Next Article in Journal
CONSMI: Contrastive Learning in the Simplified Molecular Input Line Entry System Helps Generate Better Molecules
Next Article in Special Issue
Facilitated Unidirectional Electron Transmission by Ru Nano Particulars Distribution on MXene Mo2C@g-C3N4 Heterostructures for Enhanced Photocatalytic H2 Evolution
Previous Article in Journal
Water Exchange from the Buried Binding Sites of Cytochrome P450 Enzymes 1A2, 2D6, and 3A4 Correlates with Conformational Fluctuations
Previous Article in Special Issue
Hybrid Photoelectrocatalytic TiO2-Co3O4/Co(OH)2 Materials Prepared from Bio-Based Surfactants for Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Achieving Highly Efficient Photocatalytic Hydrogen Evolution through the Construction of g-C3N4@PdS@Pt Nanocomposites

1
School of Electronic Engineering, Nanjing Xiaozhuang University, Nanjing 211171, China
2
School of Physics and Information Engineering, Jiangsu Province Engineering Research Center of Basic Education Big Data Application, Jiangsu Second Normal University, Nanjing 210013, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(2), 493; https://doi.org/10.3390/molecules29020493
Submission received: 27 December 2023 / Revised: 13 January 2024 / Accepted: 17 January 2024 / Published: 19 January 2024
(This article belongs to the Special Issue Advances in Composite Photocatalysts)

Abstract

:
Selective supported catalysts have emerged as a promising approach to enhance carrier separation, particularly in the realm of photocatalytic hydrogen production. Herein, a pioneering exploration involves the loading of PdS and Pt catalyst onto g-C3N4 nanosheets to construct g-C3N4@PdS@Pt nanocomposites. The photocatalytic activity of nanocomposites was evaluated under visible light and full spectrum irradiation. The results show that g-C3N4@PdS@Pt nanocomposites exhibit excellent properties. Under visible light irradiation, these nanocomposites exhibit a remarkable production rate of 1289 μmol·g−1·h−1, marking a staggering 60-fold increase compared to g-C3N4@Pt (20.9 μmol·g−1·h−1). Furthermore, when subjected to full spectrum irradiation, the hydrogen production efficiency of g-C3N4@PdS@Pt-3 nanocomposites reaches an impressive 11,438 μmol·g−1·h−1, representing an eightfold enhancement compared to g-C3N4@Pt (1452 μmol·g−1·h−1) under identical conditions. Detailed investigations into the microstructure and optical properties of g-C3N4@PdS catalysts were conducted, shedding light on the mechanisms governing photocatalytic hydrogen production. This study offers valuable insights into the potential of these nanocomposites and their pivotal role in advancing photocatalysis.

Graphical Abstract

1. Introduction

Hydrogen (H2) [1], renowned as the lightest gas globally, boasts a high combustion temperature, yielding water upon combustion [2]. Thus, it stands as a pivotal clean energy source, offering promise in alleviating the dual pressures of energy scarcity and environmental pollution [3]. Diverse methods exist for hydrogen production, encompassing electric water decomposition [4], metal–acid reactions, thermal compound decomposition [5], steam reforming of natural gas [6], and the photocatalytic splitting of water [7,8,9]. While the initial three methods exhibit drawbacks, photocatalytic hydrogen production via water splitting is heralded as a green and sustainable avenue for solar energy conversion [10,11,12]. In this process, the development of a highly efficient photocatalyst remains pivotal for its wider application and industrialization. Semiconductor materials, deemed ideal photocatalysts, necessitate strong light absorption capabilities, an appropriate band structure, abundant reactive sites, and efficient carrier separation [13].
Numerous semiconductors, including metal–organic frameworks [14], metal oxide (TiO2 [15], ZnO [16], CeO2 [17], CuO [18], etc.), carbon-related compounds (GO [19], C3N4 [20], etc.), and various metal sulfides (CdS [21,22], ZnIn2S4 [23], CuS [24], etc.), serve as widely employed photocatalysts for hydrogen production. Within this realm, the graphene-like two-dimensional structure of C3N4 (g-C3N4) stands out as a typical polymer semiconductor [25]. The C-N atoms in its structure exhibit sp2 hybridization, forming a highly delocalized π-conjugated system akin to graphene’s layered structure [26]. Noteworthy for its excellent chemical and thermal stability, visible light absorption, non-toxicity, mineral abundance, and simple preparation process, g-C3N4 has garnered increasing attention in recent years across multiple domains, emerging as a focal point of research [27,28]. However, inherent photogenerated carrier recombination limits the overall photocatalytic activity of pure g-C3N4, posing a challenge for substantially enhancing its efficiency. Consequently, various strategies have been explored to improve g-C3N4′s photocatalytic efficiency, including modifications [29,30,31,32] and heterojunction construction with other semiconductors [33,34,35].
Moreover, the deposition of co-catalysts on nanomaterial surfaces has proven pivotal in facilitating charge separation by swiftly capturing electrons or holes, consequently elevating photocatalytic performance. A gamut of precious metals (such as Pt [36], Pd [37], Rh [38]), metal oxides (NiO [39], CuO [40]), and metal sulfides (MoS2 [41]) have found application as co-catalysts, augmenting photocatalytic hydrogen evolution. Notably, previous studies have highlighted PdS as a promising co-catalyst in composite structures with dual co-catalysts, as seen in Pt-PdS/Cd0.5Zn0.5S [42], PANI@CdS@PdS [43], and Pd@CdS@PdS [44]. The incorporation of PdS enhances the ability for charge separation and promotes excellent photo-stability. Moreover, the introduction of PdS as a co-catalyst reduces the activation energy and fosters surface oxidation-reduction reactions [45]. Leveraging these advantageous characteristics of PdS, we opted to utilize it as a co-catalyst to enhance the photocatalytic production of H2 in conjunction with g-C3N4. This strategy of employing PdS as a catalyst to enhance the photocatalytic activity of g-C3N4 for H2 production has not been previously reported.
Thus, this study focuses on synthesizing g-C3N4 nanosheets via secondary pyrolysis of melamine, introducing PdS as a hole-trapping agent. The concentration of PdS in g-C3N4 nanosheets was regulated by controlling the Pd and S sources. Comprehensive investigations into the microstructure, morphology, optical properties, and valence states of g-C3N4 under varied PdS concentrations were conducted. Subsequently, the photocatalytic hydrogen production performance of the g-C3N4 nanocomposites under Pt and PdS co-catalysis was evaluated under visible light and full-spectrum irradiation. Detailed scrutiny of the photocatalytic mechanism and analysis thereof revealed significant enhancement in the photocatalytic hydrogen production performance of g-C3N4@PdS@Pt nanocomposites at an optimal PdS concentration.

2. Results and Discussion

Figure 1 displays the X-ray diffraction patterns (XRD) of both g-C3N4 and g- C3N4@PdS nanocomposites with adjustable PdS content. Notably, two prominent reflections appear at 2θ = 27.45, and 12.8°. The reflection at =27.5° aligns with the (002) diffraction peak of hexagonal phase in JPCDS 87-1526 [46], signifying the distinct interlayer stacking of the conjugated aromatic groups. Conversely, the faint reflection at =12.8° corresponds to the (100) plane, indicative of the in-plane structure of tri-s-triazine units. Comparatively, no new diffraction peaks attributable to PdS emerge in the g-C3N4@PdS nanocomposites. This is mainly due to the low PdS content, and may also be due to the high dispersion of PdS on the surface of g-C3N4 nanosheets, as reported by reference [47,48]. However, there is a noticeable reduction in the intensities and broadening of the (002) and (100) peaks, evident in the magnified illustration within Figure 1. This diminished intensity suggests a disruption in the interlayer structure, potentially attributed to several factors. The first reason is the ultrasonic stripping process undergone by a g-C3N4 during nanocomposite preparation. Additionally, the insertion of PdS nanoparticles between g-C3N4 layers contributes to this alteration. Moreover, as the quantity of PdS increases, the (002) diffraction peak shifts towards a lower angle, providing further evidence of successful PdS insertion between the g-C3N4 lamellas.
Figure 2 shows the Fourier Transform Infrared spectroscopy (FTIR) of sample g-C3N4 and g-C3N4@PdS nanocomposites. It is apparent that the characteristic peaks of the g-C3N4@PdS nanocomposites remain relatively consistent when compared to the pure g-C3N4. The characteristic stretching peaks were in three main regions: 809, 1100–1700, and 3000–3400 cm−1. Specifically, the wide vibration band at 3000–3400 cm−1 signifies the stretching vibration peak associated with N-H, residual amino groups and O-H adsorbed on the surface of g-C3N4 [49,50]. And the multiple strong vibration bands within the range of 1100–1700 cm−1 arise from the unique stretching vibration peak related to the C-N heterocyclic ring [51]. Additionally, the peak at 809 cm−1 aligns with the characteristic vibration of the triazine units [52,53]. Furthermore, the figure illustrates that, as the amount of PdS increases, the vibration mode within the 809 and 3000–3400 cm−1 regions weakens, suggesting successful insertion of PdS into the g-C3N4 layer. To delve deeper into the microstructure analysis of g-C3N4@PdS catalysts, we conducted TEM characterization.
Figure 3 illustrates the transmission electron microscopy (TEM) of both g-C3N4 and g-C3N4@PdS nanocomposites. The g-C3N4 exhibits a uniform composition with consistent thickness. Upon the introduction of Pd and S sources, black nanoparticles attached to the nanosheet layer (Figure 3b). Moreover, as the quantity of Pd and S sources increases, there is a proportional rise in the number of nanoparticles, notably evident in g-C3N4@PdS-3. The distribution of these black nanoparticles appears uniform. The size of the nanoparticles is about 5 nm. Detailed examination via high-resolution TEM (HRTEM, depicted in Figure 3c) reveals a lattice fringe measuring 0.231 nm, corresponding to the (202) crystal plane of the PdS. This observation confirms that the attached nanoparticles consist of PdS. Furthermore, an escalation in the Pd and S sources leads to a higher loading capacity of PdS nanoparticles.
Figure 4 illustrates the characterization of the typical g-C3N4@PdS-3 nanocomposites through high-angle annular dark-field STEM imaging and energy dispersive spectroscopy (EDS) elemental mappings. In the high-angle annular dark-field image, the brightness levels correspond to the distribution of Pd and S in nanoparticles, and, conversely, reveal the distribution of C and N. Analysis of each element’s mapping shows a uniform distribution of C, N, Pd, and S elements within the g-C3N4@PdS-3 nanocomposites. Given the low PdS dosage, the density of Pd and S on the nanosheet is relatively low.
To analyze the chemical bond state and molecular structure on the material surface, both g-C3N4 and the g-C3N4@PdS nanocomposites underwent characterization via X-ray photoelectron spectra (XPS). All XPS spectra were calibrated by aligning the C=C binding energy position to 284.5 eV. Figure 5a illustrates the elemental composition of g-C3N4, revealing the presence of C, N, and O elements exclusively. As the quantity of Pd and S sources increases, a gradual emergence of weak binding energy associated with Pd and S elements is observed. Simultaneously, intensity of the Pd and S binding energy peaks stability intensifies. This phenomenon indicates the successful recombination of Pd and S within the g-C3N4 layer, signifying a progressive increase in PdS content. The C1s spectra were analyzed in Figure 5b, revealing four distinctive peaks through Gaussian fitting. These peaks, situated at approximately ~284.5, 286.3, 287.9, and 293.5 eV, correspond to graphitic-like carbon species (C=C or C–C), the N≡C-bond between the sp2 C atom and NH groups in the aromatic ring, sp2 hybridized carbon atoms bonded with N (N–C=N), and carbon attached to uncondensed-NH2 groups. Similarly, the N1s spectrum in Figure 5c, segmented via Gaussian fitting, displayed four peaks at 398.4 (C–N=C), 399 (N–(C)3), 400.4 (C=N–H), and 404.2 eV (π excitation), respectively. The Pd 3d spectrum showcased peaks at 336.2 and 341.6 eV (Figure 5d), attributed to Pd 3d5/2, and Pd 3d3/2, indicating the presence of Pd2+. Regarding the S 2p spectrum in Figure 5e, division into two peaks with binding energies at 161.1 and 162.3 corresponded to S2p3/2 and S2p1/2, respectively, suggesting the existence of S2− in the nanocomposites. Comparison with g-C3N4 revealed a downward shift in the binding energy of C, N, Pd, and S with increasing Pd and S sources. This shift implies an altered internal chemical bond within the nanocomposites post-formation, suggesting an interface interaction between g-C3N4 and PdS. Furthermore, the XPS valence band spectra provide insights into the valence band position (EVB, XPS) of both g-C3N4 and g-C3N4@PdS-3, measuring 1.23 and 0.87 eV, respectively. Utilizing the formula EVB, NHE = φ + EVB, XPS − 4.44, where φ signifies the work function of the XPS instrument (φ = 4.258), and EVB, NHE represents the valence band position relative to the normal hydrogen electrode (NHE), the calculated EVB, NHE are 1.05 and 0.688 eV for g-C3N4 and g-C3N4@PdS-3, respectively. These results collectively confirm that the conduction band of g-C3N4@PdS heterojunction exists in a more negative position, which inherently favors a hydrogen evolution reaction.
To determine the optical band gap of the g-C3N4 and g-C3N4@PdS nanocomposites, we conducted measurements using the ultraviolet-visible diffuse scattering spectrum, displayed in Figure 6a. This spectrum exhibits a distinct absorption edge, demonstrating a redshift in the absorption edge as the load of PdS nanoparticles increases. This indicates an enhanced light absorption capability in g-C3N4@PdS nanocomposites, visibly reflected in their color transformation. As depicted in Figure S1 of the supporting material, the transition from the whitish-yellow hue of g-C3N4 to brown reinforces the heightened light absorption capacity of the nanocomposite material. This augmented light absorption implies a more efficient utilization of light in the generation of photogenerated electrons and holes during the photocatalytic H2 production process. The band gap of the both g-C3N4 and g-C3N4@PdS nanocomposites was determined using the Tauc equation [54]: αhv = A(hv − Eg)1/2, where Eg represents the optical bandgap, α signifies the absorption coefficient, h and v denote Planck’s constant and incident light frequency, respectively, while A represents a constant. Figure 6b illustrates the correlation between hv and (αhv)2. According to the Tauc equation, the optical band gap is identified at the intersection of the linear segment of the curve with the base line. For g-C3N4, g-C3N4@PdS-1, g-C3N4@PdS-2, g-C3N4@PdS-3, and g-C3N4@PdS-4, the optical band gaps measure 2.874, 2.869, 2.866, 2.861, and 2.854 eV, respectively.
The effectiveness of g-C3N4 nanocomposites in photocatalytic water splitting was evaluated under varying PdS nanoparticle loads at a temperature of 5 °C in the presence of 20 vol % lactic acid. Lactic acid was utilized as a sacrificial agent to capture the holes in the process. Before evaluating the performance of photocatalytic H2 production, a co-catalyst, Pt, was applied to the catalyst via photodeposition, where chloroplatinic acid was used as the Pt source. Figure 7a,b display the corresponding H2 generation rate of the prepared photocatalyst under visible light (using xenon lamp irradiation with a long pass filter, λ > 420 nm) and full spectrum (xenon lamp irradiation) over time. The graphs illustrate a linear increase in H2 production for all photocatalysts with extended exposure to both light sources. Particularly under visible light, it is evident that the g-C3N4@Pt catalyst exhibits the lowest H2 production compared to g-C3N4@PdS@Pt nanocomposites. This lower H2 output is attributed to the rapid recombination of photogenerated carriers in g-C3N4. Remarkably, the g-C3N4@PdS@Pt-3 nanocomposites exhibit exceptional properties, generating up to 1289 μmol·g−1·h−1, a staggering 60-times increase compared to g-C3N4@Pt (20.9 μmol·g−1·h−1). Upon removing the filter, the H2 production efficiency notably surges under full spectrum irradiation, as depicted in Figure 7b. The H2 production rate of g-C3N4@PdS@Pt-3 reaches an impressive 11438 μmol·g−1·h−1, which is eight times higher than the H2 production efficiency of g-C3N4@Pt (1452 μmol·g−1·h−1) under identical conditions. This showcases a ninefold improvement over visible light irradiation. Figure 7c clearly depicts the trend in H2 production efficiency under the two light sources. To ascertain the chemical stability of the photocatalyst, the stability of g-C3N4@PdS@Pt nanocomposites was tested over four cycles under full spectrum irradiation (Figure 7d). The results indicate that after two cycle tests the photocatalytic H2 production performance remains largely unchanged. However, during the third cycle, there is a slight decrease in performance due to the consumption of lactic acid in the reaction solution. Table 1 outlines the comparison between the H2 production efficiency achieved by combining g-C3N4 with various catalysts and the optimal efficiency observed in this study. Evidently, the introduction of the PdS and Pt co-catalysts significantly enhances the performance of photocatalytic H2 production.
The apparent quantum efficiency (AQE) was calculated using the following equation:
A Q Y s ( % ) = 2 × N H 2 N p × 100 % = 2 × N H 2 I × A × λ h × c × 100 %
where Np, I, A, h, c, and λ represent the number of incident photons, the illumination intensity, the irradiation area of the incident light, Planck’s constant, the speed of light, and the wavelength of the incident light, respectively. Here, the monochromatic light was achieved by implementing a band-pass filter in the xenon light source outlet. According to Formula 1, the AQY of g-C3N4@PdS@Pt-3 catalyst at 365, 380, 400, 420, 450, and 500 nm were calculated and are depicted in Figure 8. As illustrated in Figure 8, the quantities of H2 produced at different wavelengths correlate with the light absorption, showcasing a decrease in AQE as the wavelengths increase. Notably, the highest AQE, recorded at 365 nm, stands at an impressive 25.2%.
The separation characteristics of carriers within the samples are investigated using PL and TRPL measurements. Figure 9a illustrates the PL spectrum at a 250 nm excitation wavelength, highlighting a primary blue luminescence peak at ≈480 nm, which originates from the transition between lone pair states in the valence band and the π* antibonding states in the conduction band [61], indicating charge recombination. In the figure, the bare g-C3N4 display the most pronounced PL peak, while the PL intensities noticeably decrease upon the integration of PdS. Notably, the g-C3N4@PdS-3 hybrid exhibits the lowest PL intensity among all the nanocomposites, consistent with the comparison of photocatalytic activities. This analysis indicates that the g-C3N4@PdS nanocomposites effectively mitigate the charge recombination in g-C3N4. Furthermore, the TRPL experiments were conducted on both bare g-C3N4 and g-C3N4@PdS-3 samples to delve deeper into the charge transmission process, as depicted in Figure 9b. The TRPL curves were fitted using a multiexponential function, and the resulting parameters are summarized in Table 2. Decay lifetimes were calculated according to Equation (2):
τ = A 1 τ 1 2 + A 2 τ 2 2 A 1 τ 1 + A 2 τ 2
where, τ1, τ2 represent the short carrier lifetime attributed to quasi-free excitons and the long component due to localized exciton recombination, respectively. A1 and A2 correspond to the percentages of the short and long component in the total lifetime. Significantly, the bare g-C3N4 (5.7917) exhibits a longer average decay lifetime compared to g-C3N4@PdS-3 (7.8609 ns). The brief fluorescence lifetime hints at the possibility of extra non-radiative attenuation pathways being activated within the g-C3N4@PdS sample. These pathways could effectively impede the recombination of photogenerated carriers. The outcomes mentioned above demonstrate that the heterojunction created by embedding PdS nanoparticles onto g-C3N4 nanosheets actively facilitates the parting of electron-hole pairs, consequently enhancing the photocatalytic efficiency.
Furthermore, this study delved into the interfacial charge transfer and separation capacities of both pure g-C3N4 and g-C3N4@PdS nanocomposites through analyses using transient photocurrent and electrochemical impedance spectroscopy. As depicted in Figure 10a, the g-C3N4@PdS-3 nanocomposites demonstrated a notably heightened photocurrent response compared to the pristine g-C3N4, indicating significantly enhanced charge separation capabilities. Additionally, electrochemical impedance spectroscopy (EIS) can also be used to assess electron mobility at the electrode interface, which usually reflects the charge transfer ability of the photocatalyst. The ESI Nyquist diagram in Figure 10b vividly illustrates that the arc radius of the g-C3N4@PdS-3 nanocomposites is markedly smaller than that of the pure g-C3N4, signifying swifter charge transfer kinetics and lower charge transfer resistance in the former. In addition, Figure 11 provides a comparison of the overpotential of g-C3N4 and g-C3N4@PdS through linear scanning voltammetry (LSV) measurements. At the reference current density of 10 mAcm−2, g-C3N4@PdS-3 exhibits a lower overpotential (−0.87V) compared to g-C3N4(−1.03V), suggesting that the presence of PdS loaded onto g-C3N4 is more favorable for facilitating H2 production than g-C3N4 alone.
It has been reported that PdS is an n-type semiconductor with a band gap of 1.6 eV. Its valence and conduction band positions are situated at 1.1 eV and −0.5 eV, respectively [62,63]. Drawing upon the preceding analyses, Figure 12 outlines a reasonable charge transfer behavior and proposes a mechanism for photocatalytic hydrogen production reaction in g-C3N4@PdS@Pt nanocomposites. Initially, the attachment of PdS uniformly onto g-C3N3 nanosheets widens the light absorption spectrum of nanocomposites. Therefore, the inclusion of PdS augments the production of photogenerated charge carriers. Additionally, the uniform distribution of PdS enhances the availability of active sites. Upon exposure to light, incident light energizes valence band electrons into the conduction band while generating holes in the valence band. Subsequently, electrons within the conduction band of g-C3N4 swiftly migrate to the conduction band of PdS, where they are captured by Pt, catalyzing a reduction reaction upon interaction with absorbed hydrated protons at the Pt site, thereby liberating hydrogen. Simultaneously, the holes present in the valence band are continually consumed with the electron donor, lactic acid in the solution. Hence, the existence of PdS enhances the quantity of photoinduced carriers and effectively facilitates their separation.

3. Experimental

3.1. Materials

Melamine (99%), sodium chloropalladate (Na2PdCl4, 99.99%), sodium sulfide (Na2S, 99%), ethanolamine (99%), lactic acid (20%), chloroplatinic acid hexahydrate (AR, Pt > 37.5%), and sodium sulfate (AR, 99%) were purchased from Aladdin Reagent Co., Ltd. (Shanghai, China) and employed directly without additional purification. Deionized water with 18 MΩ cm was used in our experiment.

3.2. Synthesis of g-C3N4@PdS Nanocomposites

The synthesis process of g-C3N4@PdS nanocomposites is presented in Scheme 1. Briefly, g-C3N4 nanosheets were synthesized via a double pyrolysis process of melamine, as detailed in our previously published literature [20]. Its color is milky yellow, as shown in Scheme 1. Subsequently, 400 mg of g-C3N4 underwent ultrasonic dispersion in 50 mL of ethanolamine for 2 h. Concurrently, a solution containing 0.01 mmol of Na2PdCl4 dissolved in 10 mL of deionized water was vigorously stirred to achieve uniformity. This solution was gradually added drop by drop to the ultrasonically dispersed g-C3N4 solution. Following 2 h of stirring, 0.6 × 10−3 M Na2S was incrementally introduced into the solution. The resultant mixture was thoroughly stirred for 12 h, washed successively with ethanol and deionized water, and then freeze-dried, resulting in the sample denoted as g-C3N4@PdS-1. To find the optimal PdS load, a series of nanocomposites were synthesized. As the quantity of the Pd source was incrementally altered to 0.02, 0.03, and 0.04 mmol (with the corresponding S source being adjusted to 1.2 × 10−3 M, 2.4 × 10−3 M, and 3.6 × 10−3 M, respectively), the synthesized samples were sequentially designated as g-C3N4@PdS-2, g-C3N4@PdS-3, and g-C3N4@PdS-4. Scheme 1 depicts the color of g-C3N4@PdS nanocomposites, exhibiting a light brown hue. This coloration signifies an enhancement in the nanocomposites’ capability to absorb visible light.

3.3. Characterization

XRD analysis was performed using a Bruker D8 Advance instrument (Bruker, Saarbrucken, Germany) to assess the evolution of the crystal structure following the incorporation of PdS. Morphologies and microstructural investigations were conducted using Talos F200X G2 TEM JEOL (Thermo Scientific, Waltham, MA, USA, FEI company, Hillsboro, OR, USA), and HRTEM (Thermo Scientific, Waltham, MA, USA, FEI company, Hillsboro, OR, USA) operated at an accelerated voltage of 200 KV. Additionally, super-X model EDS accompanied TEM analysis to investigate the element distribution. For TEM analysis, a 1 mg catalyst dispersed in ethanol underwent ultrasonication for 10 min, was deposited onto a copper grid, naturally dried, and then examined. The functional group characteristics of the synthesized materials were analyzed via FTIR, Thermo Scientific Nicolet iS20 (Thermo Scientific, Waltham, MA, USA). Furthermore, changes in the valence state and band structure of the elements in the nanocomposite were explored using XPS via a PHI 5000 Versaprobe Ⅲ (Spectra Research Corporation, Mississauga, ON, USA) spectroscopy instrument, utilizing monochromatic Al Kα radiation. Ultraviolet-visible diffuse reflectance spectra were acquired using a Hitachi UH4150 (Hitachi High-Tech Corporation, Tokyo, Japan) equipped with an integrating sphere. Steady-state PL was investigated using the Hitachi F7000 spectrofluorometer (Hitachi High-Tech Corporation, Tokyo, Japan), employing an excitation wavelength of 250 nm. Evaluation of photocurrent performance, EIS and LSV were conducted using a three-electrode electrochemical workstation (CHI660E, Chen Hua, Shanghai, China). The reference electrode used was Ag/AgCl, while the counter electrode utilized a Pt plate. To prepare the working electrode, 5 mg of the catalyst was ultrasonically dispersed in 1 mL of ethanol along with 20 μL of nafion solution, forming a uniform solution. This catalyst-containing solution was subsequently deposited onto the FTO glass and dried to form the working electrode. The electrolyte employed was a 0.5 M aqueous solution of Na2SO4. A bias voltage of 0.3 V was added to the photocurrent test. The potential of the electrode of EIS test voltage was 0.24 V.

3.4. Evaluation of Photocatalytic H2 Production Activity

The procedure for measuring photocatalytic H2 production involved using a reaction flask filled with 30 mg of the photocatalyst alongside a 20% lactic acid aqueous solution (10 mL) serving as a sacrificial agent. This was combined with 100 mL of deionized water and a 3% wt Pt, employed as a co-catalyst, with chloroplatinic acid as the Pt source. The mixture was subjected to ultrasonic dispersion for 30 min to ensure the formation of a uniformly dispersed suspension. This suspension was then transferred into a quartz reactor connected to an on-line trace gas analysis system (Labsolar-6A, Beijing Perfectlight, Beijing, China). A constant temperature water-cooling system was used to maintain the reaction solution at 5 °C. To guarantee the complete elimination of air, the system and the reactor were evacuated several times. After the vacuum extraction, the reactor underwent a 30 min exposure to a 300 W xenon arc light source to facilitate the reduction of Pt. Subsequently, Pt was loaded onto the g-C3N4 and g-C3N4@PdS nanocomposites. The resulting photodeposited nanocomposites were sequentially labeled as g-C3N4@Pt, g-C3N4@PdS@Pt-1, g-C3N4@PdS@Pt-2, g-C3N4@PdS@Pt-3, and g-C3N4@PdS@Pt-4. After light deposition, the system was vacuumed again in preparation for photocatalytic H2 production experiments. The photocatalytic H2 production experiments commenced with the reactor being irradiated under visible light (using xenon lamp irradiation with a long pass cut-off filter, λ > 420 nm) and full spectrum (xenon lamp irradiation). Following irradiation, the concentration of photocatalytically produced H2 was assessed using an online gas chromatograph (Fuli instruments, Zhejiang, China, GC9720PLUS) equipped with a thermal conductive detector.

4. Conclusions

In this paper, a g-C3N4@PdS nanocomposite with varying concentration of PdS was prepared via a straightforward method. Comprehensive investigations into the microstructure, morphology, band structure, element distribution, and optical properties of these catalysts were conducted. XPS analysis unambiguously confirmed the successful loading of PdS onto the g-C3N4 layer, while TEM imaging revealed the uniform distribution of PdS nanoparticles on the g-C3N4 layer. The crystal structure of the resultant g-C3N4@PdS nanocomposite remained largely unchanged, attributed to the low PdS content. The study extensively evaluated the photocatalytic performance of these nanocomposites under both visible light and full spectrum irradiation. Encouragingly, the g-C3N4@PdS composites exhibited significantly enhanced photocatalytic hydrogen production. Notably, the hydrogen production rate of g-C3N4@PdS@Pt-3 nanocomposites surpassed that of g-C3N4@Pt by 60-times under visible light and 8-times under full spectrum irradiation. Characterization through various methods including photocurrent response curve, EIS, LVS, PL, and TRPL decay curves unveiled that the structure of the composite facilitated an accelerated transfer of photogenerated carriers, thereby augmenting the photocatalytic hydrogen production rate. The incorporation of PdS enhances light absorption and enhances the efficiency of carrier transfer, thereby contributing to the improved performance of the g-C3N4@PdS@Pt nanocomposite.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29020493/s1.

Author Contributions

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

Funding

This work was jointly supported by the National Natural Science Foundation of China (No. 12204245), the Natural Science Foundation of the Jiangsu Higher Education Institutions of China (No. 23KJB140004).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Lubitz, W.; Tumas, W. Hydrogen:  An overview. Chem. Rev. 2007, 107, 3900–3903. [Google Scholar] [CrossRef] [PubMed]
  2. Florin, N.H.; Harris, A.T. Hydrogen production from biomass coupled with carbon dioxide capture: The implications of thermodynamic equilibrium. Int. J. Hydrog. Energy 2007, 32, 4119–4134. [Google Scholar] [CrossRef]
  3. Acharya, R.; Parida, K. A review on TiO2/g-C3N4 visible-light-responsive photocatalysts for sustainable energy generation and environmental remediation. J. Environ. Chem. Eng. 2020, 8, 103896. [Google Scholar] [CrossRef]
  4. Rossmeisl, J.; Qu, Z.-W.; Zhu, H.; Kroes, G.-J.; Nørskov, J.K. Electrolysis of water on oxide surfaces. J. Electroanal. Chem. 2007, 607, 83–89. [Google Scholar] [CrossRef]
  5. Grochala, W.; Edwards, P.P. Thermal decomposition of the non-interstitial hydrides for the storage and production of hydrogen. Chem. Rev. 2004, 104, 1283–1316. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, M.; Al-Subhi, K.; Al-Rajhi, A.; Al-Maktoumi, A.; Izady, A.; Al-Hinai, A. Numerical evaluation of hydrogen production by steam reforming of natural gas. Adv. Geo-Energy Res. 2023, 7, 141–151. [Google Scholar] [CrossRef]
  7. Hu, Y.H. A highly efficient photocatalyst-Hydrogenated black TiO2 for the photocatalytic splitting of water. Angew. Chem. Int. Ed. 2012, 51, 12410–12412. [Google Scholar] [CrossRef]
  8. Maeda, K. Photocatalytic water splitting using semiconductor particles: History and recent developments. J. Photochem. Photobiol. C Photochem. Rev. 2011, 12, 237–268. [Google Scholar] [CrossRef]
  9. Ismail, A.A.; Bahnemann, D.W. Photochemical splitting of water for hydrogen production by photocatalysis: A review. Sol. Energy Mater. Sol. Cells 2014, 128, 85–101. [Google Scholar] [CrossRef]
  10. Shi, Y.; Yang, A.-F.; Cao, C.-S.; Zhao, B. Applications of MOFs: Recent advances in photocatalytic hydrogen production from water. Coord. Chem. Rev. 2019, 390, 50–75. [Google Scholar] [CrossRef]
  11. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  12. Li, T.; Tsubaki, N.; Jin, Z. S-scheme heterojunction in photocatalytic hydrogen production. J. Mater. Sci. Technol. 2024, 169, 82–104. [Google Scholar] [CrossRef]
  13. Yang, Y.; Zhou, C.; Wang, W.; Xiong, W.; Zeng, G.; Huang, D.; Zhang, C.; Song, B.; Xue, W.; Li, X. Recent advances in application of transition metal phosphides for photocatalytic hydrogen production. Chem. Eng. J. 2021, 405, 126547. [Google Scholar] [CrossRef]
  14. Hu, N.; Cai, Y.; Li, L.; Wang, X.; Gao, J. Amino-functionalized titanium based metal-organic framework for photocatalytic hydrogen production. Molecules 2022, 27, 4241. [Google Scholar] [CrossRef] [PubMed]
  15. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B Environ. 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  16. Ramírez-Ortega, D.; Guerrero-Araque, D.; Acevedo-Peña, P.; Reguera, E.; Calderon, H.A.; Zanella, R. Enhancing the photocatalytic hydrogen production of the ZnO–TiO2 heterojunction by supporting nanoscale Au islands. Int. J. Hydrog. Energy 2021, 46, 34333–34343. [Google Scholar] [CrossRef]
  17. Mansingh, S.; Padhi, D.; Parida, K. Enhanced photocatalytic activity of nanostructured Fe doped CeO2 for hydrogen production under visible light irradiation. Int. J. Hydrogen Energy 2016, 41, 14133–14146. [Google Scholar] [CrossRef]
  18. Kum, J.M.; Yoo, S.H.; Ali, G.; Cho, S.O. Photocatalytic hydrogen production over CuO and TiO2 nanoparticles mixture. Int. J. Hydrogen Energy 2013, 38, 13541–13546. [Google Scholar] [CrossRef]
  19. Hernández-Majalca, B.; Meléndez-Zaragoza, M.; Salinas-Gutiérrez, J.; López-Ortiz, A.; Collins-Martínez, V. Visible-light photo-assisted synthesis of GO-TiO2 composites for the photocatalytic hydrogen production. Int. J. Hydrogen Energy 2019, 44, 12381–12389. [Google Scholar] [CrossRef]
  20. Ma, L.; Jiang, W.; Lin, C.; Xu, L.; Zhu, T.; Ai, X. CdS Deposited In Situ on g-C3N4 via a Modified Chemical Bath Deposition Method to Improve Photocatalytic Hydrogen Production. Molecules 2023, 28, 7846. [Google Scholar] [CrossRef]
  21. Cheng, L.; Xiang, Q.; Liao, Y.; Zhang, H. CdS-based photocatalysts. Energy Environ. Sci. 2018, 11, 1362–1391. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Wang, X.; Ren, X.; Luo, S.; Huang, H.; Chen, R.; Shao, S.; Liu, D.; Gao, J.; Gui, J.; et al. Building rapid charge transfer channel via engineering Ni coordinated flexible polymer for efficient solar hydrogen evolution. Chem. Eng. J. 2023, 456, 141032. [Google Scholar] [CrossRef]
  23. Yang, R.; Mei, L.; Fan, Y.; Zhang, Q.; Zhu, R.; Amal, R.; Yin, Z.; Zeng, Z. ZnIn2S4-Based photocatalysts for energy and environmental applications. Small Methods 2021, 5, 2100887. [Google Scholar] [CrossRef]
  24. Zhang, J.; Yu, J.; Zhang, Y.; Li, Q.; Gong, J.R. Visible light photocatalytic H2-production activity of CuS/ZnS porous nanosheets based on photoinduced interfacial charge transfer. Nano Lett. 2011, 11, 4774–4779. [Google Scholar] [CrossRef] [PubMed]
  25. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based heterostructured photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  26. Gao, J.; Zhang, F.; Xue, H.; Zhang, L.; Peng, Y.; Li, X.; Gao, Y.; Li, N.; Lei, G. In-situ synthesis of novel ternary CdS/PdAg/g-C3N4 hybrid photocatalyst with significantly enhanced hydrogen production activity and catalytic mechanism exploration. Appl. Catal. B Environ. 2021, 281, 119509. [Google Scholar] [CrossRef]
  27. Yan, S.; Li, Z.; Zou, Z. Photodegradation performance of g-C3N4 fabricated by directly heating melamine. Langmuir 2009, 25, 10397–10401. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, J.; Wang, S. A critical review on graphitic carbon nitride (g-C3N4)-based materials: Preparation, modification and environmental application. Coord. Chem. Rev. 2022, 453, 214338. [Google Scholar] [CrossRef]
  29. Li, Y.; Zhang, M.; Zhou, L.; Yang, S.; Wu, Z.; Ma, Y. Recent advances in surface-modified g-C3N4-based photocatalysts for H2 production and CO2 reduction. Acta Phys.-Chim. Sin. 2021, 37, 2009030. [Google Scholar]
  30. Patnaik, S.; Sahoo, D.P.; Parida, K. An overview on Ag modified g-C3N4 based nanostructured materials for energy and environmental applications. Renew. Sustain. Energy Rev. 2018, 82, 1297–1312. [Google Scholar] [CrossRef]
  31. Fang, S.; Xia, Y.; Lv, K.; Li, Q.; Sun, J.; Li, M. Effect of carbon-dots modification on the structure and photocatalytic activity of g-C3N4. Appl. Catal. B Environ. 2016, 185, 225–232. [Google Scholar] [CrossRef]
  32. Bai, X.; Wang, L.; Wang, Y.; Yao, W.; Zhu, Y. Enhanced oxidation ability of g-C3N4 photocatalyst via C60 modification. Appl. Catal. B Environ. 2014, 152, 262–270. [Google Scholar] [CrossRef]
  33. Li, Y.; Xia, Z.; Yang, Q.; Wang, L.; Xing, Y. Review on g-C3N4-based S-scheme heterojunction photocatalysts. J. Mater. Sci. Technol. 2022, 125, 128–144. [Google Scholar] [CrossRef]
  34. Dong, F.; Zhao, Z.; Xiong, T.; Ni, Z.; Zhang, W.; Sun, Y.; Ho, W.-K. In situ construction of g-C3N4/g-C3N4 metal-free heterojunction for enhanced visible-light photocatalysis. ACS Appl. Mater. Interfaces 2013, 5, 11392–11401. [Google Scholar] [CrossRef] [PubMed]
  35. Geng, Y.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Z-Scheme 2D/2D α-Fe2O3/g-C3N4 heterojunction for photocatalytic oxidation of nitric oxide. Appl. Catal. B Environ. 2021, 280, 119409. [Google Scholar] [CrossRef]
  36. Caux, M.; Fina, F.; Irvine, J.T.; Idriss, H.; Howe, R. Impact of the annealing temperature on Pt/g-C3N4 structure, activity and selectivity between photodegradation and water splitting. Catal. Today 2017, 287, 182–188. [Google Scholar] [CrossRef]
  37. Tong, T.; Zhu, B.; Jiang, C.; Cheng, B.; Yu, J. Mechanistic insight into the enhanced photocatalytic activity of single-atom Pt, Pd or Au-embedded g-C3N4. Appl. Surf. Sci. 2018, 433, 1175–1183. [Google Scholar] [CrossRef]
  38. Lu, R.; Hu, M.; Xu, C.; Wang, Y.; Zhang, Y.; Xu, B.; Gao, D.; Bi, J.; Fan, G. Hydrogen evolution from hydrolysis of ammonia borane catalyzed by Rh/g-C3N4 under mild conditions. Int. J. Hydrogen Energy 2018, 43, 7038–7045. [Google Scholar] [CrossRef]
  39. Liu, J.; Jia, Q.; Long, J.; Wang, X.; Gao, Z.; Gu, Q. Amorphous NiO as co-catalyst for enhanced visible-light-driven hydrogen generation over g-C3N4 photocatalyst. Appl. Catal. B Environ. 2018, 222, 35–43. [Google Scholar] [CrossRef]
  40. Zhang, Q.; Li, Y.; Zhong, J.; Li, J. Facile construction of CuO/g-C3N4 heterojunctions with promoted photocatalytic hydrogen generation behaviors. Fuel 2023, 353, 129224. [Google Scholar] [CrossRef]
  41. Li, Z.; Meng, X.; Zheng, Z. Rencent development on MoS2-based photocatalysis: A review. J. Photoch. Photobio. C. Photochem. Rev. 2018, 35, 39–55. [Google Scholar] [CrossRef]
  42. Khan, K.; Tao, X.; Shi, M.; Zeng, B.; Feng, Z.; Li, C.; Li, R. Visible-light-driven photocatalytic hydrogen production on Cd0.5Zn0.5S nanorods with an apparent quantum efficiency exceeding 80%. Adv. Funct. Mater. 2020, 30, 2003731. [Google Scholar] [CrossRef]
  43. Zhang, S.; Chen, Q.; Jing, D.; Wang, Y.; Guo, L. Visible photoactivity and antiphotocorrosion performance of PdS-CdS photocatalysts modified by polyaniline. Int. J. Hydrog. Energ. 2012, 37, 791–796. [Google Scholar] [CrossRef]
  44. Sun, Q.; Wang, N.; Yu, J.; Yu, J.C. A hollow porous CdS photocatalyst. Adv. Mater. 2018, 30, 1804368. [Google Scholar] [CrossRef] [PubMed]
  45. Li, X.-L.; Yang, G.-Q.; Li, S.-S.; Xiao, N.; Li, N.; Gao, Y.-Q.; Lv, D.; Ge, L. Novel dual co-catalysts decorated Au@HCS@PdS hybrids with spatially separated charge carriers and enhanced photocatalytic hydrogen evolution activity. Chem. Eng. J. 2020, 379, 122350. [Google Scholar] [CrossRef]
  46. Ge, L.; Zuo, F.; Liu, J.; Ma, Q.; Wang, C.; Sun, D.; Bartels, L.; Feng, P. Synthesis and efficient visible light photocatalytic hydrogen evolution of polymeric g-C3N4 coupled with CdS quantum dots. J. Phys. Chem. C 2012, 116, 13708–13714. [Google Scholar] [CrossRef]
  47. Reddy, B.M.; Reddy, E.P.; Srinivas, S. Dispersion and activity of molybdena-alumina catalysts prepared by impregnation and solid/solid wetting methods. J. Catal. 1992, 136, 50–58. [Google Scholar] [CrossRef]
  48. Meng, J.; Yu, Z.; Li, Y.; Li, Y. PdS-modified CdS/NiS composite as an efficient photocatalyst for H2 evolution in visible light. Catal. Today 2014, 225, 136–141. [Google Scholar] [CrossRef]
  49. Chen, Z.; Guo, F.; Sun, H.; Shi, Y.; Shi, W. Well-designed three-dimensional hierarchical hollow tubular g-C3N4/ZnIn2S4 nanosheets heterostructure for achieving efficient visible-light photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2022, 607, 1391–1401. [Google Scholar] [CrossRef]
  50. Shi, W.; Wang, J.; Yang, S.; Lin, X.; Guo, F.; Shi, J. Fabrication of a ternary carbon dots/CoO/g-C3N4 nanocomposite photocatalyst with enhanced visible-light-driven photocatalytic hydrogen production. J. Chem. Technol. Biotechnol. 2020, 95, 2129–2138. [Google Scholar] [CrossRef]
  51. Wang, Y.; Zhang, X.; Liu, Y.; Zhao, Y.; Xie, C.; Song, Y.; Yang, P. Crystallinity and phase controlling of g-C3N4/CdS hetrostructures towards high efficient photocatalytic H2 generation. Int. J. Hydrog. Energy 2019, 44, 30151–30159. [Google Scholar] [CrossRef]
  52. Zheng, Y.; Lin, L.; Wang, B.; Wang, X. Graphitic carbon nitride polymers toward sustainable photoredox catalysis. Angew. Chem. Int. Ed. 2015, 54, 12868–12884. [Google Scholar] [CrossRef] [PubMed]
  53. Zhao, H.; Wang, S.; He, F.; Zhang, J.; Chen, L.; Dong, P.; Tai, Z.; Wang, Y.; Gao, H.; Zhao, C. Hydroxylated carbon nanotube/carbon nitride nanobelt composites with enhanced photooxidation and H2 evolution efficiency. Carbon 2019, 150, 340–348. [Google Scholar] [CrossRef]
  54. Makuła, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef]
  55. Wang, W.; Fang, J.; Huang, X. Different behaviors between interband and intraband transitions generated hot carriers on g-C3N4/Au for photocatalytic H2 production. Appl. Surf. Sci. 2020, 513, 145830. [Google Scholar] [CrossRef]
  56. Samanta, S.; Martha, S.; Parida, K. Facile synthesis of Au/g-C3N4 nanocomposites: An inorganic/organic hybrid plasmonic photocatalyst with wnhanced hydrogen gas evolution under visible-light irradiation. ChemCatChem 2014, 6, 1453–1462. [Google Scholar] [CrossRef]
  57. Liu, M.; Xia, P.; Zhang, L.; Cheng, B.; Yu, J. Enhanced photocatalytic H2-production activity of g-C3N4 nanosheets via optimal photodeposition of Pt as cocatalyst. ACS Sustain. Chem. Eng. 2018, 6, 10472–10480. [Google Scholar] [CrossRef]
  58. Yao, Y.; Ren, G.; Li, Z.; Bai, H.; Hu, X.; Meng, X. Nitrogen vacancy-induced deposition of Pd nanoparticles onto g-C3N4 with greatly improved photocatalytic activity in H2 evolution. Sol. RRL 2021, 5, 2100145. [Google Scholar] [CrossRef]
  59. Xu, Q.; Zhu, B.; Cheng, B.; Yu, J.; Zhou, M.; Ho, W. Photocatalytic H2 evolution on graphdiyne/g-C3N4 hybrid nanocomposites. Appl. Catal. B Environ. 2019, 255, 117770. [Google Scholar] [CrossRef]
  60. Yuan, Y.-J.; Shen, Z.; Wu, S.; Su, Y.; Pei, L.; Ji, Z.; Ding, M.; Bai, W.; Chen, Y.; Yu, Z.-T.; et al. Liquid exfoliation of g-C3N4 nanosheets to construct 2D-2D MoS2/g-C3N4 photocatalyst for enhanced photocatalytic H2 production activity. Appl. Catal. B Environ. 2019, 246, 120–128. [Google Scholar] [CrossRef]
  61. Das, D.; Shinde, S.L.; Nanda, K.K. Temperature-dependent photoluminescence of g-C3N4: Implication for temperature sensing. ACS Appl. Mater. Interfaces 2016, 8, 2181–2186. [Google Scholar] [CrossRef]
  62. Yan, H.; Yang, J.; Ma, G.; Wu, G.; Zong, X.; Lei, Z.; Shi, J.; Li, C. Visible-light-driven hydrogen production with extremely high quantum efficiency on Pt–PdS/CdS photocatalyst. J. Catal. 2009, 266, 165–168. [Google Scholar] [CrossRef]
  63. Ferrer, I.J.; Díaz-Chao, P.; Pascual, A.; Sánchez, C. An investigation on palladium sulphide (PdS) thin films as a photovoltaic material. Thin Solid Film. 2007, 515, 5783–5786. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of g-C3N4 and g-C3N4@PdS nanocomposites. The illustration is a magnification of the (002) diffraction peak.
Figure 1. XRD pattern of g-C3N4 and g-C3N4@PdS nanocomposites. The illustration is a magnification of the (002) diffraction peak.
Molecules 29 00493 g001
Figure 2. FTIR transmission spectra of g-C3N4 and g-C3N4 varying the PdS concentration.
Figure 2. FTIR transmission spectra of g-C3N4 and g-C3N4 varying the PdS concentration.
Molecules 29 00493 g002
Figure 3. TEM morphology of (a) g-C3N4, (b) g-C3N4@PdS-1, (c) g-C3N4@PdS-3 and (d) g-C3N4@PdS-4 nanocomposites. The inset in (c) is the corresponding high-magnification TEM morphology.
Figure 3. TEM morphology of (a) g-C3N4, (b) g-C3N4@PdS-1, (c) g-C3N4@PdS-3 and (d) g-C3N4@PdS-4 nanocomposites. The inset in (c) is the corresponding high-magnification TEM morphology.
Molecules 29 00493 g003
Figure 4. Element mapping of g-C3N4@PdS-3 nanocomposites: (a) high-angle annular dark-field STEM image, (b) C element, (c) N element, (d) Pd element, (e) S element, and (f) overlay diagram of element mapping results.
Figure 4. Element mapping of g-C3N4@PdS-3 nanocomposites: (a) high-angle annular dark-field STEM image, (b) C element, (c) N element, (d) Pd element, (e) S element, and (f) overlay diagram of element mapping results.
Molecules 29 00493 g004
Figure 5. XPS spectra of g-C3N4, g-C3N4@PdS-2, g-C3N4@PdS-3, and g-C3N4@PdS-4: (a) survey spectra, (b) C1s, (c) N 1s, (d) Pd 3d, (e) S 2p, and (f) valence band.
Figure 5. XPS spectra of g-C3N4, g-C3N4@PdS-2, g-C3N4@PdS-3, and g-C3N4@PdS-4: (a) survey spectra, (b) C1s, (c) N 1s, (d) Pd 3d, (e) S 2p, and (f) valence band.
Molecules 29 00493 g005
Figure 6. (a) UV-vis spectra, and (b) the corresponding Taus plots of the g-C3N4 and g-C3N4@PdS nanocomposites.
Figure 6. (a) UV-vis spectra, and (b) the corresponding Taus plots of the g-C3N4 and g-C3N4@PdS nanocomposites.
Molecules 29 00493 g006
Figure 7. The relationship of the amount of photocatalytic H2 evolution with time under the irradiation of a xenon lamp light source: (a) with long pass filter (λ > 420 nm) and (b) without filter; (c) the photocatalytic H2 evolution rates of g-C3N4@Pt and g-C3N4@PdS@Pt photocatalysts with various loading amounts of PdS under visible light and full spectrum irradiation; (d) cycling stability test of the photocatalytic H2 evolution for g-C3N4@PdS@Pt-3 under full spectrum irradiation.
Figure 7. The relationship of the amount of photocatalytic H2 evolution with time under the irradiation of a xenon lamp light source: (a) with long pass filter (λ > 420 nm) and (b) without filter; (c) the photocatalytic H2 evolution rates of g-C3N4@Pt and g-C3N4@PdS@Pt photocatalysts with various loading amounts of PdS under visible light and full spectrum irradiation; (d) cycling stability test of the photocatalytic H2 evolution for g-C3N4@PdS@Pt-3 under full spectrum irradiation.
Molecules 29 00493 g007
Figure 8. The UV-vis absorption spectrum and AQE data of the g-C3N4@PdS@Pt-3 catalyst.
Figure 8. The UV-vis absorption spectrum and AQE data of the g-C3N4@PdS@Pt-3 catalyst.
Molecules 29 00493 g008
Figure 9. (a) PL spectra of g-C3N4, g-C3N4@PdS-1, g-C3N4@PdS-2, g-C3N4@PdS-3, and g-C3N4@PdS-4 nanocomposites; (b) TRPL decay curves of g-C3N4 and g-C3N4@PdS-3 nanocomposites.
Figure 9. (a) PL spectra of g-C3N4, g-C3N4@PdS-1, g-C3N4@PdS-2, g-C3N4@PdS-3, and g-C3N4@PdS-4 nanocomposites; (b) TRPL decay curves of g-C3N4 and g-C3N4@PdS-3 nanocomposites.
Molecules 29 00493 g009
Figure 10. (a) Transient photocurrent responses spectra under Xe lamp, (b) EIS Nyquist spectra of g-C3N4 and g-C3N4@PdS-3 nanocomposites.
Figure 10. (a) Transient photocurrent responses spectra under Xe lamp, (b) EIS Nyquist spectra of g-C3N4 and g-C3N4@PdS-3 nanocomposites.
Molecules 29 00493 g010
Figure 11. LSV curves of the g-C3N4 and g-C3N4@PdS nanocomposites.
Figure 11. LSV curves of the g-C3N4 and g-C3N4@PdS nanocomposites.
Molecules 29 00493 g011
Figure 12. Proposed photocatalytic mechanism for the g-C3N4@PdS@Pt nanocomposites.
Figure 12. Proposed photocatalytic mechanism for the g-C3N4@PdS@Pt nanocomposites.
Molecules 29 00493 g012
Scheme 1. Schematic depiction of the synthesis process for g-C3N4@PdS nanocomposites.
Scheme 1. Schematic depiction of the synthesis process for g-C3N4@PdS nanocomposites.
Molecules 29 00493 sch001
Table 1. Comparative survey of photocatalytic H2 evolution performance: g-C3N4@PdS@Pt-3 versus other reported photocatalysts.
Table 1. Comparative survey of photocatalytic H2 evolution performance: g-C3N4@PdS@Pt-3 versus other reported photocatalysts.
CatalystSynthesis MethodDosage (mg)Type of Light SourceSacrificial ReagentH2 Evolution Rate (μmol·g−1·h−1)Refs.
Au@g-C3N4Solution–precipitation method20365 nm wavelength light excitation10% Triethanolamine530[55]
Au@g-C3N4Facile deposition–precipitation method20125W medium pressure visible-light Hg Lamp 10% Triethanolamine177.4[56]
Pt@g-C3N4Photodeposition of Pt8300 W xenon Lamp8 mL of TEOA solution4210.8[57]
Pd-NVs-C3N4Photoreduction method100300 W xenon Lamp20 vol% methanol 287.9[58]
graphdiyne/g-C3N4Calcination method50300 W xenon Lamp (λ > 420 nm)15% Triethanolamine39.6[59]
MoS2@g-C3N4Probe sonication-assisted liquid exfoliation method50300 W xenon Lamp (λ > 420 nm)0.1M Triethanolamine1155[60]
g-C3N4@PdS@Pt-3Precipitation method30300 W Xe arc Lamp (λ > 420 nm)20% lactic acid aqueous solution1289This work
300 W Xe arc Lamp11,438
Table 2. Dynamic analysis of emission decay for g-C3N4, andg-C3N4@PdS-3 nanocomposites.
Table 2. Dynamic analysis of emission decay for g-C3N4, andg-C3N4@PdS-3 nanocomposites.
Sampleτ1 (ns)A1 (%)τ2 (ns)A2 (%)τ (ns)
g-C3N42.171457.519.602342.497.8609
g-C3N4@PdS-31.500452.846.845547.165.7917
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ma, L.; Lin, C.; Jiang, W.; Yan, S.; Jiang, H.; Song, X.; Ai, X.; Cao, X.; Ding, Y. Achieving Highly Efficient Photocatalytic Hydrogen Evolution through the Construction of g-C3N4@PdS@Pt Nanocomposites. Molecules 2024, 29, 493. https://doi.org/10.3390/molecules29020493

AMA Style

Ma L, Lin C, Jiang W, Yan S, Jiang H, Song X, Ai X, Cao X, Ding Y. Achieving Highly Efficient Photocatalytic Hydrogen Evolution through the Construction of g-C3N4@PdS@Pt Nanocomposites. Molecules. 2024; 29(2):493. https://doi.org/10.3390/molecules29020493

Chicago/Turabian Style

Ma, Ligang, Chao Lin, Wenjun Jiang, Shun Yan, Huilin Jiang, Xiang Song, Xiaoqian Ai, Xiaoxiao Cao, and Yihuan Ding. 2024. "Achieving Highly Efficient Photocatalytic Hydrogen Evolution through the Construction of g-C3N4@PdS@Pt Nanocomposites" Molecules 29, no. 2: 493. https://doi.org/10.3390/molecules29020493

Article Metrics

Back to TopTop