Next Article in Journal
Novel Heterostructures of Noble Plasmonic Metals/Ga-Substituted Hydrotalcite for Solar Light Driven Photocatalysis toward Water Purification
Next Article in Special Issue
The Progress of Metal-Organic Framework for Boosting CO2 Conversion
Previous Article in Journal
A Theoretical and Experimental Approach to the Analysis of Hydrogen Generation and Thermodynamic Behavior in an In Situ Heavy Oil Upgrading Process Using Oil-Based Nanofluids
Previous Article in Special Issue
In Situ Growth of Self-Supporting MOFs-Derived Ni2P on Hierarchical Doped Carbon for Efficient Overall Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Advanced Synthesis of MOFs-Based Materials in Photocatalytic HER in Recent Three Years

1
School of Materials Science and Engineering, National Institute for Advanced Materials, TKL of Metal and Molecule-Based Material Chemistry, Nankai University, Tianjin 300350, China
2
Haihe Laboratory of Sustainable Chemical Transformations, Tianjin 300192, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1350; https://doi.org/10.3390/catal12111350
Submission received: 13 October 2022 / Revised: 29 October 2022 / Accepted: 31 October 2022 / Published: 2 November 2022
(This article belongs to the Special Issue Metal-Organic Framework Based Catalysts for Energy Applications)

Abstract

:
Since the advent of metal–organic frameworks (MOFs), researchers have paid extensive attention to MOFs due to their determined structural composition, controllable pore size, and diverse physical and chemical properties. Photocatalysis, as a significant application of MOFs catalysts, has developed rapidly in recent years and become a research hotspot continuously. Various methods and approaches to construct and modify MOFs and their derivatives can not only affect the structure and morphology, but also largely determine their properties. Herein, we summarize the advanced synthesis of MOFs-based materials in the field of the photocatalytic decomposition of water to produce hydrogen in the recent three years. The main contents include the overview of the novel synthesis strategies in four aspects: internal modification and structure optimization of MOFs materials, MOFs/semiconductor composites, MOFs/COFs-based hybrids, and MOFs-derived materials. In addition, the problems and challenges faced in this direction and the future development goals were also discussed. We hope this review will help deepen the reader’s understanding and promote continued high-quality development in this field.

Graphical Abstract

1. Introduction

Since the 21st century, the world energy crisis has become increasingly urgent mainly due to the growing shortage of non-renewable energy reserves such as coal, oil, and natural gas [1,2,3]. It is crucial to find clean renewable energy, especially one that can be produced on an industrial scale [4,5]. Hydrogen (H2) is regarded as one of the most ideal substitutes for carbon-based energy sources because of its high calorific value and lack of pollution from its combustion products [6,7]. At present, the electrocatalytic decomposition of water is mostly used in industrial hydrogen production, but this undoubtedly brings a large amount of waste of electrical energy [8,9]. As solar energy is inexhaustible, it would be a satisfactory advantage to fully utilize and convert solar energy to catalyze the hydrogen evolution reaction [10,11]. In the early days of this field, some complexes of precious metals were explored as photosensitizers. Although they were highly capable of capturing light, these complexes were too expensive to meet the economic principle [12,13,14], which promoted the researchers to develop some low-cost materials.
With the continuous in-depth research of the theory and practice of photoelectric chemistry, a multitude of semiconductor materials with superb light trapping abilities emerged, such as CdS [15,16,17,18,19], TiO2 [20,21], C3N4 [22,23,24,25], etc. However, these classical semiconductor materials also had some inevitable shortcomings and deficiencies, despite that they were generally low toxicity, cheap, and easy to synthesize. In the case of CdS, serious photocorrosion would occur in the process of photocatalysis, which reduced the available components and made it difficult to maintain a high efficiency catalysis for a long-term reaction [26,27,28]. As far as TiO2 was concerned, it mainly absorbed ultraviolet light and was not sensitive to visible light, which was obviously not the optimal way to utilize solar energy [29,30]. C3N4 did not have the above-mentioned disadvantages as a kind of novel organic semiconductor material, but its apparent quantum yields (AQYs) in photocatalysis were not satisfactory, usually no more than 10% [31,32]. Therefore, it is a prerequisite for researchers to make unremitting efforts to extend such semiconductor materials from design to practice.
In the past three decades, metal–organic frameworks (MOFs) have attracted the attention of many chemists due to their controllable pore size, redox ability of central ions, and changeful ligands [33,34,35,36,37,38,39,40,41,42,43,44]. Many literature reports in the fields of gas separation and adsorption [45,46,47,48], fluorescence recognition [49,50,51,52], electrocatalysis [53,54,55], and photocatalysis [56,57,58] have confirmed this hotspot. In the aspect of photocatalytic (hydrogen evolution reaction) HER, MOFs are particularly outstanding, and the orientations are mainly focused on the following three aspects. Firstly, the ligands and central ions of MOFs usually play various roles in the photocatalytic process, which makes reasonable design and adjustment of internal structures become one of the most common modification methods for MOFs. Secondly, the integration of MOFs and other types of materials is one of the frequent techniques, which can not only make full use of the stable skeleton of MOFs, but also make the supported substances evenly dispersed to improve the utilization of the cocatalyst. Thirdly, MOFs can be appointed as precursors to synthesize and prepare some substances with special morphology or composition. These corresponding measures give MOFs and their derivatives or composites better development potential and vitality in the field of photocatalytic HER for practical application.
The unique property of one substance is inseparable from its peculiar structure. Therefore, continuous progress in the synthesis and modification strategies for MOFs and their derived materials is the cornerstone of sustainable and efficient development. Most of the published reviews may only focus on one aspect of MOFs material modification and lack a systematic summary. Based on the concept of “from inside to outside”, we reviewed the synthetic progress of MOFs materials in the field of photocatalytic HER in the recent three years, from the internal modification of MOFs materials to the participation of MOFs materials as independent components in composite materials, and then to the derivatives of MOFs materials. We hope that this timely review will help readers gain an in-depth understanding about the relevant advances in this field and promote its rapid development with high quality.

2. Internal Modification and Structure Optimization of MOFs

Reasonable internal modification and structure optimization of MOFs may not only make the photosensitive units orderly, but also shorten the distance between the photosensitive units and the catalytic centers, so as to improve the efficiency of photogenerated electrons separation and accelerate the rate of hydrogen evolution [59,60].
Jiang et al. proposed a facile strategy for the construction of a deployable coordination microenvironment based on MOFs [61]. In this work, the earth-abundant metals (such as Ni2+, Co2+, and Cu2+) were fixed to the Zr6-oxo cluster in the form of single-atom catalysts (SACs) via a rapid and facile microwave-assisted method (Figure 1). The adjacent -O/OHx groups in the Zr6-oxo cluster of MOFs provided the lone pair electrons and the charge balance to anchor the additional single metal atoms. The atomically dispersed metal site was in close proximity to the photosensitive unit (the linker), which greatly accelerated the transfer of photogenerated electrons and thus facilitated the redox reactions. Therefore, the optimized Ni1-S/MOF had a unique Ni(I) microenvironment and exhibited excellent photocatalytic H2 evolution performance, which was 270 times higher than that of the pure MOF and far exceeded the other Ni1-X/MOF counterparts. This work unequivocally demonstrated the great advantage of MOFs in preparing high-content SACs with the proximity of variable microenvironments to photosensitive junctions, thereby promoting electrons transfer and facilitating photocatalysis.
In another example of structure optimization of MOFs, Jiang et al. discussed the role of linker engineering in MOFs for dark photocatalysis [62]. Figure 2 shows the dark photocatalysis based on MIL-125 and MIL-125 with different functional groups (MIL-125-X, X = NH2, NO2, Br). Notably, the introduction of different functional groups on the connector enabled the resulting MOFs to have markedly various activities. The dynamic and thermodynamic investigations manifested that the generation and the lifetime of Ti3+ intermediates were the most critical factors affecting their properties, due to the electron-donating/-withdrawing effect of the functional groups. The time-dependent density-functional theory (TD-DFT) calculation indicated that the introduction of an electron-donating group was beneficial, which helped to lengthen the distance between the photogenerated electrons and holes and improved the separation of these pairs. According to the investigation, this was the first study to systematically regulate the dark photocatalytic hydrogen production process of MOFs-based materials. The relevant works gave a new reference to the lifetime adjustment of electron relay for enhancing the dark photocatalysis.
The stripping of MOFs single crystals into 2D layered materials could increase the surface area and shorten the transfer path of electrons, which was regarded as an excellent medication method. Zhang and his co-authors [63] reported a water-stable nickel-based MOF single crystal (Ni-TBAPy-SC) and its exfoliated nanobelts (Ni-TBAPy-NB), which could bear a wide range of pH environments (Figure 3a). The optimized hydrogen production rate of Ni-TBAPy-NB could reach 98 μmol·h−1 (5 mmol·h−1·g−1) with an apparent quantum efficiency (AQE) of 8.0% at 420 nm, which was 164 times higher than that of Ni-TBAPy-SC. Based on the DFT calculations, the transfer of photogenerated electrons from the H4TBAPy to the [Ni3O16] cluster was thermodynamically permissible (Figure 3b). In addition, both the reduced [Ni3O16] and [Ni3O16] cluster had excellent properties of water absorption. In another report about 2D MOFs, Duan et al. [64] synthesized 2D indium-based porphyrin MOF cubic nanosheets (2D In-TCCP NS) via a surfactant-assisted method (Figure 4a). The corresponding characterization of 2D In-TCCP NS (powder XRD, SEM, TEM, and AFM) and 3D In-TCCP bulk (powder XRD and SEM) are shown in Figure 4b–f. The 2D In-TCCP NS showed great chemical stability in the range of 2-11 in acidic and basic solutions (Figure 4g). In the photocatalytic tests, the 2D In-TCCP NS exhibited a hydrogen evolution rate of 67.97 μmol·g−1·h−1, which was 11.5 times higher than that of 3D In-TCCP bulk (5.87 μmol·g−1·h−1). It was worth noting that there was not an obvious activity decrease after 40 h of photocatalysis, which reflected the practical commercial application.

3. MOFs/Semiconductor Composites

Some MOFs were not sensitive to light (especially in visible light), which led to their poor light absorption ability. The mutual combination with semiconductor materials could not only make full use of the stable skeleton of MOFs, but also improve the utilization rate of semiconductor materials [65,66,67,68]. Whereas, after they recombined or coexisted, some complicated issues such as the mechanism of interface electrons transfer between the photosensitizer and cocatalyst were not very unambiguous, and the uniform distribution of the cocatalyst on the surface or in the pores of MOFs needs to be solved urgently.
In 2020, Jiang et al. [69] incorporated the monodisperse, small-size and noble-metal-free transitional metal phosphides (TMPs; such as Ni2P, Ni12P5) into a classical MOF (UiO-66-NH2) for photocatalytic HER (Figure 5). The TEM images of Ni2P@UiO-66-NH2 (the inset is the size distribution of Ni2P NPs) and Pt@UiO-66-NH2 (the inset is the size distribution of Pt NPs) are shown in Figure 6a–d. Strikingly, the Ni12P5@UiO-66-NH2 exhibited a hydrogen evolution rate of 293.2 μmol·g−1·h−1, which was about 80 times that of the parent MOF. Both thermodynamic and kinetic studies revealed that the TMPs had similar behaviors to Pt, which could greatly accelerate the photogenerated charge transfer from the linker to the cluster and promote its separation, thereby reducing the activation energy of H2 generation. Alternatively, the relevant results manifested that Pt was thermodynamically favorable, yet Ni2P was kinetically preferred for H2 evolution. This work has laid a solid foundation for the development of novel non-precious metal composite photocatalysts.
The surfactant coating on the metal cocatalyst had a great regulating effect on the microenvironment of the catalytic site, which had been largely ignored. In 2021, Jiang and his co-authors [70] reported a series of Pt cocatalysts with adjustable microenvironments, including polyvinylpyrrolidone (PVP)-capped Pt nanoparticles (PtPVP), Pt with partially removed PVP (PtrPVP), and clean Pt without PVP (Pt) were encapsulated into UiO-66-NH2 to obtain PtPVP@ UiO-66-NH2, PtrPVP@UiO-66-NH2, and Pt@UiO-66-NH2 (Figure 7a,b). The experimental results indicated that the PVP intervention had a negative effect on the transfer of photogenerated electrons, and this negative effect was ameliorated after the removal of a portion of PVP, thus increasing the rate of hydrogen precipitation. Inspired by this rule, the Pt-Fc@UiO-66-NH2 with the best activity was synthesized via introducing an electron mediator of Fc into Pt@UiO-66-NH2. In this research, the effects of interfacial surfactants between the photosensitizer and cocatalyst on electron transfer kinetics and related activities were thoroughly investigated, which also provided vital implications for the microenvironment regulation of photocatalytic catalytic centers. In contrast, PVP played a positive role in an example reported by Li and his co-authors [71] for constructing 2D-on-3D MOFs materials. They achieved the “anti-epitaxial growth” pattern of foreign MOF nucleus on the (111) facets of the UiO-66-NH2 octahedron seed with the assistance of PVP and successfully extended this method to Zn, Cd, Co, and Ni elements for building this type of 2D-on-3D heterojunction (Figure 8a). The influence of the presence or absence of PVP on the control of product morphology was also discussed in detail (Figure 8b). The photocatalytic performance of Cu 2D-on-3D exhibited over one order of magnitude times that of the “dimensionality-identical” MOF. Especially, the apparent quantum efficiency (AQE) of 2D-on-3D Cu heterostructures could reach 12.04% with an excitation wavelength of 380 nm. Thanks to the ultrathin cover layers, the faster electron transfer and more efficient electron–hole pairs separation promoted the photocatalysis. The emergence of these two diametrically opposed situations fully illustrated and verified the vital role and the influence of the surfactants, such as PVP, in the construction of the composite materials.
Encapsulation of semiconductor materials in the internal holes of MOFs had higher electron transport and separation efficiency than encapsulation on the surface, which will be more conducive to hydrogen production. In Maji’s article [72], a MOF-based composite where CdS NPs were confined inside the nanosized pores of Zr-based MOF-808 was synthesized and named CdS@MOF-808. The authors anchored L-cysteine in the nanospace of MOF-808 by post-synthetic ligand change, where it could capture Cd2+ ions from the aqueous solution, which further facilitated the in situ growth of CdS NPs in MOFs pores (Figure 9). The CdS4@MOF-808 achieved an excellent visible-light-driven H2 evolution rate of 10.41 mmol·h−1·g−1, which was about 60-fold higher than that of similar MOFs composites (CdS NPs simply supported on the surface of MOFs). Powerful studies had shown that the key factor leading to the large difference in the photocatalytic performance between these two types MOFs composites was the rapid and effective transfer from CdS to Zr4+ clusters in nanospace, which inhibited the electron–hole pairs recombination. Thus, this research revealed the critical role of the NPs photosensitizer stabilized near the catalytic site for the development of efficient H2 evolution photocatalysts in MOFs-based systems.
In the field of the photocatalytic decomposition of water, the construction of semiconductor heterojunctions usually requires energy matching and band bending, so that the formed heterojunctions are thermodynamically permissible for hydrogen generation. However, the band bending had not been demonstrated in MOF photocatalysts. Jiang’s research group [73] expounded and demonstrated this effect in MOFs. In this work, a representative MOF (MIL-125-NH2) was integrated with metal oxides (MoO3 and V2O5) with appropriate work functions and energy levels to provide the corresponding MOF composite (Figure 10a,b). The consequence of the surface photovoltage displayed that the band bending of the MOF composite led to the generation of the built-in electric field of MIL-125-NH2, which promoted the charge separation. In the end, the MOF composites exhibited 56 and 42 times that of the pristine MOF in photocatalytic HER. This literature had a novel perspective, elucidating the band bending of MOF for the first time and supporting its semiconductor-like properties, which would greatly promote the practical applications of MOF in photocatalysis.
Inspired by the electron transfer channel in natural photosynthesis, a novel strategy of multi-stepwise charge transfer was proposed by Jiang and his co-authors [74] for the ingenious design of photocatalysts. The NM@OM/TiO2 dual heterojunction was constructed via depositing TiO2 nanoparticles on the surface of a core-shell NM@OM hetero-skeleton and the preparation procedure of the NM@OM/TiO2 is shown in Figure 11a. The obtained NM@OM/TiO2 photocatalyst could reach the H2 evolution rate of 7.108 mmol·g−1·h−1, which was much superior to that of NM/TiO2, NM@OM, TiO2, OM, and NM. In the NM@OM/TiO2 dual heterojunction, the charge transfer path from NM (−1.27 eV), OM (−1.10 eV) to TiO2 (−0.99 eV) and the hole transfer path from TiO2 (2.35 eV), OM (2.23 eV) to NM (1.41 eV) were established (Figure 11b). This three-step electron transfer process, which imitated natural fabrication, had thermodynamic advantages over most heterojunctions composed of two materials. This work might give guidance for the rational design of high performance photocatalysts and stimulate interest in charge transfer enhancement.

4. MOFs/COFs-Based Hybrids

The coupling of MOFs and COFs could not only make full use of the 2D layered plane structure of COFs as a solid substrate, but also maximized the advantages of the high porosity and easy mass transfer of MOFs [75,76,77,78]. Achieving efficient and stable hydrogen evolution of MOFs/COFs-based hybrids without the assistance of precious metals as cocatalysts was still an arduous and meaning task.
Wang et al. [79] reported the synthesis of a new series of Cu3(HHTP)2-MOF/Tp-Pa-1-COF hybrids with differences in MOF content (Figure 12). The as-synthesized MOF/COF hybrids exhibited close interactions through the coordination of Cu ions with the carbonyl oxygen and enamine nitrogen groups in Tp-Pa-1. In the optimal conditions, the hydrogen generation rate of the photocatalysts with a Cu2(HHTP)2:Tp-Pa-1 ratio of 1:15 could achieve 1.76 mmol·h−1·g−1 under visible-light illumination, which was about 93-fold higher than that of the pure Tp-Pa-1. It was worth noting that such a superior hydrogen production rate was even higher than the performance of 3% Pt/Tp-Pa-1. This relevant research provided some inspiration for the design and construction of porous heterostructure photocatalyst materials for efficient visible-light-driven HER in the absence of precious metals.
As an attractive artificial photocatalyst, titanium metal–organic frameworks (Ti-MOFs) exhibited great potentials in the field of solar energy conversion due to their good photo-redox activities (similar to TiO2), which could be appointed as the antenna to absorb visible light. Although many efforts had been made to develop Ti-MOFs with a high photocatalytic performance, their conversion rates of solar energy were still unsatisfactory. In Ma’s literature [80], they implemented a strategy of covalent integration to synthesize a series of multivariate Ti-MOF/COF hybrid materials PdTCCP@PCN-415(NH2)/TpPa (composites 1, 2, and 3) with excellent visible-light utilization (Figure 13). The experimental results manifested that material 2 presented a H2 evolution rate of 13.98 mmol·h−1·g−1 with a TON value of 24,508 after irradiation of 120 h (TOF = 204·h−1). Importantly, this reaction rate was the highest H2 production performance of MOFs-/COFs-based photocatalysts reported thus far. This paper contributed a novel covalent integration strategy for the construction of MTV-Ti-MOF/COF hybrid materials, which might promote the design and application of other types of MTV-MOF/COF in photocatalysis.
In Pan’s work [81], aldehyde-modified MOFs (NH2-MIL-101(Fe)) were covalently bonded with covalent organic frameworks (Schiff-base network-1, SNW-1) to form the NH2-MIL-101(Fe)@SNW-1 (MS-x) heterojunction via a post-synthetic covalent modification method (Figure 14a). Under the simulated sunlight, the MS-0.8 exhibited great photodegradation pollutant ability (93% in 90 min for TC and 97.2% in 90 min for RMB) and photocatalytic H2 production performance (1949.56 μmol·g−1·h−1). The formation of -C=N- bonds between the NH2-MIL-101(Fe) and SNW-1, accelerated the separation of photogenerated electrons and ameliorated the recombination of electron–hole pairs. The corresponding tests indicated that the matched band structure of NH2-MIL-101(Fe) and SNW-1 could generate a type-II heterojunction (Figure 14b), which explained why the composite system had such superior photocatalytic performance.
Similarly, using the Schiff-base condensation reaction, Jiang et al. [82] in situ constructed a core-shell MOF@COF hetero-framework materials (NH2-UiO-66@TFPT-DETH) via a two-step method (Figure 15a). The SEM image of NH2-UiO-66 and the SEM and TEM images of U@TDE4 (NH2-UiO-66@TFPT-DETH4) are displayed in Figure 15b,d, where both the large differences in morphology between the two substances and the core-shell structure of U@TDE4 could be clearly observed. The as-synthesized photocatalyst exhibited an excellent H2 production rate of 7178 μmol·g−1·h−1, which was quite competitive, especially compared to the pristine COF materials reported. In addition, there were no obvious decreases in long-term photocatalytic test for 12 h, which indicated that the U@TDE4 had outstanding reaction stability. The MOF@COF hetero-skeleton had a high surface area with the coexistence of micropores and mesopores, which extended the visible-light absorption and facilitated the efficient exciton dissolution and transfer. This work might give the inspiration for the design of hetero-framework photocatalysts and their application in solar energy conversion.

5. MOF-Derived Materials

MOFs are composed of regularly arranged structures that extend indefinitely in 3D space, which gives reprocessed MOF-derived materials an additional high regularity. The similar composition of retained components, orderly morphology change, brief and efficient post-processing, and other characteristics make MOFs become a kind of crucial precursor for the preparation of those materials with special needs [83,84]. After pyrolysis, MOFs can provide a variety of derivatives, including metal oxides, metal composites, metal carbides, and metal nitrides. In addition, the photo-response of these derivatives can be enhanced via doping some other atoms. MOF derivatives significantly expand the range of catalysts available and avoid some of the drawbacks associated with the direct utilization of MOFs as the catalysts.
Because of the composition diversity of MOFs, MOF-derived polymetallic compounds are easy to prepare. Nevertheless, there are few reports on the synthesis and preparation of MOF-derived porous metal materials at lower temperatures due to the collapse of the pore structure (prolonged treatment at high temperatures) during the synthesis process. Porous ternary metal nitrides (Ni3FeN) derived from Prussian blue analogue (PBA) were synthesized via a simple two-step oxidation rapid nitride method reported by Yang and his co-authors (Figure 16a) and used in photocatalytic hydrogen evolution [85]. When the Eosin-Y was appointed as a dye photosensitizer, the porous ternary Ni3FeN photocatalyst exhibited the excellent H2 production rate of 16.96 mmol·g−1·h−1 with an AQE of ~3.03% at 520 nm, which was more than 3.45, 3.83, and 1.31 times that of the PBA, NiFeOx, and Prussian-blue-derived Fe3N samples. On the one hand, the reaction kinetics of the photocatalytic HER was ameliorated through the well-proportioned adsorption of the Eosin-Y on the porous Ni3FeN nanocube. On the other hand, the superior conductivity of ternary Ni3FeN boosted the separation and transfer of photogenerated electrons from the Eosin-Y and enhanced the HER performance of the Ni3FeN/Eosin-Y photocatalytic system (Figure 16b). This report was expected to give a novel guideline for MOF-derived ternary metal nitrides to realize the efficient conversion from solar energy to chemical energy.
In Qin’s paper [86], they described the competitive evolution, the morphological and structural changes from Zn-based crystals to amorphous particles. The controlled contribution of organic linkers selectively derived six Zn-CPPs with multivariate characteristics (Figure 17). The FESEM images and the powder XRD patterns of CPPS-derived hollow ZnO microtubes and microspheres are displayed in Figure 18a–c. Based on the diversity of these substructures, hollow ZnO particles were generated via self-pyrolysis and effectively modified by ultrathin doped nanosheets (Figure 18d). The as-synthesized double-sided heterojunctions offered fully covered active sites and brought the efficient photo-excited electron transfer nanochannels (Figure 18e), which gave the superior H2 evolution rate of 4512.5 μmol·h−1·g−1 with stable cyclability. In addition, the basic steps of HER on the O-ZnIn2S4/ZnO heterostructure were determined by DFT calculation (see the original text), which could also be used with activity descriptors to further understand the HER performance of the present photocatalytic system. Herein, this current work opened a new paradigm for the synthesis of custom metal oxides and refined composites for efficient and functional photocatalytic applications.
In other research on the regulation of electron transfer pathways reported by Jin et al. [87], the CdS-MoS2 dumbbell structure was primarily synthesized by a solvothermal method to make the photo-excited electrons flow along the one-dimensional axis. The confinement effect of MOF nanoparticles produced by the silica hollow hexagonal prism greatly expanded the spectral absorption range of the composite photocatalyst. After that, the CdS-In2O3 S-scheme was successfully constructed through a simple electrostatically driven self-assembly method (Figure 19a). The authors attributed the superior photocatalytic activity (198.58 mmol·h−1·g−1) and stability (without obvious decrease for 20 h in four cycles) of this composite system to two aspects. Firstly, the formation of an internal electric field drove the photogenerated electrons in In2O3 to move to CdS, which accelerated the separation of electron–hole pairs. Secondly, the combination of In2O3 with the side wall of CdS-MoS2 dumbbells weakened the surface oxidation kinetics, thus inhibiting the photocorrosion reaction and giving such a good performance (Figure 19b).
In contrast to the method mentioned above in which the electrons were separated in multiple dimensions, Dai et al. [88] inhibited the photogenerated carrier recombination via introducing C with high electronic conduction between the heterojunctions. A hierarchical C/HT-In2O3/ZnIn2S4 heterostructure was fabricated via growing thin-layered ZnIn2S4 on the surface of carbon-coated hollow tubular In2O3 (C/HT-In2O3) derived from In-MOF (Figure 20a). In the optimal conditions, the hybrid system exhibited a hydrogen evolution rate of 920.5 μmol/m2, which was about 13.2 and 6.6 times higher than that of pristine C/HT-In2O3 and ZnIn2S4, respectively. The comparison of interfacial electron transfer pathways for straddling type (In2O3/ZnInS4) and staggered model (C/HT-In2O3/ZnInS4) is displayed in Figure 20b. The EXAFS analysis and density functional theory (DFT) calculation revealed that the faster charge transfer and significantly enhanced photocatalytic HER performance could be attributed to the narrowed band gap of C/HT-In2O3 and the formation of the staggered heterostructure between C/HT-In2O3 and ZnIn2S4. This work manifested that the feasibility of establishing coordinated In-N-In sites at the carbon-coated HT-In2O3/ZnIn2S4 heterostructure interface to facilitate charge transfer and introduced an ideal photo-activated HER catalyst.
If bimetallic MOFs were appointed as precursors, proper regulation of the ratio of the two metals in the MOFs and subsequent correct reprocessing were both vital. In Jin’s work [89], the bimetallic CeCo-MOF was synthesized via a hydrothermal method, and various composite catalysts were obtained from different treatment methods including phosphating, oxidation, and sulfurization (Figure 21). After reasonable adjustment of the Ce:Co ratio, it was finally determined that when the ratio of Ce:Co was 1:2, the composite phosphating catalyst had the best performance, and the photocatalytic HER rate could reach 8049 μmol·g−1·h−1. They attributed the superior reactivity of the as-synthesized CeCo1:2/P catalyst to three factors. Firstly, the Ce and Co sources were uniformly dispersed using bimetallic CeCo-MOF synthesized by a one-pot solvothermal method. Secondly, phosphating ameliorated the photo-absorption performance of the catalyst, which provided the prerequisite conditions for the generation of photogenerated electron–hole pairs. Thirdly, the close contact between CeO1.66 and CoP ensured the smooth transfer of photogenerated electrons between the catalysts. This report might provide novel insights for post-processing bimetallic MOF materials and their further applications in photocatalytic HER.

6. Conclusions and Prospects

In this review, we summarized the research advances of MOFs catalysts in the field of photocatalytic HER in the past three years. The contents mainly focused on the preparation and synthesis of these catalysts, which included four separate aspects: internal modification and structure optimization of MOFs materials, MOFs/semiconductor composites, MOFs/COFs-based hybrids, and MOFs-derived materials. The MOFs-based reaction system and corresponding HER rates mentioned in this review are listed in Table 1, and the unit of H2 evolution rate was unified as μmol∙g−1∙h−1. Compared with semiconductor materials and molecular catalysts, although the corresponding research about MOF materials started relatively late, their excellent potential in this field was quite stunning. Nevertheless, some unavoidable shortcomings remained, and the challenges were complicated, meaning that necessary efforts would contribute to the long-term and sustained development of this field.
(1) In most of the examples that have been reported in the field of photocatalysis, the MOFs that were mainly focused on were In [90,91,92], Ti [93,94], and Zr [95,96] MOFs, while there were few reports based on traditional transition metals such as Fe, Ni, and Cu, etc. This reduced the universality of synthetic methods and limited the large-scale utilization of similar materials. Some universal MOFs such as UiO-66, ZIF-8, and MOF-67, with temperature tolerance and acid-base environment tolerance need to be expanded and developed, which would lay a satisfactory foundation for building a broader MOF platform. Furthermore, some MOFs with high crystallinity can be regarded as a kind of special semiconductor material with band gap, so some methods of constructing heterojunctions can be referred to load semiconductor materials on MOFs [97,98].
(2) In terms of light sources, most laboratory studies use 300 W Xe lamps to simulate visible light rather than direct sunlight. In general, the amount of hydrogen produced by direct sunlight irradiation was usually smaller than produced by simulated visible light. Therefore, it was necessary to test the catalytic performance of the as-synthesized MOF-based materials under outdoor or direct light conditions. Alternatively, the single input amount of catalyst did not reach the gram level. The catalyst used in the experiments was usually between 5–50 mg, which greatly hindered the expansion for the practical application of this kind of catalyst. The reason for this phenomenon might be due to the inability of larger mass catalysts to disperse uniformly during photocatalytic reactions in mixed solutions, leading to agglomeration, which means that the catalyst per unit mass cannot be fully utilized. Hence, how to increase the amount of catalyst to a gram level reaction and how to carry out the hydrogen evolution reaction in real sunlight were two urgent problems to be solved in the process of the industrialization of this field.
(3) Although the high reactivity of photocatalytic HER was the most pursued goal of researchers, the long-term reaction stability and cyclic stability were also of great significance. There was no systematic evaluation of cyclic stability in some articles, and the characterization and further analysis of the cycled catalysts were also ambiguous. For example, the morphology of some catalysts changed greatly after the reaction, which might be the main reason for the decrease in the catalytic cycle reaction efficiency, but some authors did not discuss this part in detail. The superior cycle stability of these photocatalysts would lay a solid foundation for their subsequent industrial development.
(4) We should acknowledge that there were still many unknown processes or speculative results when it came to mechanism verification. This was related to the complexity of the composite system, which included the interaction between catalyst and reaction solution, the structure-activity relationship of catalysts themselves, whether photogenerated electron transfer between photosensitizer and cocatalyst was thermodynamically allowed, etc. More detailed and clear explanations of the mechanism might require advanced characterization techniques, more accurate theoretical calculations [99], and unremitting efforts of relevant researchers.
(5) For MOF derivatives and their composite systems, researchers should focus on the original driving forces for their formation or co-existence, including coordination bonds, covalent bonds, intermolecular or electrostatic forces, etc., which will provide guidance for the development and utilization of novel MOFs-based materials in a broader sense. In addition, when some newly synthesized MOFs are appointed as precursors, it will be more economical if the cost can be reduced and the preparation process can be simplified. This will make the advantages of MOFs more obvious and prominent when compared with the pure inorganic salts as precursors.
(6) Last but not the least, the solvents of many reaction systems are not pure water. The addition of organic solvents not only increases the cost, but also introduces the instability of the reaction. The addition of organic solvents with strong solubility may change the reaction system from heterogeneous to homogeneous. The comparison of activity and stability of homogeneous or heterogeneous systems is also vital for MOFs composite systems. At the same time, the recyclability of MOFs is very important and should be taken seriously. In addition, for some photocatalytic systems involving electron sacrificial agents, such as ascorbic acid, that can dissociate hydrogen ions, the hydrogen atoms in the precipitated hydrogen may be ionized by ascorbic acid, so some necessary verification experiments are needed.

Author Contributions

Writing—original draft preparation, H.Z.; validation, C.L.; writing—review and editing, Y.L.; resources, J.P.; project administration, funding acquisition, X.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant 22035003 and Grant 22201137), Nature Science Fund of Tianjin, China (Grant 19JCZDJC37200), Fundamental Research Funds for the Central Universities (63223020), and the Haihe Laboratory of Sustainable Chemical Transformations (YYJC202101).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tian, D.; Denny, S.; Li, K.; Wang, H.; Kattel, S.; Chen, J. Density functional theory studies of transition metal carbides and nitrides as electrocatalysts. Chem. Soc. Rev. 2021, 50, 12338. [Google Scholar] [CrossRef] [PubMed]
  2. Guerra, O.; Eichman, J.; Denholm, P. Optimal energy storage portfolio for high and ultrahigh carbon-free and renewable power systems. Energy Environ. Sci. 2021, 14, 5132. [Google Scholar] [CrossRef]
  3. Lv, J.; Xie, J.; Mohamed, A.; Zhang, X.; Wang, Y. Photoelectrochemical energy storage materials: Design principles and functional devices towards direct solar to electrochemical energy storage. Chem. Soc. Rev. 2022, 51, 1511. [Google Scholar] [CrossRef]
  4. Luo, Q.; Liu, P.; Fu, L.; Hu, Y.; Yang, L.; Wu, W.; Kong, X.; Jiang, L.; Wen, L. Engineered Cellulose Nanofiber Membranes with Ultrathin Low-Dimensional Carbon Material Layers for Photothermal-Enhanced Osmotic Energy Conversion. ACS Appl. Mater. Interfaces 2022, 14, 13223–13230. [Google Scholar] [CrossRef] [PubMed]
  5. Chen, Z.; Mian, M.R.; Lee, S.-J.; Chen, H.; Zhang, X.; Kirlikovali, K.O.; Shulda, S.; Melix, P.; Rosen, A.S.; Parilla, P.A.; et al. Fine-Tuning a Robust Metal-Organic Framework toward Enhanced Clean Energy Gas Storage. J. Am. Chem. Soc. 2021, 143, 18838–18843. [Google Scholar] [CrossRef]
  6. Lin, B.; Ma, B.; Chen, J.; Zhou, Y.; Zhou, J.; Yan, X.; Xue, C.; Luo, X.; Liu, Q.; Wang, J.; et al. Sea-urchin-like ReS2 nanosheets with charge edge-collection effect as a novel cocatalyst for high-efficiency photocatalytic H2 evolution. Chin. Chem. Lett. 2022, 33, 943–947. [Google Scholar] [CrossRef]
  7. Alves, H.; Frachoni, B.; Nunes, B.; Teixeira, P.; Paniago, R.; Bahnemann, D.; Paterno, L.; Patrocinio, A. Highly Stable Au/Hexaniobate Nanocomposite Prepared by a Green Intercalation Method for Photoinduced H2 Evolution Applications. ACS Appl. Energy Mater. 2022, 5, 8371–8380. [Google Scholar] [CrossRef]
  8. Bui, V.; Kumar, A.; Bui, H.; Lee, J.; Hwang, Y.; Le, H.; Kawazoe, Y.; Lee, H. Boosting Electrocatalytic HER Activity of 3D Interconnected CoSP via Metal Doping: Active and Stable Electrocatalysts for pH-Universal Hydrogen Generation. Chem. Mater. 2020, 32, 9591–9601. [Google Scholar] [CrossRef]
  9. Sun, Z.; Liang, Y.; Wu, Y.; Yu, Y.; Zhang, B. Boosting Electrocatalytic Hydrogen-Evolving Activity of Co/CoO Heterostructured Nanosheets via Coupling Photogenerated Carriers with Photothermy. ACS Sustain. Chem. Eng. 2018, 6, 11206–11210. [Google Scholar] [CrossRef]
  10. Jiang, L.; Guo, Y.; Qi, S.; Zhang, K.; Chen, J.; Lou, Y.; Zhao, Y. Amorphous NiCoB-coupled MAPbI3 for efficient photocatalytic hydrogen evolution. Dalton Trans. 2021, 50, 17960. [Google Scholar] [CrossRef]
  11. Su, H.; Rao, C.; Zhou, L.; Pang, Y.; Lou, H.; Yang, D.; Qiu, X. Mo-Doped/Ni-supported ZnIn2S4-wrapped NiMoO4 S-scheme heterojunction photocatalytic reforming of lignin into hydrogen. Green Chem. 2022, 24, 2027. [Google Scholar] [CrossRef]
  12. Soltau, S.; Niklas, J.; Dahlberg, P.; Poluektov, O.; Tiede, D.; Mulfort, K.; Utschig, L. Aqueous light driven hydrogen production by a Ru–ferredoxin–Co biohybrid. Chem. Commun. 2015, 51, 10628–10631. [Google Scholar] [CrossRef] [PubMed]
  13. Yuan, Y.; Yu, Z.; Chen, D.; Zou, Z. Metal-complex chromophores for solar hydrogen generation. Chem. Soc. Rev. 2017, 46, 603. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, P.; Jacques, P.; Chavarot-Kerlidou, M.; Wang, M.; Sun, L.; Fontecave, M.; Artero, V. Phosphine Coordination to a Cobalt Diimine−Dioxime Catalyst Increases Stability during Light-Driven H2 Production. Inorg. Chem. 2012, 51, 2115–2120. [Google Scholar] [CrossRef] [PubMed]
  15. Luan, X.; Dai, H.; Li, Q.; Xu, F.; Mai, Y. A Hybrid Photocatalyst Composed of CdS Nanoparticles and Graphene Nanoribbons for Visible-Light-Driven Hydrogen Production. ACS Appl. Energy Mater. 2022, 5, 8621–8628. [Google Scholar] [CrossRef]
  16. Zhang, J.; Cai, P.; Lin, J. Modulation of the Band Bending of CdS by Fluorination to Facilitate Photoinduced Electron Transfer for Efficient H2 Evolution over Pt/CdS. J. Phys. Chem. C 2022, 126, 7896–7902. [Google Scholar] [CrossRef]
  17. Xie, X.; Wang, R.; Ma, Y.; Chen, J.; Shi, Z.; Cui, Q.; Li, Z.; Xu, C. Sulfate-Functionalized Core−Shell ZnO/CdS/Ag2S Nanorod Arrays with Dual-Charge-Transfer Channels for Enhanced Photoelectrochemical Performance. ACS Appl. Energy Mater. 2022, 5, 6228–6237. [Google Scholar] [CrossRef]
  18. Yang, Y.; Wu, J.; Cheng, B.; Zhang, L.; Al-Ghamdi, A.; Wageh, S.; Li, Y. Enhanced Photocatalytic H2-production Activity of CdS Nanoflower using Single Atom Pt and Graphene Quantum Dot as Dual Cocatalysts. Chin. J. Struct. Chem. 2022, 41, 2206006–2206014. [Google Scholar]
  19. Gao, R.; He, H.; Bai, J.; Hao, L.; Shen, R.; Zhang, P.; Li, Y.; Li, X. Pyrene-benzothiadiazole-based Polymer/CdS 2D/2D Organic/Inorganic Hybrid S-scheme Heterojunction for Efficient Photocatalytic H2 Evolution. Chin. J. Struct. Chem. 2022, 41, 2206031–2206038. [Google Scholar]
  20. Sahoo, S.; Mansingh, S.; Babu, P.; Parida, K. Black titania an emerging photocatalyst: Review highlighting the synthesis techniques and photocatalytic activity for hydrogen generation. Nanoscale Adv. 2021, 3, 5487. [Google Scholar] [CrossRef]
  21. Liu, J.; Li, D.; Liu, X.; Zhou, J.; Zhao, H.; Wang, N.; Cui, Z.; Bai, J.; Zhao, Y. TiO2/g-C3N4 heterojunction hollow porous nanofibers as superior visible-light photocatalysts for H2 evolution and dye degradation. N. J. Chem. 2021, 45, 22123. [Google Scholar] [CrossRef]
  22. Pan, Y.; Xiong, B.; Li, Z.; Wu, Y.; Yan, C.; Song, H. In situ constructed oxygen-vacancy-rich MoO3-x/ porous g-C3N4 heterojunction for synergistically enhanced photocatalytic H2 evolution. RSC Adv. 2021, 11, 31219. [Google Scholar] [CrossRef] [PubMed]
  23. Huang, D.; Sun, X.; Liu, Y.; Ji, H.; Liu, W.; Wang, C.; Ma, W.; Cai, Z. A carbon-rich g-C3N4 with promoted charge separation for highly efficient photocatalytic degradation of amoxicillin. Chin. Chem. Lett. 2021, 32, 2787–2791. [Google Scholar]
  24. Hu, Y.; Li, X.; Wang, W.; Deng, F.; Han, L.; Gao, X.; Feng, Z.; Chen, Z.; Huang, J.; Zeng, F. Bi and S Co-doping g-C3N4 to Enhance Internal Electric Field for Robust Photocatalytic Degradation and H2 Production. Chin. J. Struct. Chem. 2022, 41, 2206069–2206078. [Google Scholar]
  25. Tao, S.; Wan, S.; Huang, Q.; Li, C.; Yu, J.; Cao, S. Molecular Engineering of g-C3N4 with Dibenzothiophene Groups as Electron Donor for Enhanced Photocatalytic H2-Production. Chin. J. Struct. Chem. 2022, 41, 2206048–2206054. [Google Scholar]
  26. Tang, Y.; Hu, X.; Liu, C. Perfect inhibition of CdS photocorrosion by graphene sheltering engineering on TiO2 nanotube array for highly stable photocatalytic activity. Phys. Chem. Chem. Phys. 2014, 16, 25321. [Google Scholar] [CrossRef]
  27. Zhang, H.; Zhu, H.; Zhao, H.; Dou, M.; Yin, X.; Yang, H.; Li, D.; Dou, J. A novel dinuclear cobalt-bis(thiosemicarbazone) complex as a cocatalyst to enhance visible-light-driven H2 evolution on CdS nanorods and a mechanism discussion. J. Photoch. Photobio. A 2022, 426, 113771. [Google Scholar] [CrossRef]
  28. Chen, S.; Huang, D.; Xu, P.; Xue, W.; Lei, L.; Cheng, M.; Wang, R.; Liu, X.; Deng, R. Semiconductor-based photocatalysts for photocatalytic and photoelectrochemical water splitting: Will we stop with photocorrosion? J. Mater. Chem. A 2020, 8, 2286. [Google Scholar] [CrossRef]
  29. Chen, R.; Jiang, S.; Zhang, Q.; Luo, Y. Intermediate Complex-Mediated Interfacial Electron Transfer in a Radical Dianion/TiO2 Dye-Sensitized Photocatalytic System. J. Phys. Chem. Lett. 2022, 13, 8091–8096. [Google Scholar] [CrossRef]
  30. Yuan, C.; Shen, Y.; Zhu, C.; Zhu, P.; Yang, F.; Liu, J.; An, C. Ru Single-Atom Decorated Black TiO2 Nanosheets for Efficient Solar-Driven Hydrogen Production. ACS Sustain. Chem. Eng. 2022, 10, 10311–10317. [Google Scholar] [CrossRef]
  31. Liu, Y.; Shen, C.; Jiang, N.; Zhao, Z.; Zhou, X.; Zhao, S.; Xu, A. g-C3N4 Hydrogen-Bonding Viologen for Significantly Enhanced Visible-Light Photocatalytic H2 Evolution. ACS Catal. 2017, 7, 8228–8234. [Google Scholar] [CrossRef]
  32. Hong, I.; Chen, Y.; Hsu, Y.; Yong, K. Triple-Channel Charge Transfer over W18O49/Au/g-C3N4 Z-Scheme Photocatalysts for Achieving Broad-Spectrum Solar Hydrogen Production. ACS Appl. Mater. Interfaces 2021, 13, 52670–52680. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, S.; Lan, H.; Guan, G.; Yang, Q. Amino-Functionalized Microporous MOFs for Capturing Greenhouse Gases CF4 and NF3 with Record Selectivity. ACS Appl. Mater. Interfaces 2022, 14, 40072–40081. [Google Scholar] [CrossRef] [PubMed]
  34. Tarasi, S.; Ramazani, A.; Ramazani, A.; Hu, M.; Ghafghazi, S.; Tarasi, R.; Ahmadi, Y. Drug Delivery Using Hydrophilic Metal−Organic Frameworks (MOFs): Effect of Structure Properties of MOFs on Biological Behavior of Carriers. Inorg. Chem. 2022, 61, 13125–13132. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, S.; Zhang, J.; Zong, M.; Xu, J.; Wang, D.; Bu, X. Energy Level Engineering: Ru Single Atom Anchored on Mo-MOF with a [Mo8O26(im)2]4− Structure Acts as a Biomimetic Photocatalyst. ACS Catal. 2022, 12, 7960–7974. [Google Scholar] [CrossRef]
  36. Kaushal, S.; Kaur, G.; Kaur, J.; Singh, P. First transition series metal-organic frameworks: Synthesis, properties and applications. Mater. Adv. 2021, 2, 7308. [Google Scholar] [CrossRef]
  37. Li, K.; Yang, J.; Gu, J. Hierarchically Porous MOFs Synthesized by Soft-Template Strategies. Acc. Chem. Res. 2022, 55, 2235–2247. [Google Scholar] [CrossRef]
  38. Hao, J.; Geng, L.; Zheng, J.; Wei, J.; Zhang, L.; Feng, R.; Zhao, J.; Li, Q.; Pang, J.; Bu, X. Ligand Induced Double-Chair Conformation Ln12 Nanoclusters Showing Multifunctional Magnetic and Proton Conductive Properties. Inorg. Chem. 2022, 61, 3690–3696. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Xu, X.; Yan, B. A multicolor-switchable fluorescent lanthanide MOFs triggered by anti-cancer drugs: Multifunctional platform for anti-cancer drug sensing and information anticounterfeiting. J. Mater. Chem. C 2022, 10, 3576. [Google Scholar] [CrossRef]
  40. Sharp, C.; Bukowski, B.; Li, H.; Johnson, E.; Ilic, S.; Morris, A.; Gersappe, D.; Snurr, R.; Morris, J. Nanoconfinement and mass transport in metal-organic frameworks. Chem. Soc. Rev. 2021, 50, 11530. [Google Scholar] [CrossRef]
  41. Sun, Z.; Liao, Y.; Zhao, S.; Zhang, X.; Liu, Q.; Shi, X. Research progress in metal–organic frameworks (MOFs) in CO2 capture from post-combustion coalfired flue gas: Characteristics, preparation, modification and applications. J. Mater. Chem. A 2022, 10, 5174. [Google Scholar] [CrossRef]
  42. Zeeshan, M.; Shahid, M. State of the art developments and prospects of metal–organic frameworks for energy applications. Dalton Trans. 2022, 51, 1675. [Google Scholar] [CrossRef] [PubMed]
  43. Li, T.; Jia, X.; Chen, H.; Chang, Z.; Li, L.; Wang, Y.; Li, J. Tuning the Pore Environment of MOFs toward Efficient CH4/N2 Separation under Humid Conditions. ACS Appl. Mater. Interfaces 2022, 14, 15830–15839. [Google Scholar] [CrossRef]
  44. Lisensky, G.; Yaghi, O. Visualizing Pore Packing and Topology in MOFs. J. Chem. Educ. 2022, 99, 1998–2004. [Google Scholar] [CrossRef]
  45. Ahmad, M.; Castro-Muñoz, R.; Budd, P. Boosting gas separation performance and suppressing the physical aging of polymers of intrinsic microporosity (PIM-1) by nanomaterial blending. Nanoscale 2020, 12, 23333. [Google Scholar] [CrossRef]
  46. Xia, Y.; Wang, C.; Yu, M.; Bu, X. A unique 3D microporous MOF constructed by cross-linking 1D coordination polymer chains for effectively selective separation of CO2/CH4 and C2H2/CH4. Chin. Chem. Lett. 2021, 32, 1153–1156. [Google Scholar] [CrossRef]
  47. Qiao, Y.; Chang, X.; Zheng, J.; Yi, M.; Chang, Z.; Yu, M.; Bu, X. Self-Interpenetrated Water-Stable Microporous Metal−Organic Framework toward Storage and Purification of Light Hydrocarbons. Inorg. Chem. 2021, 60, 2749–2755. [Google Scholar] [CrossRef] [PubMed]
  48. Meng, M.; Liu, X.; Li, N.; Zhao, J.; Chang, Z.; Zheng, J.; Bu, X. Structural Transformation and Spatial Defect Formation of a Co(II) MOF Triggered by Varied Metal-Center Coordination Configuration. Inorg. Chem. 2020, 59, 9005–9013. [Google Scholar] [CrossRef] [PubMed]
  49. Yin, J.; Chang, Z.; Li, N.; He, J.; Fu, Z.; Bu, X. Efficient Regulation of Energy Transfer in a Multicomponent Dye-Loaded MOF for White-Light Emission Tuning. ACS Appl. Mater. Interfaces 2020, 12, 51589–51597. [Google Scholar] [CrossRef] [PubMed]
  50. Gupta, M.; Zhu, Z.; Kottilil, D.; Rath, B.; Tian, W.; Tan, Z.; Liu, X.; Xu, Q.; Ji, W.; Vittal, J. Impact of the Structural Modification of Diamondoid Cd(II) MOFs on the Nonlinear Optical Properties. ACS Appl. Mater. Interfaces 2021, 13, 60163–60172. [Google Scholar] [CrossRef]
  51. Liu, L.; Chen, Q.; Lv, J.; Li, Y.; Wang, K.; Li, J. Stable Metal-Organic Frameworks for Fluorescent Detection of Tetracycline Antibiotics. Inorg. Chem. 2022, 61, 8015–8021. [Google Scholar] [CrossRef] [PubMed]
  52. Yang, Y.; Xu, S.; Gan, Y.; Zhang, B.; Chen, L. Recent Progresses in Lanthanide Metal-organic Frameworks (Ln-MOFs) as Chemical Sensors for Ions, Antibiotics and Amino Acids. Chin. J. Struct. Chem. 2022, 41, 2211045–2211070. [Google Scholar] [CrossRef]
  53. Jin, S. How to Effectively Utilize MOFs for Electrocatalysis. ACS Energy Lett. 2019, 4, 1443–1445. [Google Scholar] [CrossRef] [Green Version]
  54. Dou, S.; Li, X.; Wang, X. Rational Design of Metal−Organic Frameworks towards Efficient Electrocatalysis. ACS Materials Lett. 2020, 2, 1251–1267. [Google Scholar] [CrossRef]
  55. Hou, X.; Jiang, T.; Xu, X.; Wang, X.; Zhou, J.; Xie, H.; Liu, Z.; Chu, L.; Huang, M. Coupling of NiFe-Based Metal-Organic Framework Nanosheet Arrays with Embedded Fe-Ni3S2 Clusters as Efficient Bifunctional Electrocatalysts for Overall Water Splitting. Chin. J. Struct. Chem. 2022, 41, 2207074–2207080. [Google Scholar]
  56. Xia, T.; Lin, Y.; Li, W.; Ju, M. Photocatalytic degradation of organic pollutants by MOFs based materials: A review. Chin. Chem. Lett. 2021, 32, 1153–1156. [Google Scholar] [CrossRef]
  57. Kim, H.; Kim, N.; Ryu, J. Porous framework-based hybrid materials for solar-to-chemical energy conversion: From powder photocatalysts to photoelectrodes. Inorg. Chem. Front. 2021, 8, 4107. [Google Scholar] [CrossRef]
  58. Zhang, H.; Sun, R.; Li, D.; Dou, J. A Review on Crystalline Porous MOFs Materials in Photocatalytic Transformations of Organic Compounds in Recent Three Years. Chin. J. Struct. Chem. 2022, 41, 2211071–2211083. [Google Scholar] [CrossRef]
  59. Unnikrishnan, V.; Zabihi, O.; Ahmadi, M.; Li, Q.; Blanchard, P.; Kiziltas, A.; Naebe, M. Metal-organic framework structure-property relationships for high-performance multifunctional polymer nanocomposite applications. J. Mater. Chem. A 2021, 9, 4348. [Google Scholar] [CrossRef]
  60. Li, X.; Rajasree, S.; Yu, J.; Deria, P. The role of photoinduced charge transfer for photocatalysis, photoelectrocatalysis and luminescence sensing in metal-organic frameworks. Dalton Trans. 2020, 49, 12892. [Google Scholar] [CrossRef]
  61. Ma, X.; Liu, H.; Yang, W.; Mao, G.; Zheng, L.; Jiang, H. Modulating Coordination Environment of Single-Atom Catalysts and Their Proximity to Photosensitive Units for Boosting MOF Photocatalysis. J. Am. Chem. Soc. 2021, 143, 12220–12229. [Google Scholar] [CrossRef] [PubMed]
  62. Pan, Y.; Wang, J.; Chen, S.; Yang, W.; Ding, C.; Waseem, A.; Jiang, H. Linker engineering in metal-organic frameworks for dark photocatalysis. Chem. Sci. 2022, 13, 6696. [Google Scholar] [CrossRef] [PubMed]
  63. Liu, L.; Du, S.; Guo, X.; Xiao, Y.; Yin, Z.; Yang, N.; Bao, Y.; Zhu, X.; Jin, S.; Feng, Z.; et al. Water-Stable Nickel Metal-Organic Framework Nanobelts for Cocatalyst-Free Photocatalytic Water Splitting to Produce Hydrogen. J. Am. Chem. Soc. 2022, 144, 2747–2754. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, Z.; Wang, Y.; Niu, B.; Liu, B.; Li, J.; Duan, W. Ultra-stable two-dimensional metal-organic frameworks for photocatalytic H2 production. Nanoscale 2022, 14, 7146. [Google Scholar] [CrossRef]
  65. Indra, A.; Acharjya, A.; Menezes, P.; Merschjann, C.; Hollmann, D.; Schwarze, M.; Aktas, M.; Friedrich, A.; Lochbrunner, S.; Thomas, A.; et al. Boosting Visible-Light-Driven Photocatalytic Hydrogen Evolution with an Integrated Nickel Phosphide-Carbon Nitride System. Angew. Chem. 2017, 129, 1675–1679. [Google Scholar] [CrossRef]
  66. Bi, W.; Zhang, L.; Sun, Z.; Li, X.; Jin, T.; Wu, X.; Zhang, Q.; Luo, Y.; Wu, C.; Xie, Y. Insight into Electrocatalysts as Co-catalysts in Efficient Photocatalytic Hydrogen Evolution. ACS Catal. 2016, 6, 4253–4257. [Google Scholar] [CrossRef]
  67. Callejas, J.; McEnaney, J.; Read, C.; Crompton, J.; Biacchi, A.; Popczun, E.; Gordon, T.; Lewis, N.; Schaak, R. Electrocatalytic and Photocatalytic Hydrogen Production from Acidic and Neutral-pH Aqueous Solutions Using Iron Phosphide Nanoparticles. ACS Nano 2014, 8, 11101–11107. [Google Scholar] [CrossRef]
  68. Cao, S.; Chen, Y.; Wang, C.; Lv, X.; Fu, W. Spectacular photocatalytic hydrogen evolution using metal-phosphide/CdS hybrid catalysts under sunlight irradiation. Chem. Commun. 2015, 51, 8708. [Google Scholar] [CrossRef] [Green Version]
  69. Sun, K.; Liu, M.; Pei, J.; Li, D.; Ding, C.; Wu, K.; Jiang, H. Incorporating Transition-Metal Phosphides into Metal-Organic Frameworks for Enhanced Photocatalysis. Angew. Chem. Int. Ed. 2020, 59, 22749–22755. [Google Scholar] [CrossRef]
  70. Xu, M.; Li, D.; Sun, K.; Jiao, L.; Xie, C.; Ding, C.; Jiang, H. Interfacial Microenvironment Modulation Boosting Electron Transfer between Metal Nanoparticles and MOFs for Enhanced Photocatalysis. Angew. Chem. Int. Ed. 2021, 60, 16372–16376. [Google Scholar] [CrossRef]
  71. Wang, Y.; Zhang, Z.; Li, J.; Yuan, Y.; Yang, J.; Xu, W.; An, P.; Xi, S.; Guo, J.; Liu, B.; et al. Two-Dimensional-on-Three-Dimensional Metal-Organic Frameworks for Photocatalytic H2 Production. Angew. Chem. 2022, 61, e202211031. [Google Scholar] [CrossRef]
  72. Ghosh, A.; Karmakar, S.; Rahimi, A.; Roy, R.; Nath, S.; Gautam, U.; Maji, T. Confinement Matters: Stabilization of CdS Nanoparticles inside a Postmodified MOF toward Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2022, 14, 25220–25231. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, C.; Xie, C.; Gao, Y.; Tao, X.; Ding, C.; Fan, F.; Jiang, H. Charge Separation by Creating Band Bending in Metal-Organic Frameworks for Improved Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2022, 61, e202204108. [Google Scholar]
  74. Chen, Y.; Yang, D.; Xin, X.; Yang, Z.; Gao, Y.; Shi, Y.; Zhao, Z.; An, K.; Wang, W.; Tan, J.; et al. Multi-stepwise charge transfer via MOF@MOF/TiO2 dual-heterojunction photocatalysts towards hydrogen evolution. J. Mater. Chem. A 2022, 10, 9717. [Google Scholar] [CrossRef]
  75. Wang, X.; Wang, Y.; Qi, H.; Chen, Y.; Guo, W.; Yu, H.; Chen, H.; Ying, Y. Humidity-Independent Artificial Olfactory Array Enabled by Hydrophobic Core-Shell Dye/MOFs@COFs Composites for Plant Disease Diagnosis. ACS Nano 2022, 16, 14297–14307. [Google Scholar] [CrossRef] [PubMed]
  76. Sun, D.; Jang, S.; Yim, S.; Ye, L.; Kim, D. Metal Doped Core–Shell Metal-Organic Frameworks@ Covalent Organic Frameworks (MOFs@COFs) Hybrids as a Novel Photocatalytic Platform. Adv. Funct. Mater. 2018, 28, 1707110. [Google Scholar] [CrossRef]
  77. Garzon-Tovar, L.; Perez-Carvajal, J.; Yazdi, A.; Hernandez-Munoz, J.; Tarazona, P.; Imaz, I.; Zamora, F.; Maspoch, D. A MOF@COF Composite with Enhanced Uptakethrough Interfacial Pore Generation. Angew. Chem. 2019, 131, 9612–9616. [Google Scholar] [CrossRef]
  78. Peng, H.; Raya, J.; Richard, F.; Baaziz, W.; Ersen, O.; Ciesielski, A.; Samorì, P. Synthesis of Robust MOFs@COFs Porous Hybrid Materials via an Aza-Diels-Alder Reaction: Towards High-Performance Supercapacitor Materials. Angew. Chem. Int. Ed. 2020, 59, 19602–19609. [Google Scholar] [CrossRef]
  79. Xue, P.; Pan, X.; Huang, J.; Gao, Y.; Guo, W.; Li, J.; Tang, M.; Wang, Z. In Situ Fabrication of Porous MOF/COF Hybrid Photocatalysts for Visible-Light-Driven Hydrogen Evolution. ACS Appl. Mater. Interfaces 2021, 13, 59915–59924. [Google Scholar] [CrossRef]
  80. Chen, C.; Xiong, Y.; Zhong, X.; Lan, P.; Wei, Z.; Pan, H.; Su, P.; Song, Y.; Chen, Y.; Nafady, A.; et al. Enhancing Photocatalytic Hydrogen Production via the Construction of Robust Multivariate Ti-MOF/COF Composites. Angew. Chem. Int. Ed. 2022, 61, e202114071. [Google Scholar]
  81. Wang, Y.; Yang, H.; Shang, Q.; Cheng, Q.; Zhou, H.; Pan, Z. Integrating hollow spherical covalent organic frameworks on NH2-MIL-101(Fe) as high performance heterogeneous photocatalysts. Environ. Sci. Nano 2022, 9, 3081–3093. [Google Scholar] [CrossRef]
  82. Chen, Y.; Yang, D.; Shi, B.; Dai, W.; Ren, H.; An, K.; Zhou, Z.; Zhao, Z.; Wang, W.; Jiang, Z. In situ construction of hydrazone-linked COF-based core-shell hetero-frameworks for enhanced photocatalytic hydrogen evolution. J. Mater. Chem. A 2020, 8, 7724. [Google Scholar] [CrossRef]
  83. Wang, S.; Guan, B.; Lou, X. Construction of ZnIn2S4−In2O3 Hierarchical Tubular Heterostructures for Efficient CO2 Photoreduction. J. Am. Chem. Soc. 2018, 140, 5037–5040. [Google Scholar] [CrossRef] [PubMed]
  84. Torad, N.; Hu, M.; Ishihara, S.; Sukegawa, H.; Belik, A.; Imura, M.; Ariga, K.; Sakka, Y.; Yamauchi, Y. Direct Synthesis of MOF-Derived Nanoporous Carbon with Magnetic Co Nanoparticles toward Efficient Water Treatment. Small 2014, 10, 2096–2107. [Google Scholar] [CrossRef]
  85. Qi, W.; Wang, C.; Yu, J.; Adimi, S.; Thomas, T.; Guo, H.; Liu, S.; Yang, M. MOF-Derived Porous Ternary Nickel Iron Nitride Nanocube as a Functional Catalyst toward Water Splitting Hydrogen Evolution for Solar to Chemical Energy Conversion. ACS Appl. Energy Mater. 2022, 5, 6155–6162. [Google Scholar] [CrossRef]
  86. Zhu, Q.; Xu, S.; Wu, W.; Qi, Y.; Lin, Z.; Li, Y.; Qin, Y. Hierarchical Hollow Zinc Oxide Nanocomposites Derived from Morphology-Tunable Coordination Polymers for Enhanced Solar Hydrogen Production. Angew. Chem. Int. Ed. 2022, 61, e202205312. [Google Scholar] [CrossRef]
  87. Zhang, L.; Jiang, X.; Jin, Z.; Tsubaki, N. Spatially separated catalytic sites supplied with the CdS-MoS2-In2O3 ternary dumbbell S-scheme heterojunction for enhanced photocatalytic hydrogen production. J. Mater. Chem. A 2022, 10, 10715. [Google Scholar] [CrossRef]
  88. Zhang, Q.; Zhang, J.; Wang, X.; Li, L.; Li, Y.; Dai, W. In-N-In Sites Boosting Interfacial Charge Transfer in CarbonCoated Hollow Tubular In2O3/ZnIn2S4 Heterostructure Derived from In-MOF for Enhanced Photocatalytic Hydrogen Evolution. ACS Catal. 2021, 11, 6276–6289. [Google Scholar] [CrossRef]
  89. Li, T.; Zhang, L.; Li, X.; Wang, X.; Jin, Z. Design and synthesis of phosphating bimetallic CeCo-MOF for substantially improved photocatalytic hydrogen evolution. J. Mater. Chem. C 2022, 10, 8750. [Google Scholar] [CrossRef]
  90. Zhuang, G.; Fang, Q.; Wei, J.; Yang, C.; Chen, M.; Lyu, Z.; Zhuang, Z.; Yu, Y. Branched In2O3 Mesocrystal of Ordered Architecture Derived from the Oriented Alignment of a Metal-Organic Framework for Accelerated Hydrogen Evolution over In2O3-ZnIn2S4. ACS Appl. Mater. Interfaces 2021, 13, 9804–9813. [Google Scholar] [CrossRef]
  91. Hou, Q.; Li, X.; Pi, Y.; Xiao, J. Construction of In2S3@NH2-MIL-68(In)@In2S3 Sandwich Homologous Heterojunction for Efficient CO2 Photoreduction. Ind. Eng. Chem. Res. 2020, 59, 20711–20718. [Google Scholar] [CrossRef]
  92. Bariki, R.; Das, K.; Pradhan, S.; Prusti, B.; Mishra, B. MOF-Derived Hollow Tubular In2O3/MIIIn2S4 (MII: Ca, Mn, and Zn) Heterostructures: Synergetic Charge-Transfer Mechanism and Excellent Photocatalytic Performance to Boost Activation of Small Atmospheric Molecules. ACS Appl. Energy Mater. 2022, 5, 11002–11017. [Google Scholar] [CrossRef]
  93. Yuan, S.; Qin, J.; Xu, H.; Su, J.; Rossi, D.; Chen, Y.; Zhang, L.; Lollar, C.; Wang, Q.; Jiang, H.; et al. [Ti8Zr2O12(COO)16] Cluster: An Ideal Inorganic Building Unit for Photoactive Metal-Organic Frameworks. ACS Cent. Sci. 2018, 4, 105–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Nguyen, H.; Vu, T.; Le, D.; Doan, T.; Nguyen, V.; Phan, N. A Titanium-Organic Framework: Engineering of the Band-Gap Energy for Photocatalytic Property Enhancement. ACS Catal. 2017, 7, 338–342. [Google Scholar] [CrossRef]
  95. Gao, K.; Li, H.; Meng, Q.; Wu, J.; Hou, H. Efficient and Selective Visible-Light-Driven Oxidative Coupling of Amines to Imines in Air over CdS@Zr-MOFs. ACS Appl. Mater. Interfaces 2021, 13, 2779–2787. [Google Scholar] [CrossRef]
  96. Chen, P.; He, X.; Pang, M.; Dong, X.; Zhao, S.; Zhang, W. Iodine Capture Using Zr-Based Metal-Organic Frameworks (ZrMOFs): Adsorption Performance and Mechanism. ACS Appl. Mater. Interfaces 2020, 12, 20429–20439. [Google Scholar] [CrossRef]
  97. Zhang, J.; Bai, T.; Huang, H.; Yu, M.; Fan, X.; Chang, Z.; Bu, X. Metal-Organic-Framework-Based Photocatalysts Optimized by Spatially Separated Cocatalysts for Overall Water Splitting. Adv. Mater. 2020, 32, 2004747. [Google Scholar] [CrossRef]
  98. Mao, L.; Lu, B.; Shi, J.; Zhang, Y.; Kang, X.; Chen, Y.; Jin, H.; Guo, L. Rapid high-temperature hydrothermal post treatment on graphitic carbon nitride for enhanced photocatalytic H2 evolution. Catal. Today 2022, in press. [CrossRef]
  99. Soares, C.; Maurin, G.; Leitão, A. Computational Exploration of the Catalytic Degradation of Sarin and Its Simulants by a Titanium Metal-Organic Framework. J. Phys. Chem. C 2019, 123, 19077–19086. [Google Scholar] [CrossRef]
Figure 1. The illustration of the synthetic strategy for UiO-66-NH2 with Ni2+ encapsulated via an efficient microwave-assisted method and the follow-up coordination environment modulation of single Ni atom to obtain Ni1-X/MOF, in contrast to NiNP/MOF and the NiSNP/MOF via one-pot synthesized. Reproduced with permission from Ref. [61].
Figure 1. The illustration of the synthetic strategy for UiO-66-NH2 with Ni2+ encapsulated via an efficient microwave-assisted method and the follow-up coordination environment modulation of single Ni atom to obtain Ni1-X/MOF, in contrast to NiNP/MOF and the NiSNP/MOF via one-pot synthesized. Reproduced with permission from Ref. [61].
Catalysts 12 01350 g001
Figure 2. The illustration of the dark photocatalysis based on the MIL-125 and MIL-125-X (X = NH2, NO2 and Br). Reproduced with permission from Ref. [62].
Figure 2. The illustration of the dark photocatalysis based on the MIL-125 and MIL-125-X (X = NH2, NO2 and Br). Reproduced with permission from Ref. [62].
Catalysts 12 01350 g002
Figure 3. (a) The 2D structure of Ni-TBAPy-SC and the coordinated environment of middle [Ni3O16] SBU. Color: red, O; yellow, C; green, Ni; cyan, H. (b) The schematic diagram of the photocatalytic hydrogen evolution reaction for Ni-TBAPy-NB. Reproduced with permission from Ref. [63].
Figure 3. (a) The 2D structure of Ni-TBAPy-SC and the coordinated environment of middle [Ni3O16] SBU. Color: red, O; yellow, C; green, Ni; cyan, H. (b) The schematic diagram of the photocatalytic hydrogen evolution reaction for Ni-TBAPy-NB. Reproduced with permission from Ref. [63].
Catalysts 12 01350 g003
Figure 4. (a) Schematic diagram of synthesis for different morphologies of In-TCPP MOFs. (b) The powder XRD patterns for 2D In-TCPP NS and 3D In-TCPP bulk. (c) SEM, (d) TEM, and (e) AFM images for 2D In-TCPP NS. (f) The SEM image of 3D In-TCPP bulk. (g) The powder XRD patterns after soaking and stirring in acidic and basic solutions for 12 h. Reproduced with permission from Ref. [64].
Figure 4. (a) Schematic diagram of synthesis for different morphologies of In-TCPP MOFs. (b) The powder XRD patterns for 2D In-TCPP NS and 3D In-TCPP bulk. (c) SEM, (d) TEM, and (e) AFM images for 2D In-TCPP NS. (f) The SEM image of 3D In-TCPP bulk. (g) The powder XRD patterns after soaking and stirring in acidic and basic solutions for 12 h. Reproduced with permission from Ref. [64].
Catalysts 12 01350 g004
Figure 5. The schematic diagram of the synthesis steps for TMPs and TMPs@UiO-66-NH2. Reproduced with permission from Ref. [69].
Figure 5. The schematic diagram of the synthesis steps for TMPs and TMPs@UiO-66-NH2. Reproduced with permission from Ref. [69].
Catalysts 12 01350 g005
Figure 6. The TEM images of (a,b) Ni2P@UiO-66-NH2 (the inset is the size distribution of Ni2P NPs) and (c,d) Pt@UiO-66-NH2 (the inset is the size distribution of Pt NPs). Reproduced with permission from Ref. [69].
Figure 6. The TEM images of (a,b) Ni2P@UiO-66-NH2 (the inset is the size distribution of Ni2P NPs) and (c,d) Pt@UiO-66-NH2 (the inset is the size distribution of Pt NPs). Reproduced with permission from Ref. [69].
Catalysts 12 01350 g006
Figure 7. The schematic diagram of synthesis steps for various photocatalysts. (a) PtPVP@UiO-66-NH2 and PtrPVP@UiO-66-NH2; (b) Pt@UiO-66-NH2 and Pt-Fc@UiO-66-NH2. Reproduced with permission from Ref. [70].
Figure 7. The schematic diagram of synthesis steps for various photocatalysts. (a) PtPVP@UiO-66-NH2 and PtrPVP@UiO-66-NH2; (b) Pt@UiO-66-NH2 and Pt-Fc@UiO-66-NH2. Reproduced with permission from Ref. [70].
Catalysts 12 01350 g007
Figure 8. (a) The strategy exploration to form well-defined 2D-on-3D structures and to achieve the “anti-epitaxial growth” pattern of MOF nuclei. (b) The SEM image of the core satellite structure without the assistance of PVP. (ch) display the SEM and TEM images when the system is in the presence of PVP and the relevant kinetic control results. (c,d) Vdeposition >> Vgrowth, (e,f) Vdeposition ≈ Vgrowth, (g,h) Vdeposition << Vgrowth. Reproduced with permission from Ref. [71].
Figure 8. (a) The strategy exploration to form well-defined 2D-on-3D structures and to achieve the “anti-epitaxial growth” pattern of MOF nuclei. (b) The SEM image of the core satellite structure without the assistance of PVP. (ch) display the SEM and TEM images when the system is in the presence of PVP and the relevant kinetic control results. (c,d) Vdeposition >> Vgrowth, (e,f) Vdeposition ≈ Vgrowth, (g,h) Vdeposition << Vgrowth. Reproduced with permission from Ref. [71].
Catalysts 12 01350 g008
Figure 9. Illustration of the preparation of CdS@MOF-808 composites and the comparison of hydrogen production rates between CdS4@MOF-808 and CdS/MOF-808-cys. Reproduced with permission from Ref. [72].
Figure 9. Illustration of the preparation of CdS@MOF-808 composites and the comparison of hydrogen production rates between CdS4@MOF-808 and CdS/MOF-808-cys. Reproduced with permission from Ref. [72].
Catalysts 12 01350 g009
Figure 10. The schematic diagram of the photocatalytic hydrogen evolution for MIL-125-NH2 (a) before and (b) after integration with the metal oxides. Reproduced with permission from Ref. [73].
Figure 10. The schematic diagram of the photocatalytic hydrogen evolution for MIL-125-NH2 (a) before and (b) after integration with the metal oxides. Reproduced with permission from Ref. [73].
Catalysts 12 01350 g010
Figure 11. (a) The preparation procedure of the NM@OM/TiO2 dual heterojunction. (b) The possible mechanism for photocatalytic hydrogen evolution on NM@OM/TiO2 under full-spectrum illumination. Reproduced with permission from Ref. [74].
Figure 11. (a) The preparation procedure of the NM@OM/TiO2 dual heterojunction. (b) The possible mechanism for photocatalytic hydrogen evolution on NM@OM/TiO2 under full-spectrum illumination. Reproduced with permission from Ref. [74].
Catalysts 12 01350 g011
Figure 12. The illustration of synthesis steps of Cu3(HHTP)2/Tp-Pa-1 hybrid photocatalysts. Reproduced with permission from Ref. [79].
Figure 12. The illustration of synthesis steps of Cu3(HHTP)2/Tp-Pa-1 hybrid photocatalysts. Reproduced with permission from Ref. [79].
Catalysts 12 01350 g012
Figure 13. The schematic diagram of the synthesis steps for MTV-Ti-MOF/COF. Reproduced with permission from Ref. [80].
Figure 13. The schematic diagram of the synthesis steps for MTV-Ti-MOF/COF. Reproduced with permission from Ref. [80].
Catalysts 12 01350 g013
Figure 14. (a) The schematic illustration of the synthesis steps of the NH2-MIL-(101)Fe@SNW-1. (b) The mechanism diagram of photodegradation pollutant and photocatalytic H2 evolution for MS-0.8 under the visible-light irradiation. Reproduced with permission from Ref. [81].
Figure 14. (a) The schematic illustration of the synthesis steps of the NH2-MIL-(101)Fe@SNW-1. (b) The mechanism diagram of photodegradation pollutant and photocatalytic H2 evolution for MS-0.8 under the visible-light irradiation. Reproduced with permission from Ref. [81].
Catalysts 12 01350 g014
Figure 15. (a) The schematic illustration of the synthesis steps of U@TDEn (NH2-UiO-66@TFPT-DETH) core-shell hetero-frameworks. (b) The SEM image of NH2-UiO-66. (c,d) The SEM and TEM images of U@TDE4. Reproduced with permission from Ref. [82].
Figure 15. (a) The schematic illustration of the synthesis steps of U@TDEn (NH2-UiO-66@TFPT-DETH) core-shell hetero-frameworks. (b) The SEM image of NH2-UiO-66. (c,d) The SEM and TEM images of U@TDE4. Reproduced with permission from Ref. [82].
Catalysts 12 01350 g015
Figure 16. (a) The synthesis schematic illustration of the Ni3FeN. (b) A possible mechanism of photocatalytic Eosin-Y-sensitized hydrogen evolution based on the Ni3FeN catalyst. Reproduced with permission from Ref. [85].
Figure 16. (a) The synthesis schematic illustration of the Ni3FeN. (b) A possible mechanism of photocatalytic Eosin-Y-sensitized hydrogen evolution based on the Ni3FeN catalyst. Reproduced with permission from Ref. [85].
Catalysts 12 01350 g016
Figure 17. Morphology-selective formation of Zn-based coordination polymer particles with Zn2+, H2bdc, and/or salen linkers. [86].
Figure 17. Morphology-selective formation of Zn-based coordination polymer particles with Zn2+, H2bdc, and/or salen linkers. [86].
Catalysts 12 01350 g017
Figure 18. (a,b) The FESEM images and (c) the powder XRD patterns of CPPs-derived hollow ZnO microtubes and microspheres. (d) The schematic illustration of the formation for the hollow O-doped ZnIn2S4/ZnO microtubes. (e) The illustration of the hydrogen evolution reaction on O-ZnIn2S4/ZnO-T heterosurfaces. Reproduced with permission from Ref. [86].
Figure 18. (a,b) The FESEM images and (c) the powder XRD patterns of CPPs-derived hollow ZnO microtubes and microspheres. (d) The schematic illustration of the formation for the hollow O-doped ZnIn2S4/ZnO microtubes. (e) The illustration of the hydrogen evolution reaction on O-ZnIn2S4/ZnO-T heterosurfaces. Reproduced with permission from Ref. [86].
Catalysts 12 01350 g018
Figure 19. (a) The schematic illustration of the synthesis of CdS-MoS2-In2O3. (b) A possible hydrogen evolution mechanism of the CdS-MoS2-In2O3 ternary nanocomposite photocatalyst. Reproduced with permission from Ref. [87].
Figure 19. (a) The schematic illustration of the synthesis of CdS-MoS2-In2O3. (b) A possible hydrogen evolution mechanism of the CdS-MoS2-In2O3 ternary nanocomposite photocatalyst. Reproduced with permission from Ref. [87].
Catalysts 12 01350 g019
Figure 20. (a) The schematic illustration of the synthesis process for the C/HT-In2O3/ZnInS4. (b) The comparison of interfacial electron transfer pathways for straddling type (In2O3/ZnInS4) and staggered model (C/HT-In2O3/ZnInS4). Reproduced with permission from Ref. [88].
Figure 20. (a) The schematic illustration of the synthesis process for the C/HT-In2O3/ZnInS4. (b) The comparison of interfacial electron transfer pathways for straddling type (In2O3/ZnInS4) and staggered model (C/HT-In2O3/ZnInS4). Reproduced with permission from Ref. [88].
Catalysts 12 01350 g020
Figure 21. The preparation diagram of the CeO1.66/CoP and CeO2/Co3O4 catalysts. Reproduced with permission from Ref. [89].
Figure 21. The preparation diagram of the CeO1.66/CoP and CeO2/Co3O4 catalysts. Reproduced with permission from Ref. [89].
Catalysts 12 01350 g021
Table 1. The comparison of the H2 evolution rates in the articles mentioned in this review.
Table 1. The comparison of the H2 evolution rates in the articles mentioned in this review.
MOFsCocatalysts/DerivativesLight SourceSacrificial AgentsH2 Evolution
Rate (μmol∙g−1∙h−1)
Ref.
UiO-66-NH2 (5 mg)Ni SACs300 W Xe lamp
(λ > 380 nm)
TEA1360[61]
MIL-125-NH2 (10 mg)Pt300 W Xe lamp
(800 nm > λ > 200 nm)
TEOA~140[62]
Ni-TBAPy-NB (20 mg)300 W Xe lamp
(λ > 420 nm)
Ascorbic acid5000[63]
In-TCPP NS (5 mg)Pt300 W Xe lampTEA539.07[64]
UiO-66-NH2 (5 mg)Ni2P300 W Xe lamp
(λ > 380 nm)
TEA409[69]
UiO-66-NH2 (5 mg)Pt-Fc300 W Xe lamp
(λ > 380 nm)
TEOA102[70]
UiO-66-NH2 (5 mg)Cu-TCPP MOF300 W Xe lampTEOA25[71]
MOF-808 (5 mg)CdS4 NPs300 W Xe lamp
(800 nm > λ > 400 nm)
TEA10,410[72]
MIL-125-NH2 (10 mg)MoO3300 W Xe lamp
(λ > 380 nm)
TEA399[73]
NH2-MIL-125@OH-MIL-125 (8 mg)TiO2300 W Xe lampTEOA7108[74]
Cu3(HHTP)2 (10 mg)Tp-Pa-1-COF300 W Xe lamp
(λ > 400 nm)
sodium ascorbate1760[79]
PdTCPP@PCN-415(NH2) (10 mg)TpPa-COF300 W Xe lamp
(λ > 400 nm)
Ascorbic acid13,980[80]
NH2-MIL-125(Fe) (50 mg)SNW-1@Pt300 W Xe lampTEOA1949.56[81]
NH2-UiO-66 (8 mg)TFPT-DETH-COF300 W Xe lamp
(λ > 420 nm)
Ascorbic acid7178[82]
Prussian blue analoguesNi3FeN (derivative)300 W Xe lamp
(λ > 400 nm)
TEOA16,960[85]
Zn-based CPPO-ZnIn2S4/ZnO (derivative)300 W Xe lampNa2S and Na2SO34512.5[86]
In-CPPCdS-In2O3(derivative)300 W Xe lampLA198,580[87]
MIL-68C/HT-In2O3 /Zn2InS4300 W Xe lampNa2S and Na2SO3230.1 μmol/m2/h[88]
Ce/Co-MOFCeO1.66/CoP(derivative)5 W LED
(λ≥400 nm)
TEOA8040.9[89]
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, H.; Li, C.; Li, Y.; Pang, J.; Bu, X. The Advanced Synthesis of MOFs-Based Materials in Photocatalytic HER in Recent Three Years. Catalysts 2022, 12, 1350. https://doi.org/10.3390/catal12111350

AMA Style

Zhang H, Li C, Li Y, Pang J, Bu X. The Advanced Synthesis of MOFs-Based Materials in Photocatalytic HER in Recent Three Years. Catalysts. 2022; 12(11):1350. https://doi.org/10.3390/catal12111350

Chicago/Turabian Style

Zhang, Hao, Cha Li, Yang Li, Jiandong Pang, and Xianhe Bu. 2022. "The Advanced Synthesis of MOFs-Based Materials in Photocatalytic HER in Recent Three Years" Catalysts 12, no. 11: 1350. https://doi.org/10.3390/catal12111350

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