Next Article in Journal
Alkoxy-Functionalized Schiff-Base Ligation at Aluminum and Zinc: Synthesis, Structures and ROP Capability
Next Article in Special Issue
Ru-gC3N4 Catalyzed Hydrodebenzylation of Benzyl Protected Alcohol and Acid Groups Using Sodium Hypophosphite as a Hydrogen Source
Previous Article in Journal
Titania-Clay Mineral Composites for Environmental Catalysis and Photocatalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Noble-Metal-Free NixSy-C3N5 Hybrid Nanosheet with Highly Efficient Photocatalytic Performance

College of Science, Huazhong Agricultural University, Wuhan 430070, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2021, 11(9), 1089; https://doi.org/10.3390/catal11091089
Submission received: 23 August 2021 / Revised: 5 September 2021 / Accepted: 7 September 2021 / Published: 9 September 2021
(This article belongs to the Special Issue Advances in Graphitic Carbon Nitride-Based Catalysts)

Abstract

:
The construction of highly efficient, low-cost and noble-metal-free photocatalysts depends on photocatalytic technology. Recently, N-rich C3N5 has been explored as a novel carbon nitride material with a much narrower band gap (~2.2 eV) than that of traditional C3N4 (~2.7 eV). Planting noble-metal-free active sites on C3N5 to improve its photocatalytic activity is of great significance. Herein, 2D NixSy nanosheet is facially loaded on 2D C3N5 using a hydrothermal procedure under a low temperature. Due to the quick separation of photogenerated carries between C3N5 and NixSy, this inexpensive noble-metal-free NixSy-C3N5 hybrid nanosheet is highly efficient and stable as a multifunctional catalyst in various applications, including photocatalytic H2 production from water and NO removal. Impressively, the apparent quantum yield (AQY) value for H2 production reaches 37.0% (at 420 nm) on optimal NixSy-C3N5 hybrids, which is much higher than that of Pt-C3N5 material. This work opens an avenue to the fabrication of low-cost and noble-metal-free catalysts for multifunctional photocatalytic applications.

Graphical Abstract

1. Introduction

Semiconductor-based photocatalytic technology is one of the most promising strategies for solving the ever-increasing energy crisis and addressing environmental pollution as well as concerns for public health [1,2,3,4]. Although numerous efforts have been made since the pioneering work on TiO2 photoelectrodes for water splitting in 1972 [5,6,7], the industrialization of this technology is still unrealized due to its inefficiency and problems with implementation. In general, the photocatalytic conversion efficiency is influenced by light absorption, photogenerated carrier separation and surface reduction-oxidation reactions. Previous research has shown that the aforementioned three factors cannot be fully satisfied in one semiconductor, and the construction of hybrid materials, such as a design co-catalyst or active sites on semiconductors and semiconductor-based heterojunctions, is an effective strategy to combine the advantages of different materials to enhance photocatalytic efficiency [3,8,9,10,11,12,13]. Thus, the design of robust hybrid materials is an essential element in the photocatalytic field.
Of the various semiconductors used in photocatalytic investigation, carbon nitride has drawn enormous attention because of its low cost, metal-free status, suitable band positions, stability, environmental-friendliness, etc. [2,14,15,16,17]. Recently, a boron-doped and nitrogen-deficient strategy was reported to adjust the band positions of C3N4 nanosheets; Z-scheme carbon nitride-based heterostructures were fabricated with high solar-to-hydrogen efficiency (1.16% under one-sun illumination) for photocatalytic water splitting using Pt and Co(OH)2 as H2 and O2 evolving co-catalysts, respectively [18,19]. Lin et al. reported the controlling of the reactive facets based on a carbon nitride PTI/Li+Cl single crystals for photocatalytic water splitting with the assistance of Pt/Co as co-catalysts [16]. In addition, C3N4 was decorated with oxygen-vacancy-rich BiOBr sheets to form a hybrid, which exhibited excellent photocatalytic CO2 reduction and NO removal efficiency [20]. However, the performance of C3N4-based materials is seriously limited by its wide band gap (~2.7 eV). Recently, a new type of N-rich carbon nitride (C3N5) was synthesized using various precursors [21,22,23,24,25,26]. The structure of C3N5 and C3N4 is very similar, except that some terminal amino groups on C3N4 are replaced by terminal 1,2,4-triazole or 1H-1,2,3-triazole on C3N5 [24]. C3N5 could absorb visible light wavelengths longer than 600 nm and exhibited potential application in the fields of water splitting [25,26], pollutant removal [21,26], gas adsorption/sensing [25], N2 fixing [23] and photovoltaics [22]. Recently, the photocatalytic H2 production activity of C3N5 prepared from 3-amino-1,2,4-triazole was comprehensively investigated with Pt as a co-catalyst by our group [27]. However, the precious Pt is not practical for wide application. The design of low-cost noble-metal-free co-catalysts on C3N5 has not yet been reported in the literature. The unique visible light absorption and band position of C3N5 that can afford NO removal by the photocatalytic oxidation procedure, which has yet to be investigated, will be of great significance for atmosphere purification.
Herein, following the above hypothesis, 2D nickel sulfide (NixSy) is hybridized with 2D C3N5 to construct a noble-metal free hybrid (NixSy-C3N5) using a facial hydrothermal method at 120 °C. The photocatalytic H2 production and NO oxidation properties of this hybrid are evaluated to illustrate the potential application of this material. Experimental results show that NixSy-C3N5 is a highly efficient, low-cost photocatalyst for H2 production, even exceeding Pt-C3N5, and can produce sufficient active species participating in converting NO into harmless N-containing species such as NO3. The present work will promote the development of low-cost photocatalytic materials to address energy and environmental issues.

2. Results and Discussion

2.1. Characterization

Figure 1 depicts the X-ray diffraction (XRD) patterns of NixSy and the NixSy-C3N5 hybrid obtained by the facial hydrothermal method at 120 °C; NixSy-160 °C is listed for comparison. As revealed in Figure 1a, strong diffraction peaks are observed for bare NixSy material, indicating its high crystallinity, and the NixSymaterial prepared under hydrothermal conditions at 120 °C primarily contains NiS (PDF card of 65–3419), while some weak peaks corresponding to Ni3S4 (PDF card of 47–1739) can be seen. This result demonstrates that the prepared NixSy is composed of NiS and Ni3S4, with NiS as the main component. To obtain nickel sulfide with a single phase of NiS or Ni3S4, the hydrothermal temperature for nickel sulfide preparation is adjusted and is shown in Figure 1a and Figure S1. As can be seen, the nickel sulfides obtained under temperatures from 100 to 170 °C are mixed phases of NiS and Ni3S4, with the main phase changing from NiS to Ni3S4. Taking NixSy at 160 °C (NixSy −160 °C) as an example (Figure 1a), it is mainly composed of Ni3S4 but some small peaks corresponding to NiS can also be found. It should be noted that a new phase of NiS2 in Ni3S4 is observed when the hydrothermal temperature is increased to 200 °C. The above result reveals that it is difficult to prepare nickel sulfide with a single phase using this facial method. Notably, it has been reported that nickel sulfides have multiple crystalline phases (i.e., NiS, Ni3S4, Ni3S2, NiS2, Ni7S6, and Ni9S8) depending on the relative concentration of nickel and sulfur precursors in the synthetic system, and the mixed phase of these nickel sulfides have exhibited better performance [28,29,30]. Thus, the materials containing NixSy mentioned in the following sections are all synthesized under 120 °C unless otherwise specified.
As displayed in Figure 1b, two distinct peaks corresponding to (100) and (002) planes of C3N5 are observed, which is in agreement with our previous work and the related literature [23,25,27]. For the hybrid of NixSy-C3N5, the two peaks for C3N5 exhibit no obvious shift, indicating that the present hydrothermal procedure does not destroy the basic structure of C3N5. Briefly, no obvious peaks corresponding to NixSy are found for 1.0 wt% NixSy-C3N5, which is likely due to the low loading amount of NixSy. However, some weak peaks attributed to NixSy are observed when the loading amount of NixSy is increased to 3.0 wt%, indicating the successful fabrication of the NixSy-C3N5 hybrid.
Generally, C3N5 exhibits the micro morphology of 2D nanosheet according to our previous work. Figure 2 displays the TEM images of NixSy and NixSy-C3N5 hybrids. In brief, NixSy exhibits irregular 2D nanosheets (Figure 2a), which is similar to that of C3N5. HRTEM images of NixSy (Figure 2b) confirm the fact that it is composed of NiS and Ni3S4, and clear lattice fringes of 0.25 and 0.28 nm corresponding to NiS (101) and Ni3S4 (311), respectively, are observed. This result is in good agreement with the above-mentioned XRD analysis. As depicted in Figure 2c for 1.0 wt% NixSy-C3N5, the sheets of NixSy and C3N5 contact tightly with each other, indicating that the NixSy-C3N5 is a 2D/2D nanosheet hybrid and the hydrothermal procedure for NixSy preparation has no obvious influence on the structure of C3N5. C3N5 possesses an amorphous structure as shown in our previous work. The HRTEM image of 1.0 wt% NixSy-C3N5 (Figure 2d) conveys that amorphous C3N5 is combined with NixSy, for which clear lattice fringes of 0.25 and 0.28 nm corresponding to NiS (101) and Ni3S4 (311), respectively, are also be observed. The TEM results confirm the successful fabrication of the NixSy-C3N5 hybrid.
To further investigate the components of the hybrid, we performed scanning electron microscope (SEM) and Energy-dispersive X-ray spectroscopy (EDS) element mapping of 1.0 wt% NixSy-C3N5, as shown in Figure 3. As seen in the figure, NixSy-C3N5 exhibits the basic morphology of nanosheets, which is in good agreement with the results of TEM. In addition, the hybrid is composed mainly of C, N, Ni and S elements, and each element disperses uniformly in the observed region from the mapping images. The high dispersion of NixSy on C3N5 will be favorable for subsequent photocatalytic performance.
Figure 4 and Figure S2 show the XPS spectra of NixSy-C3N5 to determine the state of each element in the hybrid. As exhibited in the survey spectrum of Figure S2a, C, N, Ni and S are all observed in 1.0 wt% NixSy-C3N5, indicating the successful fabrication of the hybrid, which is in good agreement with the above results of XRD, TEM and SEM analyses. The regional spectra of C 1s and N 1s (Figure S2b,c) are similar to that of C3N5 in our previous work [27], and no obvious shift is observed for the corresponding peak position, revealing that the hydrothermal procedure for the NixSy-C3N5 preparation has no obvious influence on the structure of C3N5. The high-resolution XPS spectrum of S 2p is shown in Figure 4a, in which the peak at ~168 eV represents the residual sulfate group originating from the hydrothermal procedure. Two peaks at ~162.3 and ~163.1 eV corresponding to S 2p3/2 and S 2p1/2, respectively, are observed, demonstrating the existence of a metal-sulfur (Ni-S) bond and sulfur ion [30,31,32]. Two distinct spin-orbit peaks at ~855 (Ni 2p3/2) and 873 eV (Ni 2p1/2), along with two satellite peaks (identified as ‘‘Sat.’’), are observed in the high-resolution XPS spectrum of Ni 2p (Figure 4b), suggesting the existence of nickel ions in the NixSy-C3N5 hybrid [30,31,33]. These results further confirm that the NixSy-C3N5 hybrid was successfully constructed.
The DRS spectra of NixSy, C3N5 and NixSy-C3N5 were recorded to provide a direct comparison of their light absorption properties, which play an important role in photocatalytic performance. As revealed in Figure 5, C3N5 exhibits strong absorption in the visible light wavelength longer than 600 nm due to its narrow band gap (~2.25 eV), as shown in our previous work [27]. NixSy can absorb the whole visible light region due to its black color. In addition, the visible light absorption property of NixSy-C3N5 displays the combination of the two components, demonstrating the successful construction of the NixSy-C3N5 hybrid. The excellent visible-light-responsive ability of NixSy-C3N5 is favorable for photocatalytic applications.

2.2. Photocatalytic H2 Production Performance

The photocatalytic activity of NixSy-C3N5 was first evaluated for H2 production under visible light irradiation (λ > 420 nm), with TEOA as the sacrificial reagent. Our previous work demonstrated that no obvious H2 production was observed on pristine C3N5 due to the lack of co-catalyst [27]. In addition, the control experiment displayed that H2 production is extremely limited on pristine NixSy. As depicted in Figure 6a, the H2 production rate can be greatly elevated when 0.5 wt% NixSy is loaded on C3N5, demonstrating that NixSy can facilitate the separation of photogenerated carriers of C3N5 and serve as an active site for H2 production. This result is in good agreement with the literature, where nickel sulfide is widely taken as a catalyst for photo- or electro-catalytic water reduction [31,32,33,34,35]. The influence of NixSy loading amounts on the H2 production rate of NixSy-C3N5 was also investigated and is shown in Figure 6a. The H2 production rate increases when increasing the loading amount of NixSy, and a platform is obtained when the fraction of NixSy changes from 1.0 wt% to 3.0 wt%. However, a decrease in photoactivity is observed for 4.0 wt% NixSy-C3N5, which may be related to the influence of excessive NixSy on the visible light absorption of C3N5. Generally, the initial amount of the S source should be higher than that of the Ni source in the nickel sulfide preparation process. Thus, taking 3.0 wt% NixSy-C3N5 as an example, the influence of the initial mol ratio of Ni/S (Ni(NO3)2/TAA) on the H2 production rate was investigated and is shown in Figure S3; the optimal Ni/S reaction condition is confirmed as 1/3, and the highest H2 production rate is ~1595 μmol/h.
As mentioned above, the main phase of NixSy is NiS, while a slight amount of Ni3S4 also exists in the catalyst. To explore the main phase of nickel sulfide, which plays a dominant role in photocatalytic H2 production activity, taking 3.0 wt% NixSy-C3N5 as a reference, the sample of 3.0 wt% NixSy-C3N5-160 (prepared at 160 °C hydrothermal treatment, with the main component of NixSy being Ni3S4; see Figure 1a) was also tested for H2 production under the same photocatalytic condition. As shown in Figure 6b, the difference in H2 production rates between 3.0 wt% NixSy-C3N5 and 3.0 wt% NixSy-C3N5-160 is slight, demonstrating that both NiS and Ni3S4 in NixSy can act as electron traps and active sites to greatly enhance the photocatalytic activity for H2 production. To evaluate the efficiency of the NixSy co-catalyst, Pt nanoparticle, a widely used co-catalyst in photocatalysis, was taken as reference. As shown in Figure 6b, the H2 production rate of 3.0 wt% NixSy-C3N5 is higher than that of 1.0 wt% Pt-C3N5 (the optimal Pt loaded amount is 1.0 wt% according to our previous work [27]), NiS [33,34] and NiS/C3N4 [36,37] materials (see Table 1). This result conveys that the present NixSy may be a competitive alternative to noble metals such as Pt as a co-catalyst utilized in photocatalytic H2 production.
Taking 3.0 wt% NixSy-C3N5 as a model, the change tendency of AQY values along with the irradiated wavelength is shown in Figure 6c; the AQY values decrease gradually when the monochromatic light changes from 420 to 700 nm. This result demonstrates that the photocatalytic H2 production of NixSy-C3N5 is dominated by its visible light absorption property. The highest AQY value is 37.0% at 420 nm light irradiation, which is higher than that of Pt-C3N5 exhibited in our previous work [27]. In addition, the NixSy-C3N5 hybrid can also work under 700 nm monochromatic light irradiation with an AQY value of 0.2%. To evaluate the cyclicity and durability of the present hybrid material for H2 production, a long-term test under visible light irradiation was conducted on 3.0 wt% NixSy-C3N5. As revealed in Figure 6d, the hybrid exhibits excellent stability for H2 production in the four test runs during a total of 16 h of visible light irradiation. The stability of NixSy-C3N5 is also confirmed by the XRD and DRS curves of the photocatalytic H2 production test, as depicted in Figure 7. No obvious change in the peak number and position is observed for NixSy-C3N5 before and after the photocatalytic test. In addition, there is only a slight change in the visible light responsive property of NixSy-C3N5 after the photocatalytic test when compared with the pristine material. The nickle concentration in the solution after the photocatalytic reaction was determined to be 20.3 ppb by ICP-MS. Thus, the leaching rate of Ni was 0.14%, which is negligible. These results further confirm the structural stability of NixSy-C3N5 during the photocatalysis procedure.

2.3. Photocatalytic NO Oxidation Performance

Figure 8a depicts the photocatalytic NO oxidation performance of NixSy-C3N5 and C3N5; quick removal of NO is observed, and the removal ratio reaches equilibrium during a short (~20 min) test in the continuous flow reaction system. The control experiment shows that no obvious activity for NO removal is observed on the pristine NixSy. The NO removal efficiency on pristine C3N5 is ~35%, which can be attributed to its wide light absorption property and suitable band position for O2 molecular activation and the generation of reactive oxygen species (ROS, such as •OH, •O2 and 1O2), which participate in the subsequent photocatalytic NO oxidation. Generally, the production of •O2 and 1O2 follows the equations of O2 + e → •O2 and •O2 + h+1O2 [11,38]. In addition, the production •OH follows the procedure of •O2 + 2H+ → H2O2 → •OH, since the VB position of C3N5 is more negative than the potential of •OH/H2O (2.38 V vs. NHE, pH = 0) according to our previous work [11,27]. The corresponding NO removal value can be improved to ~40% when 0.5 and 1.0 wt% NixSy are loaded on C3N5, which is much higher than that of the C3N4 materials (see Table 2) [39,40,41]. However, a decrease in NO removal efficiency is observed with a higher loading amount of NixSy (such as 3.0 wt%). It should be noted that excessive NixSy may cover the surface of C3N5 and then influence the light absorption by C3N5. The improved photocatalytic performance is related to the electron transfer between C3N5 and NixSy, as well as the enhanced efficiency for ROS generation, as will be discussed in the following section. The durability of 1.0 wt% NixSy-C3N5 on photocatalytic NO removal was tested under 2 h of light irradiation, as shown in Figure 8b; the removal ratio is maintained in the long-term test, revealing the excellent photocatalytic stability of NixSy-C3N5. The cycling performance of 1.0 wt% NixSy-C3N5 on photocatalytic NO removal is depicted in Figure 8c; there is only a slight decrease after six test runs, demonstrating the excellent photocatalytic repeatability of the present hybrid. Generally, the products of photocatalytic NO oxidation are considered to be NO2 and NO3. The production of NO2 is directly recorded on the NOx analyzer shown in Figure 8d. For NO3 detection, the irradiated NixSy-C3N5 samples are re-dispersed in water and then filtered. Finally, the transparent filtrate is analyzed by ion chromatography with NaNO3 as a reference. As displayed in Figure 8e, the production of NO3 during NO removal is confirmed qualitatively. Notably, the selectivity for NO3 production is estimated to be ~70%, based on NO removal efficiency and residual NO2 amount (Figure 8a,d).
Generally, photogenerated h+ and the series of ROS play significant roles in the photocatalytic NO oxidation procedure. To explore active species that participate in the reaction of NO removal, a series of trapping agent experiments were conducted, in which Na2C2O4, tert-butanol (TBA), p-benzoquinone (PBQ) and β-carotene were used as scavengers for h+, •OH, •O2 and 1O2, respectively [42]. As can be seen in Figure 9a, the photocatalytic NO removal efficiency is inhibited to different extents when different trapping agents are added, demonstrating that all of the above-mentioned species can participate in the procedure for NO removal. In addition, the dominant species are •O2 and 1O2, since the NO removal efficiency is greatly inhibited when the two species are captured. Electron spin resonance (ESR) spectra were recorded to monitor the formation of the main active species of •O2 and 1O2. As revealed in Figure 9b,c, characteristic signals for DMPO-•O2 and TEMPO (product of TEMP oxidized by 1O2) were detected [42,43,44], and the signal intensity for NixSy-C3N5 was higher than that of pristine C3N5 under light irradiation, demonstrating that greater amounts of ROS can be generated on the NixSy-C3N5 hybrid. The result of the EPR analysis is in good agreement with the results of the photocatalytic performance.
BET analysis was first conducted to explore the factors affecting the high photocatalytic performance of NixSy-C3N5 compared with C3N5. As shown in Figure S4, the specific surface area of NixSy-C3N5 (10.32 m2/g) is almost equal to that of C3N5 (10.60 m2/g), showing that this factor cannot afford the enhanced photoactivity for NixSy-C3N5. Considering that nickel sulfide is often taken as an active site for H2 production in the field of photo- and electro-catalysis, the influence of NixSy on the separation of e/h+ pairs generated from C3N5 under light excitation was investigated using steady-state and time-resolved PL as well as photocurrent-time and EIS behaviors. As depicted in Figure 10a, the PL intensity of C3N5 is dramatically quenched by NixSy, demonstrating that the loading of NixSy can significantly restrain the e/h+ recombination of C3N5 under light excitation and NixSy can serve as electron traps to facilitate the photogenerated electron transfer from C3N5 to NixSy [15,45]. This viewpoint is further supported by the time-resolved fluorescence spectra (TRFS) shown in Figure 10b; the PL decay rate on NixSy-C3N5 is slower than that of pristine C3N5, and the fluorescence lifetime of NixSy-C3N5 (11.56 ns) is longer than that of pristine C3N5 (8.50 ns), demonstrating that the recombination of photogenerated e/h+ pairs on C3N5 is restrained by NixSy through efficient internal electron transfer [15,46]. The generation and separation of e/h+ pairs on NixSy-C3N5 and C3N5 is also reflected in their photocurrent-time behaviors. Figure 10c shows the change tendency of photocurrent curves of the two materials along with commutative light-on and light-off operation. The high current value was recorded when the work electrode was irradiated by the Xe-lamp, and the value vanished suddenly when the light was turned down, indicating that the current is generated after the material is excited. In addition, the generation of photocurrent is repeatable. Moreover, the photocurrent value recorded on NixSy-C3N5 is much higher than that of C3N5, demonstrating that the separation of e/h+ pairs in NixSy-C3N5 under light irradiation is greatly improved. This result is in accordance with the conclusions of the PL analysis. A smaller radius was obtained on NixSy-C3N5 from the EIS Nyquist plots, as shown in Figure 10d, revealing the lower charge transfer resistance of NixSy-C3N5 [15,47]. The above results convey that NixSy-C3N5 possesses quicker internal charge migration than bare C3N5.
Based on the above results, the mechanism for photocatalytic H2 production or NO oxidation is proposed. Under visible light irradiation, C3N5 is excited and generates e/h+ pairs. NixSy can harvest the e on the CB of C3N5, and then the trapped electrons participate in subsequent surface reactions. For H2 production, the collected electrons on NixSy react with adsorbed H2O molecules and H2 is produced. The residual h+ on the VB of C3N5 is consumed by TEOA, and subsequently the whole photocatalytic H2 production procedure is accomplished. In the process of photocatalytic NO oxidation, the activation of O2 is considered the first step according to most reported studies. The generation of ROS such as •O2, 1O2 and •OH are enhanced by NixSy, as discussed in the above section. These ROS and the residual h+ on the VB of C3N5 can then participate in the NO oxidation procedure along with the formation of NO3 and NO2.

3. Materials and Methods

3.1. Material Preparation

All chemicals used were purchased commercially and used without purification. C3N5 was prepared by heating 3-amino-1,2,4-triazole under 500 °C in air according to our previous work [27]. Generally, NixSy-C3N5 was synthesized by the hydrothermal method at 120 °C. In brief, C3N5 was dispersed in 40 mL deionized water, followed by the addition of Ni(NO3)2 and thioacetamide (TAA) with a mol ratio of 1:3, and then transferred into a 100 mL Teflon-lined autoclave. Finally, the autoclave was sealed and kept at 120 °C for 12 h. The precipitate was washed with distilled water and absolute ethanol several times after the autoclave cooled to room temperature. The obtained samples were dried in vacuum at 80 °C for 12 h. The loading amount of NixSy on C3N5 was calculated based on the initial input Ni ion mass fraction, and a series of NixSy-C3N5 hybrids with an NixSy loading amount varying between 0.5 and 4.0 wt% were prepared by changing the amount of Ni(NO3)2 and TAA accordingly. Pristine NixSy was also prepared by the same method in the absence of C3N5. Notably, all the materials containing NixSy in the following section were prepared under 120 °C hydrothermal conditions unless otherwise specified.

3.2. Catalyst Characterization

XRD patterns were recorded on a powder X-ray diffractometer with Cu Kα radiation (D8 Advance Bruker Inc., Munich, Germany). TEM and HRTEM images were acquired using FEI TALOS F200. SEM and EDS mapping images were acquired using Gemini Sigma 300. The Brunauer–Emmett–Teller (BET) specific surface area was evaluated using a nitrogen adsorption–desorption apparatus (ASAP 2040, Micrometrics Inc., Norcross, GA, USA), with all samples degassed at 120 °C for 12 h prior to measurements. The valence state of each element in the catalyst was analyzed by X-ray photoelectron spectroscopy (XPS) on a Thermo Fisher ESCALAB 250Xi X-ray photoelectron spectroscope. A Shimadzu UV-3100 recording spectrophotometer and fluorescence spectrometer (F-4600, Hitachi Inc., Tokyo, Japan) were used to record UV-vis diffuse reflectance (DRS) and PL spectra measurements, respectively. Time-resolved fluorescence spectra (TRFS) were obtained on an Edinburgh FLS 1000. Electron paramagnetic resonance (EPR) spectra were recorded on a JES-FA200 EPR Spectrometer. DMPO (5,5-dimethyl-1-pyrroline-N-oxide) was used as radical spin-trapped reagent for •OH and •O2. TEMP (2,2,6,6-tetramethylpiperidine) was used as the trapping agent for 1O2 [11].
The photocurrent and electrochemical impedance measurements were recorded on a CHI 660D electrochemical workstation (Chenhua Instrument, Shanghai, China) with a conventional three-electrode configuration. Pt foil and Ag/AgCl (saturated KCl) were used as the counter electrode and reference electrode, respectively. The working electrode was made by spreading catalyst/nafion slurry on FTO glass and 0.1 M Na2SO4 aqueous solution was used as the electrolyte.

3.3. Photocatalytic Activity Tests

The photocatalytic H2 production rate and AQY measurements were carried out on a top-irradiated reaction vessel containing catalyst and sacrificial reagent (TEOA) connected to a closed gas system with a 300 W Xe-lamp (PLS SXE300, Beijing Perfectlight Inc., China) equipped with a cutoff filter (λ > 420 nm) or band-pass filters (λ = 420, 500 nm etc.); the operation system is shown in Figure S5. The H2 production rate was detected by a GC (SP7820, TCD detector, 5 Å molecular sieve columns, and Ar carrier), and AQY values were calculated according to our previous report [27]:
AQY = 2 × (Number of evolved H2 molecules)/(Number of incident photons)
For photocatalytic NO removal, a continuous flow reactor under ambient conditions was adopted, as per our previous work [11]. The volume of the rectangular reactor was 4.5 L (30 cm × 15 cm × 10 cm). A 25 mg catalyst was dispersed into the mixture of ethanol and water and then transferred into a culture dish with a diameter of 12 cm, and the dish was then placed in the reactor after the solvent was evaporated. A 30 W visible LED (General Electric) was used as a light source. The gas (containing NO and air) flow rate through the reactor was controlled at 1000 mL/min by a mass flow controller with an initial NO concentration of ~600 ppb. The NO and NO2 concentrations were recorded on a NOx analyzer model T200 (Teledyne API). The generation of NO3 was detected by ion chromatography on Thermo DIONEX ICS-900.

4. Conclusions

In summary, a low-cost and noble-metal-free 2D NixSy-C3N5 hybrid photocatalyst was facially fabricated by a hydrothermal procedure under a low temperature and analyzed using a series of methods. The tight contact between 2D NixSy and 2D C3N5 sheets facilitates internal photogenerated electron transfer from C3N5 to NixSy, which is demonstrated by steady-state and time-resolved PL spectra as well as light-switched photocurrent behavior. The NixSy-C3N5 hybrid possesses multiple photocatalytic applications (H2 production from water and NO removal) with high activity and stability. In particular, the optimal NixSy-C3N5 photocatalyst exhibits a high AQY value of 37.0% at 420 nm, exceeding Pt-C3N5 material and revealing the potential of the NixSy in photocatalytic water splitting. This work not only demonstrates a simple method with which to construct a high-efficiency noble-metal-free hybrid photocatalyst, but it also exhibits the potential applications of such a photocatalyst in various fields.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11091089/s1, Figure S1, XRD patterns of NixSy prepared under different hydrothermal temperatures; Figure S2, XPS survey spectrum of NixSy-C3N5 (a) and high-resolution spectra of C 1s (b) and N 1s (c); Figure S3., Influence of mole ratio of Ni/S on the H2 production activity of 3.0 wt% NixSy-C3N5; Figure S4, Comparison of nitrogen adsorption–desorption isotherms between C3N5 and NixSy-C3N5; Figure S5, Photograph of photocatalytic H2 production system.

Author Contributions

Conceptualization, X.Z. and Y.Y.; methodology, C.P., J.H., L.S. and S.W.; formal analysis, L.H. and C.P.; investigation, L.H.; data curation, L.H. and C.P.; writing—original draft preparation, L.H.; writing—review and editing, X.Z. and Y.Y.; supervision, X.Z. and Y.Y.; project administration, X.Z.; funding acquisition, X.Z. and Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (21703075, 51872107, 52073110, and 22002047), Natural Science Foundation of Hubei Province (2020CFB694) and Fundamental Research Funds for the Central Universities (2662020LXPY005).

Data Availability Statement

All the data presented in this article are available at https://doi.org/10.6084/m9.figshare.16586414 (accessed on 8 September 2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, H.; He, C.; Fu, L.; Huo, J.; Zhao, C.; Li, X.; Song, Y. Capture and separation of CO2 on BC3 nanosheets: A DFT study. Chin. Chem. Lett. 2021. [Google Scholar] [CrossRef]
  2. Fu, L.; Wang, R.; Zhao, C.; Huo, J.; He, C.; Kim, K.; Zhang, W. Construction of Cr-embedded graphyne electrocatalyst for highly selective reduction of CO2 to CH4: A DFT study. Chem. Eng. J. 2021, 414, 128857. [Google Scholar] [CrossRef]
  3. Wang, R.; He, C.; Chen, W.; Zhao, C.; Huo, J. Rich B active centers in Penta-B2C as high-performance photocatalyst for nitrogen reduction. Chin. Chem. Lett. 2021. [Google Scholar] [CrossRef]
  4. Shen, R.; Ding, Y.; Li, S.; Zhang, P.; Xiang, Q.; Ng, Y.; Li, X. Constructing low-cost Ni3C/twin-crystal Zn0.5Cd0.5S heterojunction/homojunction nanohybrids for efficient photocatalytic H2 evolution. Chin. J. Catal. 2021, 42, 25–36. [Google Scholar] [CrossRef]
  5. Wageh, S.; Al-Ghamdi, A.; Jafer, R.; Li, X.; Zhang, P. A new heterojunction in photocatalysis: S-scheme heterojunction. Chin. J. Catal. 2021, 42, 667–669. [Google Scholar] [CrossRef]
  6. Wang, J.; Kuo, M.; Zeng, P.; Xu, L.; Chen, S.; Peng, T. Few-layer BiVO4 nanosheets decorated with SrTiO3: Rh nanoparticles for highly efficient visible-light-driven overall water splitting. Appl. Catal. B Environ. 2020, 279, 119377. [Google Scholar] [CrossRef]
  7. Jiang, Z.; Chen, Q.; Zheng, Q.; Shen, R.; Zhang, P.; Li, X. Constructing 1D/2D Schottky-Based Heterojunctions between Mn0.2Cd0.8S Nanorods and Ti3C2 Nanosheets for Boosted Photocatalytic H2 Evolution. Acta Phys. Chim. Sin. 2021, 37, 2010059. [Google Scholar]
  8. Shen, R.; Ren, D.; Ding, Y.; Guan, Y.; Ng, Y.; Zhang, P.; Li, X. Nanostructured CdS for efficient photocatalytic H2 evolution: A review. Sci. China Mater. 2020, 63, 2153–2188. [Google Scholar] [CrossRef]
  9. Sun, W.; Xiang, Y.; Jiang, Z.; Wang, S.; Yang, N.; Jin, S.; Sun, L.; Teng, H.; Chen, H. Designed polymeric conjugation motivates tunable activation of molecular oxygen in heterogeneous organic photosynthesis. Sci. Bull. 2021. [Google Scholar] [CrossRef]
  10. Zhang, X.; Han, L.; Chen, H.; Wang, S. Direct catalytic nitrogen oxide removal using thermal, electrical or solar energy. Chin. Chem. Lett. 2021. [Google Scholar] [CrossRef]
  11. Wang, S.; Ding, X.; Zhang, X.; Pang, H.; Hai, X.; Zhan, G.; Zhou, W.; Song, H.; Zhang, L.; Chen, H.; et al. In Situ Carbon Homogeneous Doping on Ultrathin Bismuth Molybdate: A Dual-Purpose Strategy for Efficient Molecular Oxygen Activation. Adv. Funct. Mater. 2017, 27, 1703923. [Google Scholar] [CrossRef]
  12. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  13. Xu, F.; Meng, K.; Cheng, B.; Wang, S.; Xu, J.; Yu, J. Unique S-scheme heterojunctions in self-assembled TiO2/CsPbBr3 hybrids for CO2 photoreduction. Nat. Commun. 2020, 11, 4613. [Google Scholar] [CrossRef]
  14. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.; 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]
  15. Xia, P.; Antonietti, M.; Zhu, B.; Heil, T.; Yu, J.; Cao, S. Designing Defective Crystalline Carbon Nitride to Enable Selective CO2 Photoreduction in the Gas Phase. Adv. Funct. Mater. 2019, 29, 1900093. [Google Scholar] [CrossRef]
  16. Lin, L.; Lin, Z.; Zhang, J.; Cai, X.; Lin, W.; Yu, Z.; Wang, X. Molecular-level insights on the reactive facet of carbon nitride single crystals photocatalysing overall water splitting. Nat. Catal. 2020, 3, 649–655. [Google Scholar] [CrossRef]
  17. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  18. Zhao, D.; Dong, C.; Wang, B.; Chen, C.; Huang, Y.; Diao, Z.; Li, S.; Guo, L.; Shen, S. Synergy of dopants and defects in graphitic carbon nitride with exceptionally modulated band structures for efficient photocatalytic oxygen evolution. Adv. Mater. 2019, 31, 1903545. [Google Scholar] [CrossRef]
  19. Zhao, D.; Wang, Y.; Dong, C.; Huang, Y.; Chen, J.; Xue, F.; Shen, S.; Guo, L. Boron-doped nitrogen-deficient carbon nitride-based Z-scheme heterostructures for photocatalytic overall water splitting. Nat. Energy 2021, 6, 388–397. [Google Scholar] [CrossRef]
  20. Liu, D.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Surface Engineering of g-C3N4 by Stacked BiOBr Sheets Rich in Oxygen Vacancies for Boosting Photocatalytic Performance. Angew. Chem. Int. Ed. 2020, 59, 4519–4524. [Google Scholar] [CrossRef]
  21. Zhang, J.; Jing, B.; Tang, Z.; Ao, Z.; Xia, D.; Zhu, M.; Wang, S. Experimental and DFT insights into the visible-light driving metal-free C3N5 activated persulfate system for efficient water purification. Appl. Catal. B Environ. 2021, 289, 120023. [Google Scholar] [CrossRef]
  22. Kumar, P.; Vahidzadeh, E.; Thakur, U.; Kar, P.; Alam, K.; Goswami, A.; Mahdi, N.; Cui, K.; Bernard, G.; Michaelis, V.; et al. C3N5: A Low Bandgap Semiconductor Containing an Azo-Linked Carbon Nitride Framework for Photocatalytic, Photovoltaic and Adsorbent Applications. J. Am. Chem. Soc. 2019, 141, 5415–5436. [Google Scholar] [CrossRef] [PubMed]
  23. Li, M.; Lu, Q.; Liu, M.; Yin, P.; Wu, C.; Li, H.; Zhang, Y.; Yao, S. Photoinduced Charge Separation via the Double-Electron Transfer Mechanism in Nitrogen Vacancies g-C3N5/BiOBr for the Photoelectrochemical Nitrogen Reduction. ACS Appl. Mater. Interfaces 2020, 12, 38266–38274. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, I.Y.; Kim, S.; Jin, X.; Premkumar, S.; Chandra, G.; Lee, N.S.; Mane, G.P.; Hwang, S.J.; Umapathy, S.; Vinu, A. Ordered mesoporous C3N5 with a combined triazole and triazine framework and its graphene hybrids for the oxygen reduction reaction (ORR). Angew. Chem. Int. Ed. 2018, 57, 17381–17386. [Google Scholar] [CrossRef]
  25. Mane, G.P.; Talapaneni, S.N.; Lakhi, K.S.; Ilbeygi, H.; Ravon, U.; Al-Bahily, K.; Mori, T.; Park, D.H.; Vinu, A. Highly ordered nitrogen-rich mesoporous carbon nitrides and their superior performance for sensing and photocatalytic hydrogen generation. Angew. Chem. Int. Ed. 2017, 56, 8481–8485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Liu, T.; Yang, G.; Wang, W.; Wang, C.; Wang, M.; Sun, X.; Xu, P.; Zhang, J. Preparation of C3N5 nanosheets with enhanced performance in photocatalytic methylene blue (MB) degradation and H2-evolution from water splitting. Environ. Res. 2020, 188, 109741. [Google Scholar] [CrossRef]
  27. Peng, C.; Han, L.; Huang, J.; Wang, S.; Zhang, X.; Chen, H. Comprehensive investigation on robust photocatalytic hydrogen production over C3N5. Chin. J. Catal. 2021. [Google Scholar] [CrossRef]
  28. Wang, H.; Zhang, W.; Zhang, X.; Hu, S.; Zhang, Z.; Zhou, W.; Liu, H. Multi-interface collaboration of graphene cross-linked NiS-NiS2-Ni3S4 polymorph foam towards robust hydrogen evolution in alkaline electrolyte. Nano Res. 2021. [Google Scholar] [CrossRef]
  29. Vijaya, S.; Landi, G.; Wu, J.J.; Anandan, S. Ni3S4/CoS2 mixed-phase nanocomposite as counter electrode for Pt-free dye-sensitized solar cells. J. Power Sources 2020, 478, 229068. [Google Scholar] [CrossRef]
  30. Zou, H.; He, B.; Kuang, P.; Yu, J.; Fan, K. NixSy nanowalls/nitrogen-doped graphene foam is an efficient trifunctional catalyst for unassisted artificial photosynthesis. Adv. Funct. Mater. 2018, 28, 1706917. [Google Scholar] [CrossRef]
  31. Yu, L.; Wu, L.; McElhenny, B.; Song, S.; Luo, D.; Zhang, F.; Yu, Y.; Chen, S.; Ren, Z. Ultrafast room-temperature synthesis of porous S-doped Ni/Fe (oxy) hydroxide electrodes for oxygen evolution catalysis in seawater splitting. Energy Environ. Sci. 2020, 13, 3439–3446. [Google Scholar] [CrossRef]
  32. Zhang, G.; Feng, Y.S.; Lu, W.T.; He, D.; Wang, C.Y.; Li, Y.K.; Wang, X.Y.; Cao, F.F. Enhanced catalysis of electrochemical overall water splitting in alkaline media by Fe doping in Ni3S2 nanosheet arrays. ACS Catal. 2018, 8, 5431–5441. [Google Scholar] [CrossRef]
  33. Yan, A.; Shi, X.; Huang, F.; Fujitsuka, M.; Majima, T. Efficient photocatalytic H2 evolution using NiS/ZnIn2S4 heterostructures with enhanced charge separation and interfacial charge transfer. Appl. Catal. B Environ. 2019, 250, 163–170. [Google Scholar] [CrossRef]
  34. Wang, B.; Ding, Y.; Deng, Z.; Li, Z. Rational design of ternary NiS/CQDs/ZnIn2S4 nanocomposites as efficient noble-metal-free photocatalyst for hydrogen evolution under visible light. Chin. J. Catal. 2019, 40, 335–342. [Google Scholar] [CrossRef]
  35. Zhao, X.; Feng, J.; Liu, J.; Shi, W.; Yang, G.; Wang, G.C.; Cheng, P. An Efficient, Visible-Light-Driven, Hydrogen Evolution Catalyst NiS/ZnxCd1− xS Nanocrystal Derived from a Metal–Organic Framework. Angew. Chem. Int. Ed. 2018, 130, 9938–9942. [Google Scholar] [CrossRef]
  36. Wen, J.; Li, X.; Li, H.; Ma, S.; He, K.; Xu, Y.; Fang, Y.; Liu, W.; Gao, Q. Enhanced visible-light H2 evolution of g-C3N4 photocatalysts via the synergetic effect of amorphous NiS and cheap metal-free carbon black nanoparticles as co-catalysts. Appl. Surf. Sci. 2015, 358, 204–212. [Google Scholar] [CrossRef]
  37. Zhao, H.; Zhang, H.; Cui, G.; Dong, Y.; Wang, G.; Jiang, P.; Wu, X.; Zhao, N. A photochemical synthesis route to typical transition metal sulfides as highly efficient cocatalyst for hydrogen evolution: From the case of NiS/g-C3N4. Appl. Catal. B Environ. 2018, 225, 284–290. [Google Scholar] [CrossRef]
  38. Wang, S.; Ding, X.; Yang, N.; Zhan, G.; Zhang, X.; Dong, G.; Zhang, L.; Chen, H. Insight into the effect of bromine on facet-dependent surface oxygen vacancies construction and stabilization of Bi2MoO6 for efficient photocatalytic NO removal. Appl. Catal. B Environ. 2020, 265, 118585. [Google Scholar] [CrossRef]
  39. Zhang, R.; Zhang, A.; Cao, Y.; Wang, S.; Dong, F.; Zhou, Y. Mo-doped carbon nitride homojunction to promote oxygen activation for enhanced photocatalytic performance. Chem. Eng. J. 2020, 401, 126028. [Google Scholar] [CrossRef]
  40. Cao, Y.; Zhang, R.; Zheng, Q.; Cui, W.; Liu, Y.; Zheng, K.; Dong, F.; Zhou, Y. Dual functions of O-atoms in the g-C3N4/BO0.2N0.8 interface: Oriented charge flow in-plane and separation within the interface to collectively promote photocatalytic molecular oxygen activation. ACS Appl. Mater. Interfaces 2020, 12, 34432–34440. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, Z.; Huang, Y.; Chen, M.; Shi, X.; Zhang, Y.; Cao, J.; Ho, W.; Lee, S.C. Roles of N-vacancies over porous g-C3N4 microtubes during photocatalytic NOx removal. ACS Appl. Mater. Interfaces 2019, 11, 10651–10662. [Google Scholar] [CrossRef]
  42. Yang, X.; Wang, S.; Chen, T.; Yang, N.; Jiang, K.; Wang, P.; Li, S.; Ding, X.; Chen, H. Chloridion-induced dual tunable fabrication of oxygen-deficient Bi2WO6 atomic layers for deep oxidation of NO. Chen, Chin. J. Catal. 2021, 42, 1013–1023. [Google Scholar] [CrossRef]
  43. Kou, M.; Deng, Y.; Zhang, R.; Wang, L.; Wong, P.K.; Su, F.; Ye, L. Molecular oxygen activation enhancement by BiOBr0.5I0.5/BiOI utilizing the synergistic effect of solid solution and heterojunctions for photocatalytic NO removal. Chin. J. Catal. 2020, 41, 1480–1487. [Google Scholar] [CrossRef]
  44. Shi, X.; Wang, P.; Li, W.; Bai, Y.; Xie, H.; Zhou, Y.; Ye, L. Change in photocatalytic NO removal mechanisms of ultrathin BiOBr/BiOI via NO3–adsorption. Appl. Catal. B Environ. 2019, 243, 322–329. [Google Scholar] [CrossRef]
  45. Zhang, X.; Xiao, J.; Hou, M.; Xiang, Y.; Chen, H. Robust visible/near-infrared light driven hydrogen generation over Z-scheme conjugated polymer/CdS hybrid. Appl. Catal. B Environ. 2018, 224, 871–876. [Google Scholar] [CrossRef]
  46. Zhang, X.; Peng, T.; Yu, L.; Li, R.; Li, Q.; Li, Z. Visible/near-infrared-light-induced H2 production over g-C3N4 co-sensitized by organic dye and zinc phthalocyanine derivative. ACS Catal. 2015, 5, 504–510. [Google Scholar] [CrossRef]
  47. Wang, J.; Wang, G.; Cheng, B.; Yu, J.; Fan, J. Sulfur-doped g-C3N4/TiO2 S-scheme heterojunction photocatalyst for Congo Red photodegradation. Chin. J. Catal. 2021, 42, 56–68. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) NixSy prepared using a hydrothermal procedure at 120 and 160 °C, and (b) C3N5 and NixSy-C3N5 with different NixSy loading amounts prepared using a hydrothermal procedure at 120 °C.
Figure 1. XRD patterns of (a) NixSy prepared using a hydrothermal procedure at 120 and 160 °C, and (b) C3N5 and NixSy-C3N5 with different NixSy loading amounts prepared using a hydrothermal procedure at 120 °C.
Catalysts 11 01089 g001
Figure 2. TEM and HRTEM images of NixSy (a,b) and 1.0 wt% NixSy-C3N5 (c,d).
Figure 2. TEM and HRTEM images of NixSy (a,b) and 1.0 wt% NixSy-C3N5 (c,d).
Catalysts 11 01089 g002
Figure 3. SEM and element mapping images of 1.0 wt% NixSy-C3N5.
Figure 3. SEM and element mapping images of 1.0 wt% NixSy-C3N5.
Catalysts 11 01089 g003
Figure 4. High-resolution XPS spectrum of S 2p (a) and Ni 2p (b).
Figure 4. High-resolution XPS spectrum of S 2p (a) and Ni 2p (b).
Catalysts 11 01089 g004
Figure 5. DRS spectra of NixSy, C3N5 and NixSy-C3N5.
Figure 5. DRS spectra of NixSy, C3N5 and NixSy-C3N5.
Catalysts 11 01089 g005
Figure 6. (a) Influence of NixSy loading amount on the H2 production rate of NixSy-C3N5 hybrid. (b) Comparison of H2 production rate over Pt-C3N5, NixSy-C3N5 and NixSy-C3N5-160. (c) Change tendency of AQY values along with irradiated wavelength on NixSy-C3N5. (d) Long-term test for H2 production on NixSy-C3N5. Conditions: 45 mg catalyst, 30 mL 10 wol% TEOA/H2O aqueous solution, λ > 420 nm (for a,b,d), λ = 420, 450, 500, 550, 600, 650, 700 nm (for c).
Figure 6. (a) Influence of NixSy loading amount on the H2 production rate of NixSy-C3N5 hybrid. (b) Comparison of H2 production rate over Pt-C3N5, NixSy-C3N5 and NixSy-C3N5-160. (c) Change tendency of AQY values along with irradiated wavelength on NixSy-C3N5. (d) Long-term test for H2 production on NixSy-C3N5. Conditions: 45 mg catalyst, 30 mL 10 wol% TEOA/H2O aqueous solution, λ > 420 nm (for a,b,d), λ = 420, 450, 500, 550, 600, 650, 700 nm (for c).
Catalysts 11 01089 g006
Figure 7. Comparison of (a) XRD and (b) DRS curves of pristine and light irradiated NixSy-C3N5.
Figure 7. Comparison of (a) XRD and (b) DRS curves of pristine and light irradiated NixSy-C3N5.
Catalysts 11 01089 g007
Figure 8. (a) Comparison of photocatalytic NO removal behaviors on NixSy, C3N5 and NixSy-C3N5. Durability (b) and cyclability (c) of photocatalytic NO removal on 1.0 wt% NixSy-C3N5. Detection of NO2 (d) and NO3 (e) during photocatalytic NO removal on NixSy-C3N5.
Figure 8. (a) Comparison of photocatalytic NO removal behaviors on NixSy, C3N5 and NixSy-C3N5. Durability (b) and cyclability (c) of photocatalytic NO removal on 1.0 wt% NixSy-C3N5. Detection of NO2 (d) and NO3 (e) during photocatalytic NO removal on NixSy-C3N5.
Catalysts 11 01089 g008
Figure 9. (a) Trapping agent experiments for photocatalytic NO removal. ESR spectra of DMPO-•O2 (b) and TEMPO (c) for ROS detection.
Figure 9. (a) Trapping agent experiments for photocatalytic NO removal. ESR spectra of DMPO-•O2 (b) and TEMPO (c) for ROS detection.
Catalysts 11 01089 g009
Figure 10. Comparison of steady-state (a) and time-resolved (b) fluorescence spectra as well as photocurrent-time (c) and EIS (d) behaviors of C3N5 and NixSy-C3N5.
Figure 10. Comparison of steady-state (a) and time-resolved (b) fluorescence spectra as well as photocurrent-time (c) and EIS (d) behaviors of C3N5 and NixSy-C3N5.
Catalysts 11 01089 g010
Table 1. The comparison of the photocatalytic H2 production performance of the NixSy-C3N5 with some reported catalysts.
Table 1. The comparison of the photocatalytic H2 production performance of the NixSy-C3N5 with some reported catalysts.
PhotocatalystsReaction ConditionsH2 Production
Rate μmol g−1 h−1
Ref.
Pt/C3N5300 W Xe lamp (λ > 420 nm), TEOA28,956[27]
NiS/ZnIn2S4320 W Xe lamp (λ > 420 nm), Lactic Acid3333[33]
NiS/CQDs/ZnIn2S4300 W Xe lamp (λ > 420 nm), TEOA568[34]
g-C3N4/1.5% NiS300 W Xe lamp (λ > 420 nm), TEOA395[36]
NiS/g-C3N4300 W Xe lamp (λ > 420 nm), TEOA16,400[37]
NixSy-C3N5300 W Xe lamp (λ > 420 nm), TEOA35,444this work
Table 2. The comparison of the photocatalytic NO removal performance of the NixSy-C3N5 with some reported catalysts.
Table 2. The comparison of the photocatalytic NO removal performance of the NixSy-C3N5 with some reported catalysts.
PhotocatalystsReaction ConditionsNO Removal
Efficiency (%)
Ref.
Mo-g-C3N4/g-C3N4flow reactor, metal halide lamp, >420 nm36.0[39]
g-C3N4/BO0.2N0.8flow reactor, metal halide lamp, >420 nm30.2[40]
N-Vacancies g-C3N4flow reactor, Xe lamp, >420 nm32.8[41]
NixSy-C3N5flow reactor, visible LED40.0this work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Han, L.; Peng, C.; Huang, J.; Sun, L.; Wang, S.; Zhang, X.; Yang, Y. Noble-Metal-Free NixSy-C3N5 Hybrid Nanosheet with Highly Efficient Photocatalytic Performance. Catalysts 2021, 11, 1089. https://doi.org/10.3390/catal11091089

AMA Style

Han L, Peng C, Huang J, Sun L, Wang S, Zhang X, Yang Y. Noble-Metal-Free NixSy-C3N5 Hybrid Nanosheet with Highly Efficient Photocatalytic Performance. Catalysts. 2021; 11(9):1089. https://doi.org/10.3390/catal11091089

Chicago/Turabian Style

Han, Lixiao, Cong Peng, Jinming Huang, Linhao Sun, Shengyao Wang, Xiaohu Zhang, and Yi Yang. 2021. "Noble-Metal-Free NixSy-C3N5 Hybrid Nanosheet with Highly Efficient Photocatalytic Performance" Catalysts 11, no. 9: 1089. https://doi.org/10.3390/catal11091089

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