts were critical. We prepared an Au-supported C3N4 catalyst—i.e., Au/C3N4-500(N2)—by strongly anchoring Au nanoparticles (~5 nm) onto a C3N4 matrix—which simultaneously enhanced the activity towards the photosynthesis of H2O2 and the stability when it was reused. The yield of H2O2 reached 1320 μmol L−1 on Au/C3N4-500(N2) after 4 h of light irradiation in an acidic solution (pH 3), which was higher than that (1067 μmol L−1) of the control sample Au/C3N4-500(Air) and 2.3 times higher than that of the pristine C3N4. Particularly, the catalyst Au/C3N4-500(N2) retained a much higher stability. The yield of H2O2 had a marginal decrease on the spent catalyst—i.e., 98% yield was kept. In comparison, only 70% yield was obtained from the spent control catalyst. The robust anchoring of Au onto C3N4 improved their interaction, which remarkably decreased the Au leaching when it was used and avoided the aggregation and aging of Au particles. Minimal Au leaching was detected on the spent catalyst. The kinetic analyses indicated that the highest formation rate of H2O2 was achieved on the Au/C3N4-500(N2) catalyst. The decomposition tests and kinetic behaviors of H2O2 were also carried out. These findings suggested that the formation rate of H2O2 could be a determining factor for efficient production of H2O2.

Graphical Abstract

1. Introduction

Hydrogen peroxide (H2O2)—a mild and environment-friendly oxidant—has been used as a green chemical and applied in a variety of industrial and living activities—such as papermaking, water treatment, chemical synthesis, disinfection, and sterilization [1]. In addition, H2O2 is a promising liquid fuel for H2O2-based fuel cells, because only water is emitted [2]. Compared with the widely known fuel H2, H2O2 is easier to store and is more conveniently transported with a low risk of wastage. The current industrial production of H2O2 is based on an anthraquinone method. The advantages of this are a scale-up and high productivity of H2O2. However, huge energy consumption, complicated by-products, and safety issues are still challenges being faced. It is highly desirable to pursue an efficient, mild, and sustained method as an alternative for the synthesis of H2O2.
The light-driven formation of H2O2 has received increased attention in recent decades [3,4,5,6,7,8]. The process involves light capture on the photocatalyst, charge separation under light irradiation, charge transfer to the surface of the photocatalyst, water/alcohol oxidation by photo-generated holes, and the reduction of dioxygen by photo-generated electrons [9,10,11]. In such a process, a highly efficient photocatalyst could be a core factor in order to direct the synthesis of H2O2. Titanium dioxide (TiO2), a widely studied photocatalyst, has been found to enable the formation of H2O2 [12,13,14,15,16]. Several strategies have been developed to enhance the H2O2 yield on TiO2-based photocatalysts. For example, the surface modification of TiO2 by fluorine ions was proposed [13]. The loading of noble metal (Au) and alloys (AuAg) on TiO2 were studied [14,16]. However, the robust hole oxidative capability of TiO2 accelerated the decomposition of H2O2 on the surface of TiO2, and thus lowered the H2O2 yield [17]. Additionally, the ultraviolet (UV) responsive characteristic of TiO2 limited its light absorption and utilization. The graphitic phase carbon nitride (g-C3N4)—a polymeric layered material—was explored as a photocatalyst for pollutant decomposition and water splitting, because of its relatively narrow bandgap (~2.7 eV), and its suitable valence band and conduction band levels [18,19]. Recently, C3N4-based materials have been utilized in the light synthesis of H2O2 [20,21,22,23,24,25]. For instance, Shiraishi et al. reported the selective production of H2O2 on g-C3N4 photocatalyst [20]. The H2O2 production from water and molecular oxygen was improved on a pyromellitic diimide (PDI)-modified C3N4—by the same authors [21]. The modification caused the conduction and valence band levels of C3N4 to positively shift, which was more favorable to the two-electron reduction of dioxygen needed to produce H2O2. The mesoporous g-C3N4 was also prepared by using silica sol as a soft template [22]. The effects of surface defects on photocatalytic H2O2 production were also investigated. The results indicated that the suitably large surface area favored an increased H2O2 yield. Recently, a carbon nitride-aromatic diimide-graphene nanohybrid was synthesized which demonstrated extremely high activity towards H2O2 generation—possessing a solar-to-hydrogen peroxide conversion efficiency of 0.2% [24]. Furthermore, Moon et al. reported a photochemical production of H2O2 over carbon nitride frameworks incorporated with multiple heteroelements (K, P, and O) [25].
In our previous studies, the g-C3N4 exhibited enhanced activity when combined with mixed-metal oxide (MMO) [26]. The reason for this is that the positive shift of the conduction band of C3N4 improved the two-electron reduction of dioxygen to H2O2. Additionally, the MMO plays the role of the water oxidation catalyst (WOC), which has been recognized to be active in oxygen-evolving reactions [27,28,29]. Furthermore, the H2O2 production was increased on the carbon-modified g-C3N4, which regulates the energy levels and is more suitable for the selective two-electron reduction of oxygen by photo-generated electrons [30]. Consequently, the efficient separation of photo-excited electrons and holes, and appropriate band positions are advantageous to the production of H2O2. The noble metals (e.g., Au, Ag, and Pt) revealed the reductive ability in supported catalysts by collecting electrons, similar to a reservoir. The activity is closely related to the size of the noble metal nanoparticles. Generally, when the size of the metal particles is greater than 7 nm, the activity drops sharply [31]. However, the smaller-sized metal particles possess a low Tammann temperature and high surface energy, leading to low thermal stability and the sintering phenomenon when heated [32]. In this article, we investigated Au-supported C3N4 prepared by a stabilized method. The Au/C3N4 catalyst exhibited high activity towards H2O2 production and superior stability when reused. The strong anchoring of Au nanoparticles (around 5 nm) on C3N4 inhibited the leaching and prevented the particles from agglomerating in the sintering process of the catalyst. The kinetics were also studied in order to understand the determining pathway for H2O2 production.

2. Results

2.1. Catalyst Characterization

The Au loading and distribution on the C3N4 matrix were observed by transmission electron microscopy TEM (Figure 1). The average size of Au particles was 5.0 nm (Figure 1a,b) and 7.5 nm (Figure 1d,e) in the Au/C3N4-500(N2) and Au/C3N4-500(Air) catalysts, respectively. The Au particles had a narrower size distribution on the Au/C3N4-500(N2) than on the Au/C3N4-500(Air). The high-resolution transmission electron microscopy photographs showed that the distance of the neighboring planes was 0.235 nm in both catalysts, corresponding to the d-spacing of the plane (111) in Au [32]. Au particles had more crystal defects on the Au/C3N4-500(N2) (Figure 1c) than on the Au/C3N4-500(Air) (Figure 1f). The smaller-sized Au possessed more edges, corners, and steps, which were favorable for the adsorption of reactive species and further catalytic reactions. The energy dispersive spectrometer EDS mapping of the Au/C3N4-500(N2) is presented in Figure 1g. The results suggested that Au, C, and N elements had a homogeneous distribution in the catalyst.
The phase structure of the catalysts was determined by X-ray diffractometer XRD (Figure 2). The diffractions at 2θ = 13.3° and 27.4° were indexed to the planes—namely, 100 and 002—of g-C3N4 (JCPDS 87-1526) in all samples, which originated from the in-plane repetition of tri-s-triazine and the stacking of the conjugated aromatic units, respectively [33]. The formation of Au was verified in the XRD patterns of the supported catalysts, in which the diffractions were indexed to the planes—namely, 111, 200, and 220 (JCPDS 04-0784) [34]. The loading of Au was estimated by Thermogravimetric analysis (TGA). The weight of Au was 1.05% and 1.35% in Au/C3N4-500(N2) and Au/C3N4-500(Air), respectively (Figure S1). This indicated that the loading of Au in the two catalysts had minimal difference.
The elemental valences and surface species were analyzed by X-ray photoelectron spectroscopy XPS (Figure 3), because of its surface-sensitive technique with a detection depth of less than ten nanometers. Two deconvoluted peaks, at 83.7 and 87.4 eV, were observed in the core level spectra of Au4f in Au/C3N4-500(N2), which were assigned to zerovalent Au [35,36]. The peaks showed a slight shift (0.2 eV) towards decreased binding energy in that of Au/C3N4-500(Air) (Table 1). In both catalysts, the N1s core level spectra could be deconvoluted into three peaks. In the g-C3N4 phase, the peaks at 400.6, 399.2, and 398.5 eV corresponded to N-(C)3, =N-, and C2NH species, respectively [37,38]. The binding energy of the three peaks had a slight shift (<0.2 eV) between the Au/C3N4-500(N2) and Au/C3N4-500(Air), which suggested that the nitrogen species had minimal change in both catalysts.
The optical absorption properties of the supported catalysts were characterized by UV-visible diffuse reflectance spectra (Figure 4a). These spectra were obtained by converting the reflection data that was measured using the Kubelka–Munk equation: F(R) = (1 − R)2/2R, where R is the reflectance [39]. The three Au/C3N4 samples showed enhanced absorption in both the UV and visible regions when compared with the bare C3N4. The absorption band below 400 nm in the UV region corresponded to the π–π* electron transition of the aromatic 1-, 3-, and 5-triazine compounds [40,41,42]. The localized surface plasmon resonance (LSPR) absorption of Au was apparent, peaking at ~550 nm in the Au/C3N4-250 and Au/C3N4-500(Air) samples. The Au/C3N4-500(N2) sample presented more intensive absorption throughout the spectral region measured. This suggested that the metal–support interaction differed between the Au/C3N4-500(N2) and Au/C3N4-500(Air), which could have affected the photocatalytic properties. The fluorescence spectra were measured because the fluorescence emission properties were associated with the carrier recombination (Figure 4b). The C3N4 itself exhibited an emission band that peaked at 462 nm, which was attributed to the transition between the lone pair valence band and the π* conduction band in the framework of graphitic C3N4 [43]. After the loading of Au nanoparticles, the intensity of the emission band decreased significantly in the Au/C3N4-250 and Au/C3N4-500(Air). Notably, the emission in the Au/C3N4-500(N2) was almost quenched. The quenching of fluorescence emission indicated the suppression of carrier recombination under photo-excitation. The efficient blockage of the recombination channel could have favored the carrier separation and thus stimulated the electrons and holes to involve the photocatalytic reactions on the surface of catalysts. The reduction of emission intensity was more significant in the Au/C3N4-500(N2) than in the Au/C3N4-500(Air). It could have indicated that the metal–support interaction was more intensive in the former, which was beneficial for the feasible electron transfer of photo-excited C3N4 to Au.

2.2. Photosynthesis of H2O2

The photocatalytic generation of H2O2 on the catalysts was conducted in an O2-equilibrated acidic aqueous solution (pH 3) under light irradiation. The H2O2 yield is shown (Figure 5a) as a function of time. The yield increased with time on all catalysts. The higher yield was observed on the Au/C3N4-500(N2) from the beginning of the reactions. After 240 min of irradiation, the H2O2 yield reached 1320 and 1067 μmol L−1 on the Au/C3N4-500(N2) and Au/C3N4-500(Air), respectively—both of which were higher than that on the C3N4 (582 μmol L−1) (Table 2). The yield was higher for the Au/C3N4-500(N2) than for the Au/C3N4-500(Air), although the latter had a larger Au loading—1.05% wt in Au/C3N4-500(N2) and 1.35% wt in Au/C3N4-500(Air) catalyst. We concluded that the catalytic activity of Au/C3N4-500(N2) was higher than that of Au/C3N4-500(Air). This was potentially ascribed to the smaller size of Au nanoparticles in the Au/C3N4-500(N2) (Figure 1). Similar findings were also observed in the Au- and AuAg-loaded TiO2 photocatalytsts [14,16]. The apparent quantum yield (ΦAQY) for H2O2 formation over the Au/C3N4-500(N2) was calculated using the equation ΦAQY (%) = (2 × [H2O2] formed)/(photon number entered into the reaction vessel) × 100. The ΦAQY value measured at a monochromic light of 400 nm was 3.63% (see Supplementary Materials for details).
In order to understand the effect of the light wavelength irradiated, we measured the H2O2 generation under light irradiation with a cut-off filter of 420 nm (Figure S2). After 4 h of irradiation, the H2O2 yield reached 694 μmol L−1 under the cut-off filter of 420 nm, lower than that measured without the cut-off filter on the Au/C3N4-500(N2) (1320 μmol L−1). This suggested that the irradiation of UV light contributed to the formation of H2O2 when no cut-off filter was used. This indicated that the UV irradiation possessed high energy and could stimulate the C3N4 (with a bandgap of ~2.8 eV) to generate photo-carriers. The Au nanoparticles showed the LSPR absorption band at around 550 nm. As a result, the Au/C3N4-500(N2) had intensive absorption above 500 nm, while the C3N4 showed minimal absorption in the same spectral region (>500 nm) (Figure 4a). To understand the effect of LSPR on the production of H2O2, we used a cut-off filter of 510 nm. The H2O2 yield was 10 times higher for the Au/C3N4-500(N2) (293 μmol L−1) than for the C3N4 (25 μmol L−1) after 4 h of irradiation (Figure S3). The LSPR of Au enhanced light absorption around the plasmon resonance band and contributed considerably to H2O2 production. To compare the effect of Au loading on the activity, we prepared another two samples of Au/C3N4-500(N2), with an Au loading of 1% and 3% (a nominal weight). The H2O2 production was carried out on the two catalysts under the same reaction conditions. The H2O2 yield after 4 h of irradiation was 743, 1320, and 917 μmol L−1, on 1%, 2%, and 3% Au-supported catalysts, respectively (Figure S4). This finding indicated that an appropriate Au loading led to a higher H2O2 yield. Higher Au loading might have caused larger-sized Au particles and the reduction in activity [44,45]. The Au/C3N4-500(N2) catalyst was prepared by a carbon-layer-stabilized method, where the dopamine molecule was used as a precursor of the carbon layer. In order to show the effect of dopamine, we prepared a control sample without a treatment of dopamine and named it Au/C3N4-500(N2)-without dopa. The H2O2 yield after 4 h of irradiation was 1320 and 1065 μmol L−1 on the Au/C3N4-500(N2) and Au/C3N4-500(N2)-without dopa, respectively (Figure S5). This finding verified the effect of dopamine on the activity of the catalyst.
The H2O2 concentration–time data (Figure 5a) was fitted using the equation [H2O2] = (Kf/Kd) {1-exp(-Kdt)}, where Kf and Kd were the formation and decomposition rate constants of H2O2, respectively, and t was the irradiation time [4,26]. The solid lines (Figure 5a) were the fitted curves. Based on the fitted results (Figure S6), the Kf and Kd values were listed (Table 2) and plotted (Figure 5b). Both Kf and Kd values were the smallest on the C3N4. The Kf value was the largest on the Au/C3N4-500(N2) (7.893 μmol L−1 min−1) and the Kd value was smaller on the Au/C3N4-500(N2) than on the Au/C3N4-500(Air) (0.00354 vs. 0.00446 min−1). This indicated that a larger Kf value and smaller Kd value led to a higher yield of H2O2. The formation rate of H2O2 was dominant in the process of H2O2 production. This was consistent with our previous studies on the metal oxide@C3N4 and carbon-doped C3N4 catalysts [26,30].
The decomposition behaviors of H2O2 on the catalysts were further investigated under an excessive amount of H2O2 (initial concentration of H2O2 of 5 mM) (Figure 6). The decomposition data was fitted using a pseudo-first-order equation: −dC(t)/dt = kt, where C(t) is the H2O2 concentration at certain a time, k is the decomposition rate constant, and t is the time. The fitted results showed that the decomposition rate constant followed the order of Au/C3N4-500(N2) (0.00406 min−1) > Au/C3N4-500(Air) (0.00251 min−1) > Au/C3N4-250 (0.00181 min−1) > C3N4 (0.00097 min−1) (Figure S7). The k values had the same order of magnitude as those in the production of H2O2 (Table 2). The decomposition of H2O2 on the Au-loaded C3N4 catalysts was less dependent on the concentration of H2O2. The sustained production of H2O2 was mainly dominated by its formation rate.

2.3. Stability of the Photocatalysts

The stable production of H2O2 was required on the spent catalysts. After being used for the photosynthesis of H2O2, the catalysts were separated from the solution and recycled for reuse. We studied the repeated use of the Au/C3N4 catalysts to confirm their stability. The H2O2 yield was almost the same on the spent Au/C3N4-500(N2) as on the original Au/C3N4-500(N2), having a negligible decrease of less than 2% (Figure 7). However, the H2O2 yield showed a significant reduction on the reused Au/C3N4-500(Air) and only 70% of the original yield was retained. This suggested that Au/C3N4-500(N2) had a much higher stability than Au/C3N4-500(Air) for the photosynthesis of H2O2. This was verified by the structural characterizations of the spent catalysts.
The spent catalysts were characterized by TEM, inductively-coupled plasma spectrum ICP, and XPS. The average size of Au particles increased from 5 to 7 nm in the Au/C3N4-500(N2) after reactions (Figure 8a,b). The Au particles had less agglomeration and were still highly dispersed in the C3N4 matrix. In contrast, the Au particles showed an increase in size from 7.5 to 10 nm in the Au/C3N4-500(Air) (Figure 8c,d). The Au particles exhibited obvious aggregates. In addition, the elemental analyses showed that the Au loading had a slight loss of 5.7% in the spent Au/C3N4-500(N2), while the Au remarkably lost 17.8% in the spent Au/C3N4-500(Air) (Figure S1). Such a significant leaching of Au could be a critical factor for the sharp decrease of H2O2 yield, as well as the increased size of Au particles in Au/C3N4-500(Air) (Figure 7). The XPS results indicated that the valence of Au had no change and Au still existed in the form of zero-valence. The fitted peaks in the N1s spectra had minimal shift compared to the original catalysts (Figure S8).
The flatband potentials (Efb) of the catalysts were estimated from the electrochemical Mott–Schottky plots, measured at varied frequencies. The Mott–Schottky plots exhibited positive slopes, characteristic of n-type semiconductors (Figure 9) [26]. The Efb values were derived by extrapolating the plots to the x-axis, which were −0.86 V, −0.68 V, and −0.70 V (vs. RHE) for C3N4, Au/C3N4-500(N2), and Au/C3N4-500(Air), respectively. The Efb values of Au-supported C3N4 catalysts had a minimal positive shift (~0.16 V) compared with that of C3N4, and had minimal difference between Au/C3N4-500(N2) and Au/C3N4-500(Air). Generally, the conduction band potential was approximately equal to the Efb [30]. Consequently, the conduction band levels hardly changed in the Au-supported catalysts, regardless of the stability. The bandgap of the catalysts was also determined from the Tauc plots. The Tauc plots of Au/C3N4-500(N2) and Au/C3N4-500(Air) were derived according to the equation (F(R)·hυ)2 = A(hυ-Eg), where h was Plank constant, υ was the frequency, A was a constant, and Eg was the band gap (Figure S9) [39,46,47]. The estimated bandgap was 2.83 eV and 2.85 eV for the Au/C3N4-500(N2) and Au/C3N4-500(Air), respectively. Additionally, the bandgap values had no change in the spent catalysts. From the energy level analyses, we concluded that the different activity and stability of the Au-supported C3N4 catalysts were mainly affected by the Au sizes and metal–support interactions rather than the energy band structure.

3. Materials and Methods

3.1. Materials

Melamine, isopropanol, and HClO4 were purchased from Sinopharm Chemical Reagent (Beijing Co. Ltd., Beijing, China). HAuCl4·3H2O was purchased from Shanghai Aladdin Biochemical Technology Co. Ltd. (Beijing, China) KH2PO4, K2HPO4·3H2O, H2SO4 (98%), and H2O2 (30%) were purchased from Beijing Chemical Reagent Co. Ltd. (Beijing, China).Peroxidase was purchased from Beijing Lamboride Trading Company (Beijing, China). N-diethyl-1,4-phenylenediamine sulfate (99%) was purchased from Adamas Reagent Co. Ltd. (Beijing, China) O2 (99.995%) was purchased from Hai Pu Gas Co. Ltd. (Beijing, China) All reagents were of analytical grade and were used without further purification. Deionized water was used throughout the experiments.

3.2. Catalyst Preparation

3.2.1. Preparation of g-C3N4

The g-C3N4 was prepared by heating of a melamine precursor [33]. The melamine (10 g) was added to a porcelain cup and calcined for 4 h at 520 °C (heating rate of 4 °C min−1). The product was a yellow powder.

3.2.2. Synthesis of Au-Supported C3N4 Catalysts

Au/C3N4-250 Catalyst

The Au/C3N4-250 catalyst was synthesized by a method similar to the literature [32]. Briefly, the pH value of an aqueous solution (0.02 g HAuCl4·3H2O in 60 mL water) was adjusted to 10.0 by adding 1 M NaOH. 1 g of C3N4 powder was added to the solution and heated to 70 °C for 2 h under stirring. The suspension was washed three times using deionized water by centrifugation. Finally, the product was dried for 2 h at 40 °C and subsequently calcined in air for 2 h at 250 °C. Before further use, the product was sealed in vacuum desiccator at room temperature.

Au/C3N4-500(N2) Catalyst

Firstly, the Au/C3N4-250 sample was modified with dopamine according to the previous method [48]. The 0.5 g of Au/C3N4-250 powder was added to 1 mg mL−1 dopamine-containing solution, which was pre-adjusted to pH 8.5 by a tris-buffer solution (10 mM). The suspension was stirred for 24 h, allowing for the formation of polydopamine (PDA) on the surface of solid. Subsequently, the suspension was separated by centrifugation, washed three times using deionized water and once using ethanol. The product was dried for 2 h at 40 °C. Subsequently, the dried product was annealed in a gas flow of nitrogen for 2 h (heating rate of 5 °C min−1) at 500 °C. The resulting powder was further annealed in air for 2 h at 500 °C to obtain the Au/C3N4-500(N2) catalyst.

Au/C3N4-500(Air) Catalyst

The 0.5 g of Au/C3N4-250 powder was directly annealed in air for 2 h at 500 °C to obtain the Au/C3N4 -500(Air) catalyst.

4. Conclusions

The Au nanoparticles were strongly anchored onto the C3N4 matrix by a carbon-layer-stabilized method. The as-prepared Au/C3N4-500(N2) catalyst showed higher activity towards light-driven synthesis of H2O2 than the control Au/C3N4-500(Air). This was ascribed to the smaller size of Au nanoparticles (~5 nm) and the stronger metal–support interaction on the Au/C3N4-500(N2). The H2O2 yield reached 1320 μmol L−1 after 240 min of light irradiation in an acidic solution (pH 3), 2.3 times higher than that of the pristine C3N4. The spent Au/C3N4-500(N2) catalyst exhibited a marginal loss of activity and the H2O2 yield remained >98% of the original yield. However, the H2O2 yield decreased to 70% of the original yield on the Au/C3N4-500(Air) catalyst when reused. There were primarily two reasons for this. One was the significant leaching of Au (17.8%) from the Au/C3N4-500(Air) catalyst after reactions, much greater than that (5.7%) from the Au/C3N4-500(N2) catalyst. The other was the increased sizes of Au (~10 nm) on Au/C3N4-500(Air), which led to the decrease in activity. The energy band structures of the Au/C3N4-500(N2) and the Au/C3N4-500(Air) showed minimal difference before or after the reactions. Therefore, the highly stable Au-supported C3N4 catalyst could be favorable for the sustained production of H2O2 via a two-electron reduction of dioxygen driven by solar light.

Supplementary Materials

The following characterization and photocatalytic measurements of H2O2 are available online at https://www.mdpi.com/2073-4344/8/4/147/s1: Figure S1, thermogravimetric analysis; Figure S2, the H2O2 generation using a 420 nm cut-off filter; Figure S3, the H2O2 generation using a 510 nm cut-off filter; Figure S4, the H2O2 generation on Au-supported catalysts; Figure S5, the H2O2 generation on Au/C3N4-500(N2)-without dopa; Figure S6, kinetic fitting curves of H2O2 generation; Figure S7, the fitting curves of H2O2 decomposition; Figure S8, XPS core level spectra of Au4f and N1s; and Figure S9, Tauc plots.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant 21576016, 21521005) and the National Key R&D Program of China (Grant 2017YFA0206804).

Author Contributions

Xu Xiang conceived and designed the experiments; Xiaoyu Chang and Dandan Han performed the experiments; Xu Xiang, Xiaoyu Chang, Jing He, Jun Jiao Yang, and Bing Zhang analyzed the data; Xu Xiang and Xiaoyu Chang wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Campos-Martin, J.M.; Blanco-Brieva, G.; Fierro, J.L.G. Hydrogen peroxide synthesis: An outlook beyond the anthraquinone process. Angew. Chem. Int. Ed. 2006, 45, 6962–6984. [Google Scholar] [CrossRef] [PubMed]
  2. Kato, S.; Jung, J.; Suenobua, T.; Fukuzumi, S. Production of hydrogen peroxide as a sustainable solar fuel from water and dioxygen. Energy Environ. Sci. 2013, 6, 3756–3764. [Google Scholar] [CrossRef]
  3. Isaka, Y.; Kato, S.; Hong, D.; Suenobu, T.; Yamada, Y.; Fukuzumi, S. Bottom-up and top-down methods to improve catalytic reactivity for photocatalytic production of hydrogen peroxide using a Ru-complex and water oxidation catalysts. J. Mater. Chem. A 2015, 3, 12404–12412. [Google Scholar] [CrossRef]
  4. Isaka, Y.; Oyama, K.; Yamada, Y.; Suenobu, T.; Fukuzumi, S. Photocatalytic production of hydrogen peroxide from water and dioxygen using cyano-bridged polynuclear transition metal complexes as water oxidation catalysts. Catal. Sci. Technol. 2016, 6, 681–684. [Google Scholar] [CrossRef]
  5. Thakur, S.; Kshetri, T.; Kim, N.H.; Lee, J.H. Sunlight-driven sustainable production of hydrogen peroxide using a CdS–graphene hybrid photocatalyst. J. Catal. 2017, 345, 78–86. [Google Scholar] [CrossRef]
  6. Fukushima, M.; Tatsumi, K.; Tanaka, S.; Nakamura, H. Photocatalytic production of hydrogen peroxide by tris(2,2′-bipyridine) ruthenium(II) using humic acids as electron donor. Environ. Sci. Technol. 1998, 32, 3948–3953. [Google Scholar] [CrossRef]
  7. Zhuang, H.; Yang, L.; Xu, J.; Li, F.; Zhang, Z.; Lin, H.; Long, J.; Wang, X. Robust photocatalytic H2O2 production by octahedral Cd3(C3N3S3)2 coordination polymer under visible light. Sci. Rep. 2015, 5, 16947. [Google Scholar] [CrossRef] [PubMed]
  8. Zhao, S.; Zhao, X.; Zhang, H.; Li, J.; Zhu, Y. Covalent combination of polyoxometalate and graphitic carbon nitride for light-driven hydrogen peroxide production. Nano Energy 2017, 35, 405–414. [Google Scholar] [CrossRef]
  9. Yamada, Y.; Nomura, A.; Miyahigashi, T.; Ohkubo, K.; Fukuzumi, S. Acetate induced enhancement of photocatalytic hydrogen peroxide production from qxalic acid and dioxygen. J. Phys. Chem. A 2013, 117, 3751–3760. [Google Scholar] [CrossRef] [PubMed]
  10. Aratani, Y.; Suenobu, T.; Ohkubo, K.; Yamada, Y.; Fukuzumi, S. Dual function photocatalysis of cyano-bridged heteronuclear metal complexes for water oxidation and two-electron reduction of dioxygen to produce hydrogen peroxide as a solar fuel. Chem. Commun. 2017, 53, 3473–3476. [Google Scholar] [CrossRef] [PubMed]
  11. Isaka, Y.; Yamada, Y.; Suenobu, T.; Nakagawa, T.; Fukuzumi, S. Production of hydrogen peroxide by combination of semiconductor-photocatalysed oxidation of water and photocatalytic two-electron reduction of dioxygen. RSC Adv. 2016, 6, 42041–42044. [Google Scholar] [CrossRef]
  12. Cai, R.; Kubota, Y.; Fujishima, A. Effect of copper ions on the formation of hydrogen peroxide from photocatalytic titanium dioxide particles. J. Catal. 2003, 219, 214–218. [Google Scholar] [CrossRef]
  13. Maurino, V.; Minero, C.; Mariella, G.; Pelizzetti, E. Sustained production of H2O2 on irradiated TiO2-fluoride systems. Chem. Commun. 2005, 20, 2627–2629. [Google Scholar] [CrossRef] [PubMed]
  14. Tsukamoto, D.; Shiro, A.; Shiraishi, Y.; Shiraishi, Y.; Sugano, Y.; Ichikawa, S.; Tanaka, S.; Hirai, T. Photocatalytic H2O2 production from ethanol/O2 system using TiO2 loaded with Au-Ag bimetallic alloy nanoparticles. ACS Catal. 2012, 2, 599–603. [Google Scholar] [CrossRef]
  15. Shiraishi, Y.; Kanazawa, S.; Tsukamoto, D.; Shiro, A.; Sugano, Y.; Hirai, T. Selective hydrogen peroxide formation by titanium dioxide photocatalysis with benzylic alcohols and molecular oxygen in water. ACS Catal. 2013, 3, 2222–2227. [Google Scholar] [CrossRef]
  16. Teranishi, M.; Naya, S.; Tada, H. In situ liquid phase synthesis of hydrogen peroxide from molecular oxygen using gold nanoparticle-loaded titanium(IV) dioxide photocatalyst. J. Am. Chem. Soc. 2010, 132, 7850–7851. [Google Scholar] [CrossRef] [PubMed]
  17. Li, X.; Chen, C.; Zhao, J. Mechanism of photodecomposition of H2O2 on TiO2 surfaces under visible light irradiation. Langmuir 2001, 17, 4118–4122. [Google Scholar] [CrossRef]
  18. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  19. Cai, Q.; Shen, J.; Feng, Y.; Shen, Q.; Yang, H. Template-free preparation and characterization of nanoporous g-C3N4 with enhanced visible photocatalytic activity. J. Alloys Compd. 2015, 628, 372–378. [Google Scholar] [CrossRef]
  20. Shiraishi, Y.; Kanazawa, S.; Sugano, Y.; Tsukamoto, D.; Sakamoto, H.; Ichikawa, S.; Hirai, T. Highly selective production of hydrogen peroxide on graphitic carbon nitride (g-C3N4) photocatalyst activated by visible light. ACS Catal. 2014, 4, 774–780. [Google Scholar] [CrossRef]
  21. Shiraishi, Y.; Kanazawa, S.; Kofuji, Y.; Sakamoto, H.; Ichikawa, S.; Tanaka, S.; Hirai, T. Sunlight-driven hydrogen peroxide production from water and molecular oxygen by metal-free photocatalysts. Angew. Chem. Int. Ed. 2014, 53, 13454–13459. [Google Scholar] [CrossRef] [PubMed]
  22. Shiraishi, Y.; Kofuji, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Effects of surface defects on photocatalytic H2O2 production by mesoporous graphitic carbon nitride under visible light irradiation. ACS Catal. 2015, 5, 3058–3066. [Google Scholar] [CrossRef]
  23. Kofuji, Y.; Ohkita, S.; Shiraishi, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Graphitic carbon nitride doped with biphenyl diimide: Efficient photocatalyst for hydrogen peroxide production from water and molecular oxygen by sunlight. ACS Catal. 2016, 6, 7021–7029. [Google Scholar] [CrossRef]
  24. Kofuji, Y.; Isobe, Y.; Shiraishi, Y.; Sakamoto, H.; Tanaka, S.; Ichikawa, S.; Hirai, T. Carbon nitride-aromatic diimide-graphene nanohybrids: MetalFree photocatalysts for solar-to-hydrogen peroxide energy conversion with 0.2% efficiency. J. Am. Chem. Soc. 2016, 138, 10019–10025. [Google Scholar] [CrossRef] [PubMed]
  25. Moon, G.-H.; Fujitsuka, M.; Kim, S.; Majima, T.; Wang, X.; Choi, W. Eco-Friendly photochemical production of H2O2 through O2 reduction over carbon nitride frameworks incorporated with multiple heteroelements. ACS Catal. 2017, 7, 2886–2895. [Google Scholar] [CrossRef]
  26. Wang, R.; Pan, K.; Han, D.; Jiang, J.; Xiang, C.; Huang, Z.; Zhang, L.; Xiang, X. Solar-driven H2O2 generation from H2O and O2 using earth-abundant mixed-metal oxide@carbon nitride photocatalysts. ChemSusChem 2016, 9, 2470–2479. [Google Scholar] [CrossRef] [PubMed]
  27. Li, Y.; Zhang, L.; Xiang, X.; Yan, D.; Li, F. Engineering of ZnCo-layered double hydroxide nanowalls toward high-efficiency electrochemical water oxidation. J. Mater. Chem. A 2014, 2, 13250–13258. [Google Scholar] [CrossRef]
  28. Tang, Y.; Wang, R.; Yang, Y.; Yan, D.; Xiang, X. Highly enhanced photoelectrochemical water oxidation efficiency based on triadic quantum dot/layered double hydroxide/BiVO4 photoanodes. ACS Appl. Mater. Interfaces 2016, 8, 19446–19455. [Google Scholar] [CrossRef] [PubMed]
  29. Han, D.; Liu, J.; Yang, J.; Xiang, X. In-situ conversion and catalytic properties of mixed-metal oxide catalysts for photosynthesis of hydrogen peroxide. Sci. China-Chem. 2017, 47, 465–473. [Google Scholar] [CrossRef]
  30. Wang, R.; Zhang, X.; Li, F.; Cao, D.; Pu, M.; Han, D.; Yang, J.; Xiang, X. Energy-level dependent H2O2 production on metal-free, carbon-content tunable carbon nitride photocatalysts. J. Energy Chem. 2018, 27, 343–350. [Google Scholar] [CrossRef]
  31. Tan, Y.; Liu, X.Y.; Zhang, L.; Wang, A.; Li, L.; Pan, X.; Miao, S.; Haruta, M.; Wei, H.; Wang, H.; et al. ZnAl-hydrotalcite-supported Au25 nanoclusters as precatalysts for chemoselective hydrogenation of 3-nitrostyrene. Angew. Chem. Int. Ed. 2017, 56, 2709–2713. [Google Scholar] [CrossRef] [PubMed]
  32. Zhan, W.; He, Q.; Liu, X.; Guo, Y.; Wang, Y.; Wang, L.; Guo, Y.; Borisevich, A.Y.; Zhang, J.; Lu, G.; et al. A sacrificial coating strategy toward enhancement of metal-support interaction for ultrastable Au nanocatalysts. J. Am. Chem. Soc. 2016, 138, 16130–16139. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, J.; Shen, S.; Guo, P.; Wu, P.; Guo, L. Spatial engineering of photo-active sites on g-C3N4 for efficient solar hydrogen generation. J. Mater. Chem. A 2014, 2, 4605–4612. [Google Scholar] [CrossRef]
  34. Sun, X.; Dong, S.; Wang, E. Large-Scale synthesis of micrometer-scale single-crystalline Au plates of nanometer thickness by a wet-chemical route. Angew. Chem. Int. Ed. 2004, 116, 6520–6523. [Google Scholar] [CrossRef]
  35. Leff, D.V.; Brandt, L.; Heath, J.R. Synthesis and characterization of hydrophobic, organically-soluble gold nanocrystals functionalized with primary amines. Langmuir 1996, 12, 4723–4730. [Google Scholar] [CrossRef]
  36. Park, M.; Shin, K.S.; Lee, J.W.; Kim, K. Novel fabrication of Au nanoparticle film on a polyelectrolyte-coated glass for efficient surface-enhanced raman scattering. Bull. Korean Chem. Soc. 2015, 36, 743–747. [Google Scholar]
  37. Singh, J.A.; Overbury, S.H.; Dudney, N.J.; Li, M.; Veith, G.M. Gold nanoparticles supported on carbon nitride: Influence of surface hydroxyls on low temperature carbon monoxide oxidation. ACS Catal. 2012, 2, 1138–1146. [Google Scholar] [CrossRef]
  38. Kartusch, C.; Krumeich, F.; Safonova, O.; Hartfelder, U.; Makosch, M.; Sa, J.; Bokhoven, J.A. Redispersion of gold multiple-twinned particles during liquid-phase hydrogenation. ACS Catal. 2012, 2, 1394–1403. [Google Scholar] [CrossRef]
  39. Gao, L.F.; Wen, T.; Xu, J.Y.; Zhai, X.P.; Zhao, M.; Hu, G.W.; Chen, P.; Wang, Q.; Zhang, H.L. Iron-doped carbon nitride-type polymers as homogeneous organocatalysts for visible light-driven hydrogen evolution. ACS Appl. Mater. Interfaces 2016, 8, 617–624. [Google Scholar] [CrossRef] [PubMed]
  40. Khabashesku, V.N.; Zimmerman, J.L.; Margrave, J.L. Powder synthesis and characterization of amorphous carbon nitride. Chem. Mater. 2000, 12, 3264–3270. [Google Scholar] [CrossRef]
  41. Guo, Q.; Xie, Y.; Wang, X.; Lv, S.; Hou, T.; Liu, X. Characterization of well crystallized graphitic carbon nitride nanocrystallites via a benzene thermal route at low temperatures. Chem. Phys. Lett. 2003, 380, 84–87. [Google Scholar] [CrossRef]
  42. Guo, Q.; Yang, Q.; Zhu, L.; Yi, C.; Zhang, S.; Xie, Y. A facile one pot solvothermal route to tubular forms of luminescent polymeric networks [(C3N3)2(NH)3]n. Solid State Commun. 2004, 132, 369–374. [Google Scholar] [CrossRef]
  43. Wang, S.; Li, C.; Wang, T.; Zhang, P.; Li, A.; Gong, J. Controllable synthesis of nanotube-type graphitic C3N4 and their visible-light photocatalytic and fluorescent properties. J. Mater. Chem. A 2014, 2, 2885–2890. [Google Scholar] [CrossRef]
  44. Zhan, G.; Huang, J.; Du, M.; Sun, D.; Abdul-Rauf, I.; Lin, W.; Hong, Y.; Li, Q. Liquid phase oxidation of benzyl alcohol to benzaldehyde with novel uncalcined bioreduction Au catalysts: High activity and durability. Chem. Eng. J. 2012, 187, 232–238. [Google Scholar] [CrossRef]
  45. Xu, Z.; Quintanilla, M.; Vetrone, F.; Govorov, A.O.; Chaker, M.; Ma, D. Harvesting lost photons: Plasmon and upconversion enhanced broadband photocatalytic activity in core@shell microspheres based on lanthanide-doped NaYF4, TiO2, and Au. Adv. Funct. Mater. 2015, 25, 2950–2960. [Google Scholar] [CrossRef]
  46. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.-T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [Google Scholar] [CrossRef] [PubMed]
  47. Moon, G.-H.; Kim, W.; Bokare, A.-D.; Sung, N.-H.; Choi, W. Solar production of H2O2 on reduced graphene oxide—TiO2 hybrid photocatalysts consisting of earth-abundant elements only. Energy Environ. Sci. 2014, 7, 4023–4028. [Google Scholar] [CrossRef]
  48. Chao, C.; Liu, J.; Wang, J.; Zhang, Y.; Zhang, B.; Zhang, Y.; Xiang, X.; Chen, R. Surface Modification of Halloysite Nanotubes with Dopamine for Enzyme Immobilization. ACS Appl. Mater. Interfaces 2013, 5, 10559–10564. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) TEM image; (b) particle size distribution and (c) HRTEM photograph of Au/C3N4-500(N2); (d) TEM image; (e) particle size distribution and (f) HRTEM image of Au/C3N4-500(Air); (g) EDS mapping of Au/C3N4-500(N2) (scale bar 100 nm). The elemental mappings show the distribution of Au, C, and N elements.
Figure 1. (a) TEM image; (b) particle size distribution and (c) HRTEM photograph of Au/C3N4-500(N2); (d) TEM image; (e) particle size distribution and (f) HRTEM image of Au/C3N4-500(Air); (g) EDS mapping of Au/C3N4-500(N2) (scale bar 100 nm). The elemental mappings show the distribution of Au, C, and N elements.
Catalysts 08 00147 g001
Figure 2. XRD patterns of C3N4, Au/C3N4-250, Au/C3N4-500(N2), and Au/C3N4-500(Air).
Figure 2. XRD patterns of C3N4, Au/C3N4-250, Au/C3N4-500(N2), and Au/C3N4-500(Air).
Catalysts 08 00147 g002
Figure 3. XPS core level spectra of (a) Au4f in Au/C3N4-500(N2); (b) N1s in Au/C3N4-500(N2) and (c) Au4f in Au/C3N4-500(Air); (d) N1s in Au/C3N4-500(Air).
Figure 3. XPS core level spectra of (a) Au4f in Au/C3N4-500(N2); (b) N1s in Au/C3N4-500(N2) and (c) Au4f in Au/C3N4-500(Air); (d) N1s in Au/C3N4-500(Air).
Catalysts 08 00147 g003
Figure 4. (a) Ultraviolet (UV)/Visible diffuse reflectance spectra; (b) fluorescence emission spectra of C3N4, Au/C3N4-250, Au/C3N4-500(Air), and Au/C3N4-500(N2); across an excitation wavelength of 400 nm.
Figure 4. (a) Ultraviolet (UV)/Visible diffuse reflectance spectra; (b) fluorescence emission spectra of C3N4, Au/C3N4-250, Au/C3N4-500(Air), and Au/C3N4-500(N2); across an excitation wavelength of 400 nm.
Catalysts 08 00147 g004
Figure 5. (a) The light-driven hydrogen peroxide (H2O2) generation in an O2-equilibrated aqueous solution over C3N4, Au/C3N4-250, Au/C3N4-500(Air), and Au/C3N4-500(N2). The experimental conditions: 1 g/L of photocatalyst, pH 3.0, 5% vol of 2-propanol, Xe lamp, illumination intensity of 100 mW/cm2; and (b) formation rate constant (Kf) and decomposition rate constant (Kd) for H2O2 production.
Figure 5. (a) The light-driven hydrogen peroxide (H2O2) generation in an O2-equilibrated aqueous solution over C3N4, Au/C3N4-250, Au/C3N4-500(Air), and Au/C3N4-500(N2). The experimental conditions: 1 g/L of photocatalyst, pH 3.0, 5% vol of 2-propanol, Xe lamp, illumination intensity of 100 mW/cm2; and (b) formation rate constant (Kf) and decomposition rate constant (Kd) for H2O2 production.
Catalysts 08 00147 g005
Figure 6. The photocatalytic decomposition of H2O2 (C0 = 5 mM) on the C3N4, Au/C3N4-250, Au/C3N4-500(Air), and Au/C3N4-500(N2). (a) The changes of C/C0 with irradiation time; (b) The changes of ln(C/C0) with irradiation time. The experimental conditions: 1 g/L of photocatalyst, pH 3.0, Xe lamp, and illumination intensity of 100 mW/cm2.
Figure 6. The photocatalytic decomposition of H2O2 (C0 = 5 mM) on the C3N4, Au/C3N4-250, Au/C3N4-500(Air), and Au/C3N4-500(N2). (a) The changes of C/C0 with irradiation time; (b) The changes of ln(C/C0) with irradiation time. The experimental conditions: 1 g/L of photocatalyst, pH 3.0, Xe lamp, and illumination intensity of 100 mW/cm2.
Catalysts 08 00147 g006
Figure 7. The repeated use of Au/C3N4-500(N2) and Au/C3N4-500(Air) for H2O2 generation. The experimental conditions: 1 g/L of photocatalyst, pH 3.0, 5% vol of 2-propanol, Xe lamp, illumination intensity of 100 mW/cm2, and O2-equilibrated conditions.
Figure 7. The repeated use of Au/C3N4-500(N2) and Au/C3N4-500(Air) for H2O2 generation. The experimental conditions: 1 g/L of photocatalyst, pH 3.0, 5% vol of 2-propanol, Xe lamp, illumination intensity of 100 mW/cm2, and O2-equilibrated conditions.
Catalysts 08 00147 g007
Figure 8. (a) TEM images and (b) particle-size distribution of Au/C3N4-500(N2)-reused; (c) TEM images and (d) particle-size distribution of Au/C3N4-500(Air)-reused.
Figure 8. (a) TEM images and (b) particle-size distribution of Au/C3N4-500(N2)-reused; (c) TEM images and (d) particle-size distribution of Au/C3N4-500(Air)-reused.
Catalysts 08 00147 g008
Figure 9. Mott–Schottky plots of (a) C3N4, (b) Au/C3N4-500(N2), and (c) Au/C3N4-500(Air) in an acidic solution (pH 3) in the dark. The frequency used was 1.0 kHz, 1.5 kHz, and 2.0 kHz.
Figure 9. Mott–Schottky plots of (a) C3N4, (b) Au/C3N4-500(N2), and (c) Au/C3N4-500(Air) in an acidic solution (pH 3) in the dark. The frequency used was 1.0 kHz, 1.5 kHz, and 2.0 kHz.
Catalysts 08 00147 g009
Table 1. XPS binding energy of Au4f and N1s in the Au-supported C3N4 catalysts.
Table 1. XPS binding energy of Au4f and N1s in the Au-supported C3N4 catalysts.
CatalystsAu4f/eVN1s/eV
4f5/24f7/2N-(C)3=N-C2NH
Au/C3N4-500(N2)87.483.7400.6399.2398.5
Au/C3N4-500(N2)-reused87.383.6400.6399.2398.5
Au/C3N4-500(Air)87.283.5400.6399.4398.6
Au/C3N4-500(Air)-reused87.083.3400.6399.4398.6
Table 2. The yield of hydrogen peroxide (H2O2) and the kinetic analyses.
Table 2. The yield of hydrogen peroxide (H2O2) and the kinetic analyses.
SampleH2O2 Yield at 120 min (μmol L−1)H2O2 Yield at 240 min (μmol L−1)Kf (μmol L−1 min−1)Kd (10−3 min−1)
C3N43215822.9291.61
Au/C3N4-2504737435.5145.68
Au/C3N4-500(Air)65310677.1704.46
Au/C3N4-500(N2)74713207.8933.54

Share and Cite

MDPI and ACS Style

Chang, X.; Yang, J.; Han, D.; Zhang, B.; Xiang, X.; He, J. Enhancing Light-Driven Production of Hydrogen Peroxide by Anchoring Au onto C3N4 Catalysts. Catalysts 2018, 8, 147. https://doi.org/10.3390/catal8040147

AMA Style

Chang X, Yang J, Han D, Zhang B, Xiang X, He J. Enhancing Light-Driven Production of Hydrogen Peroxide by Anchoring Au onto C3N4 Catalysts. Catalysts. 2018; 8(4):147. https://doi.org/10.3390/catal8040147

Chicago/Turabian Style

Chang, Xiaoyu, Junjiao Yang, Dandan Han, Bing Zhang, Xu Xiang, and Jing He. 2018. "Enhancing Light-Driven Production of Hydrogen Peroxide by Anchoring Au onto C3N4 Catalysts" Catalysts 8, no. 4: 147. https://doi.org/10.3390/catal8040147

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