Next Article in Journal
Analysis of the Equilibrium Distribution of Ligands in Heterogeneous Media–Approaches and Pitfalls
Previous Article in Journal
A Combination of a Genome-Wide Association Study and a Transcriptome Analysis Reveals circRNAs as New Regulators Involved in the Response to Salt Stress in Maize
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cd0.9Co0.1S Nanorods with an Internal Electric Field and Photothermal Effect Synergistically for Boosting Photocatalytic H2 Evolution

1
Henan Key Laboratory of Function Oriented Porous Materials, College of Chemistry and Chemical Engineering, Luoyang Normal University, Luoyang 471934, China
2
Green Catalysis Center, College of Chemistry, Zhengzhou University, Zhengzhou 450001, China
3
School of Materials and Environmental Engineering, Chengdu Technological University, Chengdu 611730, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(17), 9756; https://doi.org/10.3390/ijms23179756
Submission received: 10 August 2022 / Revised: 23 August 2022 / Accepted: 25 August 2022 / Published: 28 August 2022
(This article belongs to the Section Molecular Nanoscience)

Abstract

:
The paper reports a strategy to synthesize Cd0.9Co0.1S nanorods (NRs) via a one-pot solvothermal method. Remarkably, the pencil-shaped Cd0.9Co0.1S NRs with a large aspect ratio and good polycrystalline plane structure significantly shorten the photogenerated carrier transfer path and achieve fast separation. An appropriate amount of Co addition enhances visible light-harvesting and generates a photothermal effect to improve the surface reaction kinetics and increases the charge transfer rate. Moreover, the internal electric field facilitates the separation and transfer of carriers and effectively impedes their recombination. As a result, the optimized Cd0.9Co0.1S NRs yield a remarkable H2 evolution rate of 8.009 mmol·g−1·h−1, which is approximately 7.2 times higher than that of pristine CdS. This work improves the photocatalytic hydrogen production rate by tuning and optimizing electronic structures through element addition and using the photothermal synergistic effect.

1. Introduction

Hydrogen is essential in the clean energy system [1,2,3,4]. Semiconductor-based photocatalytic hydrogen production is promising to alleviate increasing energy and pollution problems [5,6,7,8,9]. Many metal sulfide-based semiconductors have been discovered for photocatalytic hydrogen evolution. Represented by CdS, exhibiting good visible light response, narrow bandgap, and morphology controllability, it is considered one of the most promising photocatalytic materials. However, because of the severe photo-corrosion and rapid recombination of the photogenerated carrier of pure CdS, its proton-reduction efficiency is limited, resulting in poor photocatalytic performance [10,11,12].
Hence, several researchers have reported various strategies to elevate the photocatalytic performance of CdS-based materials. First of all, CdS with multiple shapes, such as nanoparticles [13], nanotubes [14], nanospheres [15], and nanorods [16] have been found in many studies and showed excellent photocatalytic performance. Among them, the CdS nanorods have attracted more and more attention because of their large aspect ratio and fewer defects, which can significantly shorten the radial transfer path of carriers and achieve rapid charge separation, thus exhibiting high photocatalytic hydrogen evolution performance [17,18]. In addition, several other methods have been applied, such as constructing an intrinsic built-in electric field, defect and heterostructure engineering, and co-catalyst decoration to accelerate the separation and transfer of photogenerated electron-hole pairs [19,20,21,22]. Establishing the internal electric field in semiconductors by element doping can promote the separation and transfer of photogenerated electrons, and extract carriers, thus improving the photocatalytic efficiency. For example, the Huang group prepared a gradient P-doped CdS nanostructure via surface diffusion doping strategy [23] the Hu group reported gradient-P-doped CdS/CoP hybrid nanorods [24], and the Shangguan group explored Ni, Mo, Co-doped CdS for efficient photocatalytic hydrogen evolution from water splitting [25].
It is well known that photothermal conversion is an attractive strategy for solar light-harvesting, and the local temperature can go up in situ, which generally enhances the entropy of chemical reactions and enhances the kinetics of surface catalytic reactions to accelerate the transfer of photogenerated charges, thus increasing the photocatalytic reaction rate [26,27,28,29,30]. Especially in water splitting that involves a brutal thermodynamic uphill reaction, heat as an external force can effectively drive water photocatalytic hydrogen production [31]. The photothermal synergistic catalytic system that couples photochemical and thermochemical effects shows better performance than the traditional single catalytic system [32]. Therefore, using the photothermal effect to enhance CdS-based performance should be an excellent strategy.
Herein, pencil-tipped CdCoS NRs were prepared by a one-pot method, as presented in Scheme 1. In this work, we added an appropriate amount of Co into CdS NRs, which narrowed the pure CdS bandgap and significantly enhanced light-harvesting ability, achieving rapid charge separation efficiency. Meanwhile, element addition realized the construction of the internal electric field and the photothermal synergistic. Finally, the optimized Cd0.9Co0.1S NRs photocatalyst exhibits excellent hydrogen production efficiency under visible light illumination.

2. Results and Discussion

2.1. Structure and Morphology Analysis of Photocatalysts

XRD analyzed the phase and crystal structure of the synthesized samples. As presented in Figure 1a, hexagonal phase CdS existed in all samples. The XRD pattern of pure CdS displayed several obvious peaks at 24.8, 26.5, 28.1, 36.6, 43.6, 47.8, and 51.8°, attributed to the (100), (002), (101), (102), (110), (103), and (112) planes, respectively (PFD#41-1049) [16,33]. After Co addition, the diffraction peak shifted of Cd0.9Co0.1S NRs relative to pure CdS, and there was no obvious peak of cobalt sulfide present, indicating the change of its crystal structure. Specifically, the partially magnified XRD pattern of samples showed that the main diffraction peak positions on the CdCoS samples were slightly shifted to a higher 2θ value, which means that the lattice parameter changes [34]. The shift can be attributed to the successful incorporation of Co2+ into CdS because the effective ionic radius of Co2+ is smaller than that of Cd2+ [35,36]. Furthermore, the contents of Cd, Co, and S of the obtained CdCoS photocatalysts were further determined by ICP-MS. As presented in Table S1, the molar ratios of Co/Cd were 0.103, 0.228, and 0.319, corresponding to the samples Cd0.9Co0.1S, Cd0.8Co0.2S, and Cd0.7Co0.3S, respectively.
The SEM and TEM images give detailed information on the microscopic structures and morphologies of Cd0.9Co0.1S NRs. As the SEM images (Figure 2a,b) revealed, Cd0.9Co0.1S NRs displayed a clear crystal structure with a regular polyhedral shape. The TEM images (Figure 2c–f) of Cd0.9Co0.1S showed the nanorod-like system. It can be seen that the crystal surface is smooth after Co addition, and no obvious Co nanoparticles are found. Meanwhile, as the SAED pattern (Figure 2g) and HRTEM images (Figure 2h,i) presented, the corresponding lattice diffraction pattern of Cd0.9Co0.1S NRs with equal lattice diffraction spacing was well matched with the (101) plane of the CdS phase. The lattice fringes with interlayer distances of 0.21 and 0.24 nm were assigned to the (110) and (102) planes [37], respectively, of pristine CdS (PDF#41-1049). In addition, elemental mapping results (Figure 2j–m) and energy dispersive spectroscopy (EDS) spectra (Figure S1) of an individual Cd0.9Co0.1S NRs sample further confirm the existence of S, Cd, and Co elements in the photocatalyst, indicating the Co has been added successfully.
XPS studied the surface chemical state and internal elemental bond composition of samples. Figure 3a displayed the full-survey spectrum of Cd0.9Co0.1S NRs, indicating the existence of Cd, S, C, O, and Co elements. The S 2p spectrum of Cd0.9Co0.1S NRs in Figure 3b showed two binding energies at 160.8 and 162.0 eV, ascribed to S 2p3/2 and S 2p1/2, respectively. Simultaneously, as presented in Figure 3c, the Cd 3d spectrum of Cd0.9Co0.1S NRs can be deconvoluted into Cd 3d5/2 (404.5 eV) and Cd 3d3/2 (411.3 eV) [31,38]. Significantly, compared with pure CdS, the two peaks of S 2p in Cd0.9Co0.1S NRs shifted to the low binding energy region. Conversely, the characteristic peaks of Cd 3d shifted slightly to a higher position, which indicates that a fraction of Co atoms has substituted Cd sites to form the Co-S bond, resulting in the redistribution of charge around Cd and S atoms [37,39]. Moreover, the Co 2p spectrum (Figure 3d) of Cd0.9Co0.1S NRs can be deconvoluted into four peaks. The peaks located at 781.1 and 786.1 eV and at 796.4 and 802.2 eV corresponded to Co 2p3/2 and Co 2p1/2, respectively [40,41]. Such results further indicated that Co had been successfully added into the composites.

2.2. Photocatalytic H2 Evolution Performance

Photocatalytic H2 evolution performance was evaluated in the prepared samples. As demonstrated in Figure 4a,b, pure CdS exhibited the lowest H2 evolution rate of 1.12 mmol·g−1·h−1, gradually declining in the reaction. After Co addition, the H2 evolution rate of all CdCoS NRs was significantly enhanced. In particular, the hydrogen evolution rate of Cd0.9Co0.1S NRs reached the highest value of 8.009 mmol·g−1·h−1 (Video S1), which is about 7.2 times that of pure CdS. It surpassed most of the other CdS-based photocatalysts previously reported (Table S2). With the increase of Co addition, the hydrogen evolution rate gradually reduced because excessive added elements may become new carrier recombination centers, which hindered the proton-reduction process [42]. Therefore, the appropriate amount of Co addition is essential to enhance photocatalytic activity significantly. The apparent quantum efficiency (AQE) is an important measure of photocatalytic hydrogen production performance, equal to twice the number of evolved H2 molecules divided by the number of incident photons. The AQE of Cd0.9Co0.1S NRs under different monochromatic light irradiations were calculated. As shown in Figure 4c, it was 1.77, 9.60, and 3.80% at 365, 420, and 700 nm, respectively, which indicated that the photocatalyst had a modest utilization efficiency of light. It confirmed the crucial role of the photothermal synergistic effect in significantly improving the H2 generation performance of Cd0.9Co0.1S NRs.
Moreover, the photocatalytic H2 production stability of Cd0.9Co0.1S NRs was evaluated through a 20 h cyclic photocatalytic test, with a total of five cycles, each cycle of 4 h, and the results were shown in Figure 4d. In the first three cycles, the hydrogen production of the sample remained unchanged. In the 5th cycle reaction, the hydrogen evolution rate of Cd0.9Co0.1S NRs has decreased. However, it was still significantly higher than pure CdS, possibly due to the loss of photocatalyst quality and sacrificial agent in the reaction process. Therefore, we supplemented the sacrificial agent and photocatalyst after five cycles of stability testing. The results (Figure S4a) showed that the hydrogen production performance increased after the supplement but decreased compared to the initial examination.
It is known that the photo-corrosion of CdS is one of the main reasons affecting its performance. When the CdS is photo-corroded, the Cd2+ will be dispersed into the dispersion solution. To clarify the degree of photo-corrosion of the photocatalyst during the reaction, the Cd2+ concentration in the reaction solution of each run was monitored by ICP. As displayed in Figure S4b, the Cd2+ concentration in pure CdS increases rapidly with the progress of the reaction. Meanwhile, it can be seen from the XPS spectra (Figure S5) after the cyclic response that there is a small amount of S0 in the S 2p spectrum of pure CdS, which was attributed to the oxidation of divalent sulfur, indicating that obvious photo-corrosion occurred in pure CdS during the reaction process.
In comparison, the Cd2+ concentration in Cd0.9Co0.1S NRs has increased but was much lower than pure CdS. This result showed that Cd0.9Co0.1S NRs have a little photo-corrosion under long-time illumination, leading to its performance decline. However, compared with pure CdS, Co addition reduced photo-corrosion and improved the stability of the photocatalyst. Furthermore, the SEM images (Figure S2) and XRD spectra (Figure S6) of Cd0.9Co0.1S NRs showed that the morphology and phase of the photocatalyst did not change after cyclic tests, indicating that the photocatalyst had good stability.

2.3. Photothermal Effect

Temperature is one of the influential critical factors in photocatalytic reaction, and an appropriate temperature increase is conducive to photo-catalytic response [43,44]. To verify the photothermal synergistic effect of the photocatalytic, the surface temperature changes of pristine CdS and Cd0.9Co0.1S NRs under visible light irradiation were tracked in situ. As illustrated in Figure 5a,b, the surface temperatures of Cd0.9Co0.1S NRs increased rapidly in 300 s and reached 50.1 °C from 25.0 °C, which was higher than that of pristine CdS. Compared with pure CdS, the Cd0.9Co0.1S NRs exhibited better photothermal properties, indicating that the Co-S bond formed after proper Co addition played a crucial role in the photothermal effect of photocatalytic reaction [45,46]. In the photothermal phenomenon, as the conversion from light to heat, high temperature can generally accelerate the transfer of charge carriers and surface reaction kinetics, thus improving the photocatalytic performance [47,48,49]. To investigate the effect of temperature on photocatalytic H2 generation, the photocatalytic H2 evolution rate of Cd0.9Co0.1S NRs at different temperatures was measured. As shown in Figure 5c, the H2 evolution rate declined sharply with the decrease in reaction temperature, indicating that the excellent photothermal effect of Cd0.9Co0.1S NRs played an essential role in its better photocatalytic performance.

2.4. Optical Property and Band Structure

The UV-vis absorption spectra of the samples were measured to analyze their light-absorption properties. As displayed in Figure 6a, the pure CdS has an inherent bandgap (Eg) adsorption with a band edge of about 520 nm. Compared with pure CdS, with the increase of Co addition, the CdCoS NRs exhibited better light absorption capacity in the wavelength range of 230–520 nm, and all CdS-based photocatalysts had a redshift. Notably, CdCoS NRs showed an obvious tail absorbance in the field of 550–800 nm. It speculates that the tail absorbance in the 550–800 nm field is mainly caused by the vibration of Co atoms in the lattice. As the band gap of Cd0.7Co0.3S NRs became wider (Figure 6b), the absorbance decreased slightly. However, it was still significantly higher than pure CdS, demonstrating that Co addition can enhance the light-harvesting ability of the photocatalysts. Meanwhile, the samples showed a strong absorption peak around 740 nm, which may enable them to have a higher photothermal conversion efficiency.
The appropriate band structure of the catalyst is essential to charge-oriented transfer, with a suitable bandgap, the suitable conduction band (CB) and valance band (VB) edge potentials [50]. As displayed in Figure 6b, the bandgap energy of CdS and Cd0.9Co0.1S NRs calculated based on the Kubelka–Munk method from the Tauc plots were 2.37 and 2.25 eV, respectively. As presented in Figure 6c, the VB value of pure CdS and Cd0.9Co0.1S NRs were 1.24 and 0.80 eV, respectively. The VB value is the energy of the intersection point obtained by extending the straight part near 0 eV and extending the horizontal part less than 0 eV. Moreover, the Mott-Schottky method measured the Efb (flat-band potential) of pristine CdS and Cd0.9Co0.1S NRs. As shown in Figure S3a,b, the Mott–Schottky plots of CdS and Cd0.9Co0.1S NRs showed a positive slope, indicating that semiconductors have n-type characteristics [51], and the flat band potentials (Efb) of CdS and Cd0.9Co0.1S NRs were approximately −1.17 and −1.49 V (vs. SCE), respectively. The measured potentials can be converted to the standard hydrogen electrode (NHE) scale by the Nernst equation [52] (Equation (1)):
E(NHE) = E(SCE) + 0.24
By calculation, the Efb of CdS and Cd0.9Co0.1S NRs were −0.93 and −1.25 V (vs. NHE), respectively. Theoretically, the conduction band position of n-type semiconductors is −0.2 V negative than flat band potential [39,53]. So, the conduction band positions (ECB) of CdS and Cd0.9Co0.1S NRs were −1.13 and −1.45 eV, respectively. The formula Eg = EVB − ECB shows that the calculated results match well with the test results. Therefore, the band energy alignment of CdS and Cd0.9Co0.1S NRs can be determined (Figure 6d). It can be seen that Co addition can narrow the bandgap, and the Cd0.9Co0.1S NRs have a more negative CB, indicating that it has a higher reduction power and is more beneficial to solar hydrogen production [50,54].

2.5. Separation and Transfer of Charge Carriers

A series of photoelectron-chemical measurements were carried out to discuss further the reasons for the excellent photocatalytic performance of Cd0.9Co0.1S NRs. Generally speaking, the photoluminescence (PL) peak’s low intensity indicates the charge carriers’ low recombination efficiency [55]. It can be seen from Figure 7a that the pure CdS showed the strongest PL peak at about 390 nm, which is owing to the rapid recombination of electrons and holes. After Co addition, the P.L. signal of the sample was significantly reduced, which indicated that Co addition could effectively accelerate the charge separation and reduce its recombination efficiency. However, with the increase of Co addition, too much Co formed new active sites, further reducing the recombination of photogenerated carriers, so the sample Cd0.7Co0.3S NRs had the lowest PL signal. In addition, the time-resolved photoluminescence (TRPL) spectra showed the average lifetime of charge carriers of samples under irradiation, and the results are illustrated in Figure 7b and S7. The average lifetimes of carriers were 0.71, 1.25, 0.96, and 0.89 ns for pure CdS, Cd0.9Co0.1S, Cd0.8Co0.2S, and Cd0.7Co0.3S NRs, respectively. The Cd0.9Co0.1S NRs exhibited the longest carrier lifetime, indicating that the Co addition in Cd0.9Co0.1S can significantly improve electron-hole pairs separation efficiency [56], and a long lifetime of the carrier means more opportunities to participate in the hydrogen evolution reactions [57]. The above results indicated that Co addition could effectively capture electron-hole pairs and thus effectively prevent their recombination.
Transient photocurrent tests examined the efficiency of separating the electron-hole pairs. As presented in Figure 7c, Cd0.9Co0.1S NRs showed the highest photocurrent density among all samples, indicating the highest charge separation efficiency. Compared with pristine CdS, an appropriate amount of Co addition can accelerate the separation and transfer of charge carriers. Furthermore, electronic transmission capacity is a significant factor for catalysts, and electrochemical impedance spectroscopy (EIS) was utilized to study the resistance of the electrode/electrolyte interface of different catalysts during the photocatalytic process [58]. As can be seen clearly from Figure 7d, the radius of the Nyquist circle of the Cd0.9Co0.1S NRs was significantly smaller than that of other samples, indicating that it has the minor charge transfer resistance to ensure the fastest charge transfer rate. In summary, the above photochemical characterization results confirm that the Cd0.9Co0.1S NRs can improve charge separation, enhance charge transfer and prolong charge lifetime, thus significantly enhancing the photocatalytic performance of hydrogen evolution.

2.6. Photocatalytic Mechanism

To deeply understand the electronic regulation effect of Co addition strategy in CdS, density functional theory (DFT) calculations were performed. As shown in Figure 8a, the structure model of Cd0.9Co0.1S NRs was constructed. After Co addition, the local electronic structure of Cd0.9Co0.1S NRs was changed. Figure 8b,c showed the optimal configuration and different charge densities, and the charge of Cd0.9Co0.1S NRs also depleted or accumulated in other regions. Specifically, Co atoms caused the loss of surrounding electrons. Meanwhile, the electron density of S atoms increased (the brown and green isosurfaces depict electron accumulation and electron depletion, respectively). Therefore, an internal electric field was built in Cd0.9Co0.1S NRs, which boosted the separation of electrons and holes, and inhibited its recombination [25].
Based on the frontier molecular orbital theory, the region near the Fermi level is essential for photocatalysis [59,60]. As shown in Figure 8c, with Co addition, the bandgap of Cd0.9Co0.1S NRs 2.05 eV was narrower than that of pure CdS 2.12 eV (Figure S8), which helps the absorption of visible light. The simulation results were consistent with the experimental value. The valence band maximum (VBM) and conduction band minimum (CBM) of Cd0.9Co0.1S NRs are located at the same Γ point, indicating that the sample is a direct bandgap semiconductor [49,61]. It is noteworthy that metallic materials can generate electrons and holes through inter-band and intra-band transitions [62]. Therefore, the sample can produce electron-hole pairs more easily under visible light irradiation. Furthermore, the density of states (DOS) and partial density of states (PDOS) of Cd0.9Co0.1S NRs were displayed in Figure 8d. The partial PDOS of Cd0.9Co0.1S NRs was decomposed according to the contribution of different elements. As a result, the VBM is principally contributed by S orbital, and the CBM is mainly contributed by the hybridization of S, Cd, and Co orbitals, thus forming the Co-S bond. After absorbing photons, the Co-S bonds can act as a “bridge” for charge transfer, thus inhibiting the carrier recombination [38]. Figure 8e shows the phonon spectrum of Cd0.9Co0.1S NRs and the contribution of Co and Cd atoms. It can be seen that the Co addition generates new lattice vibration frequencies, which will lead to solar light absorptions. It is consistent with the characterization results of UV-Vis. In addition, the high-frequency vibration is mainly due to the contribution of Co atoms, which will lead to a strong thermal effect, thus increasing the emission of hydrogen and the reaction rate of hydrogen production.
Based on the above analysis, the possible mechanism for the charge carrier excitation and migration processes of Cd0.9Co0.1S NRs photocatalyst can be depicted in Scheme 2. Firstly, Co2+ partially filled with d orbital generates electron donor levels in the bandgap when it partially enters the CdS lattice, thus narrowing the bandgap. Under visible light irradiation, the narrower bandgap of Cd0.9Co0.1S NRs is more conducive to light-harvesting and generating electrons and holes. Then, Co addition caused spontaneous electron density redistribution in Cd0.9Co0.1S NRs. In detail, Co atoms caused the loss of surrounding electrons, and the electron density of S atoms increased.
Meanwhile, the EVB and Efb of Cd0.9Co0.1S NRs rose, and the band gap decreased. After illumination, due to the generation and migration of carriers, Efb equilibrates and band bends because of the generation and migration of carriers, thus forming an internal electric field. The formed internal electric field can significantly inhibit the recombination of electron-hole pairs and prolong the lifetime of the charge carrier. In addition, the extra energy is converted from chemical energy to thermal energy through photothermal reaction under light irradiation, thereby accelerating the migration rate of carriers in Cd0.9Co0.1S NRs. Eventually, the electrons gathered in the CB of Cd0.9Co0.1S NRs encounter H+ and release H2. The holes accumulated in the VB can obtain enough power to oxidate the sacrificial agent.

3. Materials and Methods

3.1. Characterization

The morphologies were determined by SEM (Tecnai G20) and TEM (FEI Talos F200S). XRD patterns were analyzed using a Bruker D8 Advance X-ray diffractometer. The element content of the samples was measured by ICP-MS (PerkinElmer NexION 300X). XPS spectra were conducted using an ESCALAB 250xi spectrometer. FT-IR spectra were recorded on Nicolet 6700 infrared spectrometer. UV-vis diffuse reflectance spectra were recorded on a UH-4150 spectrophotometer, and the bandgap was calculated using the absorption data. Photoluminescence emission spectra were recorded on the Hitachi F-7000 Fluorescence spectrophotometer. The fluorescence lifetime of the sample was measured and exponentially fitted by Edinburgh FLS1000 Photoluminescence Spectrometer. The FLIR monitored infrared thermal images of the sample.

3.2. Synthesis of CdCoS NRs

CdCoS NRs were synthesized via a one-pot solvothermal method. In a typical synthesis, firstly, 9 mmol cadmium acetate (Cd(CH3COO)2·2H2O), 12.5 mmol thioacetamide, and 1 mmol cobalt nitrate (Co(NO3)2·6H2O) were dissolved in 15 mL of ethylenediamine and 15 mL of deionized water. Then, the mixture solution was transferred into a 50 mL Teflon-lined stainless-steel autoclave and stirred at room temperature for 1 h. Finally, the mixture solution was maintained at 240 °C for 36 h. After suction filtration and vacuum drying, the obtained sample was simply denoted as Cd0.9Co0.1S (Cd/Co = 9:1). Meanwhile, under the same experimental conditions, the molar ratio of Cd(CH3COO)2·2H2O and Co(NO3)2·6H2O was changed to 8:2 and 7:3 respectively, and the obtained samples were named as Cd0.8Co0.2S (Cd/Co = 8:2) and Cd0.7Co0.3S (Cd/Co = 7:3), respectively.

3.3. Photocatalytic H2 Evolution

The photocatalytic H2 evolution experiment was carried out in a top-irradiated quartz reactor (Labsolar-6A, Perfectlight) with a circulating water system. Typically, 80 mg of photocatalysts were dispersed by ultrasound in a 100 mL Na2SO3 (0.35 M) and Na2S (0.25 M) mixture [63]. Before performing the photocatalytic experiments, the reaction vessel was purged with N2 and evacuated for at least 30 min to remove dissolved air. The 300 W Xe-lamp was used as a light source. GC-9790 (Ar as the carrier) gas chromatograph tested the separated hydrogen gas with a thermal conductive detector. Under the same conditions, the blank experiment revealed no appreciable H2 evolution without photocatalysts.

3.4. The Apparent Quantum Efficiency Analysis

The apparent quantum efficiency (AQE) was performed under the same reaction conditions except using various band-pass filters (365, 420, and 700 nm) to obtain the monochromatic light, and the following equation calculated the value of AQE:
A Q E = 2 × t h e   n u m b e r   o f   e v o l v e d   H 2   m o l e c u l e s t h e   n u m b e r   o f   i n c i d e n t   p h o t o n s × 100 %
Details on experimental materials, characterization, photocatalytic H2 evolution, and photoelectrochemical measurements are shown in the Supporting Information.

3.5. Theoretical Calculation

The Vienna Ab-initio Simulation Package (VASP) [64,65,66] based on the projector augmented wave (PAW) method was carried out for the first principle calculations. All material configurations were geometrically optimized by relaxing the ions and lattice with the Perdew–Burke–Ernzerhof (PBE) parameterization and Monkhorst–Pack k-meshes of 6 × 6 × 8 until the maximum residual ionic force is less than 10−4 eV. The cutoff energy of the plane-wave was set to 400 eV. The accurate band structures and density of occupied states were obtained using Heyd–Scuseria–Ernzerhof (HSE06) hybrid density functional [67]. The second interatomic force constants and phono structure were obtained using the Phonopy package [68] with 2 × 2 × 2 supercells. The Vaspkit software package [69] was adopted to analyze the band structures and density of occupied states, and the Vesta software [70] was used to visualize the calculation results.

4. Conclusions

In summary, through a one-pot strategy, we successfully prepared Cd0.9Co0.1S NRs with excellent H2 production performance (8.009 mmol·g−1·h−1). The pencil-shaped Cd0.9Co0.1S NRs can shorten the radial transfer path of electrons and realize rapid charge separation. Meanwhile, the Co addition significantly enhanced visible light-harvesting, reduced the bandgap, and generated a photothermal synergy to improve the charge transfer rate and surface reaction kinetics. Moreover, the internal electric field accelerates the separation and migration of charge carriers and effectively inhibits the recombination of electrons and holes. This work may provide an optimal platform to design and construct stable nano photocatalysts for solar energy conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23179756/s1. References [71,72,73,74,75,76,77,78,79,80,81] are cited in the supplementary materials.

Author Contributions

L.Z.: Conceptualization, Validation, Methodology, Writing—review & editing. M.H.: Data curation, Investigation, Writing—original draft. K.Z.: Visualization, Investigation. B.L.: Investigation, Software, Validation. H.F.: Data curation. X.F.: Supervision, Project administration, Resources. X.X.: Software; Investigation; Visualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the program for science & technology innovation team in universities of Henan province grant number [21IRTSTHN004] and the national natural science foundation of China grant number [U1804131] and the key scientific research projects of universities in Henan province grant number [22A430032].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the Program for Science & Technology Innovation Team in Universities of Henan Province (No. 21IRTSTHN004). The National Natural Science Foundation of China (No. U1804131), and Key scientific research projects of universities in Henan Province (No. 22A430032).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schultz, D.M.; Yoon, T.P. Solar synthesis: Prospects in visible light photocatalysis. Science 2014, 343, 1239176. [Google Scholar] [CrossRef] [PubMed]
  2. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012, 488, 294–303. [Google Scholar] [CrossRef] [PubMed]
  3. Uddin, N.; Zhang, H.; Du, Y.; Jia, G.; Wang, S.; Yin, Z. Structural-Phase Catalytic Redox Reactions in Energy and Environmental Applications. Adv. Mater. 2020, 32, e1905739. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef]
  5. Zhao, G.; Sun, Y.; Zhou, W.; Wang, X.; Chang, K.; Liu, G.; Liu, H.; Kako, T.; Ye, J. Superior Photocatalytic H2 Production with Cocatalytic Co/Ni Species Anchored on Sulfide Semiconductor. Adv. Mater. 2017, 29, 1703258. [Google Scholar] [CrossRef]
  6. Chava, R.K.; Do, J.; Kang, M. Strategy for improving the visible photocatalytic H2 evolution activity of 2D graphitic carbon nitride nanosheets through the modification with metal and metal oxide nanocomponents. Appl. Catal. B-Environ. 2019, 248, 538–551. [Google Scholar] [CrossRef]
  7. Wang, Y.; Zhang, J.; Balogun, M.S.; Tong, Y.; Huang, Y. Oxygen vacancy–based metal oxides photoanodes in photoelectrochemical water splitting. Mater. Today Sustain. 2022, 18, 100118. [Google Scholar] [CrossRef]
  8. Wang, Y.; Chen, D.; Zhang, J.; Balogun, M.-S.; Wang, P.; Tong, Y.; Huang, Y. Charge Relays via Dual Carbon-Actions on Nanostructured BiVO4 for High Performance Photoelectrochemical Water Splitting. Adv. Funct. Mater. 2022, 32, 2112738. [Google Scholar] [CrossRef]
  9. Huang, Y.; Long, B.; Tang, M.; Rui, Z.; Balogun, M.-S.; Tong, Y.; Ji, H. Bifunctional catalytic material: An ultrastable and high-performance surface defect CeO2 nanosheets for formaldehyde thermal oxidation and photocatalytic oxidation. Appl. Catal. B-Environ. 2016, 181, 779–787. [Google Scholar] [CrossRef]
  10. Zhou, P.; Zhang, Q.; Xu, Z.; Shang, Q.; Wang, L.; Chao, Y.; Li, Y.; Chen, H.; Lv, F.; Zhang, Q.; et al. Atomically Dispersed Co-P3 on CdS Nanorods with Electron-Rich Feature Boosts Photocatalysis. Adv. Mater. 2020, 32, e1904249. [Google Scholar] [CrossRef]
  11. Tian, L.; Min, S.; Wang, F. Integrating noble-metal-free metallic vanadium carbide cocatalyst with CdS for efficient visible-light-driven photocatalytic H2 evolution. Appl. Catal. B-Environ. 2019, 259, 118029. [Google Scholar] [CrossRef]
  12. Zhang, P.; Luan, D.; Lou, X.W.D. Fabrication of CdS Frame-in-Cage Particles for Efficient Photocatalytic Hydrogen Generation under Visible-Light Irradiation. Adv. Mater. 2020, 32, e2004561. [Google Scholar] [CrossRef] [PubMed]
  13. Xia, Y.; Cheng, B.; Fan, J.; Yu, J.; Liu, G. Unraveling Photoexcited Charge Transfer Pathway and Process of CdS/Graphene Nanoribbon Composites toward Visible-Light Photocatalytic Hydrogen Evolution. Small 2019, 15, e1902459. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, R.; Chen, S.; Ng, Y.H.; Gao, Q.; Yang, S.; Zhang, S.; Peng, F.; Fang, Y.; Zhang, S. ZnO/CdS/PbS nanotube arrays with multi-heterojunctions for efficient visible-light-driven photoelectrochemical hydrogen evolution. Chem. Eng. J. 2019, 362, 658–666. [Google Scholar] [CrossRef]
  15. Feng, R.; Wan, K.; Sui, X.; Zhao, N.; Li, H.; Lei, W.; Yu, J.; Liu, X.; Shi, X.; Zhai, M.; et al. Anchoring single Pt atoms and black phosphorene dual co-catalysts on CdS nanospheres to boost visible-light photocatalytic H2 evolution. Nano Today 2021, 37, 101080. [Google Scholar] [CrossRef]
  16. Wang, B.; Chen, C.; Jiang, Y.; Ni, P.; Zhang, C.; Yang, Y.; Lu, Y.; Liu, P. Rational designing 0D/1D Z-scheme heterojunction on CdS nanorods for efficient visible-light-driven photocatalytic H2 evolution. Chem. Eng. J. 2021, 412, 128690. [Google Scholar] [CrossRef]
  17. Zhang, M.; Hu, Q.; Ma, K.; Ding, Y.; Li, C. Pyroelectric effect in CdS nanorods decorated with a molecular Co-catalyst for hydrogen evolution. Nano Energy 2020, 73, 104801. [Google Scholar] [CrossRef]
  18. Lu, X.; Chen, W.; Yao, Y.; Wen, X.; Hart, J.N.; Tsounis, C.; Ying Toe, C.; Scott, J.; Ng, Y.H. Photogenerated charge dynamics of CdS nanorods with spatially distributed MoS2 for photocatalytic hydrogen generation. Chem. Eng. J. 2021, 420, 127709. [Google Scholar] [CrossRef]
  19. Ai, Z.; Zhang, K.; Chang, B.; Shao, Y.; Zhang, L.; Wu, Y.; Hao, X. Construction of CdS@Ti3C2@CoO hierarchical tandem p-n heterojunction for boosting photocatalytic hydrogen production in pure water. Chem. Eng. J. 2020, 383, 123130. [Google Scholar] [CrossRef]
  20. Yin, X.-L.; Li, L.-L.; Liu, M.-L.; Li, D.-C.; Shang, L.; Dou, J.-M. MoSx/CdS nano-heterostructures accurately constructed on the defects of CdS for efficient photocatalytic H2 evolution under visible light irradiation. Chem. Eng. J. 2019, 370, 305–313. [Google Scholar] [CrossRef]
  21. Zhang, T.; Meng, F.; Cheng, Y.; Dewangan, N.; Ho, G.W.; Kawi, S. Z-scheme transition metal bridge of Co9S8/Cd/CdS tubular heterostructure for enhanced photocatalytic hydrogen evolution. Appl. Catal. B-Environ. 2021, 286, 119853. [Google Scholar] [CrossRef]
  22. Shi, J.-W.; Sun, D.; Zou, Y.; Ma, D.; He, C.; Ji, X.; Niu, C. Trap-level-tunable Se doped CdS quantum dots with excellent hydrogen evolution performance without co-catalyst. Chem. Eng. J. 2019, 364, 11–19. [Google Scholar] [CrossRef]
  23. Huang, H.; Dai, B.; Wang, W.; Lu, C.; Kou, J.; Ni, Y.; Wang, L.; Xu, Z. Oriented Built-in Electric Field Introduced by Surface Gradient Diffusion Doping for Enhanced Photocatalytic H2 Evolution in CdS Nanorods. Nano Lett. 2017, 17, 3803–3808. [Google Scholar] [CrossRef] [PubMed]
  24. Guo, C.; Li, L.; Chen, F.; Ning, J.; Zhong, Y.; Hu, Y. One-step phosphorization preparation of gradient-P-doped CdS/CoP hybrid nanorods having multiple channel charge separation for photocatalytic reduction of water. J. Colloid Interface Sci. 2021, 596, 431–441. [Google Scholar] [CrossRef]
  25. Yan, J.; Wei, Z.; Xu, M.; Jiang, Z.; Shangguan, W. Polyoxometalate Template-Based Synthetic Strategy to Prepare Ni, Mo Co-Doped CdS for Efficient Photocatalytic Hydrogen Evolution from Water Splitting. Catalysts 2020, 10, 1478. [Google Scholar] [CrossRef]
  26. Guo, S.; Li, X.; Li, J.; Wei, B. Boosting photocatalytic hydrogen production from water by photothermally induced biphase systems. Nat. Commun. 2021, 12, 1343. [Google Scholar] [CrossRef]
  27. Jiang, H.; Xing, Z.; Zhao, T.; Yang, Z.; Wang, K.; Li, Z.; Yang, S.; Xie, L.; Zhou, W. Plasmon Ag nanoparticle/Bi2S3 ultrathin nanobelt/oxygen-doped flower-like MoS2 nanosphere ternary heterojunctions for promoting charge separation and enhancing solar-driven photothermal and photocatalytic performances. Appl. Catal. B-Environ. 2020, 274, 118947. [Google Scholar] [CrossRef]
  28. Li, Y.; Xue, J.; Shen, Q.; Jia, S.; Li, Q.; Li, Y.; Liu, X.; Jia, H. Construction of a ternary spatial junction in yolk–shell nanoreactor for efficient photo-thermal catalytic hydrogen generation. Chem. Eng. J. 2021, 423, 130188. [Google Scholar] [CrossRef]
  29. Guo, M.; Zhao, T.; Xing, Z.; Qiu, Y.; Pan, K.; Li, Z.; Yang, S.; Zhou, W. Hollow Octahedral Cu2-xS/CdS/Bi2S3 p-n-p Type Tandem Heterojunctions for Efficient Photothermal Effect and Robust Visible-Light-Driven Photocatalytic Performance. ACS Appl. Mater. Interfaces 2020, 12, 40328–40338. [Google Scholar] [CrossRef]
  30. Huang, Y.; Lu, Y.; Lin, Y.; Mao, Y.; Ouyang, G.; Liu, H.; Zhang, S.; Tong, Y. Cerium-based hybrid nanorods for synergetic photo-thermocatalytic degradation of organic pollutants. J. Mater. Chem. A 2018, 6, 24740–24747. [Google Scholar] [CrossRef]
  31. Lv, X.; Li, X.; Yang, C.; Ding, X.; Zhang, Y.; Zheng, Y.-Z.; Li, S.; Sun, X.; Tao, X. Large-Size, Porous, Ultrathin NiCoP Nanosheets for Efficient Electro/Photocatalytic Water Splitting. Adv. Funct. Mater. 2020, 30, 1910830. [Google Scholar] [CrossRef]
  32. Ghoussoub, M.; Xia, M.; Duchesne, P.N.; Segal, D.; Ozin, G. Principles of photothermal gas-phase heterogeneous CO2 catalysis. Energ. Environ. Sci. 2019, 12, 1122–1142. [Google Scholar] [CrossRef]
  33. Luo, M.; Liu, Y.; Hu, J.; Liu, H.; Li, J. One-pot synthesis of CdS and Ni-doped CdS hollow spheres with enhanced photocatalytic activity and durability. ACS Appl. Mater. Interfaces 2012, 4, 1813–1821. [Google Scholar] [PubMed]
  34. Deka, K.; Kalita, M.P.C. Structural phase controlled transition metal (Fe, Co, Ni, Mn) doping in CdS nanocrystals and their optical, magnetic and photocatalytic properties. J. Alloys Comp. 2018, 757, 209–220. [Google Scholar] [CrossRef]
  35. Zhang, B.; Chen, C.; Liu, J.; Qiao, W.; Zhao, J.; Yang, J.; Yu, Y.; Chen, S.; Qin, Y. Simultaneous Ni nanoparticles decoration and Ni doping of CdS nanorods for synergistically promoting photocatalytic H2 evolution. Appl. Surf. Sci. 2020, 508, 144869. [Google Scholar] [CrossRef]
  36. Wang, S.; Wang, Y.; Zhang, S.L.; Zang, S.Q.; Lou, X.W.D. Supporting Ultrathin ZnIn2S4 Nanosheets on Co/N-Doped Graphitic Carbon Nanocages for Efficient Photocatalytic H2 Generation. Adv. Mater. 2019, 31, e1903404. [Google Scholar] [CrossRef]
  37. Ning, X.; Wu, Y.; Ma, X.; Zhang, Z.; Gao, R.; Chen, J.; Shan, D.; Lu, X. A Novel Charge Transfer Channel to Simultaneously Enhance Photocatalytic Water Splitting Activity and Stability of CdS. Adv. Func. Mater. 2019, 29, 1902992. [Google Scholar] [CrossRef]
  38. Li, N.; Ding, Y.; Wu, J.; Zhao, Z.; Li, X.; Zheng, Y.Z.; Huang, M.; Tao, X. Efficient, Full Spectrum-Driven H2 Evolution Z-Scheme Co2P/CdS Photocatalysts with Co-S Bonds. ACS Appl. Mater. Interfaces 2019, 11, 22297–22306. [Google Scholar] [CrossRef]
  39. Li, S.; Wang, L.; Li, Y.; Zhang, L.; Wang, A.; Xiao, N.; Gao, Y.; Li, N.; Song, W.; Ge, L.; et al. Novel photocatalyst incorporating Ni-Co layered double hydroxides with P-doped CdS for enhancing photocatalytic activity towards hydrogen evolution. Appl. Catal. B-Environ. 2019, 254, 145–155. [Google Scholar] [CrossRef]
  40. Wu, D.; Wei, Y.; Ren, X.; Ji, X.; Liu, Y.; Guo, X.; Liu, Z.; Asiri, A.M.; Wei, Q.; Sun, X. Co(OH)2 Nanoparticle-Encapsulating Conductive Nanowires Array: Room-Temperature Electrochemical Preparation for High-Performance Water Oxidation Electrocatalysis. Adv. Mater. 2018, 30, 1705366. [Google Scholar] [CrossRef]
  41. Qiu, B.; Zhu, Q.; Du, M.; Fan, L.; Xing, M.; Zhang, J. Efficient Solar Light Harvesting CdS/Co9 S8 Hollow Cubes for Z-Scheme Photocatalytic Water Splitting. Angew. Chem. Int. Ed. Engl. 2017, 56, 2684–2688. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, W.; Wang, Y.; Liu, M.; Gao, L.; Mao, L.; Fan, Z.; Shangguan, W. In situ photodeposition of cobalt on CdS nanorod for promoting photocatalytic hydrogen production under visible light irradiation. Appl. Surf. Sci. 2018, 444, 485–490. [Google Scholar] [CrossRef]
  43. Li, X.; Zhang, H.; Liu, Y.; Duan, X.; Xu, X.; Liu, S.; Sun, H.; Wang, S. Synergy of NiO quantum dots and temperature on enhanced photocatalytic and thermophoto hydrogen evolution. Chem. Eng. J. 2020, 390, 124634. [Google Scholar] [CrossRef]
  44. Li, X.; Lin, J.; Li, J.; Zhang, H.; Duan, X.; Sun, H.; Huang, Y.; Fang, Y.; Wang, S. Temperature-Induced Variations in Photocatalyst Properties and Photocatalytic Hydrogen Evolution: Differences in UV, Visible, and Infrared Radiation. ACS Sustain. Chem. Eng. 2021, 9, 7277–7285. [Google Scholar] [CrossRef]
  45. Jin, B.; Li, Y.; Wang, J.; Meng, F.; Cao, S.; He, B.; Jia, S.; Wang, Y.; Li, Z.; Liu, X. Promoting Oxygen Evolution Reaction of Co-Based Catalysts (Co3O4, CoS, CoP, and CoN) through Photothermal Effect. Small 2019, 15, 1903847. [Google Scholar] [CrossRef]
  46. Zhao, L.; Yang, Q.; Guo, W.; Liu, H.; Ma, T.; Qu, F. Co2.67S4-Based Photothermal Membrane with High Mechanical Properties for Efficient Solar Water Evaporation and Photothermal Antibacterial Applications. ACS Appl. Mater. Interfaces 2019, 11, 20820–20827. [Google Scholar] [CrossRef]
  47. Wang, M.; Zhang, G.; Guan, Z.; Yang, J.; Li, Q. Spatially Separating Redox Centers and Photothermal Effect Synergistically Boosting the Photocatalytic Hydrogen Evolution of ZnIn2 S4 Nanosheets. Small 2021, 17, e2006952. [Google Scholar] [CrossRef]
  48. Lu, L.; Xu, X.; An, K.; Wang, Y.; Shi, F.-N. Coordination Polymer Derived NiS@g-C3N4 Composite Photocatalyst for Sulfur Vacancy and Photothermal Effect Synergistic Enhanced H2 Production. ACS Sustain. Chem. Eng. 2018, 6, 11869–11876. [Google Scholar] [CrossRef]
  49. Wang, K.; Xing, Z.; Meng, D.; Zhang, S.; Li, Z.; Pan, K.; Zhou, W. Hollow MoSe2@Bi2S3/CdS Core-Shell Nanostructure as Dual Z-Scheme Heterojunctions with Enhanced Full Spectrum Photocatalytic-Photothermal Performance. Appl. Catal. B-Environ. 2021, 281, 119482. [Google Scholar] [CrossRef]
  50. Wang, G.; Zhang, T.; Yu, W.; Si, R.; Liu, Y.; Zhao, Z. Modulating Location of Single Copper Atoms in Polymeric Carbon Nitride for Enhanced Photoredox Catalysis. ACS Catalysis 2020, 10, 5715–5722. [Google Scholar] [CrossRef]
  51. Xiao, R.; Zhao, C.; Zou, Z.; Chen, Z.; Tian, L.; Xu, H.; Tang, H.; Liu, Q.; Lin, Z.; Yang, X. In situ fabrication of 1D CdS nanorod/2D Ti3C2 MXene nanosheet Schottky heterojunction toward enhanced photocatalytic hydrogen evolution. Appl. Catal. B-Environ. 2020, 268, 118382. [Google Scholar] [CrossRef]
  52. Yang, X.; Qian, F.; Wang, Y.; Li, M.; Lu, J.; Li, Y.; Bao, M. Constructing a novel ternary composite (C16H33(CH3)3N)4W10O32/g-C3N4/rGO with enhanced visible-light-driven photocatalytic activity for degradation of dyes and phenol. Appl. Catal. B-Environ. 2017, 200, 283–296. [Google Scholar] [CrossRef]
  53. Li, X.; Yu, J.; Low, J.; Fang, Y.; Xiao, J.; Chen, X. Engineering heterogeneous semiconductors for solar water splitting. J. Mater. Chem. A 2015, 3, 2485–2534. [Google Scholar] [CrossRef]
  54. Han, M.; Lu, S.; Qi, F.; Zhu, S.; Sun, H.; Yang, B. Carbon Dots–Implanted Graphitic Carbon Nitride Nanosheets for Photocatalysis: Simultaneously Manipulating Carrier Transport in Inter- and Intralayers. Solar RRL 2020, 4, 2070041. [Google Scholar] [CrossRef]
  55. Cao, S.; Li, H.; Tong, T.; Chen, H.-C.; Yu, A.; Yu, J.; Chen, H.M. Single-Atom Engineering of Directional Charge Transfer Channels and Active Sites for Photocatalytic Hydrogen Evolution. Adv. Funct. Mater. 2018, 28, 1870224. [Google Scholar] [CrossRef]
  56. Yang, W.; Zhang, L.; Xie, J.; Zhang, X.; Liu, Q.; Yao, T.; Wei, S.; Zhang, Q.; Xie, Y. Enhanced Photoexcited Carrier Separation in Oxygen-Doped ZnIn2 S4 Nanosheets for Hydrogen Evolution. Angew. Chem. Int. Ed. Engl. 2016, 55, 6716–6720. [Google Scholar] [CrossRef] [PubMed]
  57. Li, Z.; Hou, J.; Zhang, B.; Cao, S.; Wu, Y.; Gao, Z.; Nie, X.; Sun, L. Two-dimensional Janus heterostructures for superior Z-scheme photocatalytic water splitting. Nano Energy 2019, 59, 537–544. [Google Scholar] [CrossRef]
  58. Zhang, W.; Peng, Q.; Shi, L.; Yao, Q.; Wang, X.; Yu, A.; Chen, Z.; Fu, Y. Merging Single-Atom-Dispersed Iron and Graphitic Carbon Nitride to a Joint Electronic System for High-Efficiency Photocatalytic Hydrogen Evolution. Small 2019, 15, e1905166. [Google Scholar] [CrossRef]
  59. Zhang, L.; Zhang, J.; Xia, Y.; Xun, M.; Chen, H.; Liu, X.; Yin, X. Metal-Free Carbon Quantum Dots Implant Graphitic Carbon Nitride: Enhanced Photocatalytic Dye Wastewater Purification with Simultaneous Hydrogen Production. Int. J. Mol. Sci. 2020, 21, 1052. [Google Scholar] [CrossRef]
  60. Zhang, L.; Zhang, J.; Qi, M.; Guo, X.; Feng, X. First-Principles Investigation of Single-Atom Ni–g-C3N4 as an Efficient Catalyst for Direct Reduction of NO with CO. Energ. Fuel. 2020, 34, 12792–12799. [Google Scholar] [CrossRef]
  61. Shao, B.; Liu, X.; Liu, Z.; Zeng, G.; Liang, Q.; Liang, C.; Cheng, Y.; Zhang, W.; Liu, Y.; Gong, S. A novel double Z-scheme photocatalyst Ag3PO4/Bi2S3/Bi2O3 with enhanced visible-light photocatalytic performance for antibiotic degradation. Chem. Eng. J. 2019, 368, 730–745. [Google Scholar] [CrossRef]
  62. Xu, J.; Ju, Z.; Zhang, W.; Pan, Y.; Zhu, J.; Mao, J.; Zheng, X.; Fu, H.; Yuan, M.; Chen, H.; et al. Efficient Infrared-Light-Driven CO2 Reduction Over Ultrathin Metallic Ni-doped CoS2 Nanosheets. Angew. Chem. Int. Ed. Engl. 2021, 60, 8705–8709. [Google Scholar] [CrossRef]
  63. Kumaravel, V.; Imam, M.D.; Badreldin, A.; Chava, R.K.; Do, J.Y.; Kang, M.; Abdel-Wahab, A. Photocatalytic Hydrogen Production: Role of Sacrificial Reagents on the Activity of Oxide, Carbon, and Sulfide Catalysts. Catalysts 2019, 9, 276. [Google Scholar] [CrossRef]
  64. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  65. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  66. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  67. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef]
  68. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
  69. Wang, V.; Xu, N.; Liu, J.-C.; Tang, G.; Geng, W.-T. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput. Phys. Commun. 2021, 267, 108033. [Google Scholar] [CrossRef]
  70. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  71. Li, X.; Lv, X.; Li, N.; Wu, J.; Zheng, Y.-Z.; Tao, X. One-step hydrothermal synthesis of high-percentage 1T-phase MoS2 quantum dots for remarkably enhanced visible-light-driven photocatalytic H2 evolution. Appl. Catal. B-Environ. 2019, 243, 76–85. [Google Scholar] [CrossRef]
  72. Zhao, Q.; Yao, W.; Huang, C.; Wu, Q.; Xu, Q. Effective and Durable Co Single Atomic Cocatalysts for Photocatalytic Hydrogen Production. ACS Appl. Mater. Interfaces 2017, 9, 42734–42741. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, Y.; Zeng, C.; Ai, L.; Jiang, J. Boosting charge transfer and hydrogen evolution performance of CdS nanocrystals hybridized with MoS2 nanosheets under visible light irradiation. Appl. Surf. Sci. 2019, 484, 692–700. [Google Scholar] [CrossRef]
  74. Liu, Y.; Niu, H.; Gu, W.; Cai, X.; Mao, B.; Li, D.; Shi, W. In-situ construction of hierarchical CdS/MoS2 microboxes for enhanced visible-light photocatalytic H2 production. Chem. Eng. J. 2018, 339, 117–124. [Google Scholar] [CrossRef]
  75. Li, H.; Pan, J.; Zhao, W.; Li, C. The 2D nickel-molybdenum bimetals sulfide synergistic modified hollow cubic CdS towards enhanced photocatalytic water splitting hydrogen production. Appl. Surf. Sci. 2019, 497, 143769. [Google Scholar] [CrossRef]
  76. Guo, J.; Liang, Y.; Liu, L.; Hu, J.; Wang, H.; An, W.; Cui, W. Core-shell structure of sulphur vacancies-CdS@CuS: Enhanced photocatalytic hydrogen generation activity based on photoinduced interfacial charge transfer. J. Colloid Interface Sci. 2021, 600, 138–149. [Google Scholar] [CrossRef] [PubMed]
  77. Ruan, D.; Fujitsuka, M.; Majima, T. Exfoliated Mo2C nanosheets hybridized on CdS with fast electron transfer for efficient photocatalytic H2 production under visible light irradiation. Appl. Catal. B-Environ. 2020, 264, 118541. [Google Scholar] [CrossRef]
  78. Qin, Y.; Li, H.; Lu, J.; Meng, F.; Ma, C.; Yan, Y.; Meng, M. Nitrogen-doped hydrogenated TiO2 modified with CdS nanorods with enhanced optical absorption, charge separation and photocatalytic hydrogen evolution. Chem. Eng. J. 2020, 384, 123275. [Google Scholar] [CrossRef]
  79. Zhu, Y.; Chen, J.; Shao, L.; Xia, X.; Liu, Y.; Wang, L. Oriented facet heterojunctions on CdS nanowires with high photoactivity and photostability for water splitting. Appl. Catal. B-Environ. 2020, 268, 118744. [Google Scholar] [CrossRef]
  80. Yin, X.L.; Li, L.L.; Li, D.C.; Wei, D.H.; Hu, C.C.; Dou, J.M. Room temperature synthesis of CdS/SrTiO3 nanodots-on-nanocubes for efficient photocatalytic H2 evolution from water. J. Colloid Interface Sci. 2019, 536, 694–700. [Google Scholar] [CrossRef]
  81. Luo, X.; Ke, Y.; Yu, L.; Wang, Y.; Homewood, K.P.; Chen, X.; Gao, Y. Tandem CdS/TiO2(B) nanosheet photocatalysts for enhanced H2 evolution. Appl. Surf. Sci. 2020, 515, 145970. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration of the synthetic process for Cd0.9Co0.1S nanorods structure.
Scheme 1. Schematic illustration of the synthetic process for Cd0.9Co0.1S nanorods structure.
Ijms 23 09756 sch001
Figure 1. The XRD patterns (a) and partially magnified XRD patterns of pure CdS and CdCoS NRs (b).
Figure 1. The XRD patterns (a) and partially magnified XRD patterns of pure CdS and CdCoS NRs (b).
Ijms 23 09756 g001
Figure 2. SEM images (a,b) and TEM images (cf), SAED pattern (g), HRTEM images (h,i), and elemental mappings (jm) of the Cd0.9Co0.1S NRs.
Figure 2. SEM images (a,b) and TEM images (cf), SAED pattern (g), HRTEM images (h,i), and elemental mappings (jm) of the Cd0.9Co0.1S NRs.
Ijms 23 09756 g002
Figure 3. XPS survey spectra of Cd0.9Co0.1S NRs (a), and high-resolution XPS spectra of: S 2p (b), Cd 3d (c) in the CdS and Cd0.9Co0.1S NRs, and Co 2p (d) in the Cd0.9Co0.1S NRs.
Figure 3. XPS survey spectra of Cd0.9Co0.1S NRs (a), and high-resolution XPS spectra of: S 2p (b), Cd 3d (c) in the CdS and Cd0.9Co0.1S NRs, and Co 2p (d) in the Cd0.9Co0.1S NRs.
Ijms 23 09756 g003
Figure 4. Time-dependent photocatalytic H2 evolution (a) and H2 evolution rates for all samples (b). Apparent quantum efficiency (AQE) and the UV–vis absorption spectrum of Cd0.9Co0.1S NRs (c). The AQE equals twice the number of evolved H2 molecules divided by the number of incident photons under light irradiation with specific wavelengths (365, 420, and 700 nm). Cyclic tests of photocatalytic H2 evolution for pure CdS and Cd0.9Co0.1S (d).
Figure 4. Time-dependent photocatalytic H2 evolution (a) and H2 evolution rates for all samples (b). Apparent quantum efficiency (AQE) and the UV–vis absorption spectrum of Cd0.9Co0.1S NRs (c). The AQE equals twice the number of evolved H2 molecules divided by the number of incident photons under light irradiation with specific wavelengths (365, 420, and 700 nm). Cyclic tests of photocatalytic H2 evolution for pure CdS and Cd0.9Co0.1S (d).
Ijms 23 09756 g004
Figure 5. Infrared thermography photos (a), the time-dependent temperature of the pure CdS and Cd0.9Co0.1S NRs under visible light (b), and H2 evolution rates of Cd0.9Co0.1S NRs without and with cooling the reaction solutions (c).
Figure 5. Infrared thermography photos (a), the time-dependent temperature of the pure CdS and Cd0.9Co0.1S NRs under visible light (b), and H2 evolution rates of Cd0.9Co0.1S NRs without and with cooling the reaction solutions (c).
Ijms 23 09756 g005
Figure 6. UV–vis diffuse reflectance spectra (a), corresponding Tauc plots (b), XPS valence band spectra (c), and Schematic band structures of pure CdS and Cd0.9Co0.1S NRs (d).
Figure 6. UV–vis diffuse reflectance spectra (a), corresponding Tauc plots (b), XPS valence band spectra (c), and Schematic band structures of pure CdS and Cd0.9Co0.1S NRs (d).
Ijms 23 09756 g006
Figure 7. PL spectra (a), TRPL decay spectra (b), photocurrent responses (c), and EIS plots (d) of as-prepared samples.
Figure 7. PL spectra (a), TRPL decay spectra (b), photocurrent responses (c), and EIS plots (d) of as-prepared samples.
Ijms 23 09756 g007
Figure 8. DFT calculation: side views of structural model (a), and Charge density differences of Cd0.9Co0.1S NRs (b) brown (gain electron), green (lose electron), Electronic band structure (c), and Total and projected density of states for Cd0.9Co0.1S NRs (d) (spin-up is upper half, spin-down is lower half), and Phonon spectra for Cd0.9Co0.1S NRs (e).
Figure 8. DFT calculation: side views of structural model (a), and Charge density differences of Cd0.9Co0.1S NRs (b) brown (gain electron), green (lose electron), Electronic band structure (c), and Total and projected density of states for Cd0.9Co0.1S NRs (d) (spin-up is upper half, spin-down is lower half), and Phonon spectra for Cd0.9Co0.1S NRs (e).
Ijms 23 09756 g008
Scheme 2. Diagram of the proposed photocatalytic mechanism of Cd0.9Co0.1S NRs under visible light irradiation.
Scheme 2. Diagram of the proposed photocatalytic mechanism of Cd0.9Co0.1S NRs under visible light irradiation.
Ijms 23 09756 sch002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, L.; Hong, M.; Zhang, K.; Li, B.; Fang, H.; Feng, X.; Xiao, X. Cd0.9Co0.1S Nanorods with an Internal Electric Field and Photothermal Effect Synergistically for Boosting Photocatalytic H2 Evolution. Int. J. Mol. Sci. 2022, 23, 9756. https://doi.org/10.3390/ijms23179756

AMA Style

Zhang L, Hong M, Zhang K, Li B, Fang H, Feng X, Xiao X. Cd0.9Co0.1S Nanorods with an Internal Electric Field and Photothermal Effect Synergistically for Boosting Photocatalytic H2 Evolution. International Journal of Molecular Sciences. 2022; 23(17):9756. https://doi.org/10.3390/ijms23179756

Chicago/Turabian Style

Zhang, Lilei, Manzhou Hong, Ka Zhang, Botan Li, Haipeng Fang, Xun Feng, and Xiuchan Xiao. 2022. "Cd0.9Co0.1S Nanorods with an Internal Electric Field and Photothermal Effect Synergistically for Boosting Photocatalytic H2 Evolution" International Journal of Molecular Sciences 23, no. 17: 9756. https://doi.org/10.3390/ijms23179756

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