Next Article in Journal
Numerical and Thermal Investigation of Magneto-Hydrodynamic Hybrid Nanoparticles (SWCNT-Ag) under Rosseland Radiation: A Prescribed Wall Temperature Case
Next Article in Special Issue
Mussel-Inspired Surface Modification of α-Zirconium Phosphate Nanosheets for Anchoring Efficient and Reusable Ultrasmall Au Nanocatalysts
Previous Article in Journal
4-Levels Vertically Stacked SiGe Channel Nanowires Gate-All-Around Transistor with Novel Channel Releasing and Source and Drain Silicide Process
Previous Article in Special Issue
Polybenzimidazole Confined in Semi-Interpenetrating Networks of Crosslinked Poly (Arylene Ether Ketone) for High Temperature Proton Exchange Membrane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Z-Scheme Heterojunction Catalysts Constructed by Graphitic-C3N4 and Photosensitive Metal-Organic Cages for Efficient Photocatalytic Hydrogen Evolution

1
The Key Laboratory of Low-Carbon Chemistry & Energy Conservation of Guangdong Province, School of Materials Science and Engineering, Sun Yat-sen University, Guangzhou 510006, China
2
School of Chemical Engineering and New Energy Materials, Zhuhai College of Science and Technology, Zhuhai 519041, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(5), 890; https://doi.org/10.3390/nano12050890
Submission received: 24 January 2022 / Revised: 7 February 2022 / Accepted: 23 February 2022 / Published: 7 March 2022

Abstract

:
The demand for improving the activity, durability, and recyclability of metal-organic cages (MOCs) that work as photocatalytic molecular devices in a homogeneous system has promoted research to combine them with other solid materials. An M2L4 type photosensitive metal-organic cage MOC-Q2 with light-harvesting ligands and catalytic Pd2+ centers has been synthesized and further heterogenized with graphitic carbon nitride to prepare a robust direct Z-scheme heterojunction photocatalyst for visible-light-driven hydrogen generation. The optimized g-C3N4/MOC-Q2 (0.7 wt%) sample exhibits a high H2 evolution activity of 6423 μmol g−1 h−1 in 5 h, and a total turnover number of 39,695 after 10 h, significantly superior to the bare MOC-Q2 used in the homogeneous solution and the comparison sample Pd/g-C3N4/L-4. The enhanced performances of g-C3N4/MOC-Q2 can be ascribed to its direct Z-scheme heterostructure, which effectively improves the charge separation and transfer efficiency. This work presents a rational approach of designing a binary photocatalytic system through combing micromolecular MOCs with heterogeneous semiconductors for water splitting.

Graphical Abstract

1. Introduction

It is meaningful but challenging to convert solar energy into chemical energies and to produce useful fuels such as H2 and CH4 [1,2,3]. Inspired by enzymatic photosynthesis in nature, scientists have attempted to develop photochemical molecular devices (PMDs) that can conduct catalysis reactions under appropriate conditions, and have adjustable reactivity and properties as well as highly researchable structure—activity relationships [4,5,6]. Metal-organic cages (MOCs) are a kind of discrete cage compound assembled by organic ligands and metal ions, which are named metal-organic polyhedrons (MOPs), macrocycles, containers, cages, and so on [7,8]. In particular, photosensitive MOCs that involve light-harvesting ligands and redox-active metal cation centers can transfer exciting electrons from ligands to catalytic centers fast, and thus work as PMDs, attracting extensive interest from researchers in the field of visible-light driven H2 production [9,10,11,12].
However, most photocatalytic MOC systems suffer from easy deactivation, are difficult to recycle, and have poor stability in water [13]. To overcome these drawbacks, efforts have been paid to developing composite catalysts formed by MOCs and some other solid materials for improving the photocatalytic performance and durability of MOCs. For example, Su and co-workers successfully heterogenized an as-synthesized MOC (MOC-16) by embedding it into a metal-organic framework (MOF) named ZIF-8, obtaining a heterogeneous PMD, MOC-16@CZIF [14]. MOC-16@CZIF inherited the efficiently directional electron transfer property of MOC-16, and the MOF matrix provided a platform for proton transfer. Therefore, the H2 production activity and durability of the MOC-16@CZIF was much higher than homogeneous MOC-16.
Since Wang and co-workers first used graphite phase carbon nitride (g-C3N4) for photocatalytic H2 production [15], g-C3N4 has become a widely investigated photocatalyst [16,17,18,19,20,21,22,23,24], because it is a non-toxic, low-cost, easily-prepared, and stable metal-free polymer with an appropriate energy band structure to absorb visible light and is used for many redox reactions. Nevertheless, the bare g-C3N4 has some disadvantages of a poor visible light response, low charge transfer rate, and low electron-hole separation efficiency, limiting its application in photocatalysis. Therefore, various strategies, such as morphology regulation, metal/non-metal element doping, dye sensitization, co-catalyst loading, and the formation of type II or Z-scheme heterojunction [25], have been investigated. Among these heterojunctions, type II heterojunction systems are widely used, but the photogenerated electrons in the type II heterojunction tend to accumulate on the more positive conduction band (CB), while the holes are prone to gather on the more negative valence band (VB), thus the redox activity is natively weakened thermodynamically. In contrast, the Z-scheme system [26], which simulates the electron transfer process of between photosystem I and photosystem II in nature, reserves both the higher excited energy level and the lower oxidized energy level, and thus shows a stronger redox ability than the traditional type II heterojunction [27,28].
Our group designed a non-photosensitive MOC-PC6 and a photosensitive MOC-Q1, and combined them with semiconductor matrixes including TiO2 [29] and g-C3N4 [30,31] for enhancing their performances. In 2021, we synthesized a M2L4 type photosensitive metal-organic cage MOC-Q2 with light-harvesting ligands and catalytic Pd2+ centers (Scheme 1). MOC-Q2 has been successfully heterogenized through combining it with TiO2 semiconductor through a simple sol−gel method. The as-prepared MOC-Q2-TiO2 photocatalyst showed a dramatically higher activity than the bare MOC-Q2 in a homogeneous solution [32]. To further enhance the performances of the MOC-Q2-based PMD, we herein designed a new binary Z-Scheme heterojunction system constructed with g-C3N4 and MOC-Q2. The optimal g-C3N4/MOC-Q2 (0.7 wt%) catalyst achieved a remarkable H2 evolution efficiency, giving a production rate of 6423 µmol g−1 h−1 after 5 h and a turnover number (TON) value of 39,694 based on MOC-Q2 after 10 h, representing the highest of all of the reported MOC-based PMD under visible light irradiation.

2. Materials and Methods

2.1. Materials

The solvent acetonitrile utilized in the cyclic voltammetry test was an HPLC grade reagent, and dimethyl sulfoxide was dried with A4 molecular sieves. All the other chemicals and reagents were purchased from commercial channels. Unless otherwise specified, further purification was unnecessary before use.

2.2. Characterizations

Powder X-ray diffraction (PXRD) curves were tested under Cu-Kα1 radiation (λ = 1.54056 Å) in a Rigaku Smart Lab diffractometer with Bragg−Brentano geometry (Rigaku corporation, Tokyo, Japan). Transmission electron microscopy (TEM) and element mapping were recorded by a JEM-ARM200P spherical aberration corrected transmission electron microscope (JEOL Ltd., Tokyo, Japan). Scanning electron microscope (SEM) images came from Hitachi ultra-high-resolution FE-SEM SU8010 (Hitachi High-Technologies Corporation, Tokyo, Japan). The UV−VIS absorption spectra were obtained on a Shimadzu UV-3600 spectrometer (Shimadzu Corporation, Kyoto, Japan). The infrared spectra were measured by a Frontier Fourier transform infrared (FT-IR) spectrometer with an average of 64 scans in a spectra range of 4000–450 cm−1 (PerkinElmer, Waltham, MA, United States). The steady-state and time-resolved photoluminescence (PL) spectra were tested with an Edinburgh FLSP980 fluorescence spectrometer (Edinburgh Instruments Ltd., Livingston, UK). The cyclic voltammetry (CV) curves were tested using a CH Instruments CHI760E electrochemical work station (CH Instruments Inc., Austin, TX, USA). The nitrogen adsorption−desorption isotherms were measured with a Quantachrome Auosorb-iQ2-MP analyzer (Quantachrome Instruments Inc., Boynton Beach, FL, USA). X-ray photoelectron spectroscopy (XPS) were performed in an ultrahigh vacuum chamber (ESCALAB 250Xi) using an XR6 monochromated AlKα X-ray source (hν = 1486.6 eV) with a 900 mm spot size (Thermo Fisher Scientific Inc., Waltham, MA, USA). The H2 yields were tested with a Fuli GC9790 II gas chromatography (Fuli Instruments Inc., Zhejiang, China). The apparent quantum yields (AQYs) were achieved using a monochromatic LED light source (Zolix, MLED4-1/M425, λ = 425 nm), and the photon flux was ascertained to be 1083 μmol h−1 (Zolix Instruments CO., LTD, Beijing, China). The Pd contents in all of the samples were obtained by inductively coupled plasma-atomic emission spectrometry (ICP-AES; SPECTRO CIROS VISION, spectra range: 120–800 nm, and holographic grating: 2924 line/mm) (SPECTRO Analytical Instruments, Kleve, Germany).

2.3. Synthesis and Preparation

2.3.1. Synthesis of MOC-Q2

The solution of Pd(BF4)2(MeCN)4 (44 mg, 0.1 mmol, 1 eq) in DMF (1 mL) was added into a stirring solution of L-2 (144 mg, 0.2 mmol, 2 eq) in DMF (2 mL), and the resulting solution was stirred at room temperature for 10 min and filtered. Then, the product was precipitated by adding diethyl ether, and it was washed with ethyl acetate and dried in a vacuum to give 118 mg MOC-Q2 as a brown solid (yield: 63%), which was in accordance with our previous work [32].

2.3.2. Preparation of g-C3N4

First, 10 g of urea was put in a 50 mL crucible and calcinated in a muffle furnace at 550 °C for 2 h at a heating rate of 10 °C/min. Then, the obtained solid was ground into powder and treated at 500 °C for another 2 h [30].

2.3.3. Preparation of g-C3N4/MOC-Q2

The composite material g-C3N4/MOC-Q2 was prepared using a solution dipping method. First, 2 mg MOC-Q2 was dissolved in 100 μL DMSO. Then, 50 mg g-C3N4 was ultrasonically dispersed in 10 mL acetonitrile for 1 h. Thereafter, different volumes of MOC-Q2 solution were added dropwise into the above g-C3N4 suspension, i.e., 7.5, 17.5, 25.0, and 50.0 µL. For keeping the same total volume of the mixtures, 42.5, 32.5, and 25.0 µL of DMSO was subsequently added to the above mixtures, respectively. Afterward, the mixture was vigorously agitated for 12 h at room temperature, followed by evaporation and vacuum drying. Finally, resulting samples with MOC-Q2 mass fractions of 0.3, 0.7, 1.0, and 2.0 wt% were obtained.

2.3.4. Preparation of g-C3N4/L-2 (0.7 wt%)

The preparation method was the same as that of g-C3N4/MOC-Q2 (0.7 wt%), except for replacing MOC-Q2 with L-2.

2.3.5. Preparation of Pd/g-C3N4/L-2 (0.7 wt%)

Pd nanoparticles were loaded onto g-C3N4/L-2 (0.7 wt%) using a photo-deposition method. The calculated amounts of H2PdCl4 corresponding to 0.7 wt% were added into a g-C3N4/L-2 (0.7 wt%) suspension and then stirred for 2 h under 300 W Xenon lamp irradiation. Then, the mixture was centrifuged to obtain the grey powder, which was washed with deionized water and dried under a vacuum to gain product.

2.4. Photocatalytic H2 Generation

The photocatalytic H2 generation experiments were performed in a closed quartz reactor. In detail, 5 mg of catalyst powder was added into a quartz reactor containing 2 mL triethanolamine (TEOA) and 18 mL water. The mixture was ultrasonic-dispersed and bubbled with N2 for 30 min to eliminate O2. Then, the reactor was irradiated with a Xenon lamp (300 W, λ ≥ 420 nm), and the amounts of produced H2 were detected by gas chromatography (GC). Especially, for 5 mg MOC-Q2, considering its poor solubility in the aqueous phase, it was dissolved in a mixed solution containing 16 mL DMSO, 2 mL water, and 2 mL TEOA.

2.5. Apparent Quantum Yield (AQY) Measurements for H2 Evolution

The AQY measurements were performed in a closed quartz reactor. In detail, 5 mg of catalyst powder was added into a quartz reactor containing 4.5 mL water and 0.5 mL TEOA. The mixture was ultrasonic-dispersed and bubbled with N2 for 30 min to eliminate O2. Then, the reactor was irradiated with a 425 nm monochromatic LED, and the amounts of produced H2 were detected by GC. The AQYs were measured with the following formula:
Φ = number   of   transfered   electrons number   of   incident   photons × 100 % = 2 × n ( H 2 ) n photons × 100 %

2.6. Hydroxyl Radical Trapping Experiment

First, 20 mg of g-C3N4, MOC-Q2 or g-C3N4/MOC-Q2 (0.7 wt%) was dispersed in a sealed 40 mL quartz glass bottle containing 10 mL aqueous solution containing 5 × 10−4 M terephthalic acid (TA) and 2 × 10−3 M NaOH. Then, the reactor was irradiated with a Xenon lamp (300 W, λ ≥ 420 nm) for 1 h. After the reaction, the mixture was filtered through an aqueous phase membrane filter, and the filtrate was used for the fluorescence measurement.

3. Results and Discussion

3.1. Synthesis of MOC-Q2 and g-C3N4/MOC-Q2

The synthesis of MOC-Q2 and the preparation of hybrid materials g-C3N4/MOC-Q2 and Pd/g-C3N4/L-2 are given in the Experimental Section. The successful formation and the geometric configuration of MOC-Q2 was confirmed in our previous study [32]. The actual MOC-Q2 contents of the prepared samples g-C3N4/MOC-Q2 confirmed by ICP-AES are listed in Table S1.

3.2. Characterization of g-C3N4/MOC-Q2

The PXRD patterns of g-C3N4, MOC-Q2, and the hybrid samples g-C3N4/MOC-Q2 are depicted in Figure 1. The pristine g-C3N4 and g-C3N4/MOC-Q2 materials displayed similar characteristic peak of 13.1° and 27.7°, demonstrating the (100) and (002) crystal plane of g-C3N4. Moreover, when MOC-Q2 contents gradually increased in the hybrids, the (002) peak of the g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%) samples slightly shifted to a lower angle. According to the Bragg formula, the decrease in the angle of the (002) peak indicated the interlayer spacing of g-C3N4/MOC-Q2 became larger, which can be attributed to the successful incorporation of MOC-Q2 into g-C3N4.
The SEM and TEM images displayed the stacked nanosheet structure of the g-C3N4/MOC-Q2 (Figure 2a–d). Owing to the tiny size of MOC-Q2, scarce signals related to MOC-Q2 were detected in the spectrogram. However, from the element mapping results of g-C3N4/MOC-Q2 (Figure 2e,f), it could be seen that the S and Pd elements, the exclusive elements of MOC-Q2, were detective all over the sample, thus confirming the homogeneously scattering of MOC-Q2 in the composite material.
The nitrogen adsorption−desorption tests of g-C3N4 and g-C3N4/MOC-Q2 (0.7 wt%) samples were carried out at 77 K. The isotherms are shown in Figure S1, and the BET surface areas and pore volumes are presented in Table S2. As shown in Figure S1a, g-C3N4 and g-C3N4/MOC-Q2 (0.7 wt%) exhibited H3 type hysteresis loops, which agreed with the layered structure of g-C3N4 materials. Furthermore, the pore size distribution diagram in Figure S1b showed that there were many mesopores, and a few micropores and macropores in g-C3N4. As for the g-C3N4/MOC-Q2 (0.7 wt%), there were a large number of mesopores and a small amount of macropores, but the micropores belonging to g-C3N4 disappeared. The specific surface area and pore volume in g-C3N4/MOC-Q2 (0.7 wt%) was 54.5 cm2 g−1 and 0.33 cm3 g−1, respectively, both smaller than those of g-C3N4, indicating that some pores in g-C3N4 were blocked by MOC-Q2.
The UV−VIS diffuse reflection spectra of g-C3N4, MOC-Q2, and a series of g-C3N4/MOC-Q2 samples are illustrated in Figure 3. g-C3N4 showed a strong UV absorption peak with an edge of ca. 457 nm, related to its 2.71 eV optical band gap according to the Tauc plot (Figure S2). MOC-Q2 had a wide absorption region from 300 to 600 nm, owing to its effective π−π* transition. As for the g-C3N4/MOC-Q2 composites, superimposed absorption bands of g-C3N4 and MOC-Q2 were observed. As the MOC-Q2 contents increased, the absorption intensity at a range of 400–600 nm was gradually enhanced.
To explore the interaction between MOC-Q2 and g-C3N4 in the composite material, the FT-IR spectra were measured as shown in Figure 4. The absorption peaks of MOC-Q2 were found in the region of 700 to 1600 cm−1. The peaks at 1637, 1573, 1460, 1407, 1316, and 1240 cm−1 in g-C3N4 represented the stretching vibrations of the triazine rings, while the peak at 812 cm−1 was a characteristic absorption that came from the deformation vibration of triazine rings. The g-C3N4/MOC-Q2 (2 wt%) showed some typical peaks of g-C3N4 as well, but the deformation vibration peak of the triazine ring shifted to 814 cm−1, which might be due to the π−π stacking effect that existed between MOC-Q2 and g-C3N4. In addition, the broadened and weakened peaks in the range from 3000 to 3500 cm−1 in the g-C3N4/MOC-Q2, which were ascribed to the amine stretching vibrations of g-C3N4, demonstrated that hydrogen bonds formed between the uncoordinated pyridine moieties in MOC-Q2 and the amine groups on g-C3N4.
For testifying the electron transfer feasibility of this system, the energy levels of the constituents were measured. In the CV curve of MOC-Q2 in Figure S3a,b, the onset oxidation potential (Eox) of MOC-Q2 appearing at 0.97 V vs. NHE, which corresponded to the Highest occupied molecular orbital (HOMO) energy level, was more positive than the CB edge potential (−1.03 V vs. NHE) of g-C3N4 [31], allowing for electron transfer from the CB edge of g-C3N4 to the oxidized MOC-Q2 so as to regenerate the neutral MOC-Q2. As seen from the normalized absorption and emission spectra of MOC-Q2 in Figure S3c, its E0-0 transition energy was determined to be 2.58 eV from the intersection (Table S3). Therefore, the excited state potential, which corresponded to its Lowest unoccupied molecular orbital (LUMO) level, was calculated to be −1.61 V vs. NHE judging by its Eox and E0-0 value. The reduction potential of H+ was more positive than the LUMO value of MOC-Q2, so that MOC-Q2 could conduct a proton reduction reaction via the Pd2+ catalytic sites. On the other hand, the VB edge potential (1.68 V vs. NHE) of g-C3N4 [31] was more positive than that of the sacrificial agent TEOA (0.84 ± 0.12 V vs. NHE) [33] and therefore it is thermodynamically feasible for the efficient supply of electrons from TEOA to the positive-charged VB of g-C3N4 during the reaction.

3.3. Photocatalytic H2 Evolution

The photocatalytic H2 evolution experiments were performed under visible light (λ > 420 nm) irradiation in an aqueous solution with triethanolamine (TEOA) as the sacrificial agent. As shown in Figure 5 and Table S4, when the content of MOC-Q2 on the g-C3N4 increased from 0.3 to 0.7 wt%, the photocatalytic activity enhanced from 2693 to 6423 µmol g−1 h−1 in 5 h. For g-C3N4/MOC-Q2 (0.7 wt%), the initial turnover number (TON) reached 20,955 based on the MOC-Q2 amounts, or 10,478 based on the number of Pd atoms. However, further increasing the incorporation amounts of MOC-Q2 conversely lowered the activity. For the g-C3N4/MOC-Q2 (1.0 and 2.0 wt%) samples, H2 yielded rates of 3153 and 2953 µmol g−1 h−1, respectively. We speculated that excessive amounts of MOC-Q2 would cause the accumulation of MOC-Q2 on the g-C3N4 surface, which resulted in intermolecular quenching, hindered the electron transfer efficiency, and thus decreased the catalyst activity. In addition, all the g-C3N4/MOC-Q2 samples exhibited a much higher H2 generation activity than the bare MOC-Q2 used in the homogeneous solution (855 µmol g−1 h−1), implying that the heterostructure of g-C3N4/MOC-Q2 could improve the effective charge transfer. As controls, Pd/g-C3N4/L-2 (0.7 wt%) samples with the same element composition were prepared and tested. The hydrogen production rate of Pd/g-C3N4/L-2 (0.7 wt%) was determined to be 156 µ mol g−1 h−1, far lower than that of g-C3N4/MOC-Q2 (0.7 wt%), indicating that MOC-Q2 played an essential role in the photocatalytic process, in which the ligand-to-metal charge transfer transition between L-2 and Pd2+ occurred.
To explore the durability of the MOC-Q2/g-C3N4 system, two reaction cycles were performed (Figure 6). The H2 yield in the second cycle decreased slightly, which might be due to the desorption or dissociation of a small part of MOC-Q2 during the photocatalytic reaction. A total H2 yield of 60.82 mmol g−1 was obtained, corresponding to a total TON[MOC-Q2] of 39,695 and TON[Pd] of 19,847, which was the highest of all of the reported MOC-based PMD under visible light irradiation (Table 1).

3.4. Mechanism Study

XPS measurements were performed to monitor the chemical states of Pd before and after the photocatalytic reaction (Figure S4a,b). Considering that the Pd contents in g-C3N4/MOC-Q2 (0.7 wt%) were too low to be detected, g-C3N4/MOC-Q2 (2 wt%) with high Pd contents was used for the test. For the fresh sample, two peaks located at 337.7 and 342.8 eV with an area ratio of 3:2 could be found, which could be attributed to 3d5/2 and 3d3/2 orbits of Pd2+, respectively. After reaction for 10 h, the 3d5/2 and 3d3/2 peaks of Pd2+ shifted to 336.7 and 341.9 eV. In addition, two new peaks appeared at 334.8 and 340.1 eV with an area ratio of 3:2, representing the 3d5/2 and 3d3/2 orbits of Pd0. These changes indicated that the break of Pd-N took place during the photocatalytic process, resulting in the formation of simple substance Pd0 (Pd black), which was in accordance with our previous report [31]. In addition, the XRD patterns of the spent sample showed a negligible change compared with that of the fresh sample, testifying that the crystal form of the sample did not change during the reaction (Figure S4c). In addition, the isolated feature of Pd on the spent g-C3N4/MOC-Q2 (2 wt%) sample after the photocatalytic reaction was observed by high-resolution TEM images (Figure S5).
To verify the electron transfer properties in g-C3N4/MOC-Q2, the steady-state and time-resolved PL spectra of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (0.7 wt%) were tested. As shown in Figure S6a, the bare g-C3N4 exhibited strong fluorescence centered at 465 nm under 360 nm excitation, and MOC-Q2 showed trace fluorescence. Whereas for g-C3N4/MOC-Q2 (0.7 wt%), a fluorescence peak at 465 nm, but with a significantly decreased intensity compared to g-C3N4, was observed. This result suggested that some photogenerated electrons in g-C3N4 transferred to MOC-Q2, and thus led to the reduced fluorescence intensity of g-C3N4/MOC-Q2. The fitted fluorescence lifetime from the time-resolved PL spectra (Figure S6b) were 10.1 ns for g-C3N4, 9.5 ns for g-C3N4/MOC-Q2 (0.7 wt%), and 1.2 ns for MOC-Q2, consistent with the steady-state PL results.
It is very important to confirm the Z-scheme mechanism for the heterojunction of g-C3N4/g-C3N4. In the presence of H+, the O2 in the reaction mixture can be reduced by the electron on the CB of g-C3N4 (Figure 7a), and the specific processes are listed as Equations (1)–(5) [49].
g-C3N4 + hv → hVB+ + eCB
eCB + O2 → O2
O2 + H+ → HO2
eCB + HO2• + H+ → H2O2
H2O2 + eCB → •OH + OH
Herein, we carried out the experiments of capturing •OH free radicals by terephthalic acid (TA), because TA could quickly combine with •OH to form a strongly fluorescent compound TA-OH, which has been widely used to testify the Z-scheme mechanism. As seen from Figure 8, the reaction mixture of the g-C3N4 sample with TA presented strong fluorescence at 430 nm, using 315 nm as the excitation wavelength, while that of the MOC-Q2 sample with TA had no fluorescence. As for the reaction mixture of the g-C3N4/MOC-Q2 (0.7 wt%) sample with TA, fluorescence with a medium intensity was observed. Considering that the produced •OH in the reaction mixture could only be generated by the electrons in the CB of g-C3N4, the attenuation of the fluorescence intensity came from the incorporation of MOC-Q2, which accepted some electrons in the CB of g-C3N4 and then reduced the yields of •OH (Figure 7b). Thus, it was verified that the as-prepared g-C3N4/MOC-Q2 system worked under a Z-scheme electron transfer route in the photocatalytic reaction. Therefore, we reasonably speculated the working mechanism of this system as follows: under visible light irradiation, g-C3N4 and MOC-Q2 synchronously entered an excited state and formed photogenerated electron−hole pairs. Then, the photo excited electrons on the CB of g-C3N4 combined with the holes on the HOMO of MOC-Q2, while the electrons on the LUMO of MOC-Q2 and the holes on the VB of g-C3N4 were retained to participate in the external reduction/oxidation reactions, respectively (Figure 9).

4. Conclusions

In summary, we designed and synthesized a photosensitive tridentate ligand L-2, and used four L-2 ligands with two catalytic Pd2+ centers to construct a M2L4-type molecular cage MOC-Q2. The as-prepared MOC-Q2 featured visible-light-harvesting and photo-induced LMCT transition properties, which allowed it to serve as a PMD for visible-light-driven water splitting. To overcome the instability of the homogeneous MOC catalyst and further improve the performances, we combined MOC-Q2 with g-C3N4 to prepare a direct Z-Scheme composite material g-C3N4/MOC-Q2. The optimized g-C3N4/MOC-Q2 (0.7 wt%) catalyst achieved a remarkable H2 generation activity of 6423 µmol g−1 h−1 and a TON value of 39,694 based on MOC amounts under visible light, far outperforming the homogenous MOC-Q2. The increased performance of the g-C3N4/MOC-Q2 catalyst could be owing to the superior charge-separation property of the Z-Scheme heterojunction, as well as the effective stabilizing effect through heterogenizing MOC-Q2 PMD with the polymeric g-C3N4 matrix. This research provides a promising idea to develop heterogeneous MOC-based catalysts for efficient and persistent photocatalytic water splitting.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12050890/s1, Table S1: The measured mass fractions of Pd and MOC-Q2 in g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%) catalysts; Table S2: BET surface areas and pore volumes of g-C3N4 and g-C3N4/MOC-Q2 (0.7 wt%); Table S3: The oxidation potential, E0-0, HOMO, and LUMO values of MOC-Q2; Table S4: Summary of the H2 production amounts and the corresponding TONs within 5 h; Figure S1: (a) N2 adsorption-desorption isotherms and (b) pore size distributions of g-C3N4 and g-C3N4/MOC-Q2 (0.7 wt%); Figure S2: Tauc plot of the g-C3N4 sample; Figure S3: CV curves of (a) MOC-Q2 (0.1 mM) and (b) ferrocene, and (c) normalized absorption and emission spectra of MOC-Q2; Figure S4: The XPS Pd 3d spectra of g-C3N4/MOC-Q2 (2 wt%) (a) before and (b) after 10 h photocatalysis, and (c) the XRD patterns of g-C3N4/MOC-Q2 (2 wt%) samples before and after photocatalytic reaction; Figure S5: High-resolution TEM images of the spent g-C3N4/MOC-Q2 (2 wt%) sample after photocatalytic reaction; Figure S6: (a) Steady state and (b) time-resolved PL spectra of MOC-Q2, g-C3N4, and g-C3N4/MOC-Q2 (0.7 wt%).

Author Contributions

Conceptualization, C.L. and J.L.; experiments, C.L. and S.Q.; data curation, Y.L. and X.L.; writing—review and editing, C.L., J.H. and J.L.; supervision, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the financial support from the National Natural Science Foundation Project of China (grant nos. 21975291, 21572280, and 6177050974) and the NSF of Guangdong Province (grant nos. 2019A1515011640 and 2020A1515110474).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T. State of the Art and Perspectives in Heterogeneous Catalysis of CO2 Hydrogenation to Methanol. Chem. Soc. Rev. 2020, 49, 1385–1413. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, Q.; Domen, K. Particulate Photocatalysts for Light-Driven Water Splitting: Mechanisms, Challenges, and Design Strategies. Chem. Rev. 2020, 120, 919–985. [Google Scholar] [CrossRef] [PubMed]
  3. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef] [PubMed]
  4. Inokuma, Y.; Kawano, M.; Fujita, M. Crystalline Molecular Flasks. Nat. Chem. 2011, 3, 349–358. [Google Scholar] [CrossRef]
  5. Zarra, S.; Wood, D.M.; Roberts, D.A.; Nitschke, J.R. Molecular Containers in Complex Chemical Systems. Chem. Soc. Rev. 2015, 44, 419–432. [Google Scholar] [CrossRef] [Green Version]
  6. Hardy, M.; Lutzen, A. Better Together: Functional Heterobimetallic Macrocyclic and Cage-like Assemblies. Chem. Eur. J. 2020, 26, 13332–13346. [Google Scholar] [CrossRef]
  7. Pilgrim, B.S.; Champness, N.R. Metal-Organic Frameworks and Metal-Organic Cages—A Perspective. Chempluschem 2020, 85, 1842–1856. [Google Scholar] [CrossRef]
  8. Pan, M.; Wu, K.; Zhang, J.-H.; Su, C.-Y. Chiral Metal-Organic Cages/Containers (MOCs): From Structural and Stereochemical Design to Applications. Coord. Chem. Rev. 2019, 378, 333–349. [Google Scholar] [CrossRef]
  9. Wu, K.; Li, K.; Chen, S.; Hou, Y.-J.; Lu, Y.-L.; Wang, J.-S.; Wei, M.-J.; Pan, M.; Su, C.-Y. The Redox Coupling Effect in a Photocatalytic RuII-PdII Cage with TTF Guest as Electron Relay Mediator for Visible-Light Hydrogen-Evolving Promotion. Angew. Chem. Int. Ed. Engl. 2020, 59, 2639–2643. [Google Scholar] [CrossRef]
  10. Chen, S.; Li, K.; Zhao, F.; Zhang, L.; Pan, M.; Fan, Y.-Z.; Guo, J.; Shi, J.-Y.; Su, C.-Y. A Metal-Organic Cage incorporating Multiple Light Harvesting and Catalytic Centres for Photochemical Hydrogen Production. Nat. Commun. 2016, 7, 13169. [Google Scholar] [CrossRef] [Green Version]
  11. Manbeck, G.F.; Fujita, E.; Brewer, K.J. Tetra- and Heptametallic Ru(II), Rh(III) Supramolecular Hydrogen Production Photocatalysts. J. Am. Chem. Soc. 2017, 139, 7843–7854. [Google Scholar] [CrossRef] [PubMed]
  12. Jin, Y.; Zhang, Q.; Zhang, Y.; Duan, C. Electron Transfer in the Confined Environments of Metal-Organic Coordination Supramolecular Systems. Chem. Soc. Rev. 2020, 49, 5561–5600. [Google Scholar] [CrossRef] [PubMed]
  13. Mollick, S.; Fajal, S.; Mukherjee, S.; Ghosh, S.K. Stabilizing Metal-Organic Polyhedra (MOP): Issues and Strategies. Chem. Asian J. 2019, 14, 3096–3108. [Google Scholar] [CrossRef] [PubMed]
  14. Luo, Y.-C.; Chu, K.-L.; Shi, J.-Y.; Wu, D.-J.; Wang, X.-D.; Mayor, M.; Su, C.-Y. Heterogenization of Photochemical Molecular Devices: Embedding a Metal−Organic Cage into a ZIF-8-Derived Matrix to Promote Proton and Electron Transfer. J. Am. Chem. Soc. 2019, 141, 13057–13065. [Google Scholar] [CrossRef]
  15. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A Metal-Free Polymeric Photocatalyst for Hydrogen Production from Water under Visible Light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef]
  16. Niu, P.; Zhang, L.; Liu, G.; Cheng, H.-M. Graphene-Like Carbon Nitride Nanosheets for Improved Photocatalytic Activities. Adv. Funct. Mater. 2012, 22, 4763–4770. [Google Scholar] [CrossRef]
  17. Li, Y.; Lu, Y.; Ma, Z.; Dong, L.; Jia, X.; Zhang, J. Enhancing Photocatalytic Hydrogen Production of g-C3N4 by Selective Deposition of Pt Cocatalyst. Nanomaterials 2021, 11, 3266. [Google Scholar] [CrossRef]
  18. Guo, Y.; Xu, A.; Hou, J.; Liu, Q.; Li, H.; Guo, X. Ag-Au Core-Shell Triangular Nanoprisms for Improving p-g-C3N4 Photocatalytic Hydrogen Production. Nanomaterials 2021, 11, 3347. [Google Scholar] [CrossRef]
  19. Nguyen, P.A.; Nguyen, T.K.A.; Dao, D.Q.; Shin, E.W. Ethanol Solvothermal Treatment on Graphitic Carbon Nitride Materials for Enhancing Photocatalytic Hydrogen Evolution Performance. Nanomaterials 2022, 12, 179. [Google Scholar] [CrossRef]
  20. Ge, F.; Huang, S.; Yan, J.; Jing, L.; Chen, F.; Xie, M.; Xu, Y.; Xu, H.; Li, H. Sulfur Promoted n-π* Electron Transitions in Thiophene-Doped g-C3N4 for Enhanced Photocatalytic Activity. Chin. J. Catal. 2021, 42, 450–459. [Google Scholar] [CrossRef]
  21. Bai, L.; Huang, H.; Zhang, S.; Hao, L.; Zhang, Z.; Li, H.; Sun, L.; Guo, L.; Huang, H.; Zhang, Y. Photocatalysis-Assisted Co3O4/g-C3N4 p–n Junction All-Solid-State Supercapacitors: A Bridge between Energy Storage and Photocatalysis. Adv. Sci. 2020, 7, 2001939. [Google Scholar] [CrossRef] [PubMed]
  22. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based Heterostructured Photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  23. Prasad, C.; Tang, H.; Liu, Q.; Bahadur, I.; Karlapudi, S.; Jiang, Y.J. A Latest Overview on Photocatalytic Application of g-C3N4 Based Nanostructured Materials for Hydrogen Production. Int. J. Hydrogen Energy 2020, 45, 337–379. [Google Scholar] [CrossRef]
  24. Li, Y.; Li, X.; Zhang, H.; Xiang, Q. Porous Graphitic Carbon Nitride for Solar Photocatalytic Applications. Nanoscale Horiz. 2020, 5, 765–786. [Google Scholar] [CrossRef]
  25. Cheng, L.; Zhang, H.; Li, X.; Fan, J.; Xiang, Q. Carbon-Graphitic Carbon Nitride Hybrids for Heterogeneous Photocatalysis. Small 2021, 17, 2005231. [Google Scholar] [CrossRef] [PubMed]
  26. Tachibana, Y.; Vayssieres, L.; Durrant, J.R. Artificial Photosynthesis for Solar Water-Splitting. Nat. Photonics 2012, 6, 511–518. [Google Scholar] [CrossRef]
  27. Wang, J.; Yu, Y.; Cui, J.; Li, X.; Zhang, Y.; Wang, C.; Yu, X.; Ye, J. Defective g-C3N4/Covalent Organic Framework Van Der Waals Heterojunction toward Highly Efficient S-scheme CO2 Photoreduction. Appl. Catal. B-Environ. 2022, 301, 120814. [Google Scholar] [CrossRef]
  28. Chen, X.; Wang, J.; Chai, Y.; Zhang, Z.; Zhu, Y. Efficient Photocatalytic Overall Water Splitting Induced by the Giant Internal Electric Field of a g-C3N4/rGO/PDIP Z-Scheme Heterojunction. Adv. Mater. 2021, 33, 2007479. [Google Scholar] [CrossRef]
  29. Hou, C.-P.; Chen, X.-L.; Huang, Z.-J.; Lei, Y.; Xiao, L.-M.; Huang, J.-F.; Li, S.-Y.; Liu, J.-M. Robust Heterogeneous Photocatalyst for Visible-Light-Driven Hydrogen Evolution Promotion: Immobilization of a Fluorescein Dye-Encapsulated Metal-Organic Cage on TiO2. ACS Appl. Mater. Interfaces 2021, 13, 57230–57240. [Google Scholar] [CrossRef]
  30. Wang, Y.-P.; Liu, L.; Wu, D.-J.; Guo, J.; Shi, J.-Y.; Liu, J.-M.; Su, C.-Y. Immobilization of Metal-Organic Molecular Cage on g-C3N4 Semiconductor for Enhancement of Photocatalytic H2 Generation. Chin. J. Catal. 2019, 40, 1198–1204. [Google Scholar] [CrossRef]
  31. Qin, S.; Lei, Y.; Guo, J.; Huang, J.-F.; Hou, C.-P.; Liu, J.-M. Constructing Heterogeneous Direct Z-Scheme Photocatalysts Based on Metal−Organic Cages and Graphitic-C3N4 for High-Efficiency Photocatalytic Water Splitting. ACS Appl. Mater. Interfaces 2021, 13, 25960–25971. [Google Scholar] [CrossRef] [PubMed]
  32. Qin, S.; Lei, Y.; Huang, J.-F.; Xiao, L.-M.; Liu, J.-M. A Robust Photocatalytic Hybrid Material Composed of Metal Organic Cages and TiO2 for Efficient Visible-Light-Driven Hydrogen Evolution. Chem. Asian J. 2021, 16, 2055–2062. [Google Scholar] [CrossRef] [PubMed]
  33. Sun, H.; Hoffman, M.Z. Reductive Quenching of the Excited States of Ruthenium(II) Complexes Containing 2,2′-Bipyridine, 2,2′-Bipyrazine, and 2,2′-Bipyrimidine Ligands. J. Phys. Chem. A 1994, 98, 11719–11726. [Google Scholar] [CrossRef]
  34. Ji, C.; Wang, W.; El-Sayed, E.; Liu, G.; Si, Y.; Su, K.; Ju, Z.; Wu, F.; Yuan, D. A High-Efficiency Dye-Sensitized Pt(II) Decorated Metal-Organic Cage for Visible-Light-Driven Hydrogen Production. Appl. Catal. B-Environ. 2021, 285, 119782. [Google Scholar] [CrossRef]
  35. Zhang, J.; Wei, M.; Lu, Y.; Wei, Z.; Wang, H.; Pan, M. Ultrafine Palladium Nanoparticles Stabilized in the Porous Liquid of Covalent Organic Cages for Photocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2020, 3, 12108–12114. [Google Scholar] [CrossRef]
  36. Sun, M.; Wang, Q.; Qin, C.; Sun, C.; Wang, X.; Su, Z. An Amine-Functionalized Zirconium Metal–Organic Polyhedron Photocatalyst with High Visible-Light Activity for Hydrogen Production. Chem. Eur. J. 2019, 25, 2824–2830. [Google Scholar] [CrossRef]
  37. Kumar, K.; Mane, V.; Yadav, A.; Kumbhar, A.; Boomishankar, R. Photochemical hydrogen evolution from water by a 1D-network of octahedral Ni6L8 cages. Chem. Commun. 2019, 55, 13156–13159. [Google Scholar] [CrossRef]
  38. Cai, J.; Zhao, L.; Wei, J.; He, C.; Long, S.; Duan, C. Negatively Charged Metal–Organic Hosts with Cobalt Dithiolene Species: Improving PET Processes for Light-Driven Proton Reduction through Host-Guest Electrostatic Interactions. Chem. Commun. 2019, 55, 8524–8527. [Google Scholar] [CrossRef]
  39. Nurttila, S.; Becker, R.; Hessels, J.; Woutersen, S.; Reek, J. Photocatalytic Hydrogen Evolution by a Synthetic [FeFe] Hydrogenase Mimic Encapsulated in a Porphyrin Cage. Chem. Eur. J. 2018, 24, 16395–16406. [Google Scholar] [CrossRef] [Green Version]
  40. Zhao, L.; Wei, J.; Zhang, J.; He, C.; Duan, C. Encapsulation of a Quinhydrone Cofactor in the Inner Pocket of Cobalt Triangular Prisms: Combined Light-Driven Reduction of Protons and Hydrogenation of Nitrobenzene. Angew. Chem. Int. Ed. 2017, 56, 15284–15288. [Google Scholar] [CrossRef]
  41. Zhao, L.; Wang, J.; Wu, P.; He, C.; Guo, X.; Duan, C. DHPA-Containing Cobalt-Based Redox Metal-Organic Cyclohelicates as Enzymatic Molecular Flasks for Light-Driven H2 Production. Sci. Rep. 2017, 7, 14347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Zhao, L.; Wei, J.; Zhang, F.; He, C.; Zheng, S.; Duan, C. Redox-Active Copper Triangles as an Enzymatic Molecular Flask for Light-Driven Hydrogen Production. RSC Adv. 2017, 7, 48989–48993. [Google Scholar] [CrossRef] [Green Version]
  43. Jing, X.; Yang, Y.; He, C.; Chang, Z.; Reek, J.; Duan, C. Control of Redox Events by Dye Encapsulation Applied to Light-Driven Splitting of Hydrogen Sulfide. Angew. Chem. Int. Ed. 2017, 56, 11759–11763. [Google Scholar] [CrossRef] [PubMed]
  44. Deshmukh, M.; Mane, V.; Kumbhar, A.; Boomishankar, R. Light-driven Hydrogen Evolution from Water by a Tripodal Silane Based CoII6L18 Octahedral Cage. Inorg. Chem. 2017, 56, 13286–13292. [Google Scholar] [CrossRef]
  45. Yu, H.; He, C.; Xu, J.; Duan, C.; Reek, J. Metal—Organic Redox Vehicles to Encapsulate Organic Dyes for Photocatalytic Protons and Carbon Dioxide Reduction. Inorg. Chem. Front. 2016, 3, 1256–1263. [Google Scholar] [CrossRef]
  46. Yang, L.; He, C.; Liu, X.; Zhang, J.; Sun, H.; Guo, H. Supramolecular Photoinduced Electron Transfer between a Redox-Active Hexanuclear Metal-Organic Cylinder and an Encapsulated Ruthenium(II) Complex. Chem. Eur. J. 2016, 22, 5253–5260. [Google Scholar] [CrossRef]
  47. Yang, L.; Jing, X.; He, C.; Chang, Z.; Duan, C. Redox-Active M8L6 Cubic Hosts with Tetraphenylethylene Faces Encapsulate Organic Dyes for Light-Driven H2 Production. Chem. Eur. J. 2016, 22, 18107–18114. [Google Scholar] [CrossRef]
  48. Jing, X.; He, C.; Yang, Y.; Duan, C. A Metal-Organic Tetrahedron as a Redox Vehicle to Encapsulate Organic Dyes for Photocatalytic Proton Reduction. J. Am. Chem. Soc. 2015, 137, 3967–3974. [Google Scholar] [CrossRef]
  49. Cui, Y.; Huang, J.; Fu, X.; Wang, X. Metal-Free Photocatalytic Degradation of 4-Chlorophenol in Water by Mesoporous Carbon Nitride Semiconductors. Catal. Sci. Technol. 2012, 2, 1396–1402. [Google Scholar] [CrossRef]
Scheme 1. (a) Assembly of the M2L4-type cage MOC-Q2, (b) structure of MOC-Q2, and (c) schematic representation of the g-C3N4/MOC-Q2 system for H2 evolution.
Scheme 1. (a) Assembly of the M2L4-type cage MOC-Q2, (b) structure of MOC-Q2, and (c) schematic representation of the g-C3N4/MOC-Q2 system for H2 evolution.
Nanomaterials 12 00890 sch001
Figure 1. The XRD patterns of g-C3N4 and g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%).
Figure 1. The XRD patterns of g-C3N4 and g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%).
Nanomaterials 12 00890 g001
Figure 2. (a,b) SEM, (c,d) TEM, (e) EDX, and (f) element mapping images of g-C3N4/MOC-Q2 (0.7 wt%).
Figure 2. (a,b) SEM, (c,d) TEM, (e) EDX, and (f) element mapping images of g-C3N4/MOC-Q2 (0.7 wt%).
Nanomaterials 12 00890 g002
Figure 3. UV−VIS absorption spectra of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%).
Figure 3. UV−VIS absorption spectra of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%).
Nanomaterials 12 00890 g003
Figure 4. FT-IR spectra of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (2 wt%).
Figure 4. FT-IR spectra of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (2 wt%).
Nanomaterials 12 00890 g004
Figure 5. Photocatalytic H2 evolution of MOC-Q2, Pd/g-C3N4/L-2 (0.7 wt%), and g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%) under visible light irradiation.
Figure 5. Photocatalytic H2 evolution of MOC-Q2, Pd/g-C3N4/L-2 (0.7 wt%), and g-C3N4/MOC-Q2 (0.3/0.7/1.0/2.0 wt%) under visible light irradiation.
Nanomaterials 12 00890 g005
Figure 6. H2 evolution recycle experiments of g-C3N4/MOC-Q2 (0.7 wt%) under visible light irradiation.
Figure 6. H2 evolution recycle experiments of g-C3N4/MOC-Q2 (0.7 wt%) under visible light irradiation.
Nanomaterials 12 00890 g006
Figure 7. Mechanism for electron transfer and •OH generation of (a) g-C3N4 and (b) g-C3N4/MOC-Q2.
Figure 7. Mechanism for electron transfer and •OH generation of (a) g-C3N4 and (b) g-C3N4/MOC-Q2.
Nanomaterials 12 00890 g007
Figure 8. PL spectra of the supernatants of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (0.7 wt%) samples in the presence of terephthalic acid.
Figure 8. PL spectra of the supernatants of g-C3N4, MOC-Q2, and g-C3N4/MOC-Q2 (0.7 wt%) samples in the presence of terephthalic acid.
Nanomaterials 12 00890 g008
Figure 9. Proposed direct Z-scheme mechanism of g-C3N4/MOC-Q2 photocatalyst for H2 evolution reaction.
Figure 9. Proposed direct Z-scheme mechanism of g-C3N4/MOC-Q2 photocatalyst for H2 evolution reaction.
Nanomaterials 12 00890 g009
Table 1. Performances of photocatalytic systems based on MOC for visible-light-driven hydrogen evolution in recent years.
Table 1. Performances of photocatalytic systems based on MOC for visible-light-driven hydrogen evolution in recent years.
No.ReferencePhotocatalystH2 YieldTONTime
1Appl. Catal. B-Environ. 2021, 285, 119782. [34]RhB-Zr-Bpydc-PtCl2 MOCs230 μmol/g378[Pt]22 h
2ACS Appl. Mater. Interfaces 2021, 13, 25960. (our work) [31]g-C3N4/MOC-Q183,692 μmol/g19,268[MOC-Q1]25 h
3ACS Appl. Mater. Interfaces 2021, 13, 57230. (our work) [29]FL@MOC-PC6-TiO229.8 mmol/g4356[MOC-PC-6]40 h
4Chem Asian J. 2021, 16, 2055.(our work) [32]TiO2-MOC-Q2309.8 mmol/g11,739[MOC-Q2]35 h
5ACS Appl. Energy Mater. 2020, 3, 12108. [35]PdNPs@C4R/g-C3N45487 μmol/g h20 h
6Chem. Eur. J. 2019, 25,2824. [36]Pt/ZrT-1-NH21060 μmol/g h4 h
7Angew. Chem. Int. Ed. 2019, 59, 2639. [9]MOC-16/TTF3173 μmol/μM1202[Pd]47 h
8J. Am. Chem. Soc. 2019, 141, 13057. [14]MOC-16@CZIF97.0 mmol/g35,000[Pd]24 h
9Chem. Commun. 2019, 55, 13156. [37][Ni6L8]14 μmol/μM2824[[Ni6L8]∞69 h
10Chinese J. Catal. 2019, 40, 1198.(our work) [30]MOC-16/g-C3N410.0 mmol/g517[Pd]15 h
11Chem. Commun. 2019, 55, 8524. [38]Co-NAS/Ru(bpy)32+10 mL/mM360[Co-NAS]9 h
12Chem. Eur. J. 2018, 24, 16395. [39]1·Fe4(Zn-L)60.4[1·Fe4(Zn-L)6]2 h
13Angew. Chem. Int. Ed. 2017, 56, 15284. [40]Co-TPC/QHQ0.4 mL/μM12 h
14Sci. Rep. 2017, 7, 14347. [41]Co-ZPB/Fl50 mL/mM400[Co-ZPB]4 h
15RSC Adv. 2017, 7, 48989. [42]Cu-OBP/Fl0.1 mL/μM1200[Cu-OBP]20 h
16Angew. Chem. Int. Ed. 2017, 56, 11759. [43]Ni-TFT/Fl0.3 mL/mM25,000[Ni-TFT]20 h
17Inorg. Chem. 2017, 56, 13286. [44]Co6L8/Ru(bpy)312 μmol/h43[Co6L8]2 h
18Inorg. Chem. Front. 2016, 3, 1256. [45]Ni-SSC/Fl1 mL/μM1250[Ni-SSC]8 h
19Chem. Eur. J. 2016, 22,5253. [46]Ni-YL/Ru(dcbpy)30.2 mL/μM1600[Ni-YL]5 h
20Chem. Eur. J. 2016, 22,18107. [47]Cage2/Fl0.1 mL/μM700[Cage2]15 h
21Nat. Commun. 2016, 7, 13169. [10]MOC-1650 μmol/μM635[Pd]48 h
22J. Am. Chem. Soc. 2015, 137, 3967. [48]Co-TFT/Fl1.5 mL/μM11,000[Co-TFT]15 h
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lv, C.; Qin, S.; Lei, Y.; Li, X.; Huang, J.; Liu, J. Direct Z-Scheme Heterojunction Catalysts Constructed by Graphitic-C3N4 and Photosensitive Metal-Organic Cages for Efficient Photocatalytic Hydrogen Evolution. Nanomaterials 2022, 12, 890. https://doi.org/10.3390/nano12050890

AMA Style

Lv C, Qin S, Lei Y, Li X, Huang J, Liu J. Direct Z-Scheme Heterojunction Catalysts Constructed by Graphitic-C3N4 and Photosensitive Metal-Organic Cages for Efficient Photocatalytic Hydrogen Evolution. Nanomaterials. 2022; 12(5):890. https://doi.org/10.3390/nano12050890

Chicago/Turabian Style

Lv, Chuying, Su Qin, Yang Lei, Xinao Li, Jianfeng Huang, and Junmin Liu. 2022. "Direct Z-Scheme Heterojunction Catalysts Constructed by Graphitic-C3N4 and Photosensitive Metal-Organic Cages for Efficient Photocatalytic Hydrogen Evolution" Nanomaterials 12, no. 5: 890. https://doi.org/10.3390/nano12050890

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