REVIEW ARTICLE

Metal-organic frameworks for CO2 photoreduction

  • Lei ZHANG ,
  • Junqing ZHANG
Expand
  • Department of Mechanical Engineering, University of Alaska Fairbanks, Fairbanks, AK 99775, USA

Received date: 17 Dec 2018

Accepted date: 23 Feb 2019

Published date: 15 Jun 2019

Copyright

2019 Higher Education Press and Springer-Verlag GmbH Germany, part of Springer Nature

Abstract

Metal-organic frameworks (MOFs) have attracted much attention because of their large surface areas, tunable structures, and potential applications in many areas. In recent years, MOFs have shown much promise in CO2 photoreduction. This review summarized recent research progresses in MOF-based photocatalysts for photocatalytic reduction of CO2. Besides, it discussed strategies in rational design of MOF-based photocatalysts (functionalized pristine MOFs, MOF-photosensitizer, MOF-semiconductor, MOF-metal, and MOF-carbon materials composites) with enhanced performance on CO2 reduction. Moreover, it explored challenges and outlook on using MOF-based photocatalysts for CO2 reduction.

Cite this article

Lei ZHANG , Junqing ZHANG . Metal-organic frameworks for CO2 photoreduction[J]. Frontiers in Energy, 2019 , 13(2) : 221 -250 . DOI: 10.1007/s11708-019-0629-8

Introduction

Carbon dioxide (CO2) released in the combustion of fossil fuels, such as oil, coal, and natural gas, makes up the majority of greenhouse gas emissions, which is the major contributor to global warming [1,2]. Therefore, it is imperative to develop technologies to reduce CO2 emissions. To date, the main approaches developed to reduce CO2 emissions include ① CO2 capture and sequestration and ② CO2 conversion and utilization. CO2 capture and sequestration is a set of technologies developed for reducing CO2 emissions from fossil-fueled power plants. CO2 capture is to separate CO2 from gas mixtures via chemical absorption using an agent such as monoethanol amine or physical adsorption using solid adsorbents such as activated carbons and metal organic frameworks (MOFs), membrane separation and cryogenic separation at a low temperature [3]. CO2 sequestration refers to long-term storage of CO2 in ocean, soils, plants, and geologic formations [3,4]. However, these technologies are relatively energy-intensive and thus are not cost-effective. An alternative sustainable approach to mitigating CO2 emissions is CO2 utilization and conversion. CO2 utilization describes the uses of CO2 in both physical processes such as welding medium and chemical processes such as chemical synthesis. The CO2 utilization technology has found applications in several industries such as carbonated drinks, dry ice, solvent, food preservation, refrigerant, and fire extinguisher. These direct CO2 utilization applications, however, have a small effect on the overall CO2 emission reduction due to the limited usage [5]. CO2 utilization can also be employed indirectly in industries to promote a process such as in enhanced gas recovery, enhanced oil recovery, and enhanced geothermal systems. CO2 conversion refers to the transformation of CO2 into valuable products such as fuels and chemicals. CO2 can be utilized directly as a feedstock to react with other components to form chemical products such as urea and formic acid under heat and/or pressure. CO2 can also be utilized indirectly as a building block of a chemical product [6]. It is worth noting that CO2 is a highly stable molecule. The C-O bond strength in CO2 molecule is 364 kJ/mol and the carbon atom has the highest oxidation state, therefore, CO2 conversion into valuable chemicals generally requires a significant input of energy and the use of a catalyst [7]. The main approaches to converting CO2 into valuable products include photocatalysis [8,9], chemical fixation [10], hydrogenation [11], and electrocatalysis [12]. Among these methods, photocatalytic reduction of CO2 attracts much attention because it converts CO2 into useful products by utilizing solar energy via a clean and sustainable route. Under sunlight irradiation, photocatalysts can induce CO2 reduction and convert it into fuels and chemicals, mainly including CO, HCHO, HCOOH, CH3OH, C2H5OH, and CH4, etc., which is determined by the number of electrons and protons (e-/H+) transferred in the reactions. The selectivity of product and efficiency of CO2 reduction may have been affected by the thermodynamic reduction potentials and reaction conditions. H2O is commonly used as a solvent for CO2 photocatalytic reduction because it is of low cost and is natural abundance of hydrogen. The photoreduction reactions of CO2 in aqueous solution at pH= 7 and their reduction potentials with reference to the normal hydrogen electrode (NHE) at 25°C and 1 atm are given in Table 1 [13,14].
Tab.1 Photoreduction reactions of CO2 in aqueous solution at pH= 7 and their reduction potentials with reference to normal hydrogen electrode (NHE) at 25°C and 1 atm
Reaction Thermodynamics potential (V) vs. NHE
CO2 + 2H+ + 2e-->HCOOH -0.61
CO2 + 4H+ + 4e-->HCHO+ H2O -0.52
CO2 + 2H+ + 2e-->CO+ H2O -0.48
CO2 + 6H+ + 6e-->CH3OH+ H2O -0.38
CO2 + 12H+ + 12e-->C2H5OH+ 3H2O -0.33
CO2 + 8H+ + 8e-->CH4 + 2H2O -0.24
H2O ->1/2O2 + 2H+ + 2e- +0.82
2H+ + 2e-->H2 -0.41
Photocatalysis is a process that converts solar energy to chemical energy, where solar energy is the driving force for the excitation and transfer of holes and electrons to induce oxidation and reduction reactions. The photocatalytic CO2 reduction process can be explained as follows: first, under light illumination with energy greater than the band gap of the photocatalysts, electrons are excited from the valence band (VB) to the conduction band (CB), generating an equal number of holes in the VB. Secondly, electron-hole pairs separate from each other and move to the surface of photocatalyst. Finally, the electrons reduce CO2 into chemical products, while the holes oxidize a sacrificial agent or H2O. To induce CO2 photoreduction, it requires that reduction potential and oxidation potential of the reaction is less negative and less positive than the CB edge and the VB edge of the photocatalyst, respectively.
Since Fujishima et al. reported on the feasibility of using TiO2 for photoelectron chemical water splitting under ultraviolet (UV) irradiation [15], many different photocatalysts have been developed, most of which are inorganic semiconductors such as TiO2 [16], CdS [17], ZnO [18], ZnS [19], Fe2O3 [20], g-C3N4 [21], Ag3PO4 [22] and their composites. Some cocatalysts such as Pt have also been explored [23,24]. However, the wide band gap, high recombination rate of electron‐hole pairs, low adsorption capacities for CO2, and less tunable structure of the conventional semiconductors limit their practical applications. For example, TiO2 is mainly used for UV light photocatalysis because of its wide band gap (3–3.2 eV) [25]. ZnO and CdS are not stable under irradiation in an aqueous solution, making them inactive over time [26,27]. Therefore, it is imperative to develop new photocatalytic materials with finely tunable energy band structures and high stability.
Metal-organic frameworks (MOFs), a new class of inorganic-organic hybrid material composed of inorganic metal clusters and organic linkers which are connected through coordination bonds, have attracted much interest due to their large surface areas, tunable structures and high porosity [2834]. These excellent properties enable their wide applications in many fields, such as gas storage and separation [35], catalysis [36], drug delivery [37], water treatments [38], and sensing [39]. Recently, MOFs have found applications in CO2 photoreduction, degradation of organics, and chemical synthesis as photocatalytic materials [4042]. In photocatalytic reaction, MOFs undergo a similar process to traditional semiconductors, but differently, the VB and CB are described as the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) in MOFs, respectively [36,43]. In general, the HOMO and LUMO energy levels are associated with the redox potential energy levels of the organic linker and the metal-oxo cluster, respectively [42]. The tunable organic linkers and/or metal clusters in MOFs can act as antennas to harvest light to generate electron-hole pairs for photocatalysis. Most of the photocatalytic reactions in MOFs reported to date are ascribed to a localized metal-to-ligand charge transfer (MLCT), a ligand-to-metalcharge transfer (LMCT), or a pp* transition of the aromatic ligand [36]. MOFs are utilized for photocatalytic reduction of CO2 because of their attributes as follows: ① both the organic linkers and metal clusters can serve as the light harvesting sites and they can be tailored to tune the optical absorption range of MOFs [44], ② some MOFs have a catalytic activity resulting from the catalytically active organic linkers and/or unsaturated metal sites [45,46], ③ MOFs have a large surface area and a high CO2 adsorption capacity, and the high CO2 concentration in the pores can facilitate the photocatalytic reactions, and ④ the three-dimensional porous structures and high surface areas enable MOFs to incorporate foreign photoactive species into their frame works, through which photocatalytic reactions can be enhanced by the synergistic cooperation of the metal clusters, organic linkers and the incorporated active sites [47]. Based on these unique properties, MOFs are promising photocatalysts for CO2 reduction.
MOF photocatalysts can be classified into two categories: ① Pristine MOF photocatalysts, which areal so called “opportunistic” photocatalysts. The photocatalytic properties of some pristine MOFs are a consequence of the catalytically active organic linkers and unsaturated metal sites [48], and some MOFs are semiconductors [4951] whose photocatalytic reactions are completed through LMCT. Most pristine MOFs as photocatalysts, however, have large band gaps and barely absorb visible light, which limits their practical applications. These MOFs generally have a low photon energy utilization efficiency, and are typically employed for photocatalytic degradation of organics which have high thermodynamic driving forces and low kinetic barriers [36]. ② Modified MOF photocatalysts, which refer to MOFs functionalized to increase light harvesting and decrease the recombination rate of the photo-generated charge carriers for photocatalytic activity enhancement. To date, many strategies have been developed for enhancing the photocatalytic properties of MOFs under solar illumination, including the decoration of the metal clusters and organic linkers [52,53], incorporation of foreign photocatalytic species such as metal particles, semiconductors [5456], and photosensitizers [5759].
So far, a number of good literature reviews have summarized the photocatalytic applications of MOFs [6069], most of which have discussed the role of MOFs and their performances in water splitting [7073], degradation of organic pollutants [7477], hydrogen evolution [36,7880], and CO2 conversion [36,43,66,78,8085]. Therefore, this review is mainly focused on reviewing the recent progress of MOF-based photocatalysts in CO2 photoreduction. Besides, it discusses modification strategies of MOFs and their photocatalytic activities in CO2 reduction. Moreover, it explores the challenges and future perspectives of MOFs-based photocatalysts in CO2 photoreduction.

Modification of pristine MOFs as photocatalysts

Compared to traditional inorganic semiconductors, it is more convenient to tune the optical properties and consequently the photocatalytic properties of MOFs by modification of the metal clusters or organic linkers. It is reported that the linker decoration can change the energy band gap of MOFs by shifting the photo absorption edge from the UV to visible light region [86]. Many strategies have been developed to functionalize organic linkers and metal clusters [44,8792], such as amine and photosensitizer functionalization of organic linkers and incorporation of foreign metal cations into organic linkers and metal sites (e.g., Ti–O, Fe–O, and Zr–O clusters). The performances of recent photocatalytic MOFs for CO2 photoreduction are summarized in Table 2.
Tab.2 Performances of recent photocatalytic MOFs for CO2 photoreduction
Functionalization MOF Irradiation Solvent/sacrificial agent Main product Photocatalytic reactivity Reaction time/h Ref.
MIL-125(Ti) UV MeCN/TEOA HCOO- 2.41 mmol 10 [53]
MIL-125(Ti) Visible 0
NH2-functionalized linker MIL-125(Ti)-NH2 8.14 mmol
NH2-functionalized linker NH2‐UiO‐66(Zr) Visible MeCN/TEOA HCOO- 13.2 mmol 10 [93]
(NH2)2‐UiO‐66(Zr) 20.7 mmol
NH2-modified with partial Ti cation substitution NH2-UiO-66(Zr/Ti) Visible MeCN/TEOA BNAH HCOO- 22.23 mmol 6 [94]
(NH2)2‐UiO‐66(Zr/Ti) 31.57±1.64 mmol
MIL-101(Fe) Visible MeCN/TEOA HCOO- 59.0 mmol 8 [95]
NH2-functionalized linker MIL-101(Fe)-NH2 178 mmol
MIL-53(Fe) 29.7 mmol
NH2-functionalized linker MIL-53(Fe)-NH2 46.5 mmol
MIL-88B(Fe) 9.0 mmol
NH2-modified with partial metal ion substitution MIL-88B(Fe)-NH2 30 mmol
NH2-modified with partial Ti cation substitution NH2-UiO-66(Zr/Ti) Visible MeCN/TEOA HCOO- 3.4 mmol/mol 10 [96]
NH2-UiO-66(Zr/Ti)-100-4 4.2 mmol/mol
NH2-UiO-66(Zr/Ti)-120-16 5.8 mmol/mol
Porphyrin- functionalized linker Rh-PMOF-1(Zr) Visible MeCN/TEOA HCOO- 6.1 mmol/mmol 18 [97]
Porphyrin- functionalized linker Zn/PMOF UV-visible H2O vapor CH4 10.43 mmol 4 [98]
Porphyrin- functionalized linker Al/PMOF Visible H2O/TEOA CH3OH 37.5 ppm /(g·h) [99]
Porphyrin- functionalized linker with partial Cu cation substitution Cu-Al/PMOF 262.6 ppm /(g·h)
Porphyrin- functionalized linker MOF-525 Visible MeCN/TEOA CO 64.02 mmol /(g·h) 6 [100]
CH4 6.2 mmol /(g·h)
Porphyrin- functionalized linker with
embedded
Zn cations
MOF-525-Zn CO 111.7 mmol /(g·h)
CH4 11.64 mmol /(g·h)
Porphyrin- functionalized linker with
embedded
Co cations
MOF-525-Co CO 200.6 mmol /(g·h)
CH4 36.76 mmol /(g·h)
Porphyrin- functionalized linker PCN-222 Visible MeCN/TEOA HCOO- 30 mmol 10 [101]
Photosensitizer functionalization Eu-Ru(phen)3-MOF Visible MeCN/TEA HCOO- 47 mmol 10 [102]
Photosensitizer functionalization UiO-67-ReI(CO)3(5,5′-dcbpy)Cl Visible MeCN/TEA CO TON= 5.0 6 [103]
H2 TON= 0.5
CO TON= 10.9 20
H2 TON= 2.5
ReI(CO)3(5,5′-dcbpy)Cl CO TON= 5.6 6
H2 TON= 0.3
CO TON= 7.0 20
H2 TON= 1.0
Photosensitizer functionalization Zr6(O)4(OH)4[Re(CO)3Cl(bpydb)]6 Visible MeCN/TEA CO TON= 6.44 6 [104]
H2 TON= 0.40 6
Photosensitizer functionalization UiO-67-ReI(CO)3(5,5′-dcbpy)Cl Visible TEA CO 0.5 mmol/(g·h) 6 [52]
UiO-67-ReI(CO)3(5,5′-dcbpy)Cl-NH2 (33% (mol)) CO 1.5 mmol/(g·h) 6
Photosensitizer functionalization UiO-67-Cp*Rh(5,5′- dcbpy)Cl2 (10%) Visible ACN/TEOA HCOO- TON= 47 10 [105]
H2 TON= 36
Cp*Rh(5,5′- dcbpy)Cl2 HCOO- TON= 42
H2 TON= 38
[Ru(bpy)3]Cl2 HCOO- TON= 125
H2 TON= 55
Photosensitizer functionalization MOF-253-Ru(CO)2Cl2 Visible MeCN/TEOA HCOO 0.67 mmol 8 [106]
CO 1.86 mmol
H2 0.09 mmol
Ru(bpy)2Cl2- sensitized MOF-253-Ru(CO)2Cl2 HCOO 4.84 mmol
CO 1.85 mmol
H2 0.72 mmol
MOF-253-Ru(bpy)
2Cl2
HCOO 0.46 mmol
CO 0.21 mmol
H2 0.07 mmol
Ru(bpy)2Cl2 HCOO 0.27 mmol
CO 0.18 mmol
H2 0 mmol
Photosensitizer functionalization Y[Ir(ppy)2(4,4′‐dcbpy)]2[OH] Visible MeCN/TEOA HCOO 118.8 mmol/(g·h) 6 [107]
Photosensitizer functionalization [Cd2[Ru(4,4’-dcbpy)3]·12H2O]n nanoflower Visible MeCN/TEOA HCOO 77.2 mmol/(g·h) 8 [108]
[Cd2[Ru(4,4’-dcbpy)3]·12H2O]n microflake 52.7 mmol/(g·h)
[Cd2[Ru(4,4’-dcbpy)3]·12H2O]n bulk crystals 30.6 mmol/(g·h)
Photosensitizer functionalization [Cd3[Ru(5,5′-dcbpy)3]2·2(Me2NH2)]n Visible MeCN/TEOA HCOO 67.5 mmol/(g·h) 6 [109]
[Cd[Ru(bpy)(4,4′-dcbpy)2]·3H2O]n 71.7 mmol/(g·h)
Catechol- functionalized linker UiO-66-CrIIIcatbdc Visible MeCN/TEOA/BNAH HCOO TON= 11.22±0.37 6 [110]
UiO-66-GaIIIcatbdc TON= 6.14±0.22
Anthracene- functionalized linker NNU-28 Visible MeCN/TEOA HCOO 183.3 mmol/(h·mmol) 10 [111]

Notes: ACN–acetonitrile; BNAH–1-benzyl-1,4-dihydronicotinamide; bpy–2,2’-bipyridine; catbdc–2,3-dihydroxyterephthalic acid; Cp*–pentamethylcyclopentadiene; MeCN–acetonitrile; phen–phenanthroline; TEA–trimethylamine; TEOA–triethanolamine; TON–total turnover number; 5,5′-dcbpy–2,2’-bipyridine-5,5′-dicarboxylic acid; bpydb–4,4’-(2,2’-bipyridine-5,5′-diyl)dibenzoate; ppy–2-phenylpyridine; 4,4’-dcbpy–2,2’-bipyridine-4,4’-dicarboxylate

Metal cluster nodes

MOFs have three-dimensional structures constructed from the interconnection of metal cluster nodes with organic linkers. Besides photoactive ligands, metal cluster nodes in MOFs can also drive photocatalytic CO2 reduction. To date, Ti-, Zr-, and Fe-based MOFs are among the most investigated photocatalytic MOFs whose photoactive metal nodes in metal-oxo clusters can initiate photocatalytic reaction by trapping photo-excited electrons and decreasing recombination rate of electron-hole pairs [83]. These MOFs share the similarity that they all contain metal ions with variable valence states (e.g., Ti4+/Ti3+ , Zr4+/Zr3+ , and Fe3+/Fe2+ ), which enables their effectiveness on photocatalytic reduction.
Of the Ti-based MOFs, MIL-125(Ti) (Ti8O8(OH)4(O2C-C6H4-CO2)6) is the most studied MOF, which is constructed from the Ti8O8(OH)4 secondary building units (SBUs) and 1,4-benzenediacarboxylate (BDC) ligands. Fu et al. investigated the CO2 photoreduction over MIL-125(Ti) [53]. 2.41 mmol HCOO- was detected in the acetonitrile (MeCN) solvent with TEOA under 365 nm UV light irradiation for 10 h. A number of strategies have been developed to functionalize MIL-125(Ti) to enhance the photocatalytic reduction of CO2, such as organic linker decoration and incorporation of foreign photocatalytic components into MOFs to form MOF composites. These strategies will be discussed in Section 2.2 and Section 3.
Zr-based MOFs, which have robust structures and excellent thermal and chemical stabilities [112,113], are another popular class of photocatalytic MOFs since their synthesis in 2008 [114]. The representative examples are UiO-66(Zr) and UiO-67(Zr), which are constructed by integrating Zr6O4(OH)4 SBUs with BDC ligands and 4,4′-biphenyldicarboxylic acid (BPDC) ligands, respectively [114,115]. Compared to Ti-based MOFs, Zr-based MOFs have more negative redox potential (Ti4+/Ti3+ (–0.1 V), Zr4+/Zr3+ (–1.06 V)) [116,117]. However, UiO-66 exhibits no absorption under visible light irradiation owing to the higher redox potential energy level of the Zr6O4(OH)4 SBUs in UiO-66 than the LUMO of the BDC linkers, leading to the low efficiency in LMCT and consequently low efficacy in CO2 photoreduction [94]. Partial substitution of metal cations in MOFs can introduce metal-to-metal charge transfer, which can promote photocatalytic performance especially under visible light irradiation [94,96]. A bimetallic UiO-based MOF, NH2-UiO-66(Zr/Ti), was prepared by Cohen and coworkers by partially substituting Zr in NH2-UiO-66(Zr) with Ti [94]. NH2-UiO-66(Zr/Ti) had a better performance in photocatalytic CO2 reduction under visible light irradiation compared to NH2-UiO-66(Zr). The reason for this is that Ti ions incorporated enable the SBUs to accept the electrons generated via light absorption by the organic linkers. No HCOO- was produced over the parent UiO-66(Zr)-NH2, demonstrating that Ti was critical for photocatalysis.
Sun et al. also prepared Ti-substituted NH2-UiO-66(Zr/Ti) MOFs (NH2-UiO-66(Zr/Ti)-120-16 and NH2-UiO-66(Zr/Ti)-100-4) doped with different amounts of Ti by a post-synthetic exchange method and examined their photocatalytic performance on CO2 reduction under visible light irradiation (Fig. 1(a)) [96]. NH2-UiO-66(Zr/Ti)-120-16 had an enhanced photocatalytic activity toward CO2 conversion with a yield of 5.8 mmol/mol of HCOO- after 10 h visible light irradiation in MeCN solvent with TEOA as a sacrificial agent, which was 1.7 times of that observed over NH2-UiO-66(Zr) (3.4 mmol/mol) under similar conditions. In contrast, NH2-UiO-66(Zr/Ti)-100-4 produced 4.2 mmol mol-1 of HCOO-, which was less than NH2-UiO-66(Zr/Ti)-120-16, but still higher than that over the pristine NH2-UiO-66(Zr). The enhancement in photocatalytic performance over Ti-substituted NH2-UiO-66(Zr/Ti) MOFs is associated with the increase in CO2 adsorption capacity and photocatalytic sites, both of which result from Ti doped into Zr-O clusters of NH2-UiO-66(Zr). Based on the experimental observations and theoretical studies, the mechanism for enhanced photocatalytic reactions over NH2-UiO-66(Zr/Ti) is proposed (see Fig. 1(b)). When Zr4+ centers in Zr6O4(OH)4 are partially substituted by Ti4+ to form (Ti/Zr)6O4(OH)4, the excited NH2-BDC upon visible light irradiation can transfer electrons to either Zr4+ or Ti4+ centers. The theoretical calculations show that there is a higher probability for electrons to be transferred to Ti4+ than that to Zr4+ centers, leading to the formation of (Ti3+/Zr4+)6O4(OH)4 SBUs. The Ti3+ in the excited (Ti3+/Zr4+ )6O4(OH)4 SBUs can play a role of electron donor to donate electrons to Zr4+ , leading to Ti4+–O–Zr3+ formation. As a result, the substituted Ti center in NH2-UiO-66(Zr/Ti) facilitates the interfacial charge transfer from the excited NH2-BDC to Zr–O clusters, which boosts the enhanced photocatalytic reactions over NH2-UiO-66(Zr/Ti).
Fig.1 Enhanced photoreduction of CO2 over NH2-UiO-66(Zr) induced by Ti substitution and the proposed CO2 photoreduction mechanism

Full size|PPT slide

Recently, Fe-based MOFs as photocatalysts for CO2 reduction have attracted much interest owing to the fact that the Fe–O clusters can be directly photoexcited to induce electron transfer from O2- to Fe3+ to form Fe2+ under visible light irradiation, which drives the photocatalytic reaction [95,118]. Since the Fe sites play a role of photocatalytic centers, Fe-based MOFs are capable of photocatalytically reducing CO2 under visible light irradiation in the absence of LMCT. Wang et al. reported a series of Fe-based MOFs, MIL-101(Fe), MIL-53(Fe), MIL-88B(Fe), all of which had a photocatalytic activity for CO2 reduction to produce HCOO- under visible light irradiation [95]. It showed that the photocatalytic activity of MIL-101(Fe), MIL-53(Fe) and MIL-88B(Fe) toward CO2 conversion into HCOO- in MeCN solvent with TEOA as a sacrificial reactant was 59.0, 29.7, and 9.0 mmol, respectively, after visible light irradiation for 8 h. Of the three investigated Fe-based MOFs, MIL-101(Fe) had the best photocatalytic performance attributing to the existence of the unsaturated Fe sites in its structure that were absent in MIL-53(Fe) and MIL-88B(Fe). MIL-53(Fe) had a better activity than MIL-88B(Fe) attributing to its higher CO2 adsorption capacity (13.5 g/cm3) than MIL-88B(Fe) (10.4 g/cm3) which might be ascribed to its one dimensional framework structure which possessed a better electro-conductivity. The photocatalytic activities of the three Fe-based MOFs were enhanced by amine functionalization. The HCOO- yield of 178, 46.5, and 30 mmol was produced over NH2-MIL-101(Fe), NH2-MIL-53(Fe), and NH2-MIL-88B(Fe), respectively, under the same photocatalytic conditions, which were higher in comparison to the parent Fe-MOFs. This photocatalytic activity enhancement is caused by the existence of the dual excitation pathways: excitation of NH2-functionalized organic ligands to transfer electrons to the Fe center in addition to the direct excitation of Fe-O clusters (see Fig. 2).
Fig.2 Proposed CO2 photoreduction through the dual excitation pathways over amino-functionalized Fe-based MOFs (Reproduced with permission from Ref [95]. Copyright 2014, American Chemical Society)

Full size|PPT slide

Modification of organic linkers

Functionalization of organic linkers has been considered as an effective approach to increase the absorption of light and decrease the recombination rate of the photo-generated charge carriers and consequently to improve the photocatalytic reduction of CO2. To date, several different organic linker functionalization strategies have been developed for enhancing the photocatalytic performance of MOFs toward CO2 reduction under light illumination, including amine functionalization, utilization of porphyrin-based organic linkers, and photosensitizer functionalization.

Amine functionalization

Studies have shown that amine groups incorporated into the organic linkers in MOFs contribute to the CO2 adsorption capacity enhancement, adsorption region broadening, and CO2 photoreduction performance boost. The introduction of amine groups into aromatic polycarboxylates in MOFs can increase the interactions between CO2 molecules and the modified linkers [119]. The CO2 photoreduction improvement is attributed to the lone pair of electrons in amine groups, which can interact with the p*-orbitals of the benzene ring, leading to the donation of electrons to the anti-bonding orbitals. This interaction enables the formation of a new higher energy HOMO level, which leads to a broader optical absorption region and consequently enhanced CO2 photoreduction [69].
In 2012, Li and coworkers [53], for the first time, prepared an amino-functionalized MOF, NH2-MIL-125(Ti), and examined CO2 adsorption and photoreduction under visible light irradiation. NH2-MIL-125(Ti) had a higher CO2 adsorption capacity (132.2 cm3/g) in comparison to that of MIL-125(Ti) (98.6 cm3/g) due to the increased interaction between CO2 molecules and the amine-modified linkers. Moreover, amine functionalization led to a broader adsorption range of 550 nm for NH2-MIL-125(Ti) compared to MIL-125(Ti) with an adsorption range of 350 nm (see Fig. 3(a)). Photoreduction of CO2 over NH2-MIL-125(Ti) was performed in MeCN solvent with TEOA as an electron donor. After visible light irradiation for 10 h, 8.14 mmol of HCOOwas obtained, which was much higher than that of MIL-125(Ti) with no yield. The photocatalytic mechanism is proposed as follows (see Fig. 3(b)): under visible light irradiation, electrons are transferred from organic linkers to Ti4+ cations, and Ti4+ cations are reduced to Ti3+ by TEOA acting as an electron donor, leading to the formation of a long-lived excited charge separation. Ti3+ cations subsequently reduce CO2 to HCOO-. In 2013, the same research group (Li and coworkers) [93] developed another amino-functionalized MOF, NH2-UiO-66(Zr), and found that it had a higher activity for CO2 photoreduction than previously reported NH2-MIL-125(Ti) in the presence of TEOA as a sacrificial agent. Similar to previous reports, substitution of the organic linker in UiO-66(Zr) by NH2-BDC led to a broader optical absorption range of NH2-UiO-66(Zr) due to the increased interaction between the NH2-BDC linker and the Zr-O clusters. CO2 uptake of NH2-UiO-66(Zr) (68 cm3/g) was also improved in comparison to the parent UiO-66(Zr) (53 cm3/g) owing to the enhanced interactions of the NH2 functional groups with the CO2 molecules [119,120]. After visible light irradiation over NH2-UiO-66(Zr) for 10 h, 13.2 mmol of HCOO was produced. Partial substitution of the organic linker NH2-BDC by (NH2)2-BDC in NH2-UiO-66(Zr) further improved its CO2 photoreduction activity, which gave 20.7 mmol of HCOO- under the same conditions as over NH2-UiO-66(Zr) (Fig. 4). The improvement in CO2 photocatalytic reduction over the mixed NH2-UiO-66(Zr) is attributed to enhanced light absorption in the visible region and increased CO2 adsorption (71 cm3/g). Similar results were reported by Cohen and coworkers [94] who investigated the CO2 photocatalytic reduction performance over (NH2)2-UiO-66(Zr) and (NH2)2-UiO-66(Zr/Ti) prepared by the introduction of diamine-substituted ligands into (NH2)-UiO-66(Zr) and (NH2)-UiO-66(Zr/Ti), respectively. It was found that an increase of the amino group number introduced new energy levels for additional light absorption and charge transfer, leading to enhancement in photocatalytic activities.

Utilization of porphyrin-based organic linkers

Porphyrins have complex cyclic structures consisting of four pyrrole rings linked to each other by methine groups. Due to their strong interactions with CO2, high light adsorption efficiency and catalytic performance, porphyrin-based organic linkers have been incorporated into MOFs for CO2 photoreduction.
Su and coworkers [97] prepared a rhodium(III)-porphyrin zirconium MOF (Rh-PMOF-1(Zr)) using a Rh-based metalloporphyrin tetracarboxylic ligand Rh(TCPP)Cl (TCPP= tetrakis(4-carboxyphenyl) porphyrin) and ZrCl4. The CO2 adsorption and photoreduction over Rh-PMOF-1(Zr) were examined. Rh-PMOF-1(Zr) had a high CO2 adsorption capacity of 53 cm3/gat 298 K. After visible-light irradiation for 18 h, the yield of HCOO reached 6.1 mmol/mmolcat. This showed that Rh-PMOF-1(Zr) had a long-lived excited-state under vacuum at 298 K with the lifetime of 207 ms, which contributed to the improvement in the photocatalytic activity of Rh-PMOF-1(Zr). Two catalytic reactions contribute to the CO2 photocatalytic reduction to the formation of HCOO over Rh-PMOF-1(Zr): ① metalloporphyrin ligand plays a role of an antenna to harvest light, generate electrons, and transfer electrons to the zirconium oxo clusters, reducing CO2 to HCOO, and ② the rhodium-porphyrin ligands serve as photocatalytic centers toward CO2 photoreduction.
Sharifnia and coworkers [98] used TCPP as ligand and Zn(NO3)2·6H2O to prepare a porphyrin-based MOF (Zn/PMOF) and performed photocatalytic reduction of CO2 over Zn/PMOF in the presence of H2O vapor under UV-visible light. After 4 h irradiation, Zn/PMOF had a CO2 photoreduction activity with CH4 formation of 10.43mmol. Only a small amount of CH4 (10.77 mmol) was produced by extending the irradiation time to 24 h, which may have been caused by the fact that the intermediate product and/or byproduct formed after 4 h have saturated the active sites.
Huang et al. [99] synthesized two types of porphyrin-based MOFs, Al/PMOF and Cu-Al/PMOF, and compared their CO2 adsorption capacities and photoreduction of CO2 toward methanol. Al/PMOF was synthesized using TCPP as organic linker and AlCl3·6H2O while Cu-Al/PMOF was produced by doping Cu2+ into Al/PMOF. It showed that Cu2+ in Cu-Al/PMOF contributed to the enhanced photocatalytic reduction of CO2. Cu-Al/PMOF had a higher CO2 adsorption capacity (277.4 mg/g) than that of Al/PMOF (153.1 mg/g). Moreover, the methanol formation rate over Cu-Al/PMOF (262.6 ppm/(g·h)) was 7 times as high as that of Al/PMOF (37.5 ppm/(g·h)). As demonstrated by in situ Fourier transform infrared (FT-IR) spectra, CO2 could be chemically adsorbed on the Cu site in Cu-Al/PMOF, where the linear CO2 molecules would bend, thus lowering the reaction barrier and improving the photocatalytic efficiency.
Ye and coworkers [100] prepared MOF-525 (Zr6O4(OH)4(TCPP-H2)3) by integrating Zr6 clusters with porphyrin-based organic linkers. MOF-525-Co and MOF–525-Zn were also developed by the introduction of coordinatively unsaturated Co sites and Zn sites into the porphyrin units of MOF-525, respectively. As shown in Fig. 5, both MOF-525-Co (33.6 cm3/g) and MOF-525-Zn (28.1 cm3/g) had a higher CO2 adsorption capacity than that of pristine MOF-525 (25.3 cm3/g) due to the enhanced interaction between CO2 molecules and the introduced open Co and Zn metal sites. CO2 photoreduction over MOF-525, MOF-525-Co and MOF-525-Zn was performed in the presence of MeCN solvent and TEOA as an electron donor under visible light irradiation for 6 h, and two products, CO and CH4 were produced. MOF-525-Co had the highest CO evolution rate of 200.6 mmol/(g·h) (yield: 2.42 mmol), and a CH4 evolution rate of 36.76 mmol/(g·h) (yield: 0.42 mmol), followed by MOF-525-Zn (CO, 111.7 mmol/(g·h); CH4, 11.635 mmol/(g·h)) and MOF-525 (CO, 64.02  mmol/(g·h); CH4, 6.2 mmol/(g·h)). The photocatalytic activity enhancement for MOF-525-Co and MOF–525-Zn is partially ascribed to the difference in their charge separation efficiencies. MOF-525-Co and MOF-525-Zn had an enhanced efficiency in charge separation because of the lower energy at Co or Zn site, to which the electrons can be transferred from porphyrin units with a higher easiness. Furthermore, MOF-525-Co had a higher electron transfer efficiency (62.1 %) than that of MOF-525-Zn (24.9 %). These partially account for the better catalytic activity of MOF-525-Co than that of MOF-525-Zn. In addition, MOF-525-Co had a good stability and a good reproducible photocatalytic activity after three cycles (Fig. 5).
Fig.3 Enhanced visible light absorption in NH2-MIL-125(Ti) induced by amino functionality and the proposed CO2photoreduction mechanism

Full size|PPT slide

Fig.4 Amount of HCOO- produced over NH2-UiO-66(Zr) and mixed NH2-UiO-66(Zr) as a function of light irradiation time (Reproduced with permission from Ref. [93]. Copyright 2013, Wiley)

Full size|PPT slide

Fig.5 Effect of metallization on the photocatalytic behavior of MOF-525

Full size|PPT slide

Xu et al. [101] prepared a zirconium-porphyrin MOF, PCN-222 (also named as MOF-545 or MMPF-6), by employing zirconium (IV) chloride and TCPP as ligand. PCN-222 had a CO2 uptake of 35 cm3/gat 298 K and 1 atm. Photocatalytic reduction of CO2 over PCN-222 was performed in MeCN as solvent and TEOA as a sacrificial agent. After visible light irradiation for 10 h, HCOO- was produced with a yield of 30 mmol, which was much higher than that observed over the TCPP ligand alone (2.4 mmol), as shown in Fig. 6. It is proposed that upon irradiation, the TCPP in PCN-222 acts as an antenna to harvest visible light, generate electron–hole pairs, and transfer electrons to the Zr-oxo clusters toward CO2 reduction to HCOO- in the presence of TEOA as the electron donor. Two factors contribute to the enhancement in photocatalytic performance of PCN-222: ① the high CO2 adsorption capacity of PCN-222 might enable higher interaction with CO2 in MeCN, thereby promoting the photocatalytic reaction, and ② the ultrafast transient absorption and photoluminescence spectroscopy reveals that the emergence of an extremely long-lived electron trap state in PCN-222 significantly suppresses the electron-hole recombination, thus enhancing the CO2 photoreduction efficiency.
Fig.6 Left: Amount of HCOO- produced as a function of visible light irradiation time over PCN-222 (a), H2TCPP (b), no PCN-222 (c), no TEOA (d), and no CO2 (e). Right: 13C Nuclear magnetic resonance (NMR) spectra for the product obtained from reaction with 13CO2 (a) or 12CO2 (b) (Reproduced with permission from Ref. [101]. Copyright 2015, American Chemical Society)

Full size|PPT slide

Photosensitizer functionalization

Photosensitizers are able to harvest light to generate electron-hole pairs and act as catalyst sites for CO2 photoreduction. Photoactive metal complexes, such as Ru, Re, and Ir-based polypyridine units, have been widely used to functionalize MOFs to enhance their CO2 photocatalytic reaction activities by introducing catalytic active centers and photosensitive sites for visible light harvesting [121123].
Yan et al. [102] produced an Eu-Ru(phen)3-MOF (phen= phenanthroline) by integrating the triangular Ru(phen)3-derived tricarboxylate ligand as photosensitizer into Eu-MOF with Eu(III)2(m2-H2O) SBUs. Transient absorption results and theoretical calculations showed that photo-excitation of the Ru metalloligands in Eu-Ru(phen)3-MOF initiated electron transfer into the nodes to generate dinuclear [Eu(II)]2 active sites, which selectively converted CO2 to HCOO- with a yield of 47 mmol in 10 h in the presence of MeCN and TEOA under visible light irradiation. This was higher than that observed over the Ru(phen)3-derived tricarboxylate acid metalloligandalone under the same condition.
Lin and coworkers [103] incorporated ReI(CO)3(5,5′-dcbpy)Cl (at 4 wt% doping level, 5,5′‐dcbpy= 2,2’-bipyridine-5,5′-dicarboxylic acid) into the UiO-67 framework built from Zr6(µ3-O)43-OH)4 SBUs and BPDC ligands, and tested the photocatalytic reduction of CO2 in MeCN solvent and trimethylamine (TEA) as a sacrificial agent under visible light. The total turnover number (TON) of CO and H2 over the as-doped UiO-67 reached 5.0 and 0.5, respectively, under visible light irradiation of 6 h. After 20 h, CO-TON and H2-TON reached 10.9 and 2.5, respectively, which were higher than that observed for the bare ReI(CO)3(5,5′-dcbpy)Cl (CO-TON= 7.0, H2-TON= 1.0) under the same condition because of the decomposition of ReI(CO)3(5,5′-dcbpy)Cl under irradiation. This enhanced TON is proposed to be associated with the stabilization effect of the active-site isolation of the immobilized Re-based catalysts in the framework. Lately, the same group produced another photosensitizer-functionalized MOF, Zr6(O)4(OH)4[Re(CO)3Cl(bpydb)]6(MOF-1), by integrating the elongated linear (bpy)Re(CO)3Cl-containing dicarboxylate ligand (bpydb= 4,4’-(2,2’-bipyridine-5,5′-diyl)dibenzoate; bpy= 2,2’-bipyridine) with Zr63-O)43-OH)4 SBUs (Fig. 7) [104]. Under the same experimental condition as described above, CO-TON and H2-TON reached 6.44 and 0.4 in 6 h, respectively, which were higher than those of the homogeneous counterpart (CO-TON= 1.12 and H2-TON 0.18). Ryu et al. [52] embedded ReI(CO)3(5,5′-dcbpy)Cl and amine (-NH2) functional group within UiO-67 (dented as Re-MOF-NH2) and varied the ratio of the-NH2 functional groups from 0 to 80 mol%. Photocatalytic CO2 conversion was performed in the presence of TEA under visible light. The results showed that Re-MOF-NH2 incorporated with 33 mol% of-NH2 functional groups had the highest photocatalytic CO2 conversion rate (1.5 mmol/(g·h)) to CO, which was three times as that observed for Re-MOF without-NH2 incorporation (0.5 mmol/(g·h)). The enhancement in photocatalytic activity is caused by the induced different bond lengths for Re-CO in ReI(CO)3(5,5′-dcbpy)Cl by the incorporation of-NH2 functional groups, which endows the intermolecular stabilization of carbamate with CO2, thus boosting the photocatalytic activity. Fontecave and coworkers [105] functionalized UiO-67 MOF by replacing 5%–35% of BPDC linkers with Cp*Rh(5,5′-bpydb)Cl2 (Cp* = pentamethylcyclopentadiene) (named as Cp*Rh@UiO-67). Under visible light irradiation for 10 h in the presence of acetonitrile (ACN) and TEOA, HCOO--TON and H2-TON of 10%-Cp*Rh@UiO-67 reached 47 and 36, respectively, with [Ru(bpy)3]Cl2 as a photosensitizer.
Fig.7 Crystal structure of Zr6(O)4(OH)4[Re(CO)3Cl(bpydb)]6 (MOF-1)

Full size|PPT slide

Li and coworkers [106] incorporated Ru(CO)2Cl2 into MOF-253 (Al(OH)(5,5′‐dcbpy)), named as MOF-253-Ru(CO)2Cl2, and enhanced the CO2 photoreduction by producing 0.67 mmol of HCOO, 1.86 mmol of CO as well as 0.09 mmol H2 after irradiation under visible light for 8 h in MeCN and TEOA, whereas no products were detected over pristine MOF-253 under the same condition. The performance of MOF-253-Ru(CO)2Cl2 was further improved via photosensitizer (Ru(bpy)2Cl2) functionalization, by enhancing the light absorption in the visible light region. The HCOO, CO, and H2 produced in 8 h over sensitized MOF-253-Ru(CO)2Cl2 (Ru(bpy)2Cl2/Ru-complex was 1: 2) was 4.84, 1.85, and 0.72 mmol, respectively, which was much higher than those observed over the non-sensitized MOF-253-Ru(CO)2Cl2 under the same condition. In the sensitized MOF-253-Ru(CO)2Cl2, Ru(CO)2Cl2 reacted with the surface N,N-chelated sites to form MOF-253-supported Ru(bpy)2(X2bpy)2+, which extended the light absorption edge to 630 nm, wider than that of MOF-253-Ru(CO)2Cl2 (470 nm), thus promoting the photocatalytic CO2 reduction. However, a decrease in the photo-reactivity of the sensitized MOF-253-Ru(CO)2Cl2 was observed with increasing the amount of the photosensitizer Ru(bpy)2Cl2, which might be attributed to the pore blocking of MOF-253 by the Ru(bpy)2Cl2.
Luo and coworkers [107] developed a photocatalytic MOF (Y[Ir(ppy)2(4,4’-dcbpy)]2[OH]) (Ir-CP, ppy: 2-phenylpyridine, 4,4’-dcbpy: 2,2’-bipyridine-4,4’-dicarboxylate) using Y(NO3)3 and Ir(ppy)2(Hdcbpy). Photoreduction of CO2 over Ir-CP was performed in MeCN and TEOA under visible light irradiation. It showed that 38.0 mmol HCOO was produced in 6 h, with a product formation rate of 118.8 mmol /(g·h). The photocatalytic mechanism is proposed as follows: the [Ir(ppy)2(dcbpy)] unit in Ir-CP is excited under visible light and is reductively quenched by TEOA, and the CO2 molecules are reduced to HCOO by getting electrons from [Ir(ppy)2(dcbpy)]2– units. Lately, the same group [108,109] incorporated Ru-polypyridine complexes into MOF structures as metalloligands and produced a series of photocatalytic MOFs. Ru-MOF, with a chemical formula of [Cd2[Ru(4,4’-dcbpy)3]·12H2O]n, was constructed from [Cd2(CO2)6] SBUs and [Ru(4,4’-dcbpy)3]4–metalloligands [108]. The morphologies of Ru-MOF can be tuned to form nanoflowers, microflakes, and bulk crystals structures by controlling the reactants concentration, as shown in Fig. 8. In the mixture of MeCN/TEOA, Ru-MOF nanoflowers had the highest HCOO formation rate of 77.2 mmol/(g·h) under visible light irradiation for 8 h, followed by microflakes (52.7 mmol/(g·h)) and bulk crystals (30.6 mmol/(g·h)). The Ru-MOF nanoflowers had a better photocatalytic activity than their bulk counterparts because of their high visible light harvesting and long-lasting excited-state originated from their large surface area (8.08 m2/g; microflake: 1.33 m2/g) and high energy transfer efficiency. After this work, Luo and coworkers [109] synthesized two Ru-polypyridine-functionalized MOFs, [Cd3[Ru(5,5′-dcbpy)3]2·2(Me2NH2)]n and [Cd[Ru(bpy)(4,4′-dcbpy)2]·3H2O]n, with non-interpenetrated and interpenetrated structures, respectively.CO2 photoreduction over the two MOFs was conducted in the presence of MeCN and TEOA under visible light irradiation. It showed that the HCOO- production over [Cd3[Ru(5,5′-dcbpy)3]2·2(Me2NH2)]n and [Cd[Ru(bpy)(4,4′-dcbpy)2]·3H2O]n reached 16.1 and 17.2 mmol in 6 h (with a production rate of 67.5 and 71.7 mmol/(g·h)), respectively. The difference in photocatalytic performance of the two MOFs may have been associated with their structural stabilities. [Cd3[Ru(5,5′udcbpy)3]2·2(Me2NH2)]n has a non-interpenetrated structure in which the porous framework is supported by the coordination between metal ions and organic linkers as well as the interactions between the framework and the guest molecules in the pores. In the photocatalytic process, the exchange of guest molecules with the solvent molecules may cause the collapse of structure, leading to the decrease in photocatalytic activity. In contrast, [Cd[Ru(bpy)(4,4′-dcbpy)2]·3H2O]n with an interpenetrated structure has a higher structural stability, which endows its better photocatalytic activity than [Cd3[Ru(5,5′-dcbpy)3]2·2(Me2NH2)]n.
Fig.8 Amount of HCOO produced as a function of irradiation time of visible light over (a) nanoflowers, (b) microcrystals, and (c) bulk crystals of the Ru-MOF ([Cd2[Ru(4,4’-dcbpy)3]·12H2O]n). (d) visible light irradiation without a sample (Inset images (from top to bottom) show nanoflowers, microcrystals and bulk crystals of the Ru-MOF, respectively. Reproduced with permission from reference [108]. Copyright 2015, Royal Society of Chemistry. )

Full size|PPT slide

In addition to the ligand functionalization strategies discussed as above, catechol- [110] and anthracene-based [111] organic linkers were also utilized to functionalize MOF structures, which all had high photocatalytic performances on CO2 reduction.

MOF composite photocatalysts

MOFs have shown much potential on CO2 photoreduction thanks to their large surface areas, high CO2 uptake, as well as tunable structures and optical properties. To further improve the photocatalytic performance of MOFs, photosensitizers, semiconductors, metals and carbon materials have been incorporated into MOF structures to promote photo-generated electrons transfer and charge separation processes. The performances of MOF composites as photocatalysts for CO2 reduction are summarized in Table 3.
Tab.3 Performances of recent photocatalytic MOF composites for CO2photoreduction
Strategy MOF composite Irradiation Solvent/
sacrificial agent
Main product Photocatalytic reactivity Reaction time/h Ref.
Photosensitizer incorporation Co-ZIF-9/[Ru(bpy)3]Cl2·6H2O Visible MeCN/H2O/
TEOA
CO 41.8 mmol 0.5 [57]
H2 29.9 mmol
Co-MOF-74/[Ru(bpy)3]Cl2·6H2O CO 11.7 mmol
H2 7.3 mmol
Mn-MOF-74/[Ru(bpy)3]Cl2·6H2O CO 1.5 mmol
H2 2.9 mmol
Zn-ZIF-8/[Ru(bpy)3]Cl2·6H2O CO 2.1 mmol
H2 2.4 mmol
Zr-UiO-66-NH2/[Ru(bpy)3]Cl2·6H2O CO 1.2 mmol
H2 2.2 mmol
Co-ZIF-67 /[Ru(bpy)3]Cl2·6H2O Visible MeCN/H2O/TEOA CO 29.6 mmol 0.5 [59]
H2 14.8 mmol
Zn-ZIF-9/[Ru(bpy)3]Cl2·6H2O CO 1.8 mmol
H2 2.0 mmol
Cu-HKUST-1/[Ru(bpy)3]Cl2·6H2O CO 1.2 mmol
H2 1.5 mmol
Fe-MIL-101-NH2/[Ru(bpy)3]Cl2·6H2O CO 4.7 mmol
H2 2.1 mmol
Zr-UiO-66-NH2/[Ru(bpy)3]Cl2·6H2O CO 0.9 mmol
H2 1.2 mmol
UiO-67-Mn(5,5′‐dcbpy) (CO)3Br)/Ru(dmb)3(PF6)2 Visible DMF/TEOA/ BNAH HCOO TON= 50 4 [58]
TON= 110 18
UiO-67-Mn(5,5′‐dcbpy)(CO)3Br) (without a photosensitizer) TON= 18 18
Mn(5,5′‐dcbpy)(CO)3Br)/
Ru(dmb)3(PF6)2
TON= 32 4
TON= 57 18
Mn(bpy)(CO)3Br)/Ru(dmb)3(PF6)2 TON= 35 4
TON= 70 18
UiO-67-5,5′‐dcbpy)/Ru(dmb)3(PF6)2 TON= 38 18
[Ru(dmb)3]2+ HCOO TON= 33 18
Semiconductor incorporation ZIF-8/TiO2 (ZIF-8 growth step on TiO2 film was repeated twice) UV H2O vapor CO 0.53 mmol/(g·h) 5 [124]
CH4 0.18 mmol/(g·h)
ZIF-8/Ti/TiO2 nanotube UV-visible Na2SO4
(0.1 mol L-1)
C2H5OH 10 mmol/L 3 [125]
CH3OH 0.7 mmol/L
Co-ZIF-9/TiO2 (mass ratio of Co-ZIF-9 in composite is 0.03) UV-visible H2O vapor CO 8.79 mmol 10 [54]
CH4 0.99 mmol
H2 1.30 mmol
TiO2 CO 3.58 mmol
CH4 0.60 mmol
H2 0.63 mmol
Co-ZIF-9 CO 0
CH4 0
H2 0
Physical mixture of TiO2 and Co-ZIF-9 with the mass ratio of 0.03:0.97 CO 3.86 mmol
CH4 0.42 mmol
H2 0.56 mmol
Cu-BTC/TiO2 UV H2O vapor CH4 2.64 mmol/(g TiO2·h) 4 [126]
TiO2 CH4 0.52 mmol/(g TiO2·h)
H2 2.29 mmol/(g TiO2·h)
Cu-BTC CH4 0
H2 0
Cu-BTC/TiO2 (molar ratio of Cu-BTC to TiO2 is 3.33) N/A CO2/H2O
vapor
CO 256.38 mmol/(g TiO2·h) 8 [127]
TiO2 11.48 mmol/(gTiO2·h)
Cu-BTC 0
CPO-27-Mg/TiO2 UV H2O vapor CO 40.9 mmol/g 10 [128]
CH4 23.5 mmol/g
TiO2 CO 22.5 mmol/g
CH4 13.7 mmol/g
CPO-27-Mg CO 0
CH4 0
Physical mixture of TiO2 and CPO-27-Mg with the ratio of 6:4 H2 8.5 mmol/g
CO 18.9 mmol/g
CH4 7.1 mmol/g
NH2-UiO-66/TiO2 (with 19%(wt) NH2-UiO-66) UV-visible CO2/H2 CO 3.74 mmol/(g·h) [129]
NH2-UiO-66/TiO2 (19.5%(wt)NH2-UiO-66) 4.24 mmol/(g·h)
NH2-UiO-66/TiO2 (24.5%(wt) NH2-UiO-66) 3.37 mmol/(g·h)
NH2-UiO-66/TiO2 (36.8%(wt) NH2-UiO-66) 2.85 mmol/(g·h)
TiO2 2.85 mmol/(g·h)
NH2-UiO-66 1.50 mmol/(g·h)
Co-ZIF-9/CdS Visible MeCN/H2O /TEOA/bpy CO 50.4 mmol 1
5
[130]
H2 11.1 mmol
Co-MOF-74/CdS CO 39.6 mmol
H2 7.7 mmol
Mn-MOF-74/CdS CO 1.0 mmol
H2 2.0 mmol
Zn-ZIF-8/CdS CO 0.6 mmol
H2 0.6 mmol
Zr-UiO-66-NH2/CdS CO 0.4 mmol
H2 0.3 mmol
CdS CO 0.5 mmol
H2 0.6 mmol
Co-ZIF-9 CO 0
H2 0
UiO-66-NH2/Cd0.2Zn0.8S (10%(wt)UiO-66-NH2) Visible Na2S/Na2SO3 H2 4591.6 mmol/(g·h) [55]
CH3OH 4.1 mmol/(g·h)
UiO-66-NH2/Cd0.2Zn0.8S (20%(wt)UiO-66-NH2) H2 5846.5 mmol/(g·h)
CH3OH 6.8 mmol/(g·h)
UiO-66-NH2/Cd0.2Zn0.8S (30%(wt)UiO-66-NH2) H2 5235.9 mmol/(g·h)
CH3OH 5.9 mmol/(g·h)
UiO-66-NH2/Cd0.2Zn0.8S (40%(wt)UiO-66-NH2) H2 4922.7 mmol/(g·h)
CH3OH 5.3 mmol/(g·h)
Cd0.2Zn0.8S H2 2804.2 mmol/(g·h)
CH3OH 2.0 mmol/(g·h)
UiO-66-NH2 H2 0
CH3OH 0
Co-ZIF-9/mesoporous g-C3N4 Visible MeCN/H2O /TEOA/bpy CO 20.8 mmol 2 [131]
H2 3.3 mmol
Co-ZIF-9 CO 0
H2 0
g-C3N4 CO 0
H2 0
ZIF-8/g-C3N4 nanotubes (molar ratio of g-C3N4 nanotubes to ZIF-8 is 10) UV-visible CO2/H2O
vapor
CH3OH 0.64 mmol/(g·h) 1 [56]
ZIF-8/g-C3N4 nanotubes (molar ratio of g-C3N4 nanotubes to ZIF-8 is 8) 0.75 mmol/(g·h)
ZIF-8/g-C3N4 nanotubes (molar ratio of g-C3N4 nanotubes to ZIF-8 is 5) 0.45 mmol/(g·h)
ZIF-8/g-C3N4 nanotubes (molar ratio of g-C3N4 nanotubes to ZIF-8 is 2) 0.31 mmol/(g·h)
ZIF-8/g-C3N4 nanotubes (molar ratio of g-C3N4 nanotubes to ZIF-8 is 1) 0.16 mmol/(g·h)
g-C3N4 nanotubes 0.49 mmol/(g·h)
Bulk g-C3N4 0.24 mmol/(g·h)
ZIF-8 nanocrystals 0
UiO-66/g-C3N4 nanosheets Visible MeCN/TEOA CO 9.9 mmol/(g g-C3N4·h) 6 [132]
UiO-66/bulk g-C3N4 3.2 mmol/(g g-C3N4·h)
g-C3N4 nanosheets 2.9 mmol/(g g-C3N4·h)
bulk g-C3N4 2.0 mmol/(g g-C3N4·h)
UiO-66 0
BIF-20/g-C3N4 nanosheets (10%(wt) g-C3N4 nanosheets) Visible MeCN/TEOA CO 3.42 mmol 6 [133]
CH4 1.12 mmol
BIF-20/g-C3N4 nanosheets (15%(wt) g-C3N4 nanosheets) CO 4.86 mmol
CH4 1.45 mmol
BIF-20/g-C3N4 nanosheets (20%(wt) g-C3N4 nanosheets) CO 6.12 mmol
CH4 1.76 mmol
BIF-20/g-C3N4 nanosheets (25%(wt) g-C3N4 nanosheets) CO 5.14 mmol
CH4 1.51 mmol
ZIF-8/Zn2GeO4 (25%(wt) ZIF-8) N/A Na2SO3 CH3OH 0.22 mmol/(g·h) 11 [134]
Metal incorporation Pt/NH2-MIL-125(Ti) Visible MeCN/TEOA HCOO 12.96 mmol 8 [135]
H2 235 mmol
Au/NH2-MIL-125(Ti) HCOO 9.06 mmol
H2 40.2 mmol
NH2-MIL-125(Ti) HCOO 10.75 mmol
H2 0
1%(wt) Co/NH2-MIL-125(Ti) Visible MeCN/TEOA HCOO 384.2 mmol 10 [136]
2%(wt) Co/NH2-MIL-125(Ti) 321.8 mmol
3%(wt) Co/NH2-MIL-125(Ti) 239.4 mmol
NH2-MIL-125(Ti) 162.8 mmol
Ag⊂Re3-MOF (16 nm thick Re3-MOF) Visible MeCN/TEOA CO TON ≈ 2.8 48 [137]
ReI(CO)3(5,5′‐dcbpy)Cl TON ≈ 1.7
Carbon materials incorporation 1%(wt) UiO-66-NH2/graphene Visible DMF/TEOA/H2O HCOO 12.3 mmol 4 [138]
CH4 0.25 mmol
H2 15.2 mmol
1.5%(wt) UiO-66-NH2/graphene HCOO 21.2 mmol
CH4 0.59 mmol
H2 13.9 mmol
2%(wt) UiO-66-NH2/graphene HCOO 33.5 mmol
CH4 0.90 mmol
H2 13.2 mmol
2.5%(wt) UiO-66-NH2/graphene HCOO 14.9 mmol
CH4 0.51 mmol
H2 15.1 mmol
3%(wt) UiO-66-NH2/graphene HCOO 8.6 mmol
CH4 0.19 mmol
H2 16.8 mmol
UiO-66-NH2 HCOO 3.1 mmol
CH4 0.11 mmol
H2 16.9 mmol
Graphene HCOO 0
CH4 0
H2 0
UiO-66-NH2/graphene (hydrothermal synthesis) HCOO 16.1 mmol
H2 20.4 mmol
Al-PMOF/5%(wt) NH2-rGO Visible MeCN/TEOA HCOO 685.6 mmol/(g·h) 6 [139]
Al-PMOF/15%(wt) NH2-rGO 479.8 mmol/(g·h)
Al-PMOF/25%(wt) NH2-rGO 476.4 mmol/(g·h)
Al-PMOF 165.3 mmol/(g·h)

Notes: CAN–acetonitrile; BNAH–1-benzyl-1,4-dihydronicotinamide; bpy–2,2’-bipyridine; BTC–benzene-1,3,5-tricarboxylate; Cp*–pentamethylcyclopentadiene; DMF–N, N-dimethylformamide; MeCN–acetonitrile; phen–phenanthroline; TEA–trimethylamine; TEOA–triethanolamine; TON–total turnover number; 5,5′‐dcbpy–2,2’-bipyridine-5,5′-dicarboxylic acid; bpydb –4,4’-(2,2’-bipyridine-5,5′-diyl)dibenzoate; ppy–2-phenylpyridine; 4,4′‐dcbpy–2,2’-bipyridine-4,4’-dicarboxylate; dmb–4,4′-dimethyl-2,2′-bipyridine

MOF composites incorporated with photosensitizers

As discussed in Section 2.2.3, metal-complex photosensitizers have been used to functionalize MOFs to enhance the photocatalytic activity. Photosensitizers can also be incorporated into MOFs to produce MOF composites for CO2 photoreduction.
Ru-based photosensitizer has been incorporated into several MOFs to enhance CO2 photocatalytic reduction [5759]. Wang and coworkers [57] investigated the performance of a series of MOFs (Co-ZIF-9, Co-MOF-74, Mn-MOF-74, Zn-ZIF-8, Zr-UiO-66-NH2) in conjunction with [Ru(bpy)3]Cl2·6H2O toward CO2 photocatalytic reduction, where the MOF acts as a co-catalyst and [Ru(bpy)3]Cl2·6H2O acts as a photosensitizer. Photoreduction of CO2 was performed in MeCN and H2O solvent with TEOA as a sacrificial agent. After visible light irradiation for 0.5 h, 41.8 mmol of CO and 29.9 mmol of H2 were obtained over Co-ZIF-9 incorporated with [Ru(bpy)3]Cl2·6H2O, which outperformed the counterparts with a lower yield of CO and H2 under the same condition (as shown in Table 3). Co-ZIF-9 is a microporous cobalt-containing benzimidazolate MOF. The superior performance of Co-ZIF-9 with regard to CO2 photocatalytic reduction is a result of the synergetic effect of the imidazolate-based ligand which has a strong interaction with CO2 molecules for CO2 adsorption and cobalt with electron-mediating functions. Lately, the same group incorporated [Ru(bpy)3]Cl2·6H2O as a photosensitizer into Co-ZIF-67 acting as a co-catalyst [59]. Under the similar reaction condition, 37 mmol of CO and 13 mmol of H2 were produced over Co-ZIF-67/[Ru(bpy)3]Cl2·6H2O in 0.5 h under visible light irradiation. This was higher than that observed over Zn-ZIF-8/[Ru(bpy)3]Cl2·6H2O (1.8 mmol of CO and 2.0 mmol of H2 production in 0.5 h under the same condition), where the co-catalyst Zn-ZIF-8 has the same organic linker (2-methylimidazole) as Co-ZIF-67 but different metal site (Zn-based). This further confirmed the effect ofcobalton the photocatalytic reaction.
Cohen and coworkers [58] incorporated a photosensitizer [Ru(dmb)3]2+ (dmb= 4,4′-dimethyl-2,2′-bipyridine) into UiO-67 which was functionalized with a Mn+bipyridine complex, Mn(5,5′‐5,5′‐dcbpy)-(CO)3Br for CO2 photocatalytic reduction. In the mixture of N, N-dimethylformamide (DMF), TEOA and BNAH, HCOO- was produced with the TON of 110 under visible light irradiation for 18 h, which outperformed the UiO-67-Mn(5,5′‐dcbpy)-(CO)3Br without a photosensitizer [Ru(dmb)3]2+ and the homogeneous reference systems (Fig. 9). The superior photocatalytic performance is attributed to the isolated active sites in MOF framework, which endows the high stability of the catalyst and impedes the dimerization of the singly reduced Mn complex.
Fig.9 TON of HCOO as a function of reaction time over UiO-67-Mn(bpy)(CO)3Br (red), Mn(bpy)(CO)3Br (green), Mn(bpydc)(CO)3Br (blue), UiO-67-bpydc (black), no added Mncomplex or MOF (only Ru2+ , brown), and UiO-67-Mn(bpy)(CO)3Br without added Ru2+ (gray) (Reproduced with permission from Ref. [58]. Copyright 2015, American Chemical Society)

Full size|PPT slide

MOF composites incorporated with semiconductors

Coupling MOFs with semiconductors has been shown to be another approach to reducing the recombination rate of the photo-generated charge carries and consequently enhancing the photocatalytic performance. Several types of semiconductors such as TiO2, CdS and graphitic C3N4 have been embedded into MOFs to form heterogeneous structures which retain the properties of both the semiconductors and MOFs that are beneficial to CO2 photocatalytic reactions.

MOF-TiO2 composites

Several studies reported the development of ZIF-8/TiO2 composites for CO2 photoreduction. Zhang and coworkers [124] integrated ZIF-8 into the TiO2 film grid and conducted CO2 photocatalytic reactions in H2O as an electron donor without the use of a sacrificial agent under UV irradiation for 5 h. The amount of ZIF-8 in ZIF-8/TiO2 composite was varied by repeating the ZIF-8 growth step on the TiO2 film from one to three times, which are denoted as TiMOF-1, TiMOF-2, and TiMOF-3, respectively. TiMOF-3 with the highest amount of ZIF-8 had the highest CO2 adsorption uptake, followed by TiMOF-2, TiMOF-1, and TiO2. TiMOF-2 had the best performance with a CO yield of 0.53 mmol/(g·h) and a CH4 yield of 0.18 mmol/(g·h), which was 38% and 157% higher than that observed over pure TiO2 film under the same condition. TiMOF-2 outperformed TiMOF-3 which had the highest CO2 adsorption capacity due to the fact that the large amount of ZIF-8 in TiMOF-3 covered the TiO2 film, impeding the photo excitation process of TiO2. It is proposed that both TiO2 and ZIF-8 can be activated to generate photo-induced charge carries under UV irradiation, while TiO2 is more effective in the photoexcitation process than ZIF-8 because of its higher photoactivity and narrower band gap (-0.45 eV vs. -0.5 eV for ZIF-8). Electrons can also transfer from TiO2 to ZIF-8 and be involved in the photoreduction process over ZIF-8.
Cardoso and coworkers [125] developed a MOF-based Ti/TiO2 composite photocatalyst by growing ZIF-8 thin films on Ti/TiO2 nanotube (NT) electrodes using a layer-by-layer process. Spectroscopic and voltammetric assays revealed that the CO2 adsorbed on ZIF-8 formed stable carbamates. The photoelectrocatalytic reduction of CO2 over Ti/TiO2NT-ZIF-8 electrodes was performed in Na2SO4 (pH 4.5) saturated with CO2 at a constant potential of+0.1 V under UV-visible light irradiation. 10 mmol/L of ethanol and 0.7 mmol/L of methanol were produced in 3 h, which increased around 20 and 430 times, respectively, compared to the values observed over Ti/TiO2NT due to the low CO2 adsorption capacity of Ti4+ species in the absence of ZIF-8.
Co-ZIF-9/TiO2 composites with different mass ratios of Co-ZIF-9 (named as ZIFx/T; where x represents the mass ratios of Co-ZIF-9 in the composite, which equals 0.01, 0.03, 0.10, 0.20, 0.30, 0.40, and 0.60) were synthesized via an in situ synthetic method by Ye and coworkers for CO2 photoreduction [54]. Of the ZIFx/T composites, ZIF0.03/T had the best photocatalytic performance with a CO yield of 8.79 mmol, CH4 of 0.99 mmol and H2 of 1.30 mmol under UV-visible light irradiation for 10 h, which was higher than that of the pure TiO2 (3.58 mmol of CO, 0.60 mmol of CH4 and 0.63 mmol of H2) and Co-ZIF-9 (no CO, CH4 or H2 was detected). ZIF0.03/T also had a better photocatalytic performance than the physical mixture of TiO2 and Co-ZIF-9 at the same mass ratio of 0.03:0.97 (where 3.86 mmol of CO, 0.42 mmol of CH4, and 0.56 mmol of H2 were produced) due to the better charge separation. It was found that when the mass ratios of Co-ZIF-9 was higher than 0.1, the photocatalytic activity decreased with increasing the Co-ZIF-9 content in ZIFx/T, which might be attributed to the heavier charge recombination.
Pipelzadeh et al. [140] reported the CO2 photoreduction to CH4 and CO over a ZIF-8/TiO2 composite with core-shell structure in a photoreactor equipped with simulated sunlight at a constant pressure (CP, 5 bar) and an intentionally controlled pressure swing (PS) at 50 mL/min (PS-50) and 100 mL/min (PS-100). A high CO yield was achieved in the PS mode at a production rate of 13.2 mmol/(g·h) (PS-50, 80% increase than CP-50 operation) and 15.6 mmol/(g·h) (PS-100, 30% increase than CP-100). The PS mode had a better promotion effect on CO production than CH4. Continuous alteration of the reactants and product adsorption/desorption over the photocatalyst in the PS mode was beneficial to the regeneration of Ti3+ active sites, which contributed to the enhancement of photoreduction activity [17]. In addition, a higher gas flow rate facilitated CO removal to avoid catalyst poisoning. The calcination of ZIF-8/TiO2 composite at 300°C further increased the yield of CO with 45.16 mmol/(g·h) under PS-100 condition, which was higher than that observed in the CP-100 mode (33.46 mmol/(g·h)) because of the higher structural stability of the ZIF-8/TiO2 composite.
Composite photocatalysts composed of Cu-BTC (also named as HKUST-1, BTC= benzene-1,3,5-tricarboxylate) and TiO2 were also developed for CO2 photocatalytic reduction. Ye and coworkers [126] synthesized a Cu-BTC/TiO2 composite with a core-shell structure (see morphologies in Fig. 10) and evaluated the CO2 photocatalytic performance. Under UV irradiation for 4 h, the production rate of CH4 and H2 from CO2 over the bare TiO2 reached 0.52 and 2.29 mmol/(gTiO2·h), respectively, whereas the formation rate of CH4 over the Cu-BTC/TiO2 composite reached 2.64 mmol/(gTiO2·h) with no H2 detected. No product was produced over Cu-BTC, because its conjugated structure did not favor charge separation. This indicates that the CH4 yield and the selectivity of CH4 to H2 over Cu-BTC/TiO2 composite are significantly improved compared to the bare TiO2 and Cu-BTC in photocatalytic reduction. Under UV irradiation, the TiO2 in Cu-BTC/TiO2 composite is photoexcited to generate charge pairs and the electrons can be effectively transferred to Cu-BTC, as demonstrated by the ultrafast spectroscopy. This facilitates the charge separation in TiO2 and supplies active electrons to CO2 molecules adsorbed on Cu-BTC, leading to an enhancement in photocatalytic activity of Cu-BTC/TiO2 composite. Theoretical simulations demonstrate that the activation-energy barrier for CO2 on the Cu sites in Cu-BTC will be lowered upon receiving the photo-excited electrons from TiO2, enabling the CO2 reduction occurring on the Cu sites of Cu-BTC and the enhanced selectivity of CH4 to H2 production. Lately, Wang and coworkers [127] synthesized Cu-BTC/TiO2 composites in microdroplets via an aerosol route. Similarly, Cu-BTC/TiO2 composites had a higher photocatalytic performance than the pure TiO2 and Cu-BTC. The yield of CO from CO2 conversion over Cu-BTC/TiO2 composites increased as a function of the molar ratio of Cu-BTC to TiO2 ranging from 0 (i.e., bare TiO2) to 3.33 (see Fig. 11), where the production rate of CO over 3.33Cu-BTC/TiO2 increased to 256.35 mmol/(gTiO2·h) as compared to the bare TiO2 with a CO formation rate of 11.48 mmol/(gTiO2·h). As demonstrated by the in situ diffuse reflectance infrared Fourier transform spectrometer (DRIFTS) analysis, the enhancement in the photoreduction performance of the Cu-BTC/TiO2 composites may have been caused by the improved adsorption of reactants on the catalyst.
Fig.10 Core‐shell structures of Cu(BTC)/TiO2

Full size|PPT slide

Fig.11 CO2 photoreduction analysis of TiO2 and HKUST-1/TiO2 composites (CO yield over the HKUST-1/TiO2 composites and pure TiO2) Inset image: CO yield peak time of the HKUST-1/TiO2 composites and pure TiO2 (Reproduced with permission from Ref. [127]. Copyright 2017, American Chemical Society)

Full size|PPT slide

In addition to ZIF and Cu-based MOFs, TiO2 was also incorporated into other MOF structures to form MOF-TiO2 composites for CO2 photoreduction. Li and coworkers [128] combined TiO2 with CPO-27-Mg (also named as Mg2(DOBDC), DOBDC= 2,5-dioxido-1,4-benzenedicarboxylate) to produce a CPO-27-Mg/TiO2 composite via a hydrothermal self-assembly method. There exists a high concentration of open alkaline metal sites (Mg2+) in CPO-27-Mg structure, endowing the high CO2 adsorption capacity of CPO-27-Mg. Under UV irradiation for 10 h, 40.9 mmol/g of CO and 23.5 mmol/g of CH4 were produced over CPO-27-Mg/TiO2, which were higher than those observed over pure TiO2 (22.5 mmol/g of CO and 13.7 mmol/g of CH4). The enhanced performance of CPO-27-Mg/TiO2 composite on CO2 photoreduction is attributed to its high CO2 adsorption capacity and the existence of open Mg2+ metal sites. The CO2 photoreduction over a physical mixture of TiO2 and CPO-27-Mg (The ratio of TiO2 to CPO-27-Mg was 6:4.) was also performed under the similar condition, and 8.5 mmol/g of H2, 18.9 mmol/g of CO, and 7.1 mmol/g of CH4 were produced, which were lower than those produced over the CPO-27-Mg/TiO2 composite. This demonstrates the indispensable effect of the strong interaction between CPO-27-Mg and TiO2 in CPO-27-Mg/TiO2 composite for the CO2 photoreduction. TiO2 nanosheets were coupled with NH2-UiO-66 using an in situ growth strategy by Petit and coworkers [129]. The content of NH2-UiO-66 in the composite was varied from 19 wt% to 37 wt%. NH2-UiO-66/TiO2 had a better performance in photo reducing CO2 to CO than their single moiety. This improvement is a result of the enhanced abundance of long-lived charge carriers and high CO2 adsorption capacity in NH2-UiO-66/TiO2 composites.

MOF-CdS composites

CdS semiconductor has been widely used for CO2 photoreduction as a photocatalyst [141143]. Wang et al. [130] incorporated CdS into Co-ZIF-9 and the as-obtained Co-ZIF-9/CdS composite had a good photocatalytic activity toward CO2 conversion to CO under visible light irradiation. CO2 photoreduction was performed in MeCN and H2O solvent with TEOA as a sacrificial agent and bipyridine (bpy) as an assistant for electron transfer. After visible light irradiation for 1 h, 50.4 mmol of CO and 11.1 mmol of H2 were produced over Co-ZIF-9/CdS composite, which outperformed its counterparts with lower yields of CO and H2 and pure CdS semiconductor (0.5 mmol of CO and 1.6 mmol of H2) under the same condition(as shown in Table 3). The photocatalytic mechanism was proposed as follows: under visible light irradiation, the CdS semiconductor was excited and charge carriers were generated. The photo-generated electrons transferred to Co-ZIF-9 and reduced the CO2 molecules adsorbed on Co-ZIF-9 to CO. Meanwhile the protons existed in the reaction system were also reduced to H2 by the excited electrons. Lately, Su et al. [55] prepared a series of UiO-66-NH2/Cd0.2Zn0.8S composites with different UiO-66-NH2 contents using a solvothermal method. CO2 photoreduction was performed over UiO-66-NH2/Cd0.2Zn0.8S composites under visible light irradiation, which all had an enhanced photocatalytic activity in comparison to their single components. The UiO-66-NH2/Cd0.2Zn0.8S composite with a UiO-66-NH2 content of 20 wt% had the best photocatalytic performance with a H2 production rate of 5846.5 mmol/(g·h) and a CH3OH production rate of 6.8 mmol/(g·h). The efficient charge separation and transfer between Cd0.2Zn0.8S and UiO-66-NH2 contributed to the enhanced photocatalytic activity of UiO-66-NH2/Cd0.2Zn0.8S composites.

MOF-graphitic C3N4 composites

Graphitic carbon nitrides (g-C3N4) with different morphologies and structures have been integrated with MOFs to improve the CO2 photoreduction activity. Wang and coworkers [131] coupled mesoporous g-C3N4 with Co-ZIF-9, which acted as a light harvester and co-catalyst, respectively, to fabricate a Co-ZIF-9/g-C3N4 composite. The Co-ZIF-9/g-C3N4 composite efficiently catalyzed CO2 to CO and H2 under visible light irradiation. 20.8 mmol of CO and 3.3 mmol of H2 were obtained over the Co-ZIF-9/g-C3N4 composite in 2 h, whereas no product was detected over the pristine Co-ZIF-9 and g-C3N4. Liu and coworkers [56] developed a series of ZIF-8/g-C3N4 composites by growing different contents of ZIF-8 nanoclusters on the surface of g-C3N4 nanotubes. The ZIF-8/g-C3N4 composites had an increased CO2 adsorption capacity than g-C3N4 nanotubes without sacrificing the light absorption capacity owing to the incorporation of ZIF-8 nanoclusters. Because of the high CO2 capture capacity of ZIF-8 and the promoted charge separation efficiency from the g-C3N4 nanotubes, the ZIF-8/g-C3N4 composites had an enhanced photocatalytic performance on CO2 reduction, where the highest production rate of methanol reached 0.75 mmol/(g·h) over the ZIF-8/g-C3N4 composite in which the mass ratio of g-C3N4 nanotubes to ZIF-8 was 8 under light irradiation for 1 h. Under similar conditions, g-C3N4 nanotubes and bulk g-C3N4 had a production rate of methanol of 0.49 and 0.24 mmol/(g·h), respectively, whereas no methanol was produced over the pure ZIF-8 nanocrystals. In addition, g-C3N4 nanosheets were combined with UiO-66 [132] and BIF-20 (a zeolite-like porous boron imidazolate framework) [133] MOFs to form MOF/g-C3N4composites, which all had an enhanced performance on CO2 photocatalytic reduction.
In addition to TiO2, CdS and C3N4, another type of semiconductor was also incorporated into MOFs to generate a composite photocatalyst for CO2 photocatalytic reduction. Wang and coworkers [134] developed a ZIF-8/Zn2GeO4 composite by growing ZIF-8 nanoparticles on Zn2GeO4 nanorods. The ZIF-8/Zn2GeO4 composite inherited both the high CO2 adsorption capacity of ZIF-8 nanoparticles and the high crystallinity of Zn2GeO4 nanorods. The ZIF-8/Zn2GeO4 composite with 25 wt% ZIF-8 had a CO2 adsorption capacity of 15.5 cm3/g, which was higher than the pure Zn2GeO4 nanorods (4.9 cm3/g) due to the high CO2 adsorption ability of ZIF-8. After 11 h of light irradiation in Na2SO3, the production of methanol at a rate of 0.22 mmol/(g·h) over the ZIF-8(25wt%)/Zn2GeO4 composite was observed. The yield of methanol over the ZIF-8(25 wt%)/Zn2GeO4 composite had a 62% increase in comparison to the pure Zn2GeO4 nanorods under light irradiation for 10 h. The enhanced CO2 photoreduction performance of ZIF-8/Zn2GeO4 composite may have been resulted from the high CO2 adsorption capacity of ZIF-8 and the higher light response.

MOF composites incorporated with metals

Because of their high Fermi energy levels, noble metal nanoparticles can effectively separate photo-generated charge pairs of photocatalysts [144,145]. Therefore, another approach for promoting the CO2 photocatalytic reduction is to dope noble metal nanoparticles into MOFs structures to decrease the recombination rate of the photo-generated electrons and holes. Li and coworkers [135] synthesized M-doped NH2-MIL-125(Ti) (M= Pt and Au) and conducted CO2 photocatalytic reaction in saturated CO2 with TEOA as a sacrificial agent under visible light irradiation. Both H2 and HCOO- were produced over M/NH2-MIL-125(Ti), while no H2 but only HCOO- was produced over bare NH2-MIL-125(Ti). The reason for this is that the electron-trapping effect of noble metals promotes the hydrogen evolution. In addition, it is noted that Pt and Au had different photocatalytic performances on the production of HCOO-. Compared to bare NH2-MIL-125(Ti) (yield of HCOO-: 10.75mmol), Pt/NH2-MIL-125(Ti) had a higher production yield of HCOO- (12.96 mmol), while Au/NH2-MIL-125(Ti) had a lower production of HCOO- (9.06 mmol) under visible light irradiation for 8 h. As demonstrated by ESR studies and DFT calculations, hydrogen spillover from Pt to the bridging oxygen linked to Ti atoms occurred in Pt/NH2-MIL-125(Ti), leading to the Ti3+ formation and boosting the hydrogen-assisted CO2 reduction to HCOO-. In contrast, it was difficult to achieve the hydrogen spillover from Au to the NH2-MIL-125(Ti) framework in Au/NH2-MIL-125(Ti), and thus resulting in a lower HCOO- formation over Au/NH2-MIL-125(Ti). Fu et al. [136] doped different contents of Co (from 1 to 3 wt%) into NH2-MIL-125(Ti) and produced Co/NH2-MIL-125(Ti) composites for CO2 photoreduction under visible light irradiation. 1 wt% Co/NH2-MIL-125(Ti) had the best performance on CO2 photoreduction with an HCOO- formation of 384.2 mmol in 10 h, which was 2-fold higher than that observed over pure NH2-MIL-125(Ti) (162.8 mmol) due to the enhanced visible-light harvesting and electron transfer stemmed from the addition of Co.
Yaghi and coworkers [137] functionalized UiO-67 with Re complexes, ReI(CO)3(5,5′‐dcbpy)Cl photosensitizer, of various densities (Ren-MOF, n = 0, 1, 2, 3, 5, 11, 16, and 24 complexes per unit cell) and found that the photocatalytic activity of the MOF system can be controlled by controlling the density of Re complexes in the framework, in which Re3-MOF had the best performance on CO2-to-CO conversion. The photocatalytic activity was further enhanced by coating a 16 nm layer of Re3-MOF onto Ag nanocubes (Ag⊂Re3-MOF, see Fig. 12), where a 7-fold improvement of CO2-to-CO reduction compared to pure Re3-MOF under visible light irradiation was achieved. Both Re complexes and Ag nanocubes contributed to the CO2 photocatalytic activity enhancement of Ag⊂Re3-MOF. Because the quadrupolar localized surface plasmon resonance (LSPR) scattering peak (lmax~ 480 nm) of Ag nanocube overlapped with the absorption range of ReI(CO)3(5,5′‐dcbpy)Cl (400 nm<l<550 nm) in the visible region [146,147] and Ag⊂Re3-MOF structure inherited the LSPR features of Ag cores, the photoactive Re metal sites within MOF shell were spatially localized into a strong electromagnetic field induced by LSPR of the Ag nanocubes for photocatalytic enhancement.
Fig.12 Structures of Ren-MOF and Ag⊂Ren-MOF for plasmon-enhanced photocatalytic CO2 conversion

Full size|PPT slide

MOF composites incorporated with carbon materials

Still, another important strategy to develop active photocatalysts for CO2 photoreduction is the incorporation of carbon materials such as graphene and its derivatives into MOFs. Graphene can act as a photosensitizer to extend the light adsorption region from the UV to the visible light region. Moreover, the excellent conductivity of graphene can facilitate the rapid transfer of the photo-generated electrons and suppress the electron-hole recombination, and subsequently boost the photocatalytic activity.
Li and coworkers [138] synthesized UiO-66-NH2/graphene composites via microwave-assisted in situ growth of different amounts (1–3 wt%) of UiO-66-NH2 nanocrystals onto graphene. As compared to the pure UiO-66-NH2 (3.1 mmol of HCOO-, 0.11 mmol of CH4, and 16.9 mmol of H2 were produced.) and the UiO-66-NH2/graphene synthesized via a traditional hydrothermal route (16.1 mmol of HCOO- and 20.4mmol of H2 were produced.), the as-obtained UiO-66-NH2/graphene composites had a better performance on CO2 photoreduction activity and selectivity under visible light irradiation. Of the UiO-66-NH2/graphene composites as prepared, 2 wt% UiO-66-NH2/graphene composite had the best performance, where 33.5mol of HCOO-, 0.9 mmol of CH4 and 13.2 mmol of H2 were produced in 4 h. The performance enhancement is attributed to both the fine particle size and high dispersion of UiO-66-NH2 nanocrystals, which enables more light trapping to generate charge pairs and shortens the electron transfer pathway to facilitate electron transfer. Additionally, the strong UiO-66-NH2/ graphene interaction can also effectively accelerate electrons transfer efficiency to enhance photocatalytic activity for CO2 reduction.
Do and coworkers [139] incorporated different contents of amine-functionalized reduced graphene oxide (NH2-rGO) (5–25 wt%) into a TCPP-based MOF (Al/PMOF) to form Al-PMOF/NH2-rGO composites as photocatalysts for CO2 reduction. Al-PMOF/NH2-rGO composites had an enhanced photocatalytic activity for CO2 reduction, where the HCOO- formation rate reached 685.6 mmol /(g·h) over Al-PMOF/5 wt% NH2-rGO under visible light irradiation in 6 h, which was much higher than the HCOO- production rate observed over pure Al-PMOF (165.3 mmol /(g·h)). The photocatalytic mechanism is proposed as follows: upon visible light irradiation, TCPP is responsible for light harvesting and is excited to produce photo-generated electron-hole pairs, and the electrons are transferred from TCPP to graphene which acts as an electron acceptor. The electrons transferred from graphene reduce the adsorbed CO2 to HCOO- in the presence of TEOA that serves as a hydrogen source.
Most recently, MOF with a special microstructure in CO2 photoreduction was reported. Lin and coworkers [148] prepared Ni-based MOF (Ni2(OH)2BDC) monolayers (Ni MOLs) and examined the performance for photoreduction of CO2 under visible light irradiation with [Ru(bpy)3]Cl2⋅6H2O as a photosensitizer and TEOA as an electron donor. After 2 h reaction in pure CO2, Ni MOLs had a CO production rate of 12.5 mmol/h and a H2 production rate of 0.28 mmol/h, whereas bulk Ni MOFs had a lower CO production rate of 7.23 mmol/h. In diluted CO2, Ni MOLs had a CO selectivity of 96.8 %, which outperformed most of the reported systems in diluted CO2. It was proposed that the strong affinity of Ni MOLs to CO2 molecules enabled their high CO2 adsorption ability and stabilized the initial Ni‐CO2 adducts, thus promoting CO2-to-CO conversion. In addition, weak affinity of Ni MOLs to H2O impeded the transfer of protons, thereby reducing the H2 formation.

Conclusions and outlook

In recent years, MOFs have attracted great attention and showed much potential as photocatalysts for CO2 reduction because of their super-high surface areas, tunable structures, and high CO2 adsorption capacity. This review summarizes the recent research progress in the development of MOFs and MOF-based composite photocatalysts for the photocatalytic reduction of CO2. Several strategies in improving light harvesting, CO2 adsorption and charge separation have been discussed, which provides guidelines for rational design of MOF-based photocatalysts with enhanced performance on CO2 reduction under visible light irradiation. Although great progress has been made in the development of MOF-based photocatalysts for CO2 reduction, there still exist some challenges and large potential need to be explored. For instance, ① to date several thousands of MOFs with different structures have been developed, however, only several types of MOFs such as ZIFs, UiO- and MIL-based MOFs are under exploration as photocatalysts for CO2 reduction. More efforts need to be made in investigation of many other MOF systems with active catalytic centers, which may be promising for photocatalysis. ② Most MOFs have a poor stability in aqueous solution and suffer structural collapse, which limit their practical applications in catalysis processes where water is involved. Therefore, it is imperative to develop novel MOF photocatalysts which have an excellent chemical stability in aqueous solution and meanwhile retain the photocatalytic functionality. ③ The current research achievements in photocatalytic reduction of CO2 over MOF-based photocatalysts have not yet met the requirement for large-scale industrial applications, therefore, exploration in further improvement of CO2 photoreduction is required. So far, many novel materials such as two-dimensional (2D) nanomaterials like graphene and 2D semiconductors have been developed; integrating these novel nanomaterials into MOF structures may be considered as a promising strategy to enhance the photocatalytic performance of MOFs. It is believed that with the rapid development progress of MOF materials and other novel photocatalytic materials, MOF-based photocatalysts will have even greater potential to fulfill the requirements for practical applications in heterogeneous photocatalysis in the future.

Acknowledgments

This work was supported by a startup fund from the University of Alaska Fairbanks.
1
Pearson P N, Palmer M R. Atmospheric carbon dioxide concentrations over the past 60 million years. Nature, 2000, 406(6797): 406695

DOI

2
Quadrelli R, Peterson S. The energy–climate challenge: recent trends in CO2 emissions from fuel combustion. Energy Policy, 2007, 35(11): 5938–5952

DOI

3
Song C. CO2 conversion and utilization: an overview. In: Song C, eds. CO2 Conversion and Utilization. Washington, DC: ACS Symposium Series, 2002, 809, 2–30

DOI

4
Herzog H J, Drake E M. Carbon dioxide recovery and disposal from large energy systems. Annual Review of Energy and the Environment, 1996, 21(1): 145–166

DOI

5
Muradov N. Industrial Utilization of CO2: A Win–Win Solution. New York: Springer New York, 2014, 325–383

6
Rafiee A, Rajab Khalilpour K, Milani D, Panahi M. Trends in CO2 conversion and utilization: a review from process systems perspective. Journal of Environmental Chemical Engineering, 2018, 6(5): 5771–5794

DOI

7
Wang B, Chen W, Song Y, Li G, Wei W, Fang J, Sun Y. Recent progress in the photocatalytic reduction of aqueous carbon dioxide. Catalysis Today, 2018, 311: 23–39

DOI

8
Yu Y, Zheng W, Cao Y. TiO2–Pd/C composited photocatalyst with improved photocatalytic activity for photoreduction of CO2 into CH4. New Journal of Chemistry, 2017, 41(8): 3204–3210

DOI

9
Sneddon G, Greenaway A, Yiu H H P. The potential applications of nanoporous materials for the adsorption, separation, and catalytic conversion of carbon dioxide. Advanced Energy Materials, 2014, 4(10): 1301873

DOI

10
North M, Pasquale R, Young C. Synthesis of cyclic carbonates from epoxides and CO2. Green Chemistry, 2010, 12(9): 1514–1539

DOI

11
Li W, Wang H, Jiang X, Zhu J, Liu Z, Guo X, Song C. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RSC Advances, 2018, 8(14): 7651–7669

DOI

12
Raciti D, Wang C. Recent advances in CO2 reduction electrocatalysis on copper. ACS Energy Letters, 2018, 3(7): 1545–1556

DOI

13
Tahir M, Amin N S. Advances in visible light responsive titanium oxide-based photocatalysts for CO2 conversion to hydrocarbon fuels. Energy Conversion and Management, 2013, 76: 194–214

DOI

14
Matsubara Y, Grills D C, Kuwahara Y. Thermodynamic aspects of electrocatalytic CO2 reduction in acetonitrile and with an ionic liquid as solvent or electrolyte. ACS Catalysis, 2015, 5(11): 6440–6452

DOI

15
Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature, 1972, 238(5358): 23837

DOI

16
Wang M, Ioccozia J, Sun L, Lin C, Lin Z. Inorganic-modified semiconductor TiO2 nanotube arrays for photocatalysis. Energy & Environmental Science, 2014, 7(7): 2182–2202

DOI

17
Bao N, Shen L, Takata T, Domen K. Self-templated synthesis of nanoporous CdS nanostructures for highly efficient photocatalytic hydrogen production under visible light. Chemistry of Materials, 2008, 20(1): 110–117

DOI

18
Ong C B, Ng L Y, Mohammad A W. A review of ZnO nanoparticles as solar photocatalysts: synthesis, mechanisms and applications. Renewable & Sustainable Energy Reviews, 2018, 81: 536–551

DOI

19
Lee G J, Wu J J. Recent developments in ZnS photocatalysts from synthesis to photocatalytic applications—a review. Powder Technology, 2017, 318: 8–22

DOI

20
Mishra M, Chun D M. α-Fe2O3 as a photocatalytic material: a review. Applied Catalysis A, General, 2015, 498: 126–141

DOI

21
Wen J, Xie J, Chen X, Li X. A review on g-C3N4-based photocatalysts. Applied Surface Science, 2017, 391: 72–123

DOI

22
Luo L, Li Y, Hou J, Yang Y. Visible photocatalysis and photostability of Ag3PO4 photocatalyst. Applied Surface Science, 2014, 319: 332–338

DOI

23
Dong C, Lian C, Hu S, Deng Z, Gong J, Li M, Liu H, Xing M, Zhang J. Size-dependent activity and selectivity of carbon dioxide photocatalytic reduction over platinum nanoparticles. Nature Communications, 2018, 9(1): 1252

DOI

24
Xing M, Zhou Y, Dong C, Cai L, Zeng L, Shen B, Pan L, Dong C, Chai Y, Zhang J, Yin Y. Modulation of the reduction potential of TiO2–x by fluorination for efficient and selective CH4 generation from CO2 photoreduction. Nano Letters, 2018, 18(6): 3384–3390

DOI

25
Zhang H, Liu G, Shi L, Liu H, Wang T, Ye J. Engineering coordination polymers for photocatalysis. Nano Energy, 2016, 22: 149–168

DOI

26
Meissner D, Memming R, Kastening B. Photoelectrochemistry of cadmium sulfide. 1. Reanalysis of photocorrosion and flat-band potential. Journal of Physical Chemistry, 1988, 92(12): 3476–3483

DOI

27
Bahnemann D W, Kormann C, Hoffmann M R. Preparation and characterization of quantum size zinc oxide: a detailed spectroscopic study. Journal of Physical Chemistry, 1987, 91(14): 3789–3798

DOI

28
Zhang L, Hu Y H. Desorption of dimethylformamide from Zn4O(C8H4O4)3 framework. Applied Surface Science, 2011, 257(8): 3392–3398

DOI

29
Hu Y H, Zhang L. Amorphization of metal-organic framework MOF-5 at unusually low applied pressure. Physical Review. B, 2010, 81(17): 174103

DOI

30
Zhang L, Hu Y H. A systematic investigation of decomposition of nano Zn4O(C8H4O4)3 metal-organic framework. Journal of Physical Chemistry C, 2010, 114(6): 2566–2572

DOI

31
Zhang L, Hu Y H. Strong effects of higher-valent cations on the structure of the zeolitic Zn(2-methylimidazole)2 framework (ZIF-8). Journal of Physical Chemistry C, 2011, 115(16): 7967–7971

DOI

32
Zhang L, Hu Y H. Structure distortion of Zn4O13C24H12 framework (MOF-5). Materials Science and Engineering B, 2011, 176(7): 573–578

DOI

33
Zhang L, Hu Y H. Observation of ZnO nanoparticles outside pores of nano Zn4O(C8H4O4)3 metal–organic framework. Physics Letters [Part A], 2011, 375(13): 1514–1517

DOI

34
Loera-Serna S, Zarate-Rubio J, Medina-Velazquez D Y, Zhang L, Ortiz E. Encapsulation of urea and caffeine in Cu3(BTC)2 metal–organic framework. Surface Innovations, 2016, 4(2): 76–87

DOI

35
Hu Y H, Zhang L. Hydrogen storage in metal–organic frameworks. Advanced Materials, 2010, 22(20): E117–E130

DOI

36
Zhang T, Lin W. Metal–organic frameworks for artificial photosynthesis and photocatalysis. Chemical Society Reviews, 2014, 43(16): 5982–5993

DOI

37
Wu M X, Yang Y W. Metal–organic framework (MOF)-based drug/cargo delivery and cancer therapy. Advanced Materials, 2017, 29(23): 1606134

DOI

38
Chowdhury T, Zhang L, Zhang J, Aggarwal S. Removal of arsenic(III) from aqueous solution using metal organic framework-graphene oxide nanocomposite. Nanomaterials (Basel, Switzerland), 2018, 8(12): 1062

DOI

39
Kreno L E, Leong K, Farha O K, Allendorf M, Van Duyne R P, Hupp J T. Metal–organic framework materials as chemical sensors. Chemical Reviews, 2012, 112(2): 1105–1125

DOI

40
Wang Y, Huang N Y, Shen J Q, Liao P Q, Chen X M, Zhang J P. Hydroxide ligands cooperate with catalytic centers in metal–organic frameworks for efficient photocatalytic CO2 reduction. Journal of the American Chemical Society, 2018, 140(1): 38–41

DOI

41
He J, Zhang Y, He J, Zeng X, Hou X, Long Z. Enhancement of photoredox catalytic properties of porphyrinic metal–organic frameworks based on titanium incorporation via post-synthetic modification. Chemical Communications, 2018, 54(62): 8610–8613

DOI

42
Horiuchi Y, Toyao T, Saito M, Mochizuki K, Iwata M, Higashimura H, Anpo M, Matsuoka M. Visible-light-promoted photocatalytic hydrogen production by using an amino-functionalized Ti(IV) metal–organic framework. Journal of Physical Chemistry C, 2012, 116(39): 20848–20853

DOI

43
Maina J W, Pozo-Gonzalo C, Kong L, Schütz J, Hill M, Dumée L F. Metal organic framework based catalysts for CO2 conversion. Materials Horizons, 2017, 4(3): 345–361

DOI

44
Nasalevich M A, Goesten M G, Savenije T J, Kapteijn F, Gascon J. Enhancing optical absorption of metal–organic frameworks for improved visible light photocatalysis. Chemical Communications, 2013, 49(90): 10575–10577

DOI

45
Jiang D, Mallat T, Krumeich F, Baiker A. Copper-based metal-organic framework for the facile ring-opening of epoxides. Journal of Catalysis, 2008, 257(2): 390–395

DOI

46
Hasegawa S, Horike S, Matsuda R, Furukawa S, Mochizuki K, Kinoshita Y, Kitagawa S. Three-dimensional porous coordination polymer functionalized with amide groups based on tridentate ligand: selective sorption and catalysis. Journal of the American Chemical Society, 2007, 129(9): 2607–2614

DOI

47
Wang J L, Wang C, Lin W. Metal–organic frameworks for light harvesting and photocatalysis. ACS Catalysis, 2012, 2(12): 2630–2640

DOI

48
Llabrés i Xamena F X, Casanova O, Galiasso Tailleur R, Garcia H, Corma A. Metal organic frameworks (MOFs) as catalysts: a combination of Cu2+ and Co2+ MOFs as an efficient catalyst for tetralin oxidation. Journal of Catalysis, 2008, 255(2): 220–227

DOI

49
Llabrés i Xamena F X, Corma A, Garcia H. Applications for metal-organic frameworks (MOFs) as quantum dot semiconductors. Journal of Physical Chemistry C, 2007, 111(1): 80–85

DOI

50
Gao J, Miao J, Li P Z, Teng W Y, Yang L, Zhao Y, Liu B, Zhang Q. A p-type Ti(iv)-based metal–organic framework with visible-light photo-response. Chemical Communications, 2014, 50(29): 3786–3788

DOI

51
Shen L, Liang S, Wu W, Liang R, Wu L. CdS-decorated UiO-66(NH2) nanocomposites fabricated by a facile photodeposition process: an efficient and stable visible-light-driven photocatalyst for selective oxidation of alcohols. Journal of Materials Chemistry. A, 2013, 1(37): 11473–11482

DOI

52
Ryu U J, Kim S J, Lim H K, Kim H, Choi K M, Kang J K. Synergistic interaction of Re complex and amine functionalized multiple ligands in metal-organic frameworks for conversion of carbon dioxide. Scientific Reports, 2017, 7(1): 612

DOI

53
Fu Y, Sun D, Chen Y, Huang R, Ding Z, Fu X, Li Z. An amine-functionalized titanium metal–organic framework photocatalyst with visible-light-induced activity for CO2 reduction. Angewandte Chemie International Edition, 2012, 51(14): 3364–3367

DOI

54
Yan S, Ouyang S, Xu H, Zhao M, Zhang X, Ye J. Co-ZIF-9/TiO2 nanostructure for superior CO2 photoreduction activity. Journal of Materials Chemistry. A, 2016, 4(39): 15126–15133

DOI

55
Su Y, Zhang Z, Liu H, Wang Y. Cd0.2Zn0.8S@UiO-66–NH2 nanocomposites as efficient and stable visible-light-driven photocatalyst for H2 evolution and CO2 reduction. Applied Catalysis B: Environmental, 2017, 200: 448–457

DOI

56
Liu S, Chen F, Li S, Peng X, Xiong Y. Enhanced photocatalytic conversion of greenhouse gas CO2 into solar fuels over g-C3N4 nanotubes with decorated transparent ZIF-8 nanoclusters. Applied Catalysis B: Environmental, 2017, 211: 1–10

DOI

57
Wang S, Yao W, Lin J, Ding Z, Wang X. Cobalt imidazolate metal–organic frameworks photosplit CO2 under mild reaction conditions. Angewandte Chemie International Edition, 2014, 53(4): 1034–1038

DOI

58
Fei H, Sampson M D, Lee Y, Kubiak C P, Cohen S M. Photocatalytic CO2 reduction to formate using a Mn(I) molecular catalyst in a robust metal–organic framework. Inorganic Chemistry, 2015, 54(14): 6821–6828

DOI

59
Qin J, Wang S, Wang X. Visible-light reduction CO2 with dodecahedral zeolitic imidazolate framework ZIF-67 as an efficient co-catalyst. Applied Catalysis B: Environmental, 2017, 209: 476–482

DOI

60
Huang Y B, Liang J, Wang X S, Cao R. Multifunctional metal–organic framework catalysts: synergistic catalysis and tandem reactions. Chemical Society Reviews, 2017, 46(1): 126–157

DOI

61
Wang S, Wang X. Multifunctional metal–organic frameworks for photocatalysis. Small, 2015, 11(26): 3097–3112

DOI

62
Yu X, Wang L, Cohen S M. Photocatalytic metal–organic frameworks for organic transformations. CrystEngComm, 2017, 19(29): 4126–4136

DOI

63
Navarro Amador R, Carboni M, Meyer D. Photosensitive titanium and zirconium metal organic frameworks: current research and future possibilities. Materials Letters, 2016, 166: 327–338

DOI

64
Liang Z, Qu C, Guo W, Zou R, Xu Q. Pristine metal–organic frameworks and their composites for energy storage and conversion. Advanced Materials, 2018, 30(37): 1702891

DOI

65
Sun D, Li Z. Robust Ti- and Zr-based metal-organic frameworks for photocatalysis. Chinese Journal of Chemistry, 2017, 35(2): 135–147

DOI

66
Shen L, Liang R, Wu L. Strategies for engineering metal-organic frameworks as efficient photocatalysts. Chinese Journal of Catalysis, 2015, 36(12): 2071–2088

DOI

67
Zhu J, Li P Z, Guo W, Zhao Y, Zou R. Titanium-based metal–organic frameworks for photocatalytic applications. Coordination Chemistry Reviews, 2018, 359: 80–101

DOI

68
Santaclara J G, Kapteijn F, Gascon J, van der Veen M A. Understanding metal–organic frameworks for photocatalytic solar fuel production. CrystEngComm, 2017, 19(29): 4118–4125

DOI

69
Nasalevich M A, van der Veen M, Kapteijn F, Gascon J. Metal–organic frameworks as heterogeneous photocatalysts: advantages and challenges. CrystEngComm, 2014, 16(23): 4919–4926

DOI

70
Song F, Li W, Sun Y. Metal–organic frameworks and their derivatives for photocatalytic water splitting. Inorganics, 2017, 5(3): 40

DOI

71
Wang W, Xu X, Zhou W, Shao Z. Recent progress in metal-organic frameworks for applications in electrocatalytic and photocatalytic water splitting. Advancement of Science, 2017, 4(4): 1600371

DOI

72
Yan Y, He T, Zhao B, Qi K, Liu H, Xia B Y. Metal/covalent–organic frameworks-based electrocatalysts for water splitting. Journal of Materials Chemistry. A, 2018, 6(33): 15905–15926

DOI

73
Meyer K, Ranocchiari M, van Bokhoven J A. Metal organic frameworks for photo-catalytic water splitting. Energy & Environmental Science, 2015, 8(7): 1923–1937

DOI

74
Pi Y, Li X, Xia Q, Wu J, Li Y, Xiao J, Li Z. Adsorptive and photocatalytic removal of persistent organic pollutants (POPs) in water by metal-organic frameworks (MOFs). Chemical Engineering Journal, 2018, 337: 351–371

DOI

75
Wu Z, Yuan X, Zhang J, Wang H, Jiang L, Zeng G. Photocatalytic decontamination of wastewater containing organic dyes by metal–organic frameworks and their derivatives. ChemCatChem, 2017, 9(1): 41–64

DOI

76
Wang C C, Li J R, Lv X L, Zhang Y Q, Guo G. Photocatalytic organic pollutants degradation in metal–organic frameworks. Energy & Environmental Science, 2014, 7(9): 2831–2867

DOI

77
Jiang D, Xu P, Wang H, Zeng G, Huang D, Chen M, Lai C, Zhang C, Wan J, Xue W. Strategies to improve metal organic frameworks photocatalyst’s performance for degradation of organic pollutants. Coordination Chemistry Reviews, 2018, 376: 449–466

DOI

78
Dhakshinamoorthy A, Li Z, Garcia H. Catalysis and photocatalysis by metal organic frameworks. Chemical Society Reviews, 2018, 47(22): 8134–8172

DOI

79
Zhu B, Zou R, Xu Q. Metal–organic framework based catalysts for hydrogen evolution. Advanced Energy Materials, 2018, 8(24): 1801193

DOI

80
Fang Y, Ma Y, Zheng M, Yang P, Asiri A M, Wang X. Metal–organic frameworks for solar energy conversion by photoredox catalysis. Coordination Chemistry Reviews, 2018, 373: 83–115

DOI

81
Chen Y, Wang D, Deng X, Li Z. Metal–organic frameworks (MOFs) for photocatalytic CO2 reduction. Catalysis Science & Technology, 2017, 7(21): 4893–4904

DOI

82
Li Y, Xu H, Ouyang S, Ye J. Metal–organic frameworks for photocatalysis. Physical Chemistry Chemical Physics, 2016, 18(11): 7563–7572

DOI

83
Li R, Zhang W, Zhou K. Metal–organic-framework-based catalysts for photoreduction of CO2. Advanced Materials, 2018, 30(35): 1705512

DOI

84
Wang C C, Zhang Y Q, Li J, Wang P. Photocatalytic CO2 reduction in metal–organic frameworks: a mini review. Journal of Molecular Structure, 2015, 1083: 127–136

DOI

85
Qiu J, Zhang X, Feng Y, Zhang X, Wang H, Yao J. Modified metal-organic frameworks as photocatalysts. Applied Catalysis B: Environmental, 2018, 231: 317–342

DOI

86
Gascon J, Hernández-Alonso M D, Almeida A R, van Klink G P M, Kapteijn F, Mul G. Isoreticular MOFs as efficient photocatalysts with tunable band gap: an operando FTIR study of the photoinduced oxidation of propylene. ChemSusChem, 2008, 1(12): 981–983

DOI

87
Barkhordarian A A, Kepert C J. Two new porous UiO-66-type zirconium frameworks: open aromatic N-donor sites and their post-synthetic methylation and metallation. Journal of Materials Chemistry. A, 2017, 5(11): 5612–5618

DOI

88
Hendon C H, Tiana D, Fontecave M, Sanchez C, D’arras L, Sassoye C, Rozes L, Mellot-Draznieks C, Walsh A. Engineering the optical response of the Titanium-MIL-125 metal–organic framework through ligand functionalization. Journal of the American Chemical Society, 2013, 135(30): 10942–10945

DOI

89
Pham H Q, Mai T, Pham-Tran N N, Kawazoe Y, Mizuseki H, Nguyen-Manh D. Engineering of band gap in metal–organic frameworks by functionalizing organic linker: a systematic density functional theory investigation. Journal of Physical Chemistry C, 2014, 118(9): 4567–4577

DOI

90
Yang H, He X W, Wang F, Kang Y, Zhang J. Doping copper into ZIF-67 for enhancing gas uptake capacity and visible-light-driven photocatalytic degradation of organic dye. Journal of Materials Chemistry, 2012, 22(41): 21849–21851

DOI

91
Yang L M, Fang G Y, Ma J, Pushpa R, Ganz E. Halogenated MOF-5 variants show new configuration, tunable band gaps and enhanced optical response in the visible and near infrared. Physical Chemistry Chemical Physics, 2016, 18(47): 32319–32330

DOI

92
Nguyen H L, Vu T T, Le D, Doan T L H, Nguyen V Q, Phan N T S. A Titanium–organic framework: engineering of the band-gap energy for photocatalytic property enhancement. ACS Catalysis, 2017, 7(1): 338–342

DOI

93
Sun D, Fu Y, Liu W, Ye L, Wang D, Yang L, Fu X, Li Z. Studies on photocatalytic CO2 reduction over NH2-Uio-66(Zr) and its derivatives: towards a better understanding of photocatalysis on metal–organic frameworks. Chemistry–A European Journal, 2013, 19(42): 14279–14285

DOI

94
Lee Y, Kim S, Kang J K, Cohen S M. Photocatalytic CO2 reduction by a mixed metal (Zr/Ti), mixed ligand metal–organic framework under visible light irradiation. Chemical Communications, 2015, 51(26): 5735–5738

DOI

95
Wang D, Huang R, Liu W, Sun D, Li Z. Fe-based MOFs for photocatalytic CO2 reduction: role of coordination unsaturated sites and dual excitation pathways. ACS Catalysis, 2014, 4(12): 4254–4260

DOI

96
Sun D, Liu W, Qiu M, Zhang Y, Li Z. Introduction of a mediator for enhancing photocatalytic performance via post-synthetic metal exchange in metal–organic frameworks (MOFs). Chemical Communications, 2015, 51(11): 2056–2059

DOI

97
Liu J, Fan Y Z, Li X, Wei Z, Xu Y W, Zhang L, Su C Y. A porous rhodium(III)-porphyrin metal-organic framework as an efficient and selective photocatalyst for CO2 reduction. Applied Catalysis B: Environmental, 2018, 231: 173–181

DOI

98
Sadeghi N, Sharifnia S, Sheikh Arabi M.A porphyrin-based metal organic framework for high rate photoreduction of CO2 to CH4 in gas phase. Journal of CO2 Utilization, 2016, 16: 450–457

99
Liu Y, Yang Y, Sun Q, Wang Z, Huang B, Dai Y, Qin X, Zhang X. Chemical adsorption enhanced CO2 capture and photoreduction over a copper porphyrin based metal organic framework. ACS Applied Materials & Interfaces, 2013, 5(15): 7654–7658

DOI

100
Zhang H, Wei J, Dong J, Liu G, Shi L, An P, Zhao G, Kong J, Wang X, Meng X, Zhang J, Ye J. Efficient visible-light-driven carbon dioxide reduction by a single-atom implanted metal–organic framework. Angewandte Chemie International Edition, 2016, 55(46): 14310–14314

DOI

101
Xu H Q, Hu J, Wang D, Li Z, Zhang Q, Luo Y, Yu S H, Jiang H L. Visible-light photoreduction of CO2 in a metal–organic framework: boosting electron–hole separation via electron trap states. Journal of the American Chemical Society, 2015, 137(42): 13440–13443

DOI

102
Yan Z H, Du M H, Liu J, Jin S, Wang C, Zhuang G L, Kong X J, Long L S, Zheng L S. Photo-generated dinuclear {Eu(II)}2 active sites for selective CO2 reduction in a photosensitizing metal-organic framework. Nature Communications, 2018, 9(1): 3353

DOI

103
Wang C, Xie Z, deKrafft K E, Lin W. Doping metal–organic frameworks for water oxidation, carbon dioxide reduction, and organic photocatalysis. Journal of the American Chemical Society, 2011, 133(34): 13445–13454

DOI

104
Huang R, Peng Y, Wang C, Shi Z, Lin W. A rhenium-functionalized metal–organic framework as a single-site catalyst for photochemical reduction of carbon dioxide. European Journal of Inorganic Chemistry, 2016, 2016(27): 4358–4362

DOI

105
Chambers M B, Wang X, Elgrishi N, Hendon C H, Walsh A, Bonnefoy J, Canivet J, Quadrelli E A, Farrusseng D, Mellot-Draznieks C, Fontecave M. Photocatalytic carbon dioxide reduction with rhodium-based catalysts in solution and heterogenized within metal–organic frameworks. ChemSusChem, 2015, 8(4): 603–608

DOI

106
Sun D, Gao Y, Fu J, Zeng X, Chen Z, Li Z. Construction of a supported Ru complex on bifunctional MOF-253 for photocatalytic CO2 reduction under visible light. Chemical Communications, 2015, 51(13): 2645–2648

DOI

107
Li L, Zhang S, Xu L, Wang J, Shi L X, Chen Z N, Hong M, Luo J. Effective visible-light driven CO2 photoreduction via a promising bifunctional iridium coordination polymer. Chemical Science (Cambridge), 2014, 5(10): 3808–3813

DOI

108
Zhang S, Li L, Zhao S, Sun Z, Hong M, Luo J. Hierarchical metal–organic framework nanoflowers for effective CO2 transformation driven by visible light. Journal of Materials Chemistry. A, 2015, 3(30): 15764–15768

DOI

109
Zhang S, Li L, Zhao S, Sun Z, Luo J. Construction of interpenetrated ruthenium metal–organic frameworks as stable photocatalysts for CO2 reduction. Inorganic Chemistry, 2015, 54(17): 8375–8379

DOI

110
Lee Y, Kim S, Fei H, Kang J K, Cohen S M. Photocatalytic CO2 reduction using visible light by metal-monocatecholato species in a metal–organic framework. Chemical Communications, 2015, 51(92): 16549–16552

DOI

111
Chen D, Xing H, Wang C, Su Z. Highly efficient visible-light-driven CO2 reduction to formate by a new anthracene-based zirconium MOF via dual catalytic routes. Journal of Materials Chemistry. A, 2016, 4(7): 2657–2662

DOI

112
Schaate A, Roy P, Godt A, Lippke J, Waltz F, Wiebcke M, Behrens P. Modulated synthesis of Zr-based metal–organic frameworks: from nano to single crystals. Chemistry–A European Journal, 2011, 17(24): 6643–6651

DOI

113
Gomes  Silva C, Luz I, Llabrés i Xamena F X, Corma A, García H. Water stable Zr–benzenedicarboxylate metal–organic frameworks as photocatalysts for hydrogen generation. Chemistry– A European Journal, 2010, 16(36): 11133–11138

DOI

114
Cavka J H, Jakobsen S, Olsbye U, Guillou N, Lamberti C, Bordiga S, Lillerud K P. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. Journal of the American Chemical Society, 2008, 130(42): 13850–13851

DOI

115
Mondloch J E, Katz M J, Planas N, Semrouni D, Gagliardi L, Hupp J T, Farha O K. Are Zr6-based MOFs water stable? Linker hydrolysis vs. capillary-force-driven channel collapse. Chemical Communications, 2014, 50(64): 8944–8946

DOI

116
Wang C C, Du X D, Li J, Guo X X, Wang P, Zhang J. Photocatalytic Cr(VI) reduction in metal-organic frameworks: a mini-review. Applied Catalysis B: Environmental, 2016, 193: 198–216

DOI

117
Dean J A. Lange’s handbook of chemistry. Materials and Manufacturing Processes, 1990, 5(4): 687–688

DOI

118
Laurier K G M, Vermoortele F, Ameloot R, De Vos D E, Hofkens J, Roeffaers M B J. Iron(III)-based metal–organic frameworks as visible light photocatalysts. Journal of the American Chemical Society, 2013, 135(39): 14488–14491

DOI

119
Torrisi A, Bell R G, Mellot-Draznieks C. Functionalized MOFs for enhanced CO2 capture. Crystal Growth & Design, 2010, 10(7): 2839–2841

DOI

120
Torrisi A, Mellot-Draznieks C, Bell R G. Impact of ligands on CO2 adsorption in metal-organic frameworks: first principles study of the interaction of CO2 with functionalized benzenes. II. Effect of polar and acidic substituents. Journal of Chemical Physics, 2010, 132(4): 044705

DOI

121
Tamaki Y, Morimoto T, Koike K, Ishitani O. Photocatalytic CO2 reduction with high turnover frequency and selectivity of formic acid formation using Ru(II) multinuclear complexes. Proceedings of the National Academy of Sciences of the United States of America, 2012, 109(39): 15673–15678

DOI

122
Sato S, Morikawa T, Kajino T, Ishitani O. A highly efficient mononuclear iridium complex photocatalyst for CO2 reduction under visible light. Angewandte Chemie International Edition, 2013, 52(3): 988–992

DOI

123
Kuramochi Y, Kamiya M, Ishida H. Photocatalytic CO2 reduction in N,N-dimethylacetamide/water as an alternative solvent system. Inorganic Chemistry, 2014, 53(7): 3326–3332

DOI

124
Huang Z, Dong P, Zhang Y, Nie X, Wang X, Zhang X.A ZIF-8 decorated TiO2 grid-like film with high CO2 adsorption for CO2 photoreduction. Journal of CO2 Utilization, 2018, 24, 369–375

125
Cardoso J C, Stulp S, de Brito J F, Flor J B S, Frem R C G, Zanoni M V B. MOFs based on ZIF-8 deposited on TiO2 nanotubes increase the surface adsorption of CO2 and its photoelectrocatalytic reduction to alcohols in aqueous media. Applied Catalysis B: Environmental, 2018, 225: 563–573

DOI

126
Li R, Hu J, Deng M, Wang H, Wang X, Hu Y, Jiang H L, Jiang J, Zhang Q, Xie Y, Xiong Y. Integration of an inorganic semiconductor with a metal–organic framework: a platform for enhanced gaseous photocatalytic reactions. Advanced Materials, 2014, 26(28): 4783–4788

DOI

127
He X, Gan Z, Fisenko S, Wang D, El-Kaderi H M, Wang W N. Rapid formation of metal–organic frameworks (MOFs) based nanocomposites in microdroplets and their applications for CO2 photoreduction. ACS Applied Materials & Interfaces, 2017, 9(11): 9688–9698

DOI

128
Wang M, Wang D, Li Z. Self-assembly of CPO-27-Mg/TiO2 nanocomposite with enhanced performance for photocatalytic CO2 reduction. Applied Catalysis B: Environmental, 2016, 183: 47–52

DOI

129
Crake A, Christoforidis K C, Kafizas A, Zafeiratos S, Petit C. CO2 capture and photocatalytic reduction using bifunctional TiO2/MOF nanocomposites under UV–vis irradiation. Applied Catalysis B: Environmental, 2017, 210: 131–140

DOI

130
Wang S, Wang X. Photocatalytic CO2 reduction by CdS promoted with a zeolitic imidazolate framework. Applied Catalysis B: Environmental, 2015, 162: 494–500

DOI

131
Wang S, Lin J, Wang X. Semiconductor–redox catalysis promoted by metal–organic frameworks for CO2 reduction. Physical Chemistry Chemical Physics, 2014, 16(28): 14656–14660

DOI

132
Shi L, Wang T, Zhang H, Chang K, Ye J. Electrostatic self-assembly of nanosized carbon nitride nanosheet onto a zirconium metal–organic framework for enhanced photocatalytic CO2 reduction. Advanced Functional Materials, 2015, 25(33): 5360–5367

DOI

133
Xu G, Zhang H, Wei J, Zhang H X, Wu X, Li Y, Li C, Zhang J, Ye J. Integrating the g-C3N4 Nanosheet with B–H bonding decorated metal–organic framework for CO2 activation and photoreduction. ACS Nano, 2018, 12(6): 5333–5340

DOI

134
Liu Q, Low Z X, Li L, Razmjou A, Wang K, Yao J, Wang H. ZIF-8/Zn2GeO4 nanorods with an enhanced CO2 adsorption property in an aqueous medium for photocatalytic synthesis of liquid fuel. Journal of Materials Chemistry. A, 2013, 1(38): 11563–11569

DOI

135
Sun D, Liu W, Fu Y, Fang Z, Sun F, Fu X, Zhang Y, Li Z. Noble metals can have different effects on photocatalysis over metal–organic frameworks (MOFs): a case study on M/NH2-MIL-125(Ti) (M=Pt and Au). Chemistry–A European Journal, 2014, 20(16): 4780–4788

DOI

136
Fu Y, Yang H, Du R, Tu G, Xu C, Zhang F, Fan M, Zhu W. Enhanced photocatalytic CO2 reduction over Co-doped NH2-MIL-125(Ti) under visible light. RSC Advances, 2017, 7(68): 42819–42825

DOI

137
Choi K M, Kim D, Rungtaweevoranit B, Trickett C A, Barmanbek J T D, Alshammari A S, Yang P, Yaghi O M. Plasmon-enhanced photocatalytic CO2 conversion within metal–organic frameworks under visible light. Journal of the American Chemical Society, 2017, 139(1): 356–362

DOI

138
Wang X, Zhao X, Zhang D, Li G, Li H. Microwave irradiation induced UIO-66–NH2 anchored on graphene with high activity for photocatalytic reduction of CO2. Applied Catalysis B: Environmental, 2018, 228: 47–53

DOI

139
Sadeghi N, Sharifnia S, Do T O. Enhanced CO2 photoreduction by a graphene–porphyrin metal–organic framework under visible light irradiation. Journal of Materials Chemistry. A, 2018, 6(37): 18031–18035

DOI

140
Pipelzadeh E, Rudolph V, Hanson G, Noble C, Wang L. Photoreduction of CO2 on ZIF-8/TiO2 nanocomposites in a gaseous photoreactor under pressure swing. Applied Catalysis B: Environmental, 2017, 218: 672–678

DOI

141
Chaudhary Y S, Woolerton T W, Allen C S, Warner J H, Pierce E, Ragsdale S W, Armstrong F A. Visible light-driven CO2 reduction by enzyme coupled CdS nanocrystals. Chemical Communications, 2012, 48(1): 58–60

DOI

142
Liu B J, Torimoto T, Yoneyama H. Photocatalytic reduction of CO2 using surface-modified CdS photocatalysts in organic solvents. Journal of Photochemistry and Photobiology A: Chemistry, 1998, 113(1): 93–97

DOI

143
Fujiwara H, Hosokawa H, Murakoshi K, Wada Y, Yanagida S, Okada T, Kobayashi H. Effect of surface structures on photocatalytic CO2 reduction using quantized CdS nanocrystallites. Journal of Physical Chemistry B, 1997, 101(41): 8270–8278

DOI

144
Nguyen N T, Altomare M, Yoo J, Schmuki P. Efficient photocatalytic H2 evolution: controlled dewetting–dealloying to fabricate site-selective high-activity nanoporous Au particles on highly ordered TiO2 nanotube arrays. Advanced Materials, 2015, 27(20): 3208–3215

DOI

145
Bouhadoun S, Guillard C, Dapozze F, Singh S, Amans D, Bouclé J, Herlin-Boime N. One step synthesis of N-doped and Au-loaded TiO2 nanoparticles by laser pyrolysis: application in photocatalysis. Applied Catalysis B: Environmental, 2015, 174–175: 367–375

DOI

146
Wu H J, Henzie J, Lin W C, Rhodes C, Li Z, Sartorel E, Thorner J, Yang P, Groves J T. Membrane-protein binding measured with solution-phase plasmonic nanocube sensors. Nature Methods, 2012, 9(12): 91189

DOI

147
Tao A, Sinsermsuksakul P, Yang P. Polyhedral silver nanocrystals with distinct scattering signatures. Angewandte Chemie International Edition, 2006, 45(28): 4597–4601

DOI

148
Han B, Ou X, Deng Z, Song Y, Tian C, Deng H, Xu Y J, Lin Z. Nickel metal–organic framework monolayers for photoreduction of diluted CO2: metal-node-dependent activity and selectivity. Angewandte Chemie International Edition, 2018, 57(51): 16811–16815

DOI

Outlines

/