Next Article in Journal
Hydrogen Gas Quality for Gas Network Injection: State of the Art of Three Hydrogen Production Methods
Next Article in Special Issue
Implementation of the Digital Sales Channel in the Coatings Industry
Previous Article in Journal
Special Issue on “Thermal Safety of Chemical Processes”
Previous Article in Special Issue
Real-Time Industrial Process Fault Diagnosis Based on Time Delayed Mutual Information Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

g-C3N4 Sensitized by an Indoline Dye for Photocatalytic H2 Evolution

1
Shanghai Key Laboratory of Atmospheric Particle Pollution and Prevention (LAP3), Department of Environmental Science and Engineering, Fudan University, Shanghai 200433, China
2
Shanghai Institute of Pollution Control and Ecological Security, Shanghai 200092, China
*
Author to whom correspondence should be addressed.
Submission received: 28 May 2021 / Revised: 13 June 2021 / Accepted: 15 June 2021 / Published: 17 June 2021
(This article belongs to the Special Issue Redesign Processes in the Age of the Fourth Industrial Revolution)

Abstract

:
Protonated g-C3N4 (pCN) formed by treating bulk g-C3N4 with an aqueous HCl solution was modified with D149 dye, i.e., 5-[[4[4-(2,2-diphenylethenyl) phenyl]-1,2,3,3a,4,8b-hexahydrocyclopent[b]indol-7-yl] methylene]-2-(3-ethyl-4-oxo-2-thioxo-5-thiazolidinylidene)-4-oxo-thiazolidin-2-ylidenerhodanine, for photocatalytic water splitting (using Pt as a co-catalyst). The D149/pCN-Pt composite showed a much higher rate (2138.2 µmol·h−1·g−1) of H2 production than pCN-Pt (657.0 µmol·h−1·g−1). Through relevant characterization, the significantly high activity of D149/pCN-Pt was linked to improved absorption of visible light, accelerated electron transfer, and more efficient separation of charge carriers. The presence of both D149 and Pt was found to be important for these factors. A mechanism was proposed.

1. Introduction

Photocatalysis is an important branch of catalysis. Compared with traditional thermal catalysis, photocatalysis can only work in a narrower range of reactions, but it usually operates under milder conditions, consumes no thermal energy, and thus is of low cost [1]. Research into photocatalysis originated from the discovery of photocatalytic water splitting on TiO2 electrode [2]. H2 in industry is generally produced via steam reforming of CH4 and the water-gas shift reaction. Without relying on fossil fuels, photocatalytic water splitting is promising for producing clean H2 [3,4]. Many photocatalysts, such as CdSe/ZnSe [5], Ag/TiO2 [6], NiO/TiO2 [7], TiO2/Bi2O3 [8], MoS2/CdS [9], ZnS/Cu3P [10], Co2P/ZnIn2S4 [11], Bi2WO6/ZnIn2S4 [12], and BaTaO2N/Pt [13], were used for photocatalytic H2 evolution under visible light.
Graphitic carbon nitride (g-C3N4) is a metal-free semiconductor photocatalyst with a suitable band gap (~2.7 eV) and high chemical stability [14,15,16]. However, pristine g-C3N4 exhibits weak visible light absorption. One of the prerequisites for photocatalysts to exhibit superior performance under visible light is sufficient absorption of visible light. Therefore, several approaches have been used to enhance the visible light absorption of g-C3N4. For instance, g-C3N4 has been modified by doping with metal or nonmetal ions [17,18,19,20], forming heterojunctions [21,22,23,24,25,26], and being photosensitized with dyes [27,28,29,30,31].
Dye sensitization is a simple and effective method to broaden the response range of materials under visible light because dyes can absorb visible light [32]. For a dye-sensitized photocatalyst, the dye molecules adsorbed on the surface of a semiconductor can be excited after absorbing visible light, thus injecting electrons into the conduction band of the semiconductor. MK2/TiO2 [33], Eosin Y/TiO2 [34], N749/rGO [35], GS12/TiO2 [36], PY-1/g-C3N4 [37], PY-2/g-C3N4 [37], ZnPy/g-C3N4 [38], chlorin e6/g-C3N4 [39], protoporphyrin/g-C3N4 [40], mTHPC/g-C3N4 [41], Ppa/g-C3N4 [42], N3/g-C3N4 [31], and N719/g-C3N4 [31] exhibited enhanced performance in H2 evolution. However, more studies are needed to enrich the family of dye-sensitized g-C3N4, to find more efficient catalysts for water splitting, and to understanding fundamental aspects of these photocatalytic systems.
D149 dye, namely, 5-[[4-[4-(2,2-diphenylethenyl)phenyl]-1,2,3,3a,4,8b-hexahydrocyclopent[b]indol-7-yl]methylene]-2-(3-ethyl-4-oxo-2-thioxo-5-thiazolidinylidene)-4-oxo-3-thiazolidineacetic acid, is a typical indoline dye (Figure S1) [43]. D149 dye has been found to exhibit remarkable performance in dye-sensitized solar cells (DSSCs), such as D149/TiO2 [44], D149/BiVO4 [45], D149/SnO2 [46], and D149/ZnO [47]. D149/BiVO4 with enhanced performance in photocatalytic degradation of methylene blue was also developed [48]. However, D149/g-C3N4 hybrid used for H2 production has not been reported.
Herein, D149/pCN-Pt was developed and found to be highly active for photocatalytic water splitting. The catalysts with and without Pt were systematically tested and characterized to understand the role of D149 and Pt. Previous studies using Pt as a co-catalyst demonstrated the presence of Pt particles on g-C3N4 by using TEM and EDX mapping [31,40,41,42]. The presence of Pt particles was found to enhance the absorption of visible light and photocatalytic activity [31,40,41,42]. However, the effect of Pt particles on transient photocurrent density, electronic impedance, and photoluminescence properties of the catalysts have not often been studied. Our current work provided clear evidence for the beneficial effect of both D149 and Pt particles for enhancing the photocatalytic activity.

2. Materials and Methods

2.1. Materials

Melamine, HCl solution (36.0–38.0 wt%), and ethyl alcohol (C2H6O, >99.7%) of analytical grade were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). 5-[[4[4-(2,2-diphenylethenyl) phenyl]-1,2,3,3a,4,8b-hexahydrocyclopent[b]indol-7-yl] methylene]-2-(3-ethyl-4-oxo-2-thioxo-5-thiazolidinylidene)-4-oxo-thiazolidin-2-ylidenerhodanine (D149 dye, 95%, LC) was purchased from Shanghai Yuanye Bio-Technology Co., Ltd. (Shanghai, China).

2.2. Preparation of Protonated g-C3N4 (pCN)

To synthesize bulk g-C3N4 (bCN), 10 g of melamine was heated at 550 °C for 2 h at a rate of 5 °C ·min−1 in a muffle. Then, 1 g bCN was put into 200 mL of HCl solution (1 M) for exfoliation. The slurry was magnetically stirred for 4 h. The sediment was collected by centrifugation for 16 min, washed with deionized water, and dried at 80 °C for 15 h [42]. The powders (pCN) were prepared for further use.

2.3. Preparation of D149/pCN

First, 0.5 g pCN was added into 25 mL ethanol solution, then the turbid liquid was magnetically stirred to make it mix evenly. Subsequently, 50 mg D149 dye was added into the slurry, and the suspension liquid was magnetically stirred rapidly overnight. Then, the sediment was collected by centrifuging and drying in an oven. The obtained powders (D149/pCN) were collected.

2.4. Characterization

XRD patterns were measured by Ultimate IV powder X-ray diffractometer (Rigaku, Tokyo, Japan) using Cu Kα radiation. FT-IR spectra were measured by Thermo Scientific Nicolet IS5. TEM and EDX mapping images were collected by a JEM-2100plus electron microscope (JEOL, Tokyo, Japan). N2 adsorption-desorption isotherms were used to measure Brunauer-Emmett-Teller (BET) specific surfaces by an ASAP2460 instrument (Micromeritics Instrument Corp., Norcross, GA USA). XPS data were recorded on an ESCALAB 250XI instrument (Thermo Fisher Scientific Instrument Corp., Now York, USA).
Ultraviolet-visible diffuse reflection spectra (UV-vis DRS) were recorded by UV-3600 (Shimazu, Kyoto, Japan) to evaluate the optical properties of catalysts. In order to collect plentiful samples containing Pt for other characterizations, the samples here were magnified. An amount of 100 mg D149/pCN was put into an aqueous solution (180 mL) containing 8.1 mg H2PtCl6 · 6H2O, and 20 mL triethanolamine (TEOA) was added as a sacrificial agent. The slurry was irradiated under a 300 W Xe lamp for 4 h and was stirred magnetically. Finally, D149/pCN-Pt was collected by centrifugation and drying. pCN-Pt was prepared with the same method described above.
Transient photocurrent density curves and EIS data were recorded by a CHI 760E electrochemical workstation (CH Instrument Company, Shanghai, China) under visible light (λ > 420 nm). The details of electrode fabrication were given elsewhere [40,41,42]. Briefly, 0.09 g sample power, 0.01 g polyvinylidene fluoride (PVDF), and 20 drops N-methylpyrrolidone were homogenously mixed and spread (1 × 1 cm2) on a 1 × 2 cm2 indium tin oxide (ITO) glass. Photoluminescence (PL) spectra were measured by Edinburgh FLS1000 (Edinburgh Instruments Livingston, Edinburgh, UK) under 300 nm excitation at room temperature.

2.5. Photocatalytic Activity

Photocatalytic H2 production was tested under visible light (300 W Xe lamp, λ > 420 nm). The light intensity was 120 mW cm−2. The distance from the reaction instrument to the light source was 4 cm.
The procedure for testing H2 evolution yields was as follows [40,41,42]: 10 mg samples were put into an aqueous solution (18 mL) containing 0.81 mg H2PtCl6 · 6H2O, and then 2 mL of triethanolamine (TEOA) was added as a sacrificial agent. The sealed quartz reactor was subjected to ultrasonic treatment for 10 min. Pure N2 was blown into the reactor to remove the remaining air. Afterwards, the system was treated by ultrasound for 10 min. Eventually, the photocatalytic reaction system (PerfectLight, Labsolar-IIIAG, Beijing, China) was irradiated by a light source (PerfectLight, PLS-SXE300C, Beijing, China) with a filter (λ > 420 nm), and operated at 12 °C. The production of hydrogen was analyzed by an online gas chromatograph (GC9200, Tian Mei, shanghai, China), with the help of a circulating pump. The schematic hydrogen production system is shown in Figure S2.

3. Results

3.1. Regular Characterization

Figure 1 shows XRD patterns. pCN shows peaks at 13.1° and 27.3°, corresponding to the (100) and (002) planes of g-C3N4 (JCPDS No.87-1526), respectively [49]. D149 dye shows a few distinct peaks. D149/pCN shows the peaks of pCN and some tiny peaks corresponding to D149 dye.
Figure S3 shows FT-IR spectra. The cuspidal peak at 807 cm−1 represents the tri-s-triazine structure of g-C3N4 [50]. The peaks at 1250, 1330, 1413, 1575, and 1645 cm−1 correspond to the stretching vibration modes of C–N and C=N bonds of the heptazine heterocyclic ring (C6N7) [51]. A broad peak located from 2800 to 3600 cm−1 signifies the existence of abundant –NHx [52]. As for D149 dye, the peak at 1128 cm−1 corresponds to a C–O bond. The peaks at 1240, 1547, and 1697 cm−1 can be classified with N–C=, C=C, and C=O bonds, respectively [53]. The D149/pCN pattern is highly similar to the pCN pattern. The presence of D149 on pCN can be seen from the differential spectrum by subtracting the spectrum of pCN from D149/pCN (Figure S4). There, the positive peaks ranging from 700~1200 nm labeled by a red dotted line are consistent with the peaks in D149 (Figure S3), whereas the negative peaks between 1150 and 1900 nm are due to the decrease in pCN content after loading D149.
Figure S5 shows TEM images of pCN and D149/pCN. g-C3N4 treated with hydrochloric acid has a porous ultrathin structure (Figure S5A). Nonetheless, D149 dye loading can not be observed on the surface of pCN (Figure S5B), which is because that TEM is unable to detect organic dye molecules. STEM-EDX mapping images certify the existence of C, N, O, and S elements, indicating that D149 dye has been loaded on pCN (Figure 2).
Figure S6 shows N2 sorption data. The N2 adsorption-desorption isotherms of pCN and D149/pCN can be classified as a type IV hysteresis loop (Figure S6A), indicating that both pCN and D149/pCN have mesoporous structure. D149/pCN has a lower BET surface area (14.7 m²·g−1) than pCN (15.3 m²·g−1), probably because the covering of pCN by D149 could block some pores of its surface. The mean pore diameter of D149/pCN is 23.5 nm, less than that of pCN (37.8 nm). The total pore volumes of D149/pCN and pCN are 0.09 and 0.10 cm³·g−1, respectively (Table S1).
Figure 3 and Figure S7 show XPS spectra. XPS survey spectra show that pristine pCN mainly consists of C, N, and O elements, and D149/pCN has an extra S element (Figure S7). Figure 3A shows the C 1s spectra of pCN and D149/pCN. The peaks at 284.40 eV and 287.91 eV are C–C and N–C=N bonds, respectively [54]. In addition, there are three main peaks in N 1s spectra (Figure 3B). The peak at 398.42 eV in the spectra of the two samples represents C–N=C bonds and the two smaller peaks at 399.80 and 400.90 eV can be classified as N–(C)3 and C–N–H, respectively [55]. Moreover, O 1s spectra of pCN and D149/pCN are shown in Figure 3C. Three main peaks at 531.40, 532.00, and 533.30 eV can be observed in pCN, assigned to O–C=O, N–C–O, and C=O, respectively [56,57]. However, these three characteristic peaks of D149/pCN have slight blue shift, pointing to 530.70, 531.90, and 533.20 eV, respectively. This result may be attributed to the interaction between D149 dye and pCN. The position of S 2p peak of D149/pCN is located at 161.40 and 163.60 eV, corresponding to S 2p3/2 and S 2p1/2, respectively [58,59]. In addition, the peak at 164.90 eV corresponds to C=S bond [60], further confirming the presence of D149 dye (Figure 3D). Overall, the XPS data confirm that D149 dye has been loaded on pCN successfully.

3.2. Photocatalytic Performance

Visible-light-driven H2 production was used for assessing photocatalytic performance. The in situ formed Pt was used as a co-catalyst. TEOA was used as an electron donor. As can be seen in Figure 4A, both pCN and D149 dyes have low activity of H2 production. When loading D149 dye onto pCN, the activity of H2 evolution increases slightly, indicating that D149 dye plays a role in the photocatalytic system. In addition, there has been a significant improvement after loading Pt onto the three catalysts, confirming the critical impact of Pt in the system of photocatalytic hydrogen production. The H2 evolution yield in 4 h under visible light of D149/pCN-Pt is much higher than that of pCN-Pt. The average H2 evolution rate of D149/pCN-Pt reaches 2138.2 µmol·h−1·g−1 (Figure S8). The D149/pCN-Pt photocatalyst exhibits good recyclability (Figure 4B).

3.3. Reasons for Enhanced Activity

Figure S9 shows the XRD patterns of samples after reaction for 4 h. The pCN-Pt and D149/pCN-Pt samples here were magnified (see Section 2.4). There are characteristic peaks of Pt at 38.1° and 44.4°, confirming the existence of Pt particles (JCPDS 04-0802) [61]. In addition, the peaks corresponding to D149 dye in D149/pCN-Pt, ranging from 10° to 25°, become lower than those in D149/pCN. This is may be because a small part of the D149 dye is lost (possibly leached) after loading Pt. Figure 5 certifies the presence of D149 dye and Pt after loading Pt.
Figure 5 and Figures S10 and S11 demonstrate the presence of Pt particles on D149/pCN-Pt (collected after the recyclability test in Figure 4B) and pCN-Pt (collected after the activity test in Figure 4A), revealing that Pt was successfully loaded on the catalysts.
Figure 6 illustrates the UV-vis DRS. The absorption edge of pCN is located at 470 nm. Loading D149 onto pCN could distinctly boost the intensity of visible light absorption. Furthermore, the catalysts with Pt as a co-catalyst show stronger capacity of visible light absorption compared with those samples without Pt [26,42]. Additionally, there is an uplift from 450~700 nm in absorption line of D149/pCN. When Pt is loaded onto D149/pCN, the uplift decreases. This phenomenon is consistent with XRD results, showing the loss of some D149 (Figure S9).
Figure 7 displays the transient photocurrent density curves of the catalysts. The current density shows the order of pCN < D149/pCN < pCN-Pt < D149/pCN-Pt. The result illustrates that D149/pCN could produce more photogenerated electrons compared with original pCN. Moreover, the addition of Pt can make more photogenerated electrons transfer to the surface of the catalysts for redox reaction [62,63].
Figure 8 shows the EIS data. Smaller relative radius of the arcs can accelerate electrons transfer and make better separation efficiency of the photogenerated e-h+ pairs, thus leading to a better photocatalytic performance [64]. In Figure 8, D149/pCN has smaller arc radius compared with pCN, showing that loading D149 dye onto pCN reduces the resistance to electron transfer and prevents the recombination of e-h+ pairs. Adding Pt onto the two samples can significantly reduce the arc radius in EIS spectra, indicating that pCN-Pt and D149/pCN-Pt have better capability of charge transfer and separation efficiency of the photogenerated e-h+ pairs [62,63].
Figure 9 shows the photoluminescence (PL) data. Apparently, the PL intensity of D149/pCN is lower than that of pure pCN, indicating that the recombination of e-h+ pairs is confined due to the formation of D149/pCN. Furthermore, pCN-Pt and D149/pCN-Pt have much lower emission peaks compared with pCN and D149/pCN, which may be because Pt can not only restrict the recombination of e-h+ pairs, but also accelerate charge transfer [65,66,67].

3.4. Photocatalytic Reaction Mechanism

Based upon the above characterizations of the catalyst, a possible photocatalytic mechanism is proposed (Figure 10). The bandgap values of pCN and D149 dye can be calculated through the following equation [68]:
αhν = A (Eg)n/2
where α, h, ν, A, and Eg represent the absorption coefficient, Plancks’ constant, light frequency, proportionality constant, and bandgap energy, respectively. The n value is equal to 1 because g-C3N4 is a direct bandgap semiconductor [69]. The Eg of pCN and D149 are 2.50 and 1.58 eV, respectively (Figure S12). Obviously, the bandgap width of D149 is narrower than that of pCN. Moreover, the EVB of pCN and D149 dye obtained by XPS valence band spectra are 1.80 and 0.25 V, respectively (Figure S13). Consequently, the ECB values of pCN and D149 are −0.70 and −1.33 V, respectively.
When illuminated with visible light, the catalyst generates electrons (e) on CB and holes (h+) on VB. Since electrons are inclined to migrate from negative to positive potential, the e at the CB of D149 dye is more likely to transfer to Pt particles rather than transfer to the CB of pCN. Therefore, the e at the CB of pCN recombined with the h+ at the VB of D149 quickly due to the synergistic effect of internal electric field and electrostatic interaction [70]. Meanwhile, the h+ and e remain on the VB of pCN and CB of D149, respectively, to oxidize TEOA to TEOA+ and reduce H+ to H2 [71,72]. A Z-scheme transfer mode is feasible for this photocatalytic system. This envisaged Z-scheme mechanism is in accordance with the mechanism proposed in previous photocatalytic works, involving PDIP/g-C3N4 [73], PTCDI/g-C3N4 [74], and O-CN/g-C3N4 [75].

4. Discussion

Many studies have shown that Pt as a co-catalyst can effectively improve the performance of photocatalytic hydrogen production [27,76,77]. Some studies have also carried out the subsequent characterizations on samples containing Pt. For example, Mei et al. prepared TCPP/Pt/g-C3N4 and characterized the catalyst by PL, suggesting that the improved hydrogen production is because Pt inhibits the recombination of the photogenerated e-h+ pairs [66]. Ou et al. characterized Pt/g-C3N4 with photocurrent, EIS, and PL, and revealed that introducing Pt onto g-C3N4 leads to better charge transfer and higher efficiency in e-h+ pairs separation, thus to an improved H2 evolution rate [62]. However, in many literatures, catalysts containing Pt have not been characterized in detail, and the reason why Pt improved the activity of photocatalytic H2 evolution has not been elaborated [31,39,40,41,42]. For instance, N3/pCN-Pt and N719/pCN-Pt were characterized by TEM and STEM-EDX mapping, indicating that Pt particles were loaded [31]. N3/pCN-Pt and N719/pCN-Pt were also characterized by UV-vis DRS, revealing that the absorbance of catalysts remained well after photocatalytic reaction [31]. However, the reason why Pt can improve the performance of photocatalytic H2 production has not been explored. In this work, XRD and TEM were used to confirm that Pt was loaded onto the catalysts (Figures S9–S11, and Figure 4). UV-vis DRS, photocurrent density, EIS, and PL characterizations of catalysts with Pt were also performed (Figure 6, Figure 7, Figure 8 and Figure 9).
The photocatalytic activity follows the order of pCN < D149/pCN < pCN-Pt < D149/pCN-Pt (Figure 4A). Here, the transient photocurrent density increase in the order of pCN < D149/pCN < pCN-Pt < D149/pCN-Pt (Figure 7). The relative radius of the arcs in EIS spectra decrease in the order of pCN > D149/pCN > pCN-Pt > D149/pCN-Pt (Figure 8). The sequence of PL intensity of the four samples from high to low is pCN > D149/pCN > pCN-Pt > D149/pCN-Pt (Figure 9). Based upon the above characterizations, it can be deduced that loading D149 dye on the pCN can improve the activity of photocatalytic hydrogen production to some extent. The reason may be attributed to increased absorption of visible light, better capability of charge transfer, and separation efficiency of the photogenerated e-h+ pairs. However, it seems that the Pt can be a key factor to enhance the catalytic activity, which may result from a superior absorption of visible light and a faster electron transfer from pCN or D149 dye to Pt NPs and consequently contribute to higher activity of H2 production under visible light [78].
Table S2 shows the comparison of the photocatalytic performance of H2 production of g-C3N4 sensitized by different dyes [27,31,37,39,40,41,42,79,80,81]. It is obvious that the range of H2 evolution rates is from 39.0 to 6525.0 µmol·h−1·g−1. This study obtained the H2 evolution rate of 2138.2 µmol·h−1·g−1, which is within a reasonable range. In particular, under similar conditions, the average HER of D149/pCN-Pt is higher than that of Ce6/pCN (1275.6 µmol·h−1·g−1) [39], Pp/pCN (1153.8 µmol·h−1·g−1) [40], mTHPC/pCN (1041.4 µmol·h−1·g−1) [41], Ppa/pCN (1093.0 µmol·h−1·g−1) [42], and N3/pCN (1490.7 µmol·h−1·g−1) [31], although slightly lower than N719/pCN (2500.1 µmol·h−1·g−1) [31]. In addition, the average HER of D149/pCN-Pt is also much higher than that of TiOF2/g-C3N4 (343.8 µmol·h−1·g−1) [26], because dye-sensitized g-C3N4 has a greater ability to absorb visible light.
It should be mentioned that if all the H2PtCl6 can be reduced in situ upon visible-light irradiation and be deposited onto the catalyst surface, then the “theoretical” Pt loading should be 3 wt.%. However, the Pt content of the magnified D149/pCN-Pt sample contains 0.033 wt.% Pt, as analyzed by ICP-OES. This information thus shows that H2PtCl6 in the slurry cannot be completely reduced under the adopted reaction condition. Even this is the case, the catalyst still shows obviously enhanced photocatalytic activity, underscoring the promoting role of Pt deposited onto the catalyst. In-depth analysis of the influence of Pt content on the activity may be conducted in the future.

5. Conclusions

In this study, we developed a novel dye-sensitized photocatalyst, and Pt was used as a co-catalyst (D149/pCN-Pt). In our study, D149 is a modena indoline dye; pCN stands for protonated g-C3N4. The photocatalyst shows better photocatalytic performance for H2 production under visible light compared with pure D149 dye and pCN. Furthermore, we also conducted a control experiment to investigate the difference in hydrogen production of the catalysts with and without Pt. In particular, the addition of Pt nanoparticles as a co-catalyst can significantly increase the hydrogen evolution rate. Through a series of characterizations of catalysts with Pt, the possible reasons for improved performance of photocatalytic hydrogen production are ascribed to the observation that Pt has enhanced absorbance of visible light, and Pt particles can accelerate charge transfer and separation of photogenerated e-h+ pairs. This study revealed the possibility of achieving improved performance of photocatalytic hydrogen production by indoline dye-sensitized g-C3N4 and discussed the reasons for enhancing the H2 evolution activity by using Pt as a co-catalyst.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/pr9061055/s1, Figure S1. Chemical structures of D149 dye. Figure S2. The schematic hydrogen production system. Figure S3. FT-IR spectra of pCN, D149, and D149/pCN. Figure S4. The differential spectra obtained by subtracting the spectrum of pCN from D149/pCN. Figure S5. TEM images of pCN (A) and D149/pCN (B). Figure S6. (A) N2 adsorption-desorption isotherms and (B) pore diameter distribution curves from BJH adsorption curves of pCN and D149/pCN. Figure S7. XPS survey spectra of pCN and D149/pCN. Figure S8. average H2 evolution rates of the three samples with and without Pt in 4 h under visible light (λ > 420 nm). Figure S9. XRD patterns of pCN, D149, D149/pCN, pCN-Pt, and D149/pCN-Pt. Figure S10. TEM images of pCN-Pt collected after reaction for one time (A) and D149/pCN-Pt collected after the recyclability test (B). Figure S11. A typical STEM image of pCN-Pt collected after reaction for one time (A) and EDX mapping images of C (B), N (C), O (D), and Pt (E). Figure S12. Band gap curves of pCN and D149. Figure S13. EVB values of pCN (A) and D149 dye (B). Table S1. BET surface area, mean pore diameter, and total pore volume of two samples. Table S2. Comparison of H2 evolution rate of dye-sensitized g-C3N4 under visible light irradiation.

Author Contributions

Investigation, most of the experimental work, data processing, original draft preparation, Y.C.; investigation, Y.L.; investigation, project administration, supervision, manuscript review and editing, Z.M. All authors have read and agreed to the published version of the manuscript.

Funding

Shanghai Institute of Pollution Control and Ecological Security.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from the authors.

Acknowledgments

Shanghai Institute of Pollution Control and Ecological Security.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hoffmann, M.R.; Martin, S.T.; Choi, W.Y.; Bahnemann, D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  2. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  3. Maeda, K.; Teramura, K.; Lu, D.; Takata, T.; Saito, N.; Inoue, Y.; Domen, K. Photocatalyst releasing hydrogen from water. Nature 2006, 440, 295. [Google Scholar] [CrossRef] [PubMed]
  4. Takata, T.; Pan, C.S.; Domen, K. Recent progress in oxynitride photocatalysts for visible-light-driven water splitting. Sci. Technol. Adv. Mater. 2015, 16, 033506. [Google Scholar] [CrossRef] [PubMed]
  5. Feng, K.T.; Xue, W.H.; Hu, X.Y.; Fan, J.; Liu, E.Z. Z-scheme CdSe/ZnSe heterojunction for efficient photocatalytic hydrogen evolution. Colloids Surf. A Physicochem. Eng. Asp. 2021, 622, 126633. [Google Scholar] [CrossRef]
  6. Barakat, N.A.M.; Erfan, N.A.; Mohammed, A.A.; Mohamed, S.E.I. Ag-decorated TiO2 nanofibers as Arrhenius equation-incompatible and effective photocatalyst for water splitting under visible light irradiation. Colloids Surf. A Physicochem. Eng. Asp. 2020, 604, 125307. [Google Scholar] [CrossRef]
  7. Zhao, X.Y.; Xie, W.Y.; Deng, Z.B.; Wang, G.; Cao, A.H.; Chen, H.M.; Yang, B.; Wang, Z.; Su, X.T.; Yang, C. Salt templated synthesis of NiO/TiO2 supported carbon nanosheets for photocatalytic hydrogen production. Colloids Surf. A Physicochem. Eng. Asp. 2020, 587, 124365. [Google Scholar] [CrossRef]
  8. Xu, D.F.; Hai, Y.; Zhang, X.C.; Zhang, S.Y.; He, R.A. Bi2O3 cocatalyst improving photocatalytic hydrogen evolution performance of TiO2. Appl. Surf. Sci. 2017, 400, 530–536. [Google Scholar] [CrossRef]
  9. Zhuge, K.X.; Chen, Z.J.; Yang, Y.Q.; Wang, J.; Shi, Y.Y.; Li, Z.Q. In-suit photodeposition of MoS2 onto CdS quantum dots for efficient photocatalytic H2 evolution. Appl. Surf. Sci. 2021, 539, 148234. [Google Scholar] [CrossRef]
  10. Rameshbabu, R.; Ravi, P.; Pecchi, G.; Delgado, E.J.; Mangalaraja, R.V.; Sathish, M. Black trumpet mushroom-like ZnS incorporated with Cu3P: Noble metal free photocatalyst for superior photocatalytic H2 production. J. Colloid Interface Sci. 2021, 590, 82–93. [Google Scholar] [CrossRef]
  11. Zhang, Q.; Wang, X.H.; Zhang, J.H.; Li, L.F.; Gu, H.J.; Dai, W.L. Hierarchical fabrication of hollow Co2P nanocages coated with ZnIn2S4 thin layer: Highly efficient noble-metal-free photocatalyst for hydrogen evolution. J. Colloid Interface Sci. 2021, 590, 632–640. [Google Scholar] [CrossRef]
  12. Chachvalvutikul, A.; Luangwanta, T.; Pattisson, S.; Hutchings, G.J.; Kaowphong, S. Enhanced photocatalytic degradation of organic pollutants and hydrogen production by a visible light-responsive Bi2WO6/ZnIn2S4 heterojunction. Appl. Surf. Sci. 2021, 544, 148885. [Google Scholar] [CrossRef]
  13. Wang, Z.; Luo, Y.; Hisatomi, T.; Vequizo, J.J.M.; Suzuki, S.; Chen, S.S.; Nakabayashi, M.; Lin, L.H.; Pan, Z.H.; Kariya, N.; et al. Sequential cocatalyst decoration on BaTaO2N towards highly-active Z-scheme water splitting. Nat. Commun. 2021, 12, 1005. [Google Scholar] [CrossRef]
  14. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic carbon nitride (g-C3N4)-based photocatalysts for artificial photosynthesis and environmental remediation: Are we a step closer to achieving sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  15. Fu, J.W.; Yu, J.G.; Jiang, C.J.; Cheng, B. g-C3N4-based heterostructured photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  16. Sudhaik, A.; Raizada, P.; Shandilya, P.; Jeong, D.-Y.; Lim, J.-H.; Singh, P. Review on fabrication of graphitic carbon nitride based efficient nanocomposites for photodegradation of aqueous phase organic pollutants. J. Ind. Eng. Chem. 2018, 67, 28–51. [Google Scholar] [CrossRef]
  17. Hu, S.Z.; Ma, L.; You, J.G.; Li, F.Y.; Fan, Z.P.; Lu, G.; Liu, D.; Gui, J.Z. Enhanced visible light photocatalytic performance of g-C3N4 photocatalysts co-doped with iron and phosphorus. Appl. Surf. Sci. 2014, 311, 164–171. [Google Scholar] [CrossRef]
  18. Lei, L.; Wang, W.J.; Wang, C.; Zhang, M.C.; Zhong, Q.; Fan, H.Q. In situ growth of boron doped g-C3N4 on carbon fiber cloth as a recycled flexible film-photocatalyst. Ceram. Int. 2021, 47, 1258–1267. [Google Scholar] [CrossRef]
  19. Wang, D.B.; Huang, X.Q.; Huang, Y.; Yu, X.; Lei, Y.; Dong, X.Y.; Su, Z.L. Self-assembly synthesis of petal-like Cl-doped g-C3N4 nanosheets with tunable band structure for enhanced photocatalytic activity. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125780. [Google Scholar] [CrossRef]
  20. Mohtasham, H.; Gholipour, B.; Rostamnia, S.; Ghiasi-Moaser, A.; Farajzadeh, M.; Nouruzi, N.; Jang, H.W.; Varma, R.S.; Shokouhimehr, M. Hydrothermally exfoliated P-doped g-C3N4 decorated with gold nanorods for highly efficient reduction of 4-nitrophenol. Colloids Surf. A Physicochem. Eng. Asp. 2021, 614, 126187. [Google Scholar] [CrossRef]
  21. Ai, C.L.; Wu, S.C.; Li, L.Y.; Lei, Y.J.; Shao, X.W. Novel magnetically separable γ-Fe2O3/Ag/AgCl/g-C3N4 composite for enhanced disinfection under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2019, 583, 123981. [Google Scholar] [CrossRef]
  22. Sun, X.Y.; Zhang, F.J.; Kong, C. Porous g-C3N4/WO3 photocatalyst prepared by simple calcination for efficient hydrogen generation under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2020, 594, 124653. [Google Scholar] [CrossRef]
  23. Bi, X.J.; Yu, S.R.; Liu, E.Y.; Liu, L.; Zhang, K.; Zang, J.; Zhao, Y. Construction of g-C3N4/TiO2 nanotube arrays Z-scheme heterojunction to improve visible light catalytic activity. Colloids Surf. A Physicochem. Eng. Asp. 2020, 603, 125193. [Google Scholar] [CrossRef]
  24. Jiang, K.; Iqbal, W.; Yang, B.; Rauf, M.; Ali, I.; Lu, X.Y.; Mao, Y.P. Noble metal-free NiCo2S4/CN sheet-on-sheet heterostructure for highly efficient visible-light-driven photocatalytic hydrogen evolution. J. Alloys Compd. 2021, 853, 157284. [Google Scholar] [CrossRef]
  25. Li, D.G.; Huang, J.X.; Li, R.B.; Chen, P.; Chen, D.N.; Cai, M.X.; Liu, H.J.; Feng, Y.P.; Lv, W.Y.; Liu, G.G. Synthesis of a carbon dots modified g-C3N4/SnO2 Z-scheme photocatalyst with superior photocatalytic activity for PPCPs degradation under visible light irradiation. J. Harzard. Mater. 2021, 401, 123257. [Google Scholar] [CrossRef]
  26. Liu, Y.F.; Ma, Z. TiOF2/g-C3N4 composite for visible-light driven photocatalysis. Colloids Surf. A Physicochem. Eng. Asp. 2021, 681, 126471. [Google Scholar] [CrossRef]
  27. Wang, P.; Guan, Z.J.; Li, Q.Y.; Yang, J.J. Efficient visible-light-driven photocatalytic hydrogen production from water by using Eosin Y-sensitized novel g-C3N4/Pt/GO composites. J. Mater. Sci. 2018, 53, 774–786. [Google Scholar] [CrossRef]
  28. Yang, Y.J.; Sun, B.W.; Qian, D.J.; Chen, M. Fabrication of multiporphyrin@g-C3N4 nanocomposites via Pd(II)-directed layer-by-layer assembly for enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2019, 478, 1027–1036. [Google Scholar] [CrossRef]
  29. Zhang, C.; Liu, J.D.; Liu, X.L.; Xu, S.A. D-π-A-type triphenylamine dye covalent-functionalized g-C3N4 for highly efficient photocatalytic hydrogen evolution. Catal. Sci. Technol. 2020, 10, 1609–1618. [Google Scholar] [CrossRef]
  30. Lu, M.; Sun, Z.; Zhang, Y.Z.; Liang, Q.; Zhou, M.; Xu, S.; Li, Z.Y. Construction of cobalt phthalocyanine sensitized SnIn4S8/g-C3N4 composites with enhanced photocatalytic degradation and hydrogen production performance. Synth. Met. 2020, 268, 116480. [Google Scholar] [CrossRef]
  31. Chen, Y.H.; Liu, Y.F.; Ma, Z. Graphitic C3N4 modified by Ru(II)-based dyes for photocatalytic H2 evolution. Colloids Surf. A Physicochem. Eng. Asp. 2021, 614, 126119. [Google Scholar] [CrossRef]
  32. Liu, X.; Li, Y.X.; Peng, S.Q.; Lai, H. Progress in visible-light photocatalytic hydrogen production by dye sensitization. Acta Phys.-Chim. Sin. 2015, 31, 612–626. [Google Scholar]
  33. Kumari, A.; Mondal, I.; Pal, U. A simple carbazole based sensitizer attached to a Nafion-coated-TiO2 photocatalyst: The impact of controlling parameters towards visible light driven H2 production. New J. Chem. 2015, 39, 713–720. [Google Scholar] [CrossRef]
  34. Shi, J.W.; Guan, X.J.; Zhou, Z.H.; Liu, H.P.; Guo, L.J. Eosin Y-sensitized nanosheet-stacked hollow-sphere TiO2 for efficient photocatalytic H2 production under visible-light irradiation. J. Nanopart. Res. 2015, 17, 252. [Google Scholar] [CrossRef]
  35. Huang, J.; Wang, D.D.; Yue, Z.K.; Li, X.; Chu, D.M.; Yang, P. Ruthenium dye N749 covalently functionalized reduced graphene oxide: A novel photocatalyst for visible light H2 evolution. J. Phy. Chem. C 2015, 119, 27892–27899. [Google Scholar] [CrossRef]
  36. Swetha, T.; Mondal, I.; Bhanuprakash, K.; Pal, U.; Singh, S.P. First study on phosphonite-coordinated ruthenium sensitizers for efficient photocatalytic hydrogen evolution. ACS Appl. Mater. Interfaces 2015, 7, 19635–19642. [Google Scholar] [CrossRef]
  37. Yu, F.T.; Wang, Z.Q.; Zhang, S.C.; Yun, K.; Ye, H.N.; Gong, X.Q.; Hua, J.L.; Tian, H. N-Annulated perylene-based organic dyes sensitized graphitic carbon nitride to form an amide bond for efficient photocatalytic hydrogen production under visible-light irradiation. Appl. Catal. B Environ. 2018, 237, 32–42. [Google Scholar] [CrossRef]
  38. Zheng, Y.; Wang, J.M.; Zhang, J.; Peng, T.Y.; Li, R.J. Syntheses of asymmetric zinc porphyrins bearing different pseudo-pyridine substituents and their photosensitization for visible-light-driven H2 production activity. Dalton Trans. 2017, 46, 8219–8228. [Google Scholar] [CrossRef] [PubMed]
  39. Liu, Y.F.; He, M.F.; Guo, R.; Fang, Z.R.; Kang, S.F.; Ma, Z.; Dong, M.D.; Wang, W.L.; Cui, L.F. Ultrastable metal-free near-infrared-driven photocatalysts for H2 production based on protonated 2D g-C3N4 sensitized with Chlorin e6. Appl. Catal. B Environ. 2020, 260, 118137. [Google Scholar] [CrossRef]
  40. Liu, Y.F.; Kang, S.F.; Cui, L.F.; Ma, Z. Boosting near-infrared-driven photocatalytic H2 evolution using protoporphyrin-sensitized g-C3N4. J. Photochem. Photobiol. A 2020, 396, 112517. [Google Scholar] [CrossRef]
  41. Liu, Y.F.; Ma, Z. g-C3N4 modified by meso-tetrahydroxyphenylchlorin for photocatalytic hydrogen evolution under visible/near-infrared light. Front. Chem. 2020, 8, 605343. [Google Scholar] [CrossRef]
  42. Liu, Y.F.; Ma, Z. g-C3N4 modified by pyropheophorbide-a for photocatalytic H2 evolution. Colloids Surf. A Physicochem. Eng. Asp. 2021, 615, 126128. [Google Scholar] [CrossRef]
  43. Falgenhauer, J.; Fiehler, F.; Richter, C.; Rudolph, M.; Schlettwein, D. Consequences of changes in the ZnO trap distribution on the performance of dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2017, 19, 16159–16168. [Google Scholar] [CrossRef] [PubMed]
  44. Lin, J.J.; Heo, Y.U.; Nattestad, A.; Shahabuddin, M.; Yamauchi, Y.; Kim, J.H. N719-and D149-sensitized 3D hierarchical rutile TiO2 solar cells-a comparative study. Phys. Chem. Chem. Phys. 2015, 17, 7208–7213. [Google Scholar] [CrossRef] [PubMed]
  45. Ordon, K.; Coste, S.; Noel, O.; El-Ghayoury, A.; Ayadi, A.; Kassiba, A.; Makowska-Janusik, M. Investigations of the charge transfer phenomenon at the hybrid dye/BiVO4 interface under visible radiation. RSC Adv. 2019, 9, 30698–30706. [Google Scholar] [CrossRef] [Green Version]
  46. Ranasinghe, C.S.K.; Jayaweera, E.N.; Kumara, G.R.A.; Rajapakse, R.M.G.; Onwona-Agyeman, B.; Perera, A.G.U.; Tennakone, K. Tin oxide based dye-sensitized solid-state solar cells: Surface passivation for suppression of recombination. Mater. Sci. Semicond. Process. 2015, 40, 890–895. [Google Scholar] [CrossRef]
  47. Tefashe, U.M.; Loewenstein, T.; Miura, H.; Schlettwein, D.; Wittstock, G. Scanning electrochemical microscope studies of dye regeneration in indoline (D149)-sensitized ZnO photoelectrochemical cells. J. Electroanal. Chem. 2010, 650, 24–30. [Google Scholar] [CrossRef]
  48. Ordon, K.; Merupo, V.I.; Coste, S.; Noel, O.; Errien, N.; Makowska-Janusik, M.; Kassiba, A. Charge-transfer peculiarities in mesoporous BiVO4 surfaces with anchored indoline dyes. Appl. Nanosci. 2018, 8, 1895–1905. [Google Scholar] [CrossRef]
  49. Qin, Y.Y.; Lu, J.; Meng, F.Y.; Lin, X.Y.; Feng, Y.H.; Yan, Y.S.; Meng, M.J. Rationally constructing of a novel 2D/2D WO3/Pt/g-C3N4 Schottky-Ohmic junction towards efficient visible-light-driven photocatalytic hydrogen evolution and mechanism insight. J. Colloid Interface Sci. 2021, 586, 576–587. [Google Scholar] [CrossRef]
  50. Zhang, X.D.; Yan, J.; Lee, L.Y.S. Highly promoted hydrogen production enabled by interfacial P-N chemical bonds in copper phosphosulfide Z-scheme composite. Appl. Catal. B Environ. 2021, 283, 119624. [Google Scholar] [CrossRef]
  51. Liu, Q.; Chen, T.X.; Guo, Y.R.; Zhang, Z.G.; Fang, X.M. Ultrathin g-C3N4 nanosheets coupled with carbon nanodots as 2D/OD composites for efficient photocatalytic H2evolution. Appl. Catal. B Environ. 2016, 193, 248–258. [Google Scholar] [CrossRef]
  52. Ma, J.L.; Jin, D.N.; Li, Y.C.; Xiao, D.Q.; Jiao, G.J.; Liu, Q.; Guo, Y.Z.; Xiao, L.P.; Chen, X.H.; Li, X.Z.; et al. Photocatalytic conversion of biomass-based monosaccharides to lactic acid by ultrathin porous oxygen doped carbon nitride. Appl. Catal. B Environ. 2021, 283, 119520. [Google Scholar] [CrossRef]
  53. Al-Attafi, K.; Jawdat, F.H.; Qutaish, H.; Hayes, P.; Al-Keisy, A.; Shim, K.; Yamauchi, Y.; Dou, S.X.; Nattestad, A.; Kim, J.H. Cubic aggregates of Zn2SnO4 nanoparticles and their application in dye-sensitized solar cells. Nano Energy 2019, 57, 202–213. [Google Scholar] [CrossRef]
  54. Nguyen, V.K.; Nguyen Thi, V.N.; Tran, H.H.; Tran Thi, T.P.; Truong, T.T.; Vo, V. A facile synthesis of g-C3N4/BaTiO3 photocatalyst with enhanced activity for degradation of methylene blue under visible light. Bull. Mater. Sci. 2021, 44, 28. [Google Scholar] [CrossRef]
  55. Shi, H.F.; Zhao, T.T.; Wang, J.B.; Wang, Y.T.; Chen, Z.; Liu, B.L.; Ji, H.F.; Wang, W.D.; Zhang, G.L.; Li, Y.G. Fabrication of g-C3N4/PW12/TiO2 composite with significantly enhanced photocatalytic performance under visible light. J. Alloys Compd. 2021, 860, 157924. [Google Scholar] [CrossRef]
  56. Teng, Z.Y.; Lv, H.Y.; Wang, C.Y.; Xue, H.G.; Pang, H.; Wang, G.X. Bandgap engineering of ultrathin graphene-like carbon nitride nanosheets with controllable oxygenous functionalization. Carbon 2017, 113, 63–75. [Google Scholar] [CrossRef]
  57. Zhang, G.G.; Lan, Z.A.; Wang, X.C. Surface engineering of graphitic carbon nitride polymers with cocatalysts for photocatalytic overall water splitting. Chem. Sci. 2017, 8, 5261–5274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Chen, Y.L.; Su, F.Y.; Xie, H.Q.; Wang, R.P.; Ding, C.H.; Huang, J.D.; Xu, Y.X.; Ye, L.Q. One-step construction of S-scheme heterojunctions of N-doped MoS2 and S-doped g-C3N4 for enhanced photocatalytic hydrogen evolution. Chem. Eng. J. 2021, 404, 126498. [Google Scholar] [CrossRef]
  59. Zhang, Y.K.; Jin, Z.L. Synergistic enhancement of hydrogen production by ZIF-67 (Co) derived Mo–Co–S modified g-C3N4/rGO photocatalyst. Catal. Lett. 2018, 149, 34–48. [Google Scholar] [CrossRef]
  60. Wang, W.C.; Tao, Y.; Du, L.L.; Wei, Z.; Yan, Z.P.; Chan, W.K.; Lian, Z.C.; Zhu, R.X.; Phillips, D.L.; Li, G.S. Femtosecond time-resolved spectroscopic observation of long-lived charge separation in bimetallic sulfide/g-C3N4 for boosting photocatalytic H2 evolution. Appl. Catal. B Environ. 2021, 282, 119568. [Google Scholar] [CrossRef]
  61. Cao, S.W.; Jiang, J.; Zhu, B.C.; Yu, J.G. Shape-dependent photocatalytic hydrogen evolution activity over a Pt nanoparticle coupled g-C3N4 photocatalyst. Phys. Chem. Chem. Phys. 2016, 18, 19457–19463. [Google Scholar] [CrossRef]
  62. Ou, M.; Wan, S.P.; Zhong, Q.; Zhang, S.L.; Wang, Y.A. Single Pt atoms deposition on g-C3N4 nanosheets for photocatalytic H2 evolution or NO oxidation under visible light. Int. J. Hydrog. Energy 2017, 42, 27043–27054. [Google Scholar] [CrossRef]
  63. Zhang, D.G.; Liu, W.B.; Wang, R.W.; Zhang, Z.T.; Qiu, S.L. Interface engineering of hierarchical photocatalyst for enhancing photoinduced charge transfers. Appl. Catal. B Environ. 2021, 283, 119632. [Google Scholar] [CrossRef]
  64. Li, S.S.; Peng, Y.N.; Hu, C.; Chen, Z.H. Self-assembled synthesis of benzene-ring-grafted g-C3N4 nanotubes for enhanced photocatalytic H2 evolution. Appl. Catal. B Environ. 2020, 279, 119401. [Google Scholar] [CrossRef]
  65. Zhang, G.G.; Lan, Z.A.; Lin, L.H.; Lin, S.; Wang, X.C. Overall water splitting by Pt/g-C3N4 photocatalysts without using sacrificial agents. Chem. Sci. 2016, 7, 3062–3066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Mei, S.K.; Gao, J.P.; Zhang, Y.; Yang, J.B.; Wu, Y.L.; Wang, X.X.; Zhao, R.R.; Zhai, X.G.; Hao, C.Y.; Li, R.X.; et al. Enhanced visible light photocatalytic hydrogen evolution over porphyrin hybridized graphitic carbon nitride. J. Colloid Interface Sci. 2017, 506, 58–65. [Google Scholar] [CrossRef]
  67. Xing, W.N.; Tu, W.G.; Ou, M.; Wu, S.Y.; Yin, S.M.; Wang, H.J.; Chen, G.; Xu, R. Anchoring active Pt2+/Pt0 hybrid nanodots on g-C3N4 nitrogen vacancies for photocatalytic H2 evolution. ChemSusChem 2019, 12, 2029–2034. [Google Scholar] [CrossRef] [PubMed]
  68. Nguyen, V.H.; Mousavi, M.; Ghasemi, J.B.; Le, Q.V.; Delbari, S.A.; Shahedi Asl, M.; Shokouhimehr, M.; Mohammadi, M.; Azizian-Kalandaragh, Y.; Sabahi Namini, A. In situ preparation of g-C3N4 nanosheet/FeOCl: Achievement and promoted photocatalytic nitrogen fixation activity. J. Colloid Interface Sci. 2021, 587, 538–549. [Google Scholar] [CrossRef]
  69. Li, X.; Jiang, H.P.; Ma, C.C.; Zhu, Z.; Song, X.H.; Wang, H.Q.; Huo, P.W.; Li, X.Y. Local surface plasma resonance effect enhanced Z-scheme ZnO/Au/g-C3N4 film photocatalyst for reduction of CO2 to CO. Appl. Catal. B Environ. 2021, 283, 119638. [Google Scholar] [CrossRef]
  70. Zhang, B.; Shi, H.X.; Yan, Y.J.; Liu, C.Q.; Hu, X.Y.; Liu, E.Z.; Fan, J. A novel S-scheme 1D/2D Bi2S3/g-C3N4 heterojunctions with enhanced H2 evolution activity. Colloids Surf. A Physicochem. Eng. Asp. 2021, 608, 125598. [Google Scholar] [CrossRef]
  71. Sui, Y.; Liu, J.H.; Zhang, Y.W.; Tian, X.K.; Chen, W. Dispersed conductive polymer nanoparticles on graphitic carbon nitride for enhanced solar-driven hydrogen evolution from pure water. Nanoscale 2013, 5, 9150–9155. [Google Scholar] [CrossRef]
  72. Shiraishi, Y.; Kofuji, Y.; Kanazawa, S.; Sakamoto, H.; Ichikawa, S.; Tanaka, S.; Hirai, T. Platinum nanoparticles strongly associated with graphitic carbon nitride as efficient co-catalysts for photocatalytic hydrogen evolution under visible light. Chem. Commun. 2014, 50, 15255–15258. [Google Scholar] [CrossRef]
  73. Chen, X.J.; Wang, J.; Chai, Y.Q.; Zhang, Z.J.; Zhu, Y.F. Efficient photocatalytic overall water splitting induced by the giant internal electric field of a g-C3N4/rGO/PDIP Z-Scheme heterojunction. Adv. Mater. 2021, 33, 2007479. [Google Scholar] [CrossRef]
  74. Chen, S.; Wang, C.; Bunes, B.R.; Li, Y.X.; Wang, C.Y.; Zang, L. Enhancement of visible-light-driven photocatalytic H2 evolution from water over g-C3N4 through combination with perylene diimide aggregates. Appl. Catal. A Gen. 2015, 498, 63–68. [Google Scholar] [CrossRef]
  75. Chen, Y.L.; Liu, X.Q.; Hou, L.; Guo, X.R.; Fu, R.W.; Sun, J.M. Construction of covalent bonding oxygen-doped carbon nitride/graphitic carbon nitride Z-scheme heterojunction for enhanced visible-light-driven H2 evolution. Chem. Eng. J. 2020, 383, 123132. [Google Scholar] [CrossRef]
  76. Speltini, A.; Scalabrini, A.; Maraschi, F.; Sturini, M.; Pisanu, A.; Malavasi, L.; Profumo, A. Improved photocatalytic H2 production assisted by aqueous glucose biomass by oxidized g-C3N4. Int. J. Hydrog. Energy 2018, 43, 14925–14933. [Google Scholar] [CrossRef]
  77. Su, Z.Z.; Zhang, B.X.; Shi, J.B.; Tan, D.X.; Zhang, F.Y.; Liu, L.F.; Tan, X.N.; Shao, D.; Yang, G.Y.; Zhang, J.L. An NH2-MIL-125 (Ti)/Pt/g-C3N4 catalyst promoting visible-light photocatalytic H2 production. Sustain. Energy Fuels 2019, 3, 1233–1238. [Google Scholar] [CrossRef]
  78. Li, Z.; Wang, Q.S.; Kong, C.; Wu, Y.Q.; Li, Y.X.; Lu, G.X. Interface charge transfer versus surface proton reduction: Which is more pronounced on photoinduced hydrogen generation over sensitized Pt cocatalyst on RGO? J. Phys. Chem. C 2015, 119, 13561–13568. [Google Scholar] [CrossRef]
  79. Takanabe, K.; Kamata, K.; Wang, X.; Antonietti, M.; Kubota, J.; Domen, K. Photocatalytic hydrogen evolution on dye-sensitized mesoporous carbon nitride photocatalyst with magnesium phthalocyanine. Phys. Chem. Chem. Phys. 2010, 12, 13020–13025. [Google Scholar] [CrossRef]
  80. Wang, Y.B.; Hong, J.D.; Zhang, W.; Xu, R. Carbon nitride nanosheets for photocatalytic hydrogen evolution: Remarkably enhanced activity by dye sensitization. Catal. Sci. Technol. 2013, 3, 1703–1711. [Google Scholar] [CrossRef]
  81. Min, S.X.; Lu, G.X. Enhanced electron transfer from the excited Eosin Y to mpg-C3N4 for highly efficient hydrogen evolution under 550 nm irradiation. J. Phys. Chem. C 2012, 116, 19644–19652. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pCN, D149, and D149/pCN.
Figure 1. XRD patterns of pCN, D149, and D149/pCN.
Processes 09 01055 g001
Figure 2. A typical STEM image of D149/pCN (A) and EDX mapping images of C (B), N (C), O (D), and S (E).
Figure 2. A typical STEM image of D149/pCN (A) and EDX mapping images of C (B), N (C), O (D), and S (E).
Processes 09 01055 g002
Figure 3. High-resolution XPS data of (A) C 1s, (B) N 1s, (C) O 1s, and (D) S 2p of pCN and D149/pCN.
Figure 3. High-resolution XPS data of (A) C 1s, (B) N 1s, (C) O 1s, and (D) S 2p of pCN and D149/pCN.
Processes 09 01055 g003
Figure 4. Photocatalytic H2 production of six samples (A), and cycling experiments of photocatalytic H2 production of D149/pCN-Pt (B).
Figure 4. Photocatalytic H2 production of six samples (A), and cycling experiments of photocatalytic H2 production of D149/pCN-Pt (B).
Processes 09 01055 g004
Figure 5. A typical STEM image of D149/pCN-Pt collected after the recyclability test (A) and EDX mapping images of C (B), N (C), O (D), S (E), and Pt (F).
Figure 5. A typical STEM image of D149/pCN-Pt collected after the recyclability test (A) and EDX mapping images of C (B), N (C), O (D), S (E), and Pt (F).
Processes 09 01055 g005
Figure 6. UV-vis DRS data of pCN, D149, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Figure 6. UV-vis DRS data of pCN, D149, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Processes 09 01055 g006
Figure 7. Transient photocurrent density of pCN, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Figure 7. Transient photocurrent density of pCN, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Processes 09 01055 g007
Figure 8. EIS spectra of pCN, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Figure 8. EIS spectra of pCN, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Processes 09 01055 g008
Figure 9. PL spectra of pCN, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Figure 9. PL spectra of pCN, D149/pCN, pCN-Pt, and D149/pCN-Pt.
Processes 09 01055 g009
Figure 10. Mechanism of photocatalytic H2 production by D149/pCN-Pt.
Figure 10. Mechanism of photocatalytic H2 production by D149/pCN-Pt.
Processes 09 01055 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Y.; Liu, Y.; Ma, Z. g-C3N4 Sensitized by an Indoline Dye for Photocatalytic H2 Evolution. Processes 2021, 9, 1055. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9061055

AMA Style

Chen Y, Liu Y, Ma Z. g-C3N4 Sensitized by an Indoline Dye for Photocatalytic H2 Evolution. Processes. 2021; 9(6):1055. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9061055

Chicago/Turabian Style

Chen, Yihang, Yanfei Liu, and Zhen Ma. 2021. "g-C3N4 Sensitized by an Indoline Dye for Photocatalytic H2 Evolution" Processes 9, no. 6: 1055. https://0-doi-org.brum.beds.ac.uk/10.3390/pr9061055

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