

MoS2/Ag/g-C3N4 Z型异质结的制备及其光催化性能
-
关键词:
- MoS2/Ag/g-C3N4
- / 异质结
- / 光催化降解
- / 降解机理
English
MoS2/Ag/g-C3N4 Z-scheme heterojunction: Preparation and photocatalytic performance
-
Key words:
- MoS2/Ag/g-C3N4
- / heterojunction
- / photocatalytic degradation
- / degradation mechanism
-
0. Introduction
Photocatalysis is an effective solution for energy and environmental problems as it can degrade organic pollutants. Many new photocatalytic materials have been developed and reported, including graphite-like carbon nitride (g-C3N4), which has gained significant attention[1-2]. This material is a promising photocatalyst due to its low cost, wide visible light response range, and high stability[3-4]. However, the photocatalytic performance of g-C3N4 is hindered by the efficiency of charge‑fast recombination and the limited range of light absorption. To efficiently develop and utilize g-C3N4, modification studies on g-C3N4 are necessary[5-6]. Recent research has shown that the properties of photocatalysts, such as photoelectric properties and charge separation, greatly affect their photocatalytic activity. Therefore, the modification of g-C3N4 has focused on improving these properties, including optical properties, charge separation, and surface charge separation[7]. For instance, the catalyst′s wide absorption region is crucial for enhancing photoactivity. Combining g-C3N4 with other semiconductors expands its light absorption range and enhances its catalytic activity. The composite photocatalysts based on g‑C3N4 show satisfactory photoresponsive properties, which effectively improve the photocatalytic performance. Improving all the properties of g‑C3N4 with a single modification is difficult. To obtain g‑C3N4‑based photocatalysts with excellent performance, it is essential to dope and compose multiple substances. Researchers have developed new photocatalysts using this strategy, such as Ag3PO4/Ag/g‑C3N4[8], ZnO/g‑C3N4[9], WO3/g‑C3N4[10], and CoFe1.9Y0.1O4/g‑C3N4[11]. However, their efficiency requires further improvement.
Metal nanoparticles, such as Au, Ag, and Cu, exhibit a surface plasmon resonance effect that broadens the catalyst′s absorption range in the visible light spectrum. When silver is introduced onto the g-C3N4 surface, it creates a metal-semiconductor heterojunction, enhancing the efficiency of charge transfer and minimizing the recombination of photogenerated electron-hole pairs. This, in turn, leads to a marked enhancement in the photocatalytic performance of g-C3N4.
MoS2 is a semiconductor photocatalyst with a narrow band gap (Eg) ranging from 1.2 to 2.0 eV, making it highly sensitive to visible light. The staggered band alignment between g-C3N4 and MoS2 enhances the separation of photogenerated carriers, thereby improving photocatalytic efficiency. Spherical MoS2, which consists of multiple nanosheets stacked together, offers a higher number of active sites, as these sites are primarily concentrated at the edges of the nanosheets.
The study of metal nanoparticles is necessary due to their significant impact on electron-hole transport and separation. In a study by Ju et al.[12], noble metals Pt, Au, and Ag have the potential to enhance the photocatalytic activity of g-C3N4/MoS2 composites through density functional theory. However, research on MoS2/Ag/g-C3N4 composites is currently limited to theoretical calculations and predictions. Composite photocatalytic materials consisting of g‑C3N4, Ag, and MoS2 have gained significant attention due to their construction and performance.
This paper presents the preparation of MoS2/Ag/g-C3N4 composites to enhance the photocatalytic activity for the degradation of organics. The addition of silver significantly prolonged the visible-light response of g-C3N4 and improved charge separation. The performance of the photocatalysts was characterized by various means. The article investigates the catalytic degradation mechanism of MoS2/Ag/g-C3N4 to reveal the reason behind the increased photocatalytic activity.
1. Experimental
1.1 Materials and reagents
The analytically pure reagents were obtained from China Pharmaceutical Group Chemical Reagent Co. These included urea, thiourea, sodium molybdate dihydrate, polyvinylpyrrolidone, ethylene glycol, methanol, rhodamine B (RhB), disodium ethylenediaminetetraacetic acid, p-benzoquinone (BQ), and concentrated hydrochloric acid. Solutions were prepared using deionized water.
1.2 Preparation of Ag/g-C3N4 and MoS2
Typically, 25.06 g of urea was placed in a crucible with a lid. The crucible was then placed in an oven, heated at a rate of 5 ℃·min-1 until the temperature reached 500 ℃, and held for 6 h. After cooling to room temperature, the obtained g-C3N4 sample was ground. Then, 2.40 g of the prepared g-C3N4 was weighed into a 100 mL glass beaker, 80 mL of methanol was added, and the mixture was ultrasonicated for 1 h to obtain suspension A. A specific amount of 0.24 mol·L-1 AgNO3 aqueous solution was added dropwise to suspension A using a constant pressure dropping funnel. The mixture was stirred for 2 h at 70 ℃ while irradiating a 500 W xenon lamp. Finally, the mixture was filtered, washed, and vacuum-dried at a temperature of 70 ℃ for 10 h to yield the desired Ag/g-C3N4 sample.
The 1.26 g of sodium molybdate dihydrate was transferred into a 50 mL beaker containing 20 mL of deionized water, and solution B was obtained after sufficient stirring. In a separate container, 1.37 g of thiourea was weighed and dissolved in 25 mL of anhydrous ethanol. The solution was stirred at a temperature of 60 ℃ on a thermostatic stirrer until the solid was completely dissolved. Using a constant-pressure funnel, solution B was added to the thiourea solution and stirred at a temperature of 60 ℃ for 1 h. Concentrated hydrochloric acid was introduced to adjust the pH to 3. The mixture was then stirred for 1 h, after which it was filtered. The precipitate obtained was washed three times with deionized water and anhydrous ethanol and subsequently dried in a vacuum oven at 60 ℃ for 24 h. Finally, the dried samples were ground to obtain the desired MoS2 sample.
1.3 Preparation of MoS2/Ag/g-C3N4
To prepare dispersion C, 0.50 g of Ag/g-C3N4 was dispersed with 20 mL of absolute ethanol in a 50 mL beaker. The mixture was stirred at room temperature for 30 min on a magnetic stirrer and then sonicated for 60 min in an ultrasonic cleaner. In a 50 mL separate beaker, 1.24 g of sodium molybdate dihydrate was dissolved in 20 mL of deionized water to obtain solution D. Finally, the two solutions had to be thoroughly mixed to achieve the desired outcome. 1.37 g thiourea was dissolved in 25 mL of anhydrous ethanol at a temperature of 60 ℃ with magnetic stirring until completely dissolved, forming solution E. Solution E was added to the previously prepared dispersion through a constant-pressure funnel. The mixture was stirred at 60 ℃ for 1 h, after which the pH was adjusted to 3 using concentrated hydrochloric acid, and was continuously stirred for an additional hour. The mixture was then filtered, and the resulting precipitate was washed three times with deionized water and three times with anhydrous ethanol. The washed precipitate was dried in a vacuum oven at 60 ℃ for 24 h. Finally, the dried material was ground to obtain the target sample.
1.4 Characterizations
The phase structures of the samples were analyzed using X-ray diffraction (XRD, Bruker D8 Advance), test conditions: Cu Kα (λ=0.154 056 nm), ceramic X-ray tube, a voltage of 40 kV, a current of 40 mA, a scanning range of 5°-80°, and a scanning speed of 5 (°)·min-1. The morphology and crystalline architecture were examined utilizing a ZEISS Sigma500 Scanning Electron Microscope (SEM, Carl Zeiss AG, Oberkochen, Germany). The morphology of the samples was analyzed using high-resolution transmission electron microscopy (HRTEM, FEI Talos-F200S, accelerating voltage: 200 kV). Elemental composition and valence band (VB) structure were investigated via X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha). The band gap was measured using UV‑Vis diffuse reflectance spectroscopy (DRS, Shimadzu UV‑3600). The electron-hole recombination efficiency of the samples was analyzed using a photoluminescence (PL) spectrometer (RF-5301PC). The electrochemical properties of the samples were determined using an electrochemical workstation (CHI660E), and free radical active species experiments were performed using electron paramagnetic resonance (EPR, JEOLJES-FA200). Ultraviolet photoelectron spectroscopy (UPS, Japan PHI 5000 Versaprobe Ⅲ) was used to determine the sample work function. The total organic carbon (TOC) CONTENT of the RhB solution was detected using the German Jena Multi3100 TOC analyzer.
1.5 Photocatalytic degradation
To improve the accuracy of the catalytic degradation experiments and to achieve an adsorption-desorption equilibrium between the catalyst and the pollutant, adsorption in the dark was conducted for 30 min. 50 mg of the sample was accurately weighed and added to 50 mL 10 mg·L-1 RhB solution. The mixture was stirred under a 500 W xenon lamp for 10 min, and 5 mL of the sample was centrifuged and filtered every 10 min to determine the absorbance of the supernatant at 550 nm. Additionally, five cyclic degradation experiments were conducted under the aforementioned experimental conditions in the absence of dark adsorption to investigate the cyclic degradation stability of the photocatalytic composites.
2. Results and discussion
2.1 Structures and morphologies
Fig. 1 displays the XRD patterns of MoS2, Ag/g-C3N4, and MoS2/Ag/g-C3N4. The MoS2 sample exhibited three distinct diffraction peaks corresponding to the (002), (100), and (110) crystal planes, respectively[13]. The diffraction pattern of Ag/g-C3N4 showed two diffraction peaks at 12.9° and 27.4°, corresponding to the (100) and (002) crystal planes of g-C3N4, respectively[14-15]. No silver diffraction peak was observed in Ag/g-C3N4, possibly due to the low diffraction intensity or the low silver content. The XRD patterns of MoS2/Ag/g-C3N4 showed the diffraction peaks of MoS2 and g-C3N4, indicating the successful preparation of the composite photocatalyst.
Figure 1
SEM and TEM images of the composite samples are shown in Fig. 2a and 2b. The larger black particles were MoS2, and the smaller grey particles loaded on the surface of the graphite phase were silver, which can be seen from the figure, indicating that the target photocatalyst has been successfully composited[16-18]. Fig. 2c shows the HRTEM image of MoS2/Ag/g-C3N4, from which it can be seen that the crystal plane spacings (d) of silver and MoS2 nanoparticles were 0.269 2 and 0.618 6 nm respectively, corresponding to the (111) crystal plane of silver and the (002) crystal plane of MoS2, respectively. In addition, the loading of silver nanoparticles can establish a fast charge transfer channel between MoS2 and g-C3N4, thus enhancing the photocatalytic activity.
Figure 2
2.2 XPS analysis
Fig. 3a shows the XPS survey spectra of the MoS2/Ag/g-C3N4 sample. MoS2/Ag/g-C3N4 was mainly composed of the elements C, N, Ag, Mo, and S. The peaks at 284.44 and 287.85 eV in Fig. 3b correspond to C—C and C—C=N in Ag/g-C3N4, respectively. Notably, in composite material MoS2/Ag/g-C3N4, it can be seen that the binding energy of C1s shifted, indicating that the composite of MoS2 and Ag/g-C3N4 was not a simple mixture but rather a stress generated between the two, and the chemical environment of the composite changed. Fig. 3c shows the high-resolution N1s XPS spectra of MoS2/Ag/g-C3N4, with the peak of the three-coordinated N—C3 located at 399.65 eV, the peak of the surface amino group (—NH2) located at 400.64 eV, the π-excitation peak of the C—N heterocyclic ring located at 404.59 eV, and the peak of the dicoordinated N species (C=N—C) located at 398.35 eV[17-18]. The S2p XPS spectra of MoS2/Ag/g-C3N4 are shown in Fig. 3d, and the characteristic peaks at 160.62 and 161.74 eV are attributed to S2p3/2 and S2p1/2, respectively, which correspond to S2- in MoS2. From the high-resolution Mo3d XPS spectrum of MoS2/Ag/g-C3N4 in Fig. 3e, the characteristic peaks at 227.73 and 230.82 eV are attributed to Mo3d5/2 and Mo3d3/2, respectively[18-19]. As shown in Fig. 3f, the binding energies of MoS2/Ag/g-C3N4 are located at 369.56 and 375.62 eV with a set of strong spectral peaks corresponding to the 3d5/2 and 3d3/2 characteristic peaks of metallic Ag monomers, respectively, indicating that the samples contain metallic Ag, which confirms that the Ag nanoparticles loaded into the composites successfully. In summary, after the combination of MoS2 and Ag/g-C3N4, the shift of element binding energy indicates a change in its chemical environment. MoS2 contacts with Ag/g-C3N4 to form a heterojunction, which is beneficial for subsequent catalytic degradation reactions.
Figure 3
2.3 Energy band structures
The composites were analyzed using DRS tests to determine their absorption and energy band structure across different wavelengths. Fig. 4a displays the UV-Vis absorption spectra of MoS2, g-C3N4, Ag/g-C3N4, and MoS2/Ag/g-C3N4. The light absorption region of Ag/g-C3N4 had a limited response of only about 550 nm, while the light absorption region of MoS2 exhibited higher absorption in the visible range. However, the UV-visible absorption wavelength range of MoS2/Ag/g-C3N4 was broader than that of the individual catalysts Ag/g-C3N4 and g-C3N4. Importantly, the formation of a Z-scheme heterojunction restricts the recombination of photogenerated electrons and holes, and the silver atoms can serve as a bridge for the photogenerated electron and hole pairs. Consequently, the MoS2/Ag/g-C3N4 composites demonstrated high catalytic activity under the same irradiation conditions[19]. Additionally, the VB (valance band)-XPS spectra of g-C3N4 and MoS2 are shown in Fig. 4b and 4c. VB-XPS detected the VB potentials of g-C3N4 and MoS2 in a small range[20-21]. The VB energies of g-C3N4 and MoS2 were 1.62 and 1.86 eV, respectively. Therefore, using the formula EVB-NHE=φ+EVB-XPS-4.44 [where EVB-NHE is the VB energy based on the normal hydrogen electrode (NHE), φ is the electron work function of the XPS analyzer with a value of 4.55 eV, and EVB-XPS is the VB energy tested by VB-XPS][22], the calculated values of the VB energies of g-C3N4 and MoS2 were 1.73 and 1.97 eV, respectively. Understanding the band gaps and energy band structures of the materials is the key to exploring the photocatalytic reaction mechanism. According to the analysis of the Mott-Schottky curves (Fig. 4d and 4e). The curve of MoS2 was positively sloped, indicating that MoS2 is an n-type semiconductor with a flat band potential of 0.08 V (vs Ag/AgCl). The curve of g-C3N4 was positively sloped, indicating that g-C3N4 is an n-type semiconductor with a flat band potential of -0.85 V (vs Ag/AgCl). Therefore, since the flat band potential of n-type semiconductor is closer to its conduction band (CB) potential, according to the formula E=E′-0.197 (where E′ is the potential relative to Ag/AgCl electrode, E is the potential relative to the NHE), the CB potentials of g-C3N4 and MoS2 could be obtained as -1.05 and -0.12 V (vs NHE), respectively[23]. The band energy levels of g-C3N4 and MoS2 are shown in Fig. 4f.
Figure 4
2.4 Electrochemical properties and photocatalytic activity
The electron-hole recombination rates of the samples were analyzed by PL spectroscopy, as shown in Fig. 5a. Ag/g-C3N4 had high fluorescence intensity and electron-hole recombination rate. However, the photogenerated carrier recombination rate and fluorescence intensity decreased after the formation of composite samples of Ag/g-C3N4 with MoS2[24-25]. This indicates that the electron-hole recombination efficiency was significantly suppressed, suggesting that the photocatalytic properties were enhanced and the performance was improved.
Figure 5
To further investigate the separation and fluorescence lifetime of photoexcited charges, time-resolved photoluminescence spectra (TRPL) were measured. The decay curves were smoothly fitted with a double-exponential equation as shown in Fig. 5b. The average lifetime (τa) was calculated according to the following Eq.1:
$ \tau_{\mathrm{a}}=\left(A_1 \tau_1^2+A_2 \tau_2^2\right)/\left(A_1 \tau_1+A_2 \tau_2\right) $ (1) where A1 and A2 are the amplitudes of distinct decay processes, while τ1 and τ2 represent the corresponding lifetimes. The fitted decay curves disclose the lifetime (τa). It is noted that the unrecombined radiative carriers get involved in surface photocatalytic reactions. Therefore, the longer the fluorescence lifetime of the material, the more conducive to carrier separation, the greater the chance of photogenerated charge participation in the photocatalytic reaction, and the stronger the catalytic activity[26]. As can be seen from Fig. 5b, the fluorescence lifetimes of Ag/g-C3N4 and MoS2/Ag/g-C3N4 were 1.78 and 2.07 ns, respectively. The fluorescence lifetime of MoS2/Ag/g-C3N4 was larger than that of Ag/g-C3N4, which suggests that the photogenerated electron-hole separation is higher due to the formation of heterojunction.
Fig. 5c illustrates the Nyquist curves for Ag/g-C3N4, MoS2, and MoS2/Ag/g-C3N4. MoS2/Ag/g-C3N4 exhibited a smaller arc radius compared to Ag/g-C3N4. This suggests a reduction in resistance alongside an enhancement in both electron-hole separation efficiency and photocatalytic performance, which is attributed to the incorporation of MoS2 onto Ag/g-C3N4[27-28].
Fig. 6a shows the results of the photocatalytic degradation experiments on the samples. The degradation rate (η) can be calculated according to η=(1-At/A0)×100% (where At is the absorbance of RhB at time t, and A0 is the absorbance of the initial RhB solution). In the absence of a photocatalyst, the degradation rate of RhB remained unchanged after 120 min of light exposure, indicating that there was weak self-degradation of the RhB solution. However, when the catalyst was added, the degradation rate of the RhB solution gradually increased with prolonged light exposure time, demonstrating the catalytic activity of the sample under illumination. Among them, the photocatalytic activity of MoS2/Ag/g-C3N4 was the highest, reaching a 98% degradation rate after 120 min. In addition, for the RhB aqueous solution, its photocatalytic degradation process can be expressed as a first-order reaction with the following reaction Eq.2:
Figure 6
$ -\ln \left(A_t/A_0{ }^{\prime}\right)=k t $ (2) where A0′ is the absorbance after adsorption equilibrium and k is the degradation rate constant. In Fig. 6b, the k value of MoS2/Ag/g-C3N4 was 0.024 min-1, which was two and three times higher than that of MoS2 (0.012 min-1) and Ag/g-C3N4 (0.008 1 min-1), respectively. Thus, the composite of MoS2 and Ag/g-C3N4 can significantly improve the photocatalytic degradation rate.
The extent of mineralization during the photocatalytic degradation of RhB by the photocatalyst can be assessed through the rate of change in total organic carbon (TOC). As shown in Fig. 6c, the mineralization rates of RhB by Ag/g-C3N4, MoS2, and MoS2/Ag/g-C3N4 within 2 h were 52%, 59%, and 70%, respectively. MoS2/Ag/g-C3N4 had the highest degree of mineralization of RhB, indicating that the formation of heterojunction enhances the catalytic activity.
2.5 Cyclic degradation stability and degradation mechanism
The cyclic degradation stability of the samples is shown in Fig. 7a and 7b, which illustrates that the degradation rate of MoS2/Ag/g-C3N4 catalyzed degradation of RhB showed a decreasing trend with the increase in the number of cyclic degradation, but the decrease was not significant. After five cycles of degradation, the degradation rate remained at about 85%. In the five- cycle degradation experiments, the composite degradation performance was significantly reduced, possibly due to the loss of catalyst during recycling or as a result of silver leaching. Cyclic degradation experiments showed that the degradation rate could still be maintained at 85.3% in the presence of loss during catalyst recycling, indicating that the composite photocatalyst had good cyclic degradation stability. In addition, the XRD patterns of MoS2/Ag/g-C3N4 before and after photocatalytic degradation are shown in Fig. 7c. The intensity of the diffraction peaks after photocatalytic degradation changed but not significantly, indicating that MoS2/Ag/g-C3N4 had a stable crystal structure during the photocatalytic degradation process.
Figure 7
The mechanism of photocatalysis in the MoS2/Ag/g-C3N4 degradation reaction system was investigated by introducing isopropanol (IPA, 20 mmol·L-1), disodium ethylenediaminetetraacetate (EDTA-2Na, 20 mmol·L-1), and benzoquinone (BQ, 20 mmol·L-1). Fig. 8a shows that the degradation rate of RhB did not decrease significantly after adding IPA [hydroxyl radical (·OH) scavenger] during the reaction process, indicating that ·OH is not an active substance in the degradation reaction process. However, the degradation rate of RhB was significantly reduced after the addition of BQ [superoxide radicals (·O2-) scavenger] during the degradation process. It was shown that ·O2- was the main active substance in the degradation reaction. Furthermore, the addition of EDTA-2Na, which acts as a scavenger for h+, led to a reduction in the degradation rate of RhB, indicating that h+ plays a role in the RhB degradation process[29-30]. To further investigate the influence of ·O2-, an EPR experiment was performed under xenon lamp irradiation. As shown in Fig. 8b, no signal was detected in the absence of light, but after 5 min of exposure to visible light, a clear peak corresponding to the ·O2- signal was observed. This confirms that the sample had a strong response to visible light and sufficient oxidizing ability to convert O2 to ·O2- [31]. The formation of the MoS2/Ag/g-C3N4 system by the Z-scheme mechanism was further confirmed by UPS. The UPS spectra of g-C3N4 and MoS2 are shown in Fig. 8c and 8d. The work functions (WF) of g-C3N4 and MoS2 were calculated from Fig. 8c and 8d to be 4.95 and 5.44 eV, respectively, indicating that MoS2/Ag/g-C3N4 should be the reduced and oxidized photocatalysts, respectively. The transfer of photogenerated electrons at the interface between semiconductors depends on their figure of merit; therefore, once g-C3N4 and MoS2 are in close contact, the photogenerated electrons in MoS2 will transfer to g-C3N4 through the heterojunction until their Fermi energy levels reach an equilibrium state[32-33]. As a result of the work function differences, the surface of g-C3N4 will accumulate a positive charge, whereas the MoS2 surface will retain a negative charge. This leads to the formation of an internal electric field at the interface between g-C3N4 and MoS2, which aids in the efficient separation of photogenerated charges. On the contrary, due to the energy band bending and internal electric field, the photogenerated electrons cannot be transferred in a way that follows the conventional type Ⅱ pathway, which is thermodynamically unfavorable for sustained charge transfer[34]. Thus, under visible light irradiation, the photogenerated electrons on the CB of MoS2 are transferred to the VB of g-C3N4 via silver nanoparticles due to the formation of an electric field and combined with the remaining holes[35]. Based on the above analysis, the photocatalytic mechanism is illustrated in Fig. 9. In this process, silver acts as a bridge for electron transport, facilitating the rapid recombination of holes in the VB of g-C3N4 with electrons in the CB of MoS2. This recombination leads to the generation of a significant number of electrons. As a result, a large number of electrons accumulate in the CB of g-C3N4, while holes gather in the VB of MoS2. This separation of charge carriers enhances photocatalytic efficiency by promoting the generation of reactive species that degrade pollutants[36]. The electrons in the g-C3N4 CB have sufficient energy to convert O2 to ·O2- [-0.33 V (vs NHE)]. Then, ·O2- can directly oxidize organic pollutants. Meanwhile, h+ in the VB of MoS2 directly oxidizes organic pollutants[37].
Figure 8
Figure 9
3. Conclusions
The paper presents the preparation of the photocatalyst MoS2/Ag/g-C3N4 Z-scheme heterojunction using a simple hydrothermal synthesis method. The photocatalytic effect of MoS2/Ag/g-C3N4 catalysts on the degradation of RhB was evaluated, and the results showed a degradation rate of 98% within 120 min. The MoS2/Ag/g-C3N4 photocatalyst exhibited excellent catalytic performance. The material′s efficient photocatalytic performance is attributed to its high separation efficiency of photogenerated carriers and effective Z-scheme electron transfer mechanism, as well as its excellent photo-oxidation-reduction ability. This study presents an effective strategy for developing applied photocatalysts with high efficiency, narrow band gap, and excellent cyclic degradation stability at low cost.
-
-
[1]
ZHAO B B, ZHONG W, CHEN F, WANG P, BIE C B, YU H G. High-crystalline g-C3N4 photocatalysts: Synthesis, structure modulation, and H2-evolution application[J]. Chin. J. Catal., 2023, 52: 127-143. doi: 10.1016/S1872-2067(23)64491-2
-
[2]
NGUYEN T K A, PHAM T T, GENDENSUREN B, OH E S, SHIN E W. Defect engineering of water-dispersible g-C3N4 photocatalysts by chemical oxidative etching of bulk g-C3N4 prepared in different calcination atmospheres[J]. J. Mater. Sci. Technol., 2022, 103: 232-243. doi: 10.1016/j.jmst.2021.07.013
-
[3]
MISHRA A, MEHTA A, BASU S, SHETTI N P, REDDY K R, AMINABHAVI T M. Graphitic carbon nitride (g-C3N4)-based metal-free photocatalysts for water splitting: A review[J]. Carbon, 2019, 149: 693-721. doi: 10.1016/j.carbon.2019.04.104
-
[4]
NIUY Q, HU F G, XU H L, ZHANG S Z, SONG B, WANG H L, LI M L, SHAO G, WANG H L, LU H X. Exploration for high performance g-C3N4 photocatalyst from different precursors[J]. Mater. Today Commun., 2023, 34: 105040. doi: 10.1016/j.mtcomm.2022.105040
-
[5]
LIU S Z, WANG S B, JIANG Y, ZHAO Z Q, JIANG G Y, SUN Z Y. Synthesis of Fe2O3 loaded porous g-C3N4 photocatalyst for photocatalytic reduction of dinitrogen to ammonia[J]. Chem. Eng. J., 2019, 373: 572-579. doi: 10.1016/j.cej.2019.05.021
-
[6]
KONG L R, WANG J G, MU X J, RUI LI, LI X X, FAN X X, SONG P, MA F C, SUN M T. Porous size dependent g-C3N4 for efficient photocatalysts: Regulation synthesizes and physical mechanism[J]. Mater. Today Energy, 2019, 13: 11-21. doi: 10.1016/j.mtener.2019.04.011
-
[7]
CUI L, LIU S L, WANG F K, LI J Y, SONGY H, SHENGY , ZOU H F. Growth of uniform g-C3N4 shells on 1D TiO2 nanofibers via vapor deposition approach with enhanced visible light photocatalytic activity[J]. J. Alloy. Compd., 2020, 826: 154001. doi: 10.1016/j.jallcom.2020.154001
-
[8]
LIU Q W, MENG Y, LIU Q M, XU M, HU Y H, CHEN S K. Synthesis of Ag3PO4/Ag/g-C3N4 composite for enhanced photocatalytic degradation of methyl orange[J]. Molecules, 2023, 28(16): 6082. doi: 10.3390/molecules28166082
-
[9]
WANG Y Y, YANG X J, LOU J H. Enhance ZnO photocatalytic performance via radiation modified g-C3N4[J]. Molecules, 2022, 27(23): 8476. doi: 10.3390/molecules27238476
-
[10]
FELLIPE M C, CAUE R, EDUARDO B A. Improving g-C3N4:WO3 Z-scheme photocatalytic performance under visible light by multivariate optimization of g-C3N4 synthesis[J]. Appl. Surf. Sci., 2021, 537: 147904. doi: 10.1016/j.apsusc.2020.147904
-
[11]
CHEN S K, JIANG D C, CAO Y M, ZENG G, CHI H B, MIAO Y X, ZHANG Y F, LI L, HE Y X, KE F, YE S. Doping and heterojunction construction dual-regulation: Magnetically recoverable CoFe1.9Y0.1O4/g-C3N4 nanosheets with enhanced visible-light-driven photocatalytic activity[J]. Mater. Lett., 2021, 287: 129275. doi: 10.1016/j.matlet.2020.129275
-
[12]
JU L, LIU C, SHI L R, SUN L. The high-speed channel made of metal for interfacial charge transfer in Z-scheme g-C3N4/MoS2 water-splitting photocatalyst[J]. Mater. Res. Express, 2019, 6: 115545. doi: 10.1088/2053-1591/ab509c
-
[13]
ZHOU D F, QIU C Q. Study on the effect of Co doping concentration on optical properties of g-C3N4[J]. Chem. Phys. Lett., 2019, 728: 70-73. doi: 10.1016/j.cplett.2019.04.060
-
[14]
MAGED S, EL-BORADY O M, EL-HOSAINY M, MAGED E K. Efficient photocatalytic reduction of p-nitrophenol under visible light irradiation based on Ag NPs loaded brown 2D g-C3N4/g-C3N4 QDs nanocomposite[J]. Environ. Sci. Pollut. Res., 2023, 30: 117909-117922. doi: 10.1007/s11356-023-30010-z
-
[15]
ZHAI N X, LUO J H, SHU P C, MEI J, LI X P, YAN W X. 1D/2D CoTe2@MoS2 composites constructed by CoTe2 nanorods and MoS2 nanosheets for efficient electromagnetic wave absorption[J]. Nano Res., 2023, 16: 10698-10706. doi: 10.1007/s12274-023-5777-9
-
[16]
LI Z C, YAN X X, TANG Z K, HUO Z Y, LI G L, JIAO L Y, LIU L M, ZHANG M, LUO J, ZHU J. Direct observation of multiple rotational stacking faults coexisting in freestanding bilayer MoS2[J]. Sci. Rep., 2017, 7: 8323. doi: 10.1038/s41598-017-07615-9
-
[17]
TRUONG N X, DIEN N T, TUE N N, KHANH D Q, MIKLÓS N, SHOAIB M, OTTÓ H. Effect of ruthenium modification of g-C3N4 in the visible-light-driven photocatalytic reduction of Cr(Ⅵ)[J]. Catalysts, 2023, 13(6): 964.
-
[18]
YU M, NIKOLAY K, LENNART V, PATRICIA J K, EMIEL J M H. Investigation of the active phase in K-promoted MoS2 catalysts for methanethiol synthesis[J]. ACS Catal., 2020, 10(3): 1838-1846.
-
[19]
SENTHILNATHAN S, GANESH K K, SUGUNRAJ S, AMMAL D M, RAJENDRAPRASAD M, SUGANTHI M, SHREE K K, SASIKUMAR P, MOHAMED A, WIREEN L T D. MoS2 modified g-C3N4 composite: A potential candidate for photocatalytic applications[J]. J. Saudi Chem. Soc., 2023, 27(5): 101717.
-
[20]
MOHAMMED A B, KHALED A, GUBRAN A, SUJAY SHEKAR G C, BASHEER M A, MAHMOOD M S A, AMAR A K, LOKANATH N K, NEPPOLIAN B, BHOJYA NAIK H S. Tailoring morphology and structure of 1D/2D isotype g-C3N4 for sonophotocatalytic hydrogen evaluation[J]. Surf. Interfaces, 2023, 42: 103511.
-
[21]
XIE H Y, WANG K, LI S L, JIN Z L. Construction of Co9S8/MoS2/Ni2P double S-scheme heterojunction for enhanced photocatalytic hydrogen evolution[J]. Surf. Interfaces, 2023, 42: 103353.
-
[22]
LI X B, XIONG J, GAO X M, MA J, CHEN Z, KANG B B, LIU J Y, LI H, FENG Z J, HUANG J T. Novel BP/BiOBr S-scheme nano-heterojunction for enhanced visible-light photocatalytic tetracycline removal and oxygen evolution activity[J]. J. Hazard. Mater., 2020, 387: 121690.
-
[23]
YAO Y, SHEN Q H, CHEN Q F, HE Y Y, JIANG L B, LIU J, LI Y, CAO X D, YANG H. Ag/AgX (X=Cl, Br, or I) nanocomposite loaded on Ag3PO4 tetrapods as a photocatalyst for the degradation of contaminants[J]. ACS Appl. Nano Mater., 2024, 7: 3711-3723.
-
[24]
YU J S, HWANG J J, LEE J Y, HA D H, KIM J H. Tuning photoluminescence spectra of MoS2 with liquid crystals[J]. Nanoscale, 2021, 13: 16641-16648.
-
[25]
ADITYA S, MAYORA V, KEUN H C, SUNG O W. Mechanistic investigations on emission characteristics from g-C3N4, g-C3N4@Pt and g-C3N4@Ag nanostructures using X-ray absorption spectroscopy[J]. Curr. Appl. Phys., 2018, 18(11): 1458-1464.
-
[26]
QI H J, WANG C Y, SHEN L P, WANG H M, LIAN Y, ZHANG H X, MA H X, ZHANG Y, ZHANG J Z. α-NiS/g-C3N4 nanocomposites for photocatalytic hydrogen evolution and degradation of tetracycline hydrochloride[J]. Catalysts, 2023, 13: 983.
-
[27]
DU H, GAO X H, MA Q X, YANG X J, ZHAO T S. Cu/PCN metal-semiconductor heterojunction by thermal reduction for photoreaction of CO2-aerated H2O to CH3OH and C2H5OH[J]. ACS Omega, 2022, 7(19): 16817-16826.
-
[28]
SHAO G L, LU Y H, HONG J H, XUE X X, HUANG J Q, XU Z Y, LU X C, JIN Y Y, LIU X, LI H M, HU S, KAZU S, HAN Z, JIANG Y, LI S S, FENG Y X, PAN A L, LIN Y C, CAO Y, LIU S. Seamlessly splicing metallic SnxMo1-xS2 at MoS2 edge for enhanced photoelectrocatalytic performance in microreactor[J]. Adv. Sci., 2020, 7(24): 2002172.
-
[29]
LI X B, KANG B B, DONG F, ZHANG Z Q, LUO X D, HAN L, HUANG J T, FENG Z J, CHEN Z, XU J L, PENG B L, WANG Z L. Enhanced photocatalytic degradation and H2/H2O2 production performance of S-pCN/WO2.72 S-scheme heterojunction with appropriate surface oxygen vacancies[J]. Nano Energy, 2021, 81: 105671.
-
[30]
DU X Y, SONG S, WANG Y T, JIN W F, DING T, TIAN Y, LI X G. Facile one-pot synthesis of defect-engineered step-scheme WO3/g-C3N4 heterojunctions for efficient photocatalytic hydrogen production[J]. Catal. Sci. Technol., 2021, 11: 2734-2744.
-
[31]
ZHANG H J, DENG L, CHEN H, ZHANG Y T, LIU Y H, CHEN Y F, ZENG H X, SHI Z. How MoS2 assisted sulfur vacancies featured Cu2S in hollow Cu2S@MoS2 nanoboxes to activate H2O2 for efficient sulfadiazine degradation[J]. Chem. Eng. J., 2022, 446: 137364.
-
[32]
SWEETY S, SEKHAR C R. Trigonal (1T) and hexagonal (2H) mixed phases MoS2 thin films[J]. Appl. Surf. Sci., 2019, 474: 227-231.
-
[33]
SHI Y N, CHEN J J, MAO Z Y, BRADLEY D F, WANG D J. Construction of Z-scheme heterostructure with enhanced photocatalytic H2 evolution for g-C3N4 nanosheets via loading porous silicon[J]. J. Catal., 2017, 356: 22-31.
-
[34]
XIONG W, DONG Y H, PAN A H. Fabricating a type Ⅱ heterojunction by growing lead-free perovskite Cs2AgBiBr6 in situ on graphite-like g-C3N4 nanosheets for enhanced photocatalytic CO2 reduction[J]. Nanoscale, 2023, 15: 15619-15625.
-
[35]
LUO X L, YUAN J H, LIU J H, HU H K, YANG Z Y, HOU X P, SUN Q, XU D. Construction of type Ⅱ ZnBi2O4/g-C3N4 heterojunction photocatalysts for efficient degradation of acid red B[J]. New J. Chem., 2023, 47: 21944-21959.
-
[36]
KEBENA G M, WU C, SZUTSEN L. Novel Ag3PO4@ZIF-8 p-n heterojunction for effective photodegradation of organic pollutants[J]. J. Water Process Eng., 2023, 52: 103586.
-
[37]
HASEEB U, ZAHID H, ABRAR A, IAN S B, REBWAR N D, ZIAUR R. MoS2 and CdS photocatalysts for water decontamination: A review[J]. Inorg. Chem. Commun., 2023, 153: 110775.
-
[1]
-
-

计量
- PDF下载量: 0
- 文章访问数: 27
- HTML全文浏览量: 3