Metal organic framework based catalysts for CO2 conversion

James W. Maina *a, Cristina Pozo-Gonzalo a, Lingxue Kong a, Jürg Schütz b, Matthew Hill c and Ludovic F. Dumée *a
aDeakin University, Institute for Frontier Materials, 75 Pigdons Road, Waurn Ponds, Vic 3216, Australia. E-mail: jmaina@deakin.edu.au; Ludovic.dumee@deakin.edu.au
bCommonwealth Scientific and Industrial Research Organization (CSIRO), 75 Pigdons Road, Waurn Ponds, Vic 3216, Australia
cCommonwealth Scientific and Industrial Research Organization (CSIRO), Research Way, Clayton, Vic 3168, Australia

Received 4th November 2016 , Accepted 10th February 2017

First published on 10th February 2017


Abstract

Metal organic frameworks (MOFs) are hybrid crystalline materials, exhibiting high specific surface areas, controllable pore sizes and surface chemistry. These properties have made MOFs attractive for a wide range of applications including gas separation, gas storage, sensing, drug delivery and catalysis. This review focuses on recent progress in the application of MOF materials as catalysts for CO2 conversion through chemical fixation, photocatalysis and electrocatalysis. In particular, this review discusses the co-relationship between the physicochemical properties of MOF materials including their catalytic performance as well as their stability and recyclability under different reaction conditions, relevant to CO2 conversion. Current modification techniques for improving MOF performance are highlighted along with the recent understanding of their electronic properties. The limitations of MOF based catalysts are also discussed and potential routes for improvement are suggested.


1. Introduction

Increased emission of carbon dioxide (CO2) from the combustion of fossil fuels is the primary cause of global warming, leading to adverse climate changes and ocean acidification.1,2 Carbon capture and storage (CCS) technologies, whereby CO2 is captured from a point source or the atmosphere, prior to being stored or injected into an underground reservoir, have been proposed as potential pathways to reduce anthropogenic emissions.2–5 Although storage in porous materials such as metal organic frameworks (MOFs)6 and zeolites has also been proposed, long-term storage in underground reservoirs is currently the leading proposal, due to their large capacity for holding this greenhouse gas. However, the high costs associated with CO2 transport and injection to the geologically suitable storage sites render such a process energy-intensive and cost prohibitive.3 In addition, to ensure that the large amounts of CO2 stored underground do not leak back to the atmosphere, complex infrastructures and frequent monitoring are required, which currently have limited feasibility and require extremely long term planning.7

The utilization of CO2 as a chemical feedstock to produce valuable chemicals and fuels is an alternative sustainable approach for mitigating the adverse effect of this greenhouse gas,8,9 offering a more sustainable business model for the CCS industry. Utilization technology may be employed directly to emission sources using appropriate materials that are able to selectively capture CO2 and facilitate catalytic conversion, or maybe used in conjunction with other CO2 capturing technologies.10–13 This greenhouse gas may be converted into valuable products through chemical fixation,14,15 hydrogenation,16 photocatalysis9,17 and electrocatalysis.8 However, CO2 utilization is limited by the high chemical stability of the C[double bond, length as m-dash]O bond (bond enthalpy +805 kJ mol−1), thus requiring high energy input for bond cleavage.9 Significant efforts are therefore currently being devoted to developing catalytic materials with a high surface area, large adsorption capacity for CO2, and both high and stable catalytic activity to facilitate the energy efficient conversion of CO2.

Amongst potential candidates, metal organic frameworks (MOFs) are a class of hybrid materials composed of metal ions or clusters coordinated by organic ligands.18 These materials exhibit exceptionally high specific surface area, tunable pore size distribution and surface chemistry, which make them attractive not only in gas and liquid surface driven capture and separation applications and heterogeneous catalysis, but also in sensing and drug delivery.1,18–21 The chemical composition and microstructure of some MOF materials that have been investigated for CO2 reduction are summarized in Table 1. Over the last two decades, MOF materials have been extensively studied for selective gas adsorption, where they have been shown to outperform other materials, such as zeolites, silica and metal oxides, in terms of adsorption capacities for CO2, as reviewed in a number of articles.1,22–26 These high adsorption capacities are attributed to their high porosity and surface area, as well as the chemical interaction between CO2 molecules and metal centers or organic ligands within MOFs.1 This high affinity for CO2 chemisorption has inspired research on MOF based catalysts for CO2 fixation and conversion.27,28 The catalytic properties of MOF materials can be readily controlled by the judicious selection of metal ions and organic ligands, by modification with catalytically active functional groups,18 and by tuning the morphology to maximize the surface area and/or enhance charge separation.29

Table 1 The chemical composition and structure of some investigated MOF materials
Entry MOF Metal Organic ligand SABET (m2 g−1) Structure Catalytic cycles Ref.
1 MOF-74 Mg, Zn, Co 2,5-Dihydroxybenzene-1,4-dicarboxylic acid 393–1027 image file: c6mh00484a-u1.tif 3 38–40
2 HKUST-1 Cu 1,3,5-Benzenetricarboxylic acid 1055 image file: c6mh00484a-u2.tif 1 41 and 42
3 MOF-505 Cu 1,3-Benzendicarboxylic acid 2713 image file: c6mh00484a-u3.tif 1 43 and 44
4 MMCF-2 Cu Custom designed azamacrocycle 450 image file: c6mh00484a-u4.tif 1 43
5 Hf-NU-1000 Hf 1,3,6,8-Tetrakis(p-benzoic acid)pyrene (H4TBAPy) 1780 image file: c6mh00484a-u5.tif 1 45
6 MIL-101 Cr, Al, Fe 1,4-Benzodicarboxylic acid 4100 image file: c6mh00484a-u6.tif 3 49 and 42
7 Ru-MOF Cd, Ru 2,2′-Bipyridine-4,4′-dicarboxylate 8.08 image file: c6mh00484a-u7.tif 4 29
8 PCN-222 Zr Tetrakis(4-carboxyphenyl)-porphyrin (H2TCPP) 1728 image file: c6mh00484a-u8.tif 1 46
9 Ni-TCPE1 Ni, Tetrakis(4-carboxyphenyl)ethylene N/A image file: c6mh00484a-u9.tif 20 47
10 Ni-TCPE2 Ni Tetrakis(4-carboxyphenyl)ethylene N/A image file: c6mh00484a-u10.tif 10 47
11 ZIF-8 Zn, Co 2-Methylimidazole 1200–1400 image file: c6mh00484a-u11.tif 3 48 and 49
12 MOF-5 (IRMOF-1) Zn 1,4-Benzodicarboxylic acid 3800 image file: c6mh00484a-u12.tif 1 50 and 42
13 MIL-53 Sc, Fe, Al, Cr 1,4-Benzodicarboxylic acid 1100 image file: c6mh00484a-u13.tif 3 51 and 52
14 MIL-68 In, Ga, V 1,4-Benzodicarboxylic acid 603 image file: c6mh00484a-u14.tif 2 53–55
15 Cu4[(C57H32N12)(COO)8]n Cu Octcarboxylate 2436 image file: c6mh00484a-u15.tif 5 56
16 [Zn6(TATAB)4(DABCO)3(H2O)3]·12DMF·9H2O Zn 4,4′,4′′-s-Triazine-1,3,5-triyl-tri-p-aminobenzoic acid, and (1,4-diazabicyclo[2.2.2]-octane) 61.4 image file: c6mh00484a-u16.tif 6 57
17 UMCM-1 Zn Terephthalic acid (H2BDC), 1,3,5-tris(4-carboxyphenyl)benzene (H3TBT) 4160 image file: c6mh00484a-u17.tif 5 58
18 ZIF-68 Zn Benzimidazolate, Nim 791.13 image file: c6mh00484a-u18.tif 3 59−61


This review focuses on the recent progress in the application of MOF based catalysts in the fixation and conversion of CO2. Although there have been a number of recent reviews on this topic,17,27,30–37 there is a need for a comprehensive review to highlight the co-relationship between the physicochemical properties of MOF materials and not only their catalytic activities, but also their stability and recyclability, under different reaction conditions relevant to CO2 conversion. This review also discusses recent modification techniques for improving the catalytic performance of MOF materials, and highlights recent understanding of their electronic properties. The review discussion is divided into four sections, where the first section gives a brief introduction on the stability of MOF based catalysts. The second section presents MOF catalysts for the chemical fixation of CO2 with epoxides, while the third part covers MOF materials that have been reported to exhibit photocatalytic activity towards CO2 conversion. The fourth section reviews electrocatalytic MOF materials. Particular challenges of MOF based catalysts as well as research opportunities for further development are highlighted and critically discussed.

2. Chemical stability of MOF catalysts

Due to the potential degradation of MOF catalysts during catalytic reactions,62–64 detailed characterization to compare their properties before and after catalysis is necessary. The bond between the metal ion and the organic linker is the weakest point in the structure, since the MOF crystals may be degraded via hydrolysis or displacement of organic ligands with solvent molecules.63,64 Depending on the properties of the metal centers such as the coordination number and the hydrophobicity or chemical structure of the organic ligand, MOF materials exhibit different levels of stability in the presence of solvents.63,65,66 Metals with high coordination numbers exhibit higher stability because they can coordinate multiple ligands, which can hold the structure together even after the cleavage of some bonds.63 In addition, the strength of the metal ligand bond varies depending on the nature of metal ions. For example, the Al–O bond across MIL-53 exhibits a higher strength of ∼520 kJ mol−1 as compared to the Zn–O bond across MOF-5, which has a bond strength of 365 kJ mol−1.62,66 Consequently, a higher activation energy (180 kJ mol−1) is required for the displacement of linkers in MIL-53 by water, as compared to MOF-5 (50 kJ mol−1).62 MOF structures with a highly hydrophobic ligand, on the other hand, exhibit high stability in hydrophilic solvents, since the solvent molecules cannot access the metal centers.64

XRD analysis after the exposure to the reaction conditions may reveal details about the integrity of the crystal structure, but may not reveal partial collapse of the pore structure, or surface degradation.62 Comparison of the morphology of the MOF catalysts before and after reactions reveals any surface degradation, while a decrease in the BET surface area maybe an indication of the partial collapse or blockage of pores.62 TGA and FTIR analysis, on the other hand, could shed light on any change in chemical composition following the catalytic reactions.

3. Chemical fixation of CO2 with epoxides to form cyclic carbonates

3.1. Introduction

CO2 captured from power plants may be utilized as a raw material for the synthesis of cyclic carbonates, which have important applications as precursors in the synthesis of polymers, and as electrolytes in lithium ion batteries and aprotic solvents.15,67–69 This is accomplished through cycloaddition reaction, a process first disclosed in 1943 by Vierling et al.,70,71 where CO2 is reacted with epoxides, in the presence of a Lewis acid catalyst and a Lewis base co-catalyst, at elevated temperature or pressure, to form a cyclic carbonate. A schematic illustration for the cycloaddition reaction is presented in Scheme 1.72
image file: c6mh00484a-s1.tif
Scheme 1 Cycloaddition of CO2 to epoxides.

As recently reviewed by Martin et al. (2015)69 most catalysts for the cycloaddition reaction require high temperature (above 100 °C) and a CO2 pressure higher than 9.9 atm to achieve a sufficient yield. Among the different catalysts, homogenous catalysts, such as metal complexes of porphyrin, salen, and salphen, display the highest activity, with high turnover numbers (TONs, i.e. moles of product formed per mole of catalyst) and turnover frequency (TOF, i.e. TON per time).15,69 For example, bifunctional MgII porphyrin with 8 tetraalkylammonium bromide groups has been reported as being among the most efficient homogeneous catalysts to date, with a TON of 138[thin space (1/6-em)]000 and a TOF of 19[thin space (1/6-em)]000 h−1.73 However this impressive performance was achieved at a high CO2 pressure of 16.8 atm and a temperature of 120 °C, which may be an impediment to an economical and environmentally friendly conversion process. Another limitation of homogenous catalysts is their high solubility in the organic solvents used during cycloaddition reactions, complicating the purification procedures and recyclability.45 Although the issue of recyclability could be addressed by mobilizing the homogeneous catalyst on solid supports such as silica74 and polymers75,76 to make heterogeneous catalysts, high temperature and/or pressure was still necessary (Table S1, ESI), and their product yield and selectivity were lower due to the inhomogeneous distribution of the catalytic sites.14 Consequently, the development of more efficient and recyclable catalysts that can operate near ambient pressure and temperature is a topic of great interest and will be hereafter discussed.

3.2. MOF materials as heterogeneous catalysts for the cycloaddition of CO2 to epoxides

Owing to their crystalline nature, chemical tunability, exceptionally high surface area, and regular pore structure, MOF materials have recently gained attention as potential recyclable alternatives to homogeneous catalysts.27 Vacant or partially coordinated metal ions are readily available Lewis acid sites to activate epoxides while the organic ligand may be functionalized with Lewis basic sites or additional metal ions to enhance the catalytic activity.27

As summarized in Table 2, a broad range of MOF catalysts for cycloaddition reactions have been reported. Herein the MOFs are divided into three categories, namely (i) MOFs with active Lewis acid metal sites, (ii) MOFs with both Lewis acid and Lewis base active sites, and (iii) MOFs with active defect sites. The cycloaddition reactions may be carried out not only in the presence of solvents such as chlorobenzene and acetonitrile, but also in the absence of solvents for substrates that are in the liquid state such as propylene oxide and epichlorohydrin.77,78 The solvents play a critical role in facilitating the efficient transport of the substrates to the catalytic sites, as well as extraction of catalytic products from the catalyst surface. However, the solvents may also retard the catalytic reactions via competitive binding at the metal centers,27 and also contribute to the degradation of the MOF catalysts.63,64 MOF materials that are able to exhibit high catalytic efficiency and recyclability in the absence of a solvent are therefore of great interest.

Table 2 Application of MOFs in the cycloaddition of CO2 to epoxides
Entry MOF catalyst Pore (Å) Substrate Reaction conditions % Yield per cycle Ref.
1st 2nd 3rd
1 Mg-MOF-74 12 Styrene oxide 100 °C, 19.7 atm, 4 h 95 95 95 40
2 Co-MOF-74 12 Styrene oxide 100 °C, 19.7 atm, 4 h 96 82
3 Hf-NU-1000 13 & 29 Styrene oxide RMT, 1 atm, 56 h 100 45
4 MMCF-2 Propylene oxide RMT, 1 atm, 48 h 95.4 43
5 MOF-505 8.3 Propylene oxide RMT, 1 atm, 48 h 48 43
6 [Zn6(TATAB)4(DABCO)3(H2O)3]·12DMF·9H2O Propylene oxide 100 °C, 1 atm, 16 h 99 99 99 57
7 BIT-103 Propylene oxide 160 °C, 30 atm, 24 h 95.2 83
8 BIT-102 Propylene oxide 160 °C, 30 atm, 24 h 89.4 83
9 BIT-101 Propylene oxide 160 °C, 30 atm, 24 h 84.7 83
10 [Cd2(Ni-L)2(H2O)4]·3DMF Chloropropylene oxide 80 °C, 19.7 atm 80 84
11 Ni-TCPE1 21 Styrene oxide 100 °C, 9.9 atm, 12 h >99 97 91 47
12 Ni-TCPE2 Styrene oxide 100 °C, 9.9 atm, 12 h 86.2 47
13 In2(OH)(btc)(Hbtc)0.4(L)0.6·3H2O Propylene oxide 80 °C, 19.7 atm, 4 h 93.9 93.9 93.9 85
14 Cu4[(C57H32N12)(COO)8]n 1.11 2-Methyloxirane RMT, 1 atm, 48 h 96 96 96 56
15 Cu4[(C57H32N12)(COO)8]n 1.11 2-Ethyloxirane, RMT, 1 atm, 48 h 83 56
16 Cu4[(C57H32N12)(COO)8]n 1.11 (Chloromethyl)oxirane RMT, 1 atm, 48 h 85 56
17 Cu4[(C57H32N12)(COO)8]n 1.11 2-(Bromomethyl)oxirane RMT, 1 atm, 48 h 88 56
18 MIL-68(In) Styrene oxide 150 °C, 7.9 atm, 8 h 39 55
19 MIL-68-NH2 Styrene oxide 150 °C, 7.9 atm, 8 h 74 53 55
20 UiO-66-NH2 Styrene oxide 100 °C, 19.7 atm 96 95 95 86
21 UMCM-1-NH2 17–28 Propylene oxide RMT, 11.8 atm, 24 h 90 90 90 77
22 MIL-101-N-(n-Bu)3Br 13–27 Propylene oxide 80 °C, 19.7 atm, 8 h 98.6 98.5 42
23 MIL-101-P(n-Bu)3Br Propylene oxide 80 °C, 19.7 atm, 8 h 99.1 98.1 42
24 HKUST-1 10 Propylene oxide 80 °C, 19.7 atm, 8 h 5.4 42
25 MOF-5 4–6 Propylene oxide 80 °C, 19.7 atm, 8 h 2.5 42
26 ZIF-8 3.4 Epochlorohydrin 80 °C, 6.9 atm, 4 h 43.7 22.7 78
27 ZIF-8-NH2 Epochlorohydrin 80 °C, 6.9 atm, 4 h 73.1 30.6 78
28 ZIF-68 3.4–8.6 Styrene oxide 120 °C, 9.9 atm, 12 h 93.3 88.3 80.9 61 and 87
29 ZIF-8 3.4 Styrene oxide 80 °C, 6.9 atm, 5 h 39.4 37 37 49
30 PCN 224 (Co) 19 Propylene oxide 100 °C, 19.7 atm, 4 h 42 33 39 88
31 Cr-MIL-101 14.5 × 16 Styrene oxide 25 °C, 8 atm, 48 h 95 64 52 81 and 89
32 Cr-MIL-101 14.5 × 16 Propylene oxide 25 °C, 8 atm, 48 h 91 92 82 26 and 43
33 ZIF-22 3 Epochlorohydrin 120 °C, 11.8 atm, 2 h 99 90
34 ZIF-22 3 Propylene oxide 120 °C, 11.8 atm, 2 h 99 90
35 ZIF-67 Epichlorohydrin 100 °C, 8 atm, 8 h 98 98 91


3.2.1. MOFs with active Lewis acid metal sites. Unsaturated metal centers within MOFs serve as Lewis acid sites to activate epoxides towards cycloaddition of CO2.14 Examples of MOFs with catalytically active metal centers include, M-MOF-74, HKUST, MOF-505, Hf-NU-1000, Fe-MIL-101, and Ni-TCPE140,45,47,79–81 (Table 1, entries 1 to 9). The catalytic efficiency of Mg-MOF-74 in the cycloaddition of CO2 to styrene oxide was investigated at 100 °C and 19.7 atm of CO2.40 An optimal yield of 95% was achieved after 4 h with 30 mg of the catalyst in chlorobenzene solvent. While most catalytic systems require a Lewis base co-catalyst, there was no co-catalyst necessary for Mg-MOF-74. The oxygen atoms from the organic linkers acted as Lewis base sites, while unsaturated Mg atoms acted as Lewis acid sites. The stability of the MOF materials was also confirmed after three repeated runs which showed no decrease in catalytic activity, while powder XRD showed no change across the diffraction patterns suggesting that the crystal structure of the MOF was stable even after the catalytic reactions.

An Hf based MOF, Hf-NU-1000 (Table 1, entry 5), has also been investigated for the cycloaddition of CO2 to epoxides under mild conditions, in the presence of a tetrabutylammonium bromide co-catalyst, with acetonitrile as the solvent.45 The MOF achieved a yield of 100% in the cycloaddition of CO2 to propylene oxide, following 26 h of reaction at 1 atm pressure and at room temperature (Table 2, entry 3). This efficiency was shown to be higher than that of other MOFs such as HKUST-1, MOF-505, MMPF-9, and MMCF-2, which achieved a yield of 49, 48, 87, and 95%, respectively, under similar experimental conditions but longer reaction durations of 48 h.43 The superior performance of the Hf-NU-1000 MOF was attributed to the high density of Hf Lewis acid sites, and the relatively large pore size (13–29 Å).

Aiming at increasing the density of Lewis acid sites, Gao et al.43 designed a metal macrocycle framework (MMCF-2), by crystal engineering of MOF-505 (Table 1, entry 3). The 3,3′,5,5′-biphenyltetracarboxylate ligands, typically present in MOF-505, were substituted with a custom designed azamacrocycle ligand (1,4,7,10-tetrazazcyclododecane-N,N′,N′′,N\-tetra-p-methylbenzoic acid), to yield a new MOF (Table 1, entry 4), where each of the six faces of the cuboctahedral cage is occupied by a CuII metallated azamacrocycle, leading to 18 Cu2+ Lewis acid sites per cage, 50% higher than its parent MOF-505, which has only 12 Cu2+ sites. In addition, the Cu2+ sites of MMCF-2 are also more accessible, since they are located towards the center of the cage, as opposed to MOF-505 where the Cu2+ sites are located at the corner of the octahedral cages. These features resulted in improved catalytic efficiency with MMCF-2 having a yield of 95.4% in the conversion of propylene oxide into propylene carbonate, which was almost twice as efficient as compared to MOF-505 (48% yield), at room temperature and 1 atm of CO2, after 48 h of reaction in a solvent free environment. The yield was also higher than those of a homogenous Cu catalyst (tactmb) and a benchmark Cu based MOF, HKUST-1, which had yields of 47.5% and 49.2%, respectively.43 The recyclability and stability of the MMCF-2 MOF were, however, not reported.

Very recently, a dual wall caged MOF, having a high density of Lewis acid sites, has been reported.57 The MOF was prepared by interpenetration of two independent but similar cages, Zn24-A and Zn24-B, leading to a higher number of Zn Lewis acid sites, with 48 Zn2+ per cage (Fig. 1a–c). Consequently, the MOF exhibited high efficiency for the cycloaddition of CO2 with propylene oxide, with up to 99% yield, at 100 °C and ambient pressure after 12 h of reaction, in the absence of a solvent and a co-catalyst (Table 2, entry 6). This yield was higher than those of zinc-trimesic acid (Zn-BTC) MOFs, BIT-103, BIT-102 and BIT-101 (Table 2, entries 7–9), which had a yield of 95, 89.4 and 84.7%, respectively, despite being tested at a higher temperature of 160 °C and a pressure of 30 atm, for a duration of 24 h.83 In addition, the MOF could be recycled up to 6 times, with only a 3% decrease in yield after the 6th run (Fig. 1d). Other than the high density of Lewis catalytic sites, the good catalytic performance was also attributed to the presence of amine groups across the organic ligand, which may increase the affinity for CO2, and facilitate the activation of the CO2 molecules upon adsorption.55


image file: c6mh00484a-f1.tif
Fig. 1 (a) View of the Zn24-A cage (the distance between opposite vertices is ca. 34.4 Å). (b) View of the Zn24-B cage (the distance between opposite vertices is ca. 30.2 Å). (c) View of the dual-walled cage. (d) The catalytic cycles for the cycloaddition of CO2 to propylene epoxides using the dual-walled cage MOF. Reproduced from ref. 39 with permission from the Royal Society of Chemistry.

Another approach for increasing the density of Lewis acid sites is the use of metal–organic complexes as ligands for constructing MOFs.18,23,24 Ren et al. in 201384 used a nickel salphen complex, Ni–H2L, as a bridging metalloligand to construct MOFs with Cd metal centers. The presence of Cd2+ and Ni2+ was found to provide synergistic and additive activation effects, resulting in an enhanced catalytic activity to give 80% yield, in the synthesis of propylene carbonate in the presence of an ammonium salt co-catalyst (Table 2, entry 10). This yield was significantly higher than the 38% yield obtained when Ni–H2L was used as a homogeneous catalyst under the same experimental conditions. The MOF could also be used for three cycles, with only a 2% decrease in the product yield in the third cycle, while XRD analysis after the catalytic reactions showed no noticeable change across the diffraction patterns.

The pore size has also been shown to play an important role in governing the efficiency of MOFs in cycloaddition reactions.14,47 MOFs having relatively large pores facilitate the reaction by enabling the efficient diffusion of the reactants and products, while small pores may retard reaction by hindering diffusion. Zhou et al.47 designed single walled metal organic framework nanotubes (Ni-TCPE1), using the tetrakis(4-carboxyphenyl)ethylene (H4TCPE) ligand with the Ni metal ion (Table 1, entry 9). The Ni-TCPE1 MOF exhibited a large cross-sectional pore structure, enabling superior catalytic performance in cycloaddition of CO2 to epoxides, in the presence of a tetrabutylammonium bromide (TBABr) co-catalyst. Catalytic studies were carried out for 12 h at 373 K, 9.9 atm CO2 and in the absence of a solvent, to obtain almost complete conversion of styrene oxide into styrene carbonate (>99%) (Table 2, entry 11). Under these conditions, the MOF had a turnover number of 2000, which was more efficient than other MOFs such as Hf-Nu-1000, Nu-1000 and Cr-MIL-101, which had TONs of 25, 11.5, and 177.6, respectively, under the same experimental conditions. Furthermore, the MOF maintained its catalytic activity for over 20 cycles (70 h), achieving a TON value of 35[thin space (1/6-em)]000. XRD analysis following the catalytic reactions confirmed that the crystal structure of the MOF was maintained. The excellent performance was attributed to the large cross-section of the nanotube channels that enabled the efficient transport of the substrate and products. The high stability was attributed to the nonplanar configuration of the H4TCPE ligand, which facilitated the formation of highly connected frameworks.

A triazole containing MOF (Cu4[(C57H32N12)(COO)8]n) for the cycloaddition of CO2 to epoxides under ambient conditions has been reported.56 The cycloaddition reactions were carried out at 1 atm CO2 pressure and room temperature, for 48 h, with a tetra-n-tertbutylammonium bromide co-catalyst in the absence of a solvent. The reaction yields were found to be dependent upon the substrate size. For example, propylene oxide with a molecular weight of 58 Da leads to a product yield of 96%, while larger substrates 1,2-epoxyoctane (Mw = 128.21 Da) and 2-ethyl-hexy glycidyl ether (Mw = 186.29) lead to dramatically lower yields of 8% and 5%, respectively. The low yield implied that the larger substrate could not enter the pore framework and that the cycloaddition reactions were therefore only limited on the surface of the MOF crystals. Under similar conditions, the MOF crystals were found to be more efficient than HKUST-1, in the cycloaddition of CO2 to 2-methyloxirane, 2-ethyloxirane, 2-(chloromethyl)oxirane, and 2-(bromomethyl)oxirane, with a yield of 65, 54, 56 and 57%, respectively, as compared to the 96, 83, 85 and 88% yields obtained with the triazole containing MOF. The recyclability of the MOF was confirmed by conducting 5 repeated cycles, with 95% yield of propylene carbonate being achieved in the 5th cycle. Powder X-ray diffraction (PXRD) analysis also demonstrated that the MOF maintained its crystal structure following the catalytic reactions.

By increasing the density of Lewis acid sites, the catalytic activity of MOFs towards the cycloaddition of CO2 to epoxides may also be enhanced. MOF crystals with large pore channels enhance the accessibility of the Lewis acid sites and facilitate the efficient diffusion of substrates and products, leading to improved catalytic performance.

3.2.2. MOFs with both Lewis acid and Lewis base active sites. MOFs may also be constructed with linkers exhibiting Lewis base functionalities to eliminate the need for a co-catalyst.5 To this end, several MOFs containing both Lewis acid metal centers and Lewis base rich organic linkers have been designed for the cycloaddition of CO2 to epoxides.55,86,92–94 The concurrent presence of the Lewis acid site and the basic acid site within the MOFs results in better performance as compared to MOFs without Lewis basic functionality.

For example, amine functionalized MIL-68(In) demonstrated an improved catalytic efficiency for the synthesis of styrene carbonate, as compared to pristine MIL-68 without such amine functionality (Table 2, entries 18 and 19).55 The reaction was carried out for 8 h at 150 °C and 7.9 atm of CO2, and in the absence of a co-catalyst with dimethylformamide (DMF) as the solvent, to obtain a conversion efficiency of 74%, which is significantly higher than the 39% conversion obtained in the absence of an amine functionality. XRD analysis after the catalytic reaction showed that the diffraction patterns of the powder remained unchanged. However, the catalytic efficiency decreased from 74% to 53% upon recycling, while BET analysis revealed a decrease in surface area from 1100 to 720 m2 g−1, following the catalytic reactions, potentially due to blockage of pores and catalytic sites by the products.

Amine functionalized UiO-66 crystals were also shown to be more efficient in the cycloaddition of CO2 to styrene oxide compared to pristine UiO-66.86 The reaction was carried out at 100 °C and 19.7 atm of CO2 in chlorobenzene solvent, in the absence of a co-catalyst. Under these experimental conditions, the amine functionalized MOF achieved a conversion of 70%, which was 22% higher than the conversion obtained with a similar amount of non-functionalized UiO-66, under similar experimental conditions.

Babu et al.77 very recently designed dual porous amine functionalized MOFs (UMCM-1-NH2,), combining meso- and microporous structures for room temperature CO2 fixation (Table 2, entry 21). The MOF materials exhibited a high product selectivity greater than 99% and a yield of 90%, in the cycloaddition of propylene oxide to propylene carbonate, in the presence of a TBAB co-catalyst at room temperature and 11.8 atm of CO2, and under solventless conditions. These MOF crystals were found to outperform other reported MOF materials tested, such as MOF-5, which gave a yield of 93% at 50 °C and 4 atm, and Cr-MIL-101 which gave a yield of 82% at room temperature and the pressure of 8 atm.45 The excellent performance was attributed to the dual porous structure, where mesoporosity allows easy accessibility and efficient transport of substrates and products, while micropores modulate the catalytic reactions. The UMCM-1-NH2 MOF also maintained its catalytic activity upon recycling, with a yield of 89% in the 5th cycle. XRD analysis of the MOF powder revealed that the crystal structure was maintained after the catalytic reactions.

Ma et al. in 201542 used post-synthesis modification to prepare quaternary ammonium and quaternary phosphorous ionic liquid functionalized MOFs, MIL-101-N-(n-Bu)3Br and MIL-101-P(n-Bu)3Br (entries 24 and 25). Owing to the synergistic effect of Br Lewis acid sites from the ionic liquids and coordinatively unsaturated Cr3+ sites, the MOFs exhibited superior catalytic activity in the cycloaddition of propylene oxide (PO), with product yields of 98.6 and 99.1 for MIL-101-P(n-Bu)3Br and MIL-101-N-(n-Bu)3Br, respectively, at 80 °C and 19.7 atm in the absence of a solvent and a co-catalyst, after 8 h of reaction. This was significantly higher than the yield obtained with Mg-MOF-74, MIL-101, HKUST and MOF-5, which gave a yield of 23.2, 20.9, 5.4, and 2.5%, respectively, under the same experimental conditions. The functionalized MOFs could also be recovered and re-used up to 3 times, with less than 1% decrease in the product yield in the third run.

Incorporating Lewis base functionalities within MOFs leads to improved performance due to their ability to act as co-catalysts. Functionalization with amine groups also results in enhanced CO2 adsorption, further leading to an enhanced yield. Among the different amine functionalized MOFs, UMCM-1-NH277 outperforms most of them such as amine functionalized MIL-6855 and UiO-66,86 due to its unique dual porous structure, in addition to its Lewis base functionality. The relatively large meso-pores enable the efficient transport of substrates and products across the porous frameworks, while micropores, owing to their smaller size, shorten the product retention duration within the reaction sites, thus reducing the probability of the formation of the secondary product, and leading to improved product selectivities.77

3.2.3. MOFs with active defect sites. Defects in MOFs, arising from crystal imperfections such as vacant metal or linker sites, may also serve as active sites for catalytic reactions.95,96 MOFs with catalytic defect sites include ZIF-8, ZIF-68 and MOF-5.14,61,78 For example, Miralda et al. in 201278 utilized ZIF-8 for the cycloaddition of CO2 to chloropropene oxide, in the absence of co-catalysts and solvents, between 70 °C and 100 °C, and obtained a maximum carbonate yield of 44% at 80 °C (Table 2, entry 26). The conversion efficiency could be improved up to 73% after functionalization of the ZIF-8 surface with amine functional groups. However, the catalysts offered poor recyclability and the particles lost their crystalline structure and catalytic performance upon recycling. In a separate work, Zhu et al.49 used ZIF-8 for the conversion of styrene oxide into styrene carbonate and obtained a conversion efficiency of 54% at 100 °C and 6.9 atm, without the need for a co-catalyst or solvent (Table 2, entry 29). XRD analysis demonstrated that the crystal structure of ZIF-8 particles could be maintained under these reaction conditions, at least for three subsequent cycles. The cycloaddition reaction is expected to take place on the surface due to the small aperture of ZIF-8 (0.34 nm), where low coordinated Zn atoms and active defect sites are located.97,98

The catalytic performance of ZIF-68 in cycloaddition reactions has also been reported, where 93.3% yield of styrene carbonate was achieved at 120 °C and 9.9 atm in the absence of a solvent and co-catalyst.61 Similar to ZIF-8, ZIF-68 crystals exhibit small pore apertures of 0.55 nm, which may prevent the reaction from taking place within the pore structures. The reaction thus takes place on the surface of the crystals where unsaturated coordinative metal ions and catalytically active structural defects are present. XRD analysis revealed that the MOF crystals were stable for at least three cycles, though there was a decrease in product yield with each subsequent use, from 93.3% for fresh MOFs to 88.3% in the second cycle and 80.9% in the third cycle. The fourth cycle recorded an even bigger drop in the yield down to 66.4%, highlighting the potential blockage or poisoning of catalytic sites.

Hwang et al.90 evaluated the performance of ZIF-22 in the cycloaddition of CO2 to epoxides, in the absence of a solvent and co-catalyst (Table 2, entry 33). The reactions were carried out using the epichlorohydrin (ECH) substrate, at 120 °C and a pressure of 11.8 atm. The MOF achieved a TOF of 155 h−1, which was significantly higher than other Zn containing MOFs, ZIF-8, ZIF-67 and ZIF-8-NH2, which had TOFs of 12, 22 and 17, respectively,61,78,99 under similar experimental conditions. The higher efficiency of ZIF-22 in comparison with other ZIFs was attributed to the presence of the extra non-coordinated N atom in the 5-azabenzimidazolate linker, which acted as a Lewis base co-catalyst and facilitated the cycloaddition reactions. XRD analysis revealed that the crystal structure of the MOFs could be maintained even after recycling 3 times.

The defective sites present across the MOF surface may act as active sites for cycloaddition reactions. However, these MOFs required high temperatures and pressure, suggesting that the catalytic activity of the defect sites is not sufficient. A potential route for increasing the catalytic performance of these MOFs is by engineering MOFs with a high density of defect sites.96,100,101 One effective strategy to increase the density of defect sites is to use fast precipitation synthesis methods during the preparation of MOFs, which denies the MOF building blocks sufficient time to adhere to the growing crystal lattice at the right place, leading to a higher density of defect sites.96,100 This technique has been used successfully to increase the density of defect sites across MOF-5.100 Another approach is to use the isostructural mixed linker, but with different side functionalities to coordinate extra metal ions, resulting in a high density of defect sites while preserving the framework topology.100 Post-synthetic treatment of MOF crystals with acid or base has also been used to introduce defect sites across MOFs,100,101 but these approaches should be employed cautiously, since they may lead to the degradation of the MOF materials, by exposing the metal centers to solvents.

3.2.4. Conclusions and prospects. Lewis acid metal centers and defect sites across metal organic frameworks act as active sites for the cycloaddition of CO2 to epoxides. However, Lewis acid sites appear to play the most important role, and their density and accessibility are critical for efficient catalytic performance. The chemical versatility of MOF materials allows the density of these active sites to be enhanced directly by employing organic ligands having multiple coordination sites for the metal ions,102 or by mixing two or more similar ligands during MOF synthesis, to lead to an interpenetrated structure, such as the dual walled cage MOF reported by Han et al.57 The rational choice of organic ligands can also be used to tune the pore size and architecture,47 which in turn influence the accessibility of the active sites as well as the diffusibility of substrates and catalytic products across the porous frameworks. Post-functionalization with Lewis base functionality such as amine groups55,103 and halide ions (e.g. Br and I)42,104 also leads to improved catalytic performance due to their ability to act as a co-catalyst, besides enhancing the CO2 adsorption capacity of the MOF materials.

Recyclability tests have also been reported, where most MOF catalysts were shown to maintain their activity and crystal integrity for 3 to 5 cycles. However, to fully assess the potential MOF catalysts, a higher number of cycles is required, while monitoring not only the changes in the crystal structure, but also the changes in the morphology, surface area and chemical composition over time. Tailoring the porous structure and optimizing the density of Lewis acid and base sites across MOF materials are expected to lead to enhanced catalytic activity towards cycloaddition reactions. However, significant effort should also be devoted to designing MOF catalysts with a high degree of connectivity and hydrophobic pores, which result in more tolerant structures under harsh reaction conditions.63,64,105

4. Photocatalytic conversion of CO2

4.1. Introduction

Another attractive approach for utilizing CO2 is its conversion into valuable fuels and chemicals via photocatalysis.17 This greenhouse gas may be converted into valuable products such as methanol, formic acid, methane and carbon monoxide, in the presence of a photocatalyst.9,17 Since photocatalysis can be accomplished using renewable solar energy, with water as the proton source, the production of valuable liquid fuels such as methanol using this approach would be particularly attractive, since the current industrial synthesis approach using the Fischer–Tropsch process is energy intensive,10 which makes the process expensive and contributes further to CO2 emission. Using methanol produced from the photocatalytic conversion of captured CO2 as fuel is also not expected to lead to net CO2 emission and will reduce the reliance on fossil fuels, further mitigating the adverse effects of global warming.

The most common photocatalysts are inorganic semiconductors such as TiO2, ZnO, CdS, and ZnS,9,17 with TiO2 being the most popular due to its ready availability, non-toxicity and long term stability. However, the catalytic performance of these materials is limited due to the rapid recombination of photo-generated electrons and holes, and low adsorption capacities for CO2.17,46,106 As recently reviewed by Ola et al. the benchmark TiO2 photocatalysts are only active under UV irradiation.6 Modifying TiO2 with organic or inorganic materials and photosensitization with organic dyes resulted in photocatalytic activity under visible light, but the product yield of most of the modified systems was low, with the product yield lower than 100 μmol per gram of catalyst.6 N doped TiO2 nanotubes exhibited the highest activity among different modified TiO2 systems, with 12[thin space (1/6-em)]475.8 μmol of formic acid per gram of catalyst,9,107 after visible light illumination for 12 h. Significant effort is currently being devoted to the development of alternative photocatalytic materials able to facilitate more efficiently the conversion of CO2 under visible light activation.52

Recently, MOF materials have generated interest as visible light active photocatalytic materials in (i) degradation of organics, (ii) chemical synthesis and (iii) reduction of CO2.19,28,108 By judicious selection of metal ions and organic ligands, the photocatalytic properties of MOFs may be modulated.19 The organic ligands may serve as antennae to harvest light and generate electrons, which are then transferred to the metal centers via linkers by metal cluster charge transfer.19,109,110

4.1.1. Electronic properties of MOF materials. The efficiency of a photocatalytic material is determined not only by the width of the bandgap but also by the absolute positions of the band edges.12,111,112 For a MOF material to be active towards CO2 conversion, the lowest unoccupied crystal orbital (LUCO) must be above the redox potential for the CO2 reduction half reaction,112,113 which depends on the product formed. For example the redox potential for the conversion of CO2 into formic acid is −3.42 eV while those for methanol and methane are −3.65 and −3.79 eV, respectively.112 While both bridging metal ions and organic ligands have been shown to have an influence on the electronic properties,109,113,114 the organic ligand has been shown to play the most important role,112,115,116 and the electronic properties of MOFs may be predicted based on the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of the organic ligand.12 Grau-Crespo et al. demonstrated through computational studies that the band gap of zeolitic imidazolate frameworks (ZIFs) could be varied from 5.2 eV when a pure imidazole ligand was used to 3.3 eV when a nitro-imidazole ligand was used. The use of mixed linkers of furanimidazole/nitroimidazole (fIm/nIm) and methylimidazole/nitroimidazole (mIm/nIm) even led to a further reduction in the band gap to 1.9 eV and 2.5 eV, respectively, while the incorporation of Co and Cu ions into the Zn/mIm/nIm MOF led to band gaps of 2.3 eV and 2.1 eV, respectively.

The metal centers should be selected such that the empty d orbitals overlap with the lowest unoccupied molecular orbital (LUMO) of the organic ligand, to facilitate efficient ligand to metal charge transfer (LMCT),117–119 and long lived charge separation. LMCT enables the MOF to exhibit a longer charge separation time, as compared to isolated organic ligands,119 and explains why some MOF materials exhibit a longer charge separation time as compared to typical semiconductors. For example MOF-5 was shown to have a charge separation time of 15 μs which was 3 times higher than that of P25 TiO2.109

The following section discusses specific examples of MOF materials that have been utilized in the photocatalytic conversion of CO2.

4.2. MOFs as photocatalysts for CO2 conversion

Li et al. investigated the MIL-125(Ti) MOF and its amine functionalized derivative, NH2-MIL-125, for the photocatalytic conversion of CO2 under visible light (Table 3, entry 1). NH2-MIL-125(Ti) exhibited an enhanced catalytic activity towards the conversion of CO2 into formate, in acetonitrile with triethanolamine (TEOA) as an electron donor, as compared to the MIL-125(Ti) without the amine functionality.120 The accumulated HCOO content, over a 10 h period, was 8.14 μmol, while no HCOO ions were formed in the absence of the amine functionality. Upon excitation, excited electrons are transferred from the amine functionalized ligand to Ti4+ clusters, reducing them to Ti3+, which subsequently transfers the excess electrons to the CO2 molecule. The valence transition between Ti4+ and Ti3+ was witnessed by the observation of color change from bright yellow to green upon excitation with visible light, signifying the reduction of Ti4+ clusters to Ti3+. However, upon introduction of CO2, the color reverted from green to the original bright yellow, as Ti3+ was oxidized to Ti4+, upon transfer of the electrons to the CO2 molecule. XRD patterns taken after the photocatalytic reaction revealed that the scattering pattern of the parent MOF was preserved, suggesting that the crystal structure of the MOF catalyst was not compromised during the photocatalytic reactions.
Table 3 Application of MOFs in the photocatalytic conversion of CO2
Entry MOF BET Light source Proton donor Product Yield Ref.
1 NH2-MIL-125(Ti) 1302 Visible TEOA Formate 16.3 μmol g−1 h−1 120
2 MIL-88B(Fe) Visible TEOA Formate 67.5 μmol h−1 52
3 NH2-MIL-88(Fe) Visible TEOA Formate 225 μmol g−1 h−1 52
4 MIL-53(Fe) Visible TEOA Formate 222.75 μmol g−1 h−1 52
5 NH2-MIL-53(Fe) Visible TEOA Formate 348.75 μmol g−1 h−1 52
6 MIL-101(Fe) Visible TEOA Formate 442.5 μmol g−1 h−1 52
7 NH2-MIL-101(Fe) Visible TEOA Formate 1335 μmol g−1 h−1 52
8 Re doped UiO-67 1092 UV TEA CO 0.545 TOF 123
9 UiO-66-CrCAT Visible TEOA Formate 20[thin space (1/6-em)]692 μmol g−1 h−1 121
10 UiO-66-GaCAT Visible TEOA Formate 11[thin space (1/6-em)]512 μmol g−1 h−1 121
11 Ru-MOF 8.08 Visible TEOA Formate 77.2 μmol g−1 cat h−1 29
12 NH2-UiO-66(Zr) 778 Visible TEOA Formate 26.4 μmol g−1 h−1 129
13 Porphyrin based MOF (Al) 1187 Visible Water Methanol 37.5 ppm g−1 h−1 124
14 Porphyrin based MOF (Al, Cu) 932 Visible Water Methanol 262.6 ppm g−1 h−1 124
15 10%-Cp*Rh@UiO-67 1650 Visible TEOA Formate 7.4 TOF 130
16 PCN-222 1728 Visible TEOA Formate 60 μmol g−1 h−1 46
17 ZIF-8/Zn2GeO4 319.5 Water Methanol 0.2218 μmol g−1 h−1 106
18 TiO2 doped HKUST-1 756 UV Water CH4 2.64 μmol g−1 h−1 128
19 Co-ZIF-9 1428 Visible TEOA CO, H2 179 TOF 131 and 132
20 Zr6(O)4(OH)4-[Re(CO)3Cl(bpydb)]6 Visible TEOA CO 1.073 TOF 133


MOFs composed of Fe metal centers, owing to the wide availability of Fe and the visible light responsiveness of Fe based photocatalysts, were also investigated.52 The study compared MIL-101(Fe), MIL-53(Fe) and MIL-88B(Fe), and their amine functionalized derivatives (Table 3, entries 2–7).52 Unlike the Ti–O clusters, the Fe–O clusters in the Fe based MOFs exhibited visible light activity towards the conversion of CO2 into formic acid, even in the absence of an amine functionality, with a yield of 59 μmol, 29.7 μmol, and 9 μmol, for MIL-101, MIL53 and MIL88, respectively, after 8 h of reaction. MIL-101(Fe) exhibited the best photocatalytic activity towards the conversion among the non-functionalized MOFs, due to the presence of uncoordinated unsaturated Fe sites in the structure. Amine functionalization was found to boost the photocatalytic activity of all the Fe based MOFs, leading to 178 μmol, 46.5 μmol, and 30 μmol, for MIL-101, MIL53 and MIL88, respectively, over the same reaction duration, due to dual excitation pathways of the NH2 functionality and Fe–O clusters. The Fe based MOFs could be used for three cycles without any noticeable loss in the yield of formic acid. The crystal structure, chemical composition and porosity were also preserved as XRD, IR, TGA and N2 adsorption of the crystals recovered after the reactions matched those of fresh samples.

MOF materials may also be modified with homogenous catalysts or photosensitizer molecules,19,121,122 to enhance their catalytic properties. Wang et al. (2011)123 prepared MOF UiO-67 doped with catalytically active Ir, Re, and Ru complexes H2L1–H2L6, with dicarboxylic acid functionalities (Table 3, entry 8). The study found [Re1(CO)3(dcbpy)Cl] (H2L4) to be the most active catalyst in the selective reduction of CO2 to CO, under visible light irradiation with trimethylamine (TEOA) as an electron donor. The catalyst was found to be 3 times more effective than homogenous catalyst Re-H2L4. The activity was attributed to the [Re1(CO)3(dcbpy)Cl] (H2L4) moiety, since UiO-67 demonstrated no activity under the same experimental conditions. The doped UiO-67 catalyst was, however, found to have poor stability as 43.6% of Re leaked into the solution during 20 h of photocatalysis. Liu et al. in 2013124 investigated the photocatalytic reduction of CO2 using Cu modified porphyrin based MOFs, with Al metal centers. The study compared the efficiency of the Cu modified MOF (SCu), with that of the parent MOF (Sp), and found SCu to be up to 7 times more efficient than Sp (262.6 ppm g−1 h−1vs. 37.5 ppm g−1 h−1), in the photo-conversion of CO2 into methanol. The improved performance was attributed to the enhanced chemical adsorption and activation of CO2 due to the presence of Cu2+. The recyclability of the catalyst was, however, not reported.

Lee et al.121 post-synthetically modified UiO-66(Zr) with a catechol functionalized organic linker (catbdc, 2,3-dihydroxyterephthalic acid), to produce UiO-66-CAT, followed by the incorporation of trivalent metal ions, Cr and Ga, at the catbdc sites. Photocatalytic reactions were carried out under visible light irradiation, where formic acid was found to be the major product, with 51.73 ± 2.64 μmol for UiO-66-CrCAT and 28.78 ± 2.52 μmol for UiO-66-GaCAT. The catbdc organic linker served as a visible light absorber to generate excited electrons, while Cr and Ga metal ions facilitated electron transfer within the framework. The use of UiO-66-CAT in the absence of the Cr and Ga metal ions did not produce any formic acid, due to the high redox potential of the Zr metal centers across UiO-66 relative to the lower unoccupied molecular orbital (LUMO) of the catbdc linker, which hinders the transfer of photo-generated electrons.117,118 Although there was no noticeable drop in the product yield for three consecutive cycles, ICP-MS analysis indicated that a small amount of Cr and Ga ions leached out after every cycle.

Flower-like hierarchical Ru-MOFs, {Cd2[Ru(dcbpy)3]·12H2O}n (dcbpy = 2,2′-bipyridine-4,4′-dicarboxylate), have also been recently reported by Zhang et al. (2015) (entry 11).29 Photocatalytic reactions were carried out in the presence of the sacrificial agent TEOA under visible light. The Ru-MOF demonstrated superior catalytic properties towards the conversion of CO2 into formate, under visible light in the presence of TEOA, with a product formation rate of 77.2 μmg−1 h−1. This efficiency was higher than those of other MOF based catalysts such as NH2-MIL-125(Ti), NH2-UiO-66(Zr) and mixed NH2-UiO-66(Zr), whose product formation rates were 16.3, 26.4 and 41.4 μmol g−1 h−1 of catalyst per hour, respectively. The value is also higher than some TiO2 based catalysts such as iodide doped TiO2 and nitrogen and nickel co-doped TiO2125 (product formation rates of 2.4 and 15.1 μm g−1 of catalyst).

The improved properties were attributed to the high visible light harvesting capacity of the MOF and the long lasting excited state. The MOF exhibited a long-lived excited state, with a luminescence lifetime of 5.49 μs, which was significantly higher than that of Ru units (483 ns), and TiO2 particles (0.16–0.55 ns)126 and TiO2 nanorods (12.02 ns),127 suggesting a low recombination rate of the photo-generated electrons and holes. The superior charge separation properties were facilitated by the unique hierarchical morphology, which enabled the efficient diffusion of generated electrons to the catalyst surface. Catalytic studies using microcrystals and bulk crystals of the same MOF produced lower formate yields of 52.7 and 30.6 μmol per g, respectively (Fig. 2), highlighting the critical role of the morphology in photocatalytic activity. The MOFs could be recycled up to 4 times with no noticeable change in the photo-catalytic activity. Inductively coupled plasma (ICP) analysis found that only 0.43% of Ru ions leaked after the reaction, while PXRD showed no apparent change in the crystal structure after the reactions.


image file: c6mh00484a-f2.tif
Fig. 2 The role of the morphology of Ru-MOFs in the generation of HCOO (a) nanoflowers, (b) microcrystals, and (c) bulk crystals. The hierarchical morphology of the nanoflowers results in more enhanced charge separation, leading to improved HCOO generation as compared to microcrystals or bulk crystals of the same MOF. Reproduced from ref. 29 with permission from the Royal Society of Chemistry.

The composition of metal centers, the organic ligand and the morphology of MOFs all play an important role in the catalytic activity. For example, MOFs with Fe–O metal centers exhibit visible light activity as opposed to MOFs with Ti–O clusters. Similar to cycloaddition reactions described in the previous sections, the modification of MOFs with amine functionality led to enhanced photocatalytic activity, due to the ability of the amine group to generate excited electrons following irradiation with light, and its high affinity for CO2. MOF catalysts with a hierarchical morphology were also shown to result in improved performance, due to the efficient transfer of photo-generated electrons to the reaction sites.

Another approach for imparting or modifying the photocatalytic properties of MOF materials is the integration of inorganic semiconductors into MOFs, to make hybrid catalyst meta-materials.

4.3. Inorganic semiconductor MOF hybrid catalysts

The integration of MOF materials with inorganic catalysts to make hybrid catalysts has the potential to result in enhanced catalytic activity, due to the synergistic effect arising from the photo-excitation properties of the inorganic semiconductors, and the higher CO2 adsorption capacities for MOFs.128

For example, with the aim of boosting the photocatalytic activity of Zn2GeO4 nanorods, Liu et al. deposited ZIF-8 nanoparticles on the nanorods, and obtained improved efficiency towards the photocatalytic conversion of CO2 into methanol, in aqueous medium, as compared to Zn2GeO4 nanorods alone (Table 3, entry 17).106 The hybrid catalysts led to a yield of 2.32 μmol g−1 over a 10 h period, which was 62% higher than 1.43 μmol g−1 obtained with Zn2GeO4 nanorods alone. The enhanced performance was attributed to improved CO2 adsorption capacity, which increased by 3.8 times following the incorporation of ZIF-8 (25 wt%) into the Zn2GeO4 nanorods. No product was detected by using ZIF-8 particles alone, clearly demonstrating the synergistic role of combining the two different materials.

Li et al., on the other hand, prepared TiO2 coated Cu3(BTC)2 MOFs, and demonstrated that charge transfer can occur between the photoexcited TiO2 and the MOF, leading to improved catalytic properties under UV irradiation (Table 3, entry 18). Methane was found to be the only product generated for the TiO2 doped MOFs, with a yield of 2.64 μmol g−1 h−1, which was significantly higher than 0.52 μmol g−1 h−1 obtained with bare TiO2.128 The Cu3(BTC)2 MOF crystals alone did not exhibit any catalytic activity under similar experimental conditions. The TiO2 doped MOF could be used for three subsequent cycles without any change in photocatalytic efficiency. TEM imaging and XRD analysis revealed that the morphology, crystal structures and composition were maintained during the photocatalytic reactions.

The improved catalytic efficiency of the ZIF-8/Zn2GeO4 and the TiO2/Cu3(BTC)2 MOF hybrid systems clearly highlights the benefits of combining these two classes of materials, to make more efficient photo-catalysts. Further work should investigate hybrid systems based on MOFs with inherent photocatalytic activity, such as MIL-101-NH2 and Ru-MOFs, which may lead to more enhanced performance due to additive photocatalytic activity between the two materials. In addition, understanding the influence of the inorganic nano-material–MOF interface on charge transfer will facilitate the design of hybrid systems with optimized charge transfer, leading to more efficient photocatalysts. The influence of the inorganic materials on the structure and long-term stability of MOFs should also be investigated.

The products formed during the photocatalytic conversion of CO2 are dependent on the number of electrons transferred and reaction pathways, which are in turn influenced by the type of catalyst used, as well as the reaction environment.9,134 For example, the formation of carbon monoxide and formic acid requires the transfer of two electrons, while methanol and methane require 6 and 8 electrons, respectively. Consequently, the formation of CO and formic acid is more thermodynamically favorable. As can be seen in Table 3, most MOF catalysts led to high product selectivity, producing only one product, as opposed to typical photocatalysts which lead to a mixture of products.9 As discussed previously, the high product selectivity of MOF based catalysts may be attributed to the confined reaction spaces that shorten the product retention duration within the reaction sites, and reduce the probability for the formation of secondary products.77 The high selectivity is also attributed to the uniform distribution of catalytic sites and microspores that ensure the uniform supply of protons to the catalytic sites.135

4.4. Conclusions and prospects

Unlike inorganic semiconductors, MOF based catalysts exhibit catalytic activity under visible light, allowing CO2 conversion with product yields similar to or better than most state of the art TiO2 based systems. The chemical composition of the MOF materials plays a critical role in the extent of their photocatalytic properties, and careful selection of the MOF building blocks is necessary for MOFs to exhibit photocatalytic properties. Organic ligands should be selected from those which have chromophores, which are responsible for absorbing light to generate excited electrons.136 The metal centers, on the other hand, should have a lower redox potential relative to the lowest unoccupied molecular orbital (LUMO) of the organic ligand to be able to accept the photo-generated electrons.117,118

The main challenge, however, will be balancing the photocatalytic properties with MOF stability, both of which are strongly influenced by the composition of the MOF building blocks. The stability in the presence of water, in particular, is of critical importance given that water is commonly used as a proton donor in the photocatalytic reduction of CO2, and also due to the possible presence of moisture in the CO2 stream, which could be significantly high if the CO2 is sourced from brown coal flue gas. The presence of other contaminants in flue gas streams, such as SOx and NOx, should also be taken into account, since they have been shown to accelerate the degradation of MOF materials.137–139 For example, MIL-125 was found to lose its structure and N2 adsorption capacity upon exposure to an aqueous environment containing SO2, although amine functionalized MIL-125 was found to be stable under similar conditions.138 While using hydrophobic ligands and metal with a high coordination number may enhance the stability in water,63,64 the building blocks may not necessarily be suited for photocatalytic activity. A potential approach is catenation, whereby two or more identical but independent frameworks are interpenetrated, by using mixed linkers and/or metal ions during MOF synthesis. By using this approach building blocks that elicit photocatalytic properties could be mixed with those that lead to high stability, to possibly result in robust photocatalytic MOFs. Catenation has also been shown to lead to more stable frameworks by providing steric hindrance to ligand displacement,63 locking labile ligands across the framework in their place.

5. Electrocatalytic conversion of CO2

5.1. Introduction

Electrocatalysis is another process by which CO2 maybe converted into useful products, such as carbon monoxide, formic acid, methanol, and methane, in the presence of an electrocatalyst.8 Similar to photocatalysis, electrocatalysis can be carried out at room temperature, and has low energy requirements, which can be directly obtained from renewable sources.3 Metal electrodes are to date the most investigated electrocatalysts for CO2 conversion due to their wide availability and suitability in electrochemical processes. Hori et al. (1994)140 studied the electrocatalytic conversion of CO2 using metal electrodes in aqueous media, and categorized them as CO forming electrodes and formate forming electrodes. Formate forming electrode materials include Pb, Hg, In, Sn, Cd, and Tl, while CO forming electrode materials include Cu, Au, Ag, Zn, Pd, Ga, Ni, and Pt. Cu was found to be unique and the most promising electrode, since it may facilitate the formation of multi-electron products, such as methane and methanol.135,141–148 The key determinant to the product selectivity has been attributed to the binding energy of the metal electrodes to CO, which is the main intermediate for CO2 conversion.149 Metals that bind CO strongly exhibit a low product yield, as the surface of the catalyst is poisoned by the CO. Metals with weak binding strength for CO, on the other hand, produce CO as the main product, since the molecules are desorbed from the catalyst surface before further reduction. Cu has an intermediate binding energy for CO, which gives it the unique ability to produce hydrocarbon products. However, Cu based electrocatalysis yields a mixture of products, and it is difficult to obtain a single product with high purity.

The product selectivity may be improved by using porous catalytic materials, which can provide confined reaction spaces for the catalytic reactions, while maintaining the uniform supply of protons.135 Potential candidates are proton conducting MOF materials, due to their high porosity and the uniform distribution of the catalytic sites across the nano-sized pores, which may restrict reaction pathways and enhance product selectivities.150 This section reviews MOF materials that have been applied in the electrocatalytic conversion of CO2. The product selectivity may be expressed by the purity of the product obtained or by faradaic efficiency, which is the mole of product formed on an electrode per charge consumed.

5.2. MOF materials as electrocatalysts for the conversion of CO2

MOF materials have been explored in a number of electrochemical processes, including electrocatalytic water splitting to produce hydrogen and oxygen evolution,151,152 and in electrochemical sensors,153 lithium ion batteries154 and the electrocatalytic reduction of CO2.150,155,156 Hinogami et al.150 first evaluated the electrocatalytic potential of copper rubeanate metal organic frameworks (CR-MOFs), for the electroconversion of CO2. By using a CR-MOF coated on carbon paper to form a working electrode, formic acid was found to be the only by-product. Using the same set up, the use of Cu metal as the working electrode in the absence of the MOF led to a mixture of products, including formic acid, methane, ethylene and ethane. The use of a CR-MOF also led to a 13 fold increase in formic acid production as compared to the Cu electrode at 1.2 V vs. SHE, probably due to the higher CO2 adsorption within the MOF crystals. The high selectivity of Cu based MOFs was also demonstrated by Kumar et al.,156 and oxalic acid with 90% purity was obtained using a Cu3(BTC)2 MOF film, supported on a glassy carbon electrode.

Cobalt-porphyrin MOF films supported on a carbon substrate have also been reported.31 The MOF materials achieved a faradaic efficiency of 76% and a turnover number of 1400 (assuming that every Co atom was catalytically active), in the conversion of CO2 into CO. In situ spectroelectrochemical analysis showed that most of the Co metal centers were accessible, and the Co(II) was reduced to Co(I) during the electrocatalysis, which subsequently reduced CO2. Although the recyclability of the MOF was not tested, XRD analysis and SEM imaging taken after 7 h of reaction revealed that the films retained their morphology as well as their crystalline structure.

Recently, Hod et al. (2015) used Fe_MOF-525 as a support to immobilize the Fe-porphyrin molecular catalyst, and obtained a high faradaic efficiency of 100% for H2 and CO (54 ± 2% for CO and 45 ± 1% for H2).157 The catalytic efficiency could be enhanced by the addition of a trifluoroethanol (TFE) proton donor, resulting in up to a 7 times increase in CO generation, with a TON of 1520, after 3.2 h. However, the turnover frequency (TOF) of the MOF was 16 times lower than that of the Fe-porphyrin homogeneous catalyst, potentially due to the limited rate of charge diffusion across the MOFs.

5.3. Conclusions and prospects

Electrocatalytic conversion of CO2 using MOF catalysts is a new area of research, where a lot is yet to be discovered. However, the few examples discussed above highlight the potential of MOF based electrocatalysts in facilitating CO2 conversion with high product selectivity. The high porosity and well defined pore channels facilitate the uniform supply of protons to catalytic sites, while also modulating the reaction pathways. Unlike cycloaddition and photocatalysis where MOF crystals can be employed directly, for electrocatalytic application MOF materials need to be deposited on the surface of an electrically conductive substrate. This brings further complexity, since the nature of the interface between the substrate and the MOF may have an influence on charge transfer, and ultimately on their catalytic performance. The nature of the interface and its influence on catalytic performance should therefore be critically investigated in future studies.

Although the reported MOF materials demonstrated high product selectivity, none of the reports discussed the recyclability of the MOF based electrocatalysts. Given the vulnerability of most MOF materials in a wide range of solvents,64 their recyclability in the electrochemical solution is critical and should be thoroughly investigated to assess their limitations. A potential means of enhancing their stability is by carbonization. Carbonized MOFs have been shown to preserve the key properties of parent MOFs such as surface area and CO2 adsorption capacities, while exhibiting higher stability.158,159 For example, the carbonized MOF MIL-88B-NH3 outperformed the commercial Pt/C electrocatalysts in the oxygen reduction reaction (ORR), in terms of stability and performance when tested in alkaline direct fuel cells, where it achieved a power density of 22.7 mW cm−2, which was 1.7 times higher than the commercial Pt/C catalyst.

6. General conclusions

In this review, the application of MOF based catalysts in the conversion of CO2 through chemical fixation, photocatalysis and electrocatalysis has been discussed. Compared to other catalytic materials used for CO2 conversion, MOF based catalysts stand out, due to their chemical versatility, higher adsorption capacity for CO2, uniform distribution of catalytic sites and confined reaction spaces. A major problem in the area of CO2 conversion is poor product selectivity, since most catalytic systems lead to a mixture of products, making the process expensive by increasing the cost of product separation. Across all three areas discussed in this review, MOFs consistently demonstrate higher product selectivity due to the uniform distribution of catalytic sites and restricted reaction spaces. Just by manipulating the composition of MOF materials, key properties relevant to CO2 conversion, including the density of active sites, pore architecture and electronic properties, can be optimized, explaining why MOFs feature prominently across all three different areas of CO2 utilization.

However, while tailoring the chemical functionality and crystal structures of MOF materials can lead to enhanced catalytic activity, their stability under different reaction conditions should be taken into account. Cross-comparison of crystal structures, chemical composition and nitrogen adsorption properties before and after catalytic reaction is critical, to assess their long-term stability. Highly interconnected frameworks should be targeted while designing new MOF catalysts, by using metal ions with a high coordination number or organic ligands with a high number of coordination sites. Since it may be challenging to get building blocks that combine both high connectivity and catalytic activity, a mixed ligand or metal ion approach should be employed to bring different functionalities into a single unit. In electrocatalysis, carbonized MOFs should be explored as they may exhibit superior performance, due to their higher stability in electrochemical solutions. Other areas for exploration should include photoelectrocatalytic conversion of CO2 using MOF based catalysts, since photoelectrocatalysis has been shown to lead to a higher yield and selectivity for methanol,160,161 as compared to either photocatalysis or electrocatalysis alone.

Abbreviations

TONTurnover number
TOFTurnover frequency
MOFsMetal organic frameworks
CCScarbon capture and storage

References

  1. J.-R. Li, Y. Ma, M. C. McCarthy, J. Sculley, J. Yu, H.-K. Jeong, P. B. Balbuena and H.-C. Zhou, Coord. Chem. Rev., 2011, 255, 1791–1823 CrossRef CAS .
  2. T. M. McDonald, J. A. Mason, X. Kong, E. D. Bloch, D. Gygi, A. Dani, V. Crocella, F. Giordanino, S. O. Odoh, W. S. Drisdell, B. Vlaisavljevich, A. L. Dzubak, R. Poloni, S. K. Schnell, N. Planas, K. Lee, T. Pascal, L. F. Wan, D. Prendergast, J. B. Neaton, B. Smit, J. B. Kortright, L. Gagliardi, S. Bordiga, J. A. Reimer and J. R. Long, Nature, 2015, 519, 303–308 CrossRef CAS PubMed .
  3. A. S. Agarwal, Y. Zhai, D. Hill and N. Sridhar, ChemSusChem, 2011, 4, 1301–1310 CrossRef CAS PubMed .
  4. A. Goeppert, M. Czaun, R. B. May, G. S. Prakash, G. A. Olah and S. Narayanan, J. Am. Chem. Soc., 2011, 133, 20164–20167 CrossRef CAS PubMed .
  5. K. S. Lackner, Eur. Phys. J.: Spec. Top., 2009, 176, 93–106 CrossRef .
  6. A. R. Millward and O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 17998–17999 CrossRef CAS PubMed .
  7. E. J. Maginn, J. Phys. Chem. Lett., 2010, 1, 3478–3479 CrossRef CAS .
  8. J. Qiao, Y. Liu, F. Hong and J. Zhang, Chem. Soc. Rev., 2014, 43, 631–675 RSC .
  9. O. Ola and M. M. Maroto-Valer, J. Photochem. Photobiol., C, 2015, 24, 16–42 CrossRef CAS .
  10. J. Kothandaraman, A. Goeppert, M. Czaun, G. A. Olah and G. K. S. Prakash, J. Am. Chem. Soc., 2016, 138, 778–781 CrossRef CAS PubMed .
  11. J. Ren, F.-F. Li, J. Lau, L. González-Urbina and S. Licht, Nano Lett., 2015, 15, 6142–6148 CrossRef CAS PubMed .
  12. R. Grau-Crespo, A. Aziz, A. W. Collins, R. Crespo-Otero, N. C. Hernández, L. M. Rodriguez-Albelo, A. R. Ruiz-Salvador, S. Calero and S. Hamad, Angew. Chem., Int. Ed., 2016, 55, 16012–16016 CrossRef CAS PubMed .
  13. M. Pérez-Fortes, J. C. Schöneberger, A. Boulamanti and E. Tzimas, Appl. Energy, 2016, 161, 718–732 CrossRef .
  14. M. H. Beyzavi, C. J. Stephenson, Y. Liu, O. Karagiaridi, J. T. Hupp and O. K. Farha, Frontiers in Energy Research, 2015, 2, 63 CrossRef .
  15. M. North, R. Pasquale and C. Young, Green Chem., 2010, 12, 1514–1539 RSC .
  16. F. Rahmani, M. Haghighi, P. Estifaee and M. R. Rahimpour, J. Nat. Gas Sci. Eng., 2012, 7, 60–74 CrossRef CAS .
  17. W. Tu, Y. Zhou and Z. Zou, Adv. Mater., 2014, 26, 4607–4626 CrossRef CAS PubMed .
  18. H. Furukawa, K. E. Cordova, M. O'Keeffe and O. M. Yaghi, Science, 2013, 341, 1230444 CrossRef PubMed .
  19. S. Wang and X. Wang, Small, 2015, 11, 3097–3112 CrossRef CAS PubMed .
  20. C.-Y. Sun, C. Qin, X.-L. Wang and Z.-M. Su, Expert Opin. Drug Delivery, 2013, 10, 89–101 CrossRef CAS PubMed .
  21. N. A. Khan, Z. Hasan and S. H. Jhung, J. Hazard. Mater., 2013, 244–245, 444–456 CrossRef CAS PubMed .
  22. M. Shah, M. C. McCarthy, S. Sachdeva, A. K. Lee and H.-K. Jeong, Ind. Eng. Chem. Res., 2012, 51, 2179–2199 CrossRef CAS .
  23. B. Li, H. Wang and B. Chen, Chem. – Asian J., 2014, 9, 1474–1498 CrossRef CAS PubMed .
  24. Y. Fong Yeong, L. Sze Lai, K. Keong Lau and M. Shariff Az, J. Appl. Sci., 2014, 14, 1161–1167 CrossRef .
  25. C. Zhang and W. J. Koros, J. Phys. Chem. Lett., 2015, 6, 3841–3849 CrossRef CAS PubMed .
  26. A. R. Millward and O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 17998–17999 CrossRef CAS PubMed .
  27. M. H. Beyzavi, C. J. Stephenson, Y. Liu, O. Karagiaridi, J. T. Hupp and O. K. Farha, Frontiers in Energy Research, 2015, 2, 63 CrossRef .
  28. C.-C. Wang, Y.-Q. Zhang, J. Li and P. Wang, J. Mol. Struct., 2015, 1083, 127–136 CrossRef CAS .
  29. S. Zhang, L. Li, S. Zhao, Z. Sun, M. Hong and J. Luo, J. Mater. Chem. A, 2015, 3, 15764–15768 CAS .
  30. H. He, J. A. Perman, G. Zhu and S. Ma, Small, 2016, 12, 6309–6324 CrossRef CAS PubMed .
  31. N. Kornienko, Y. Zhao, C. S. Kley, C. Zhu, D. Kim, S. Lin, C. J. Chang, O. M. Yaghi and P. Yang, J. Am. Chem. Soc., 2015, 137, 14129–14135 CrossRef CAS PubMed .
  32. A. Dhakshinamoorthy, A. M. Asiri and H. García, Angew. Chem., Int. Ed., 2016, 55, 5414–5445 CrossRef CAS PubMed .
  33. A. Morozan and F. Jaouen, Energy Environ. Sci., 2012, 5, 9269–9290 CAS .
  34. R. Navarro Amador, M. Carboni and D. Meyer, Mater. Lett., 2016, 166, 327–338 CrossRef CAS .
  35. Y. Li, H. Xu, S. Ouyang and J. Ye, Phys. Chem. Chem. Phys., 2016, 18, 7563–7572 RSC .
  36. T. Zhang and W. Lin, Chem. Soc. Rev., 2014, 43, 5982–5993 RSC .
  37. L. Zeng, X. Guo, C. He and C. Duan, ACS Catal., 2016, 6, 7935–7947 CrossRef CAS .
  38. N. L. Rosi, J. Kim, M. Eddaoudi, B. Chen, M. O'Keeffe and O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 1504–1518 CrossRef CAS PubMed .
  39. C. Palomino Cabello, G. Gómez-Pozuelo, M. Opanasenko, P. Nachtigall and J. Čejka, ChemPlusChem, 2016, 81, 828–835 CrossRef CAS .
  40. D.-A. Yang, H.-Y. Cho, J. Kim, S.-T. Yang and W.-S. Ahn, Energy Environ. Sci., 2012, 5, 6465–6473 CAS .
  41. K.-S. Lin, A. K. Adhikari, C.-N. Ku, C.-L. Chiang and H. Kuo, Int. J. Hydrogen Energy, 2012, 37, 13865–13871 CrossRef CAS .
  42. D. Ma, B. Li, K. Liu, X. Zhang, W. Zou, Y. Yang, G. Li, Z. Shi and S. Feng, J. Mater. Chem. A, 2015, 3, 23136–23142 CAS .
  43. W. Y. Gao, Y. Chen, Y. Niu, K. Williams, L. Cash, P. J. Perez, L. Wojtas, J. Cai, Y. S. Chen and S. Ma, Angew. Chem., 2014, 53, 2615–2619 CrossRef CAS PubMed .
  44. B. Chen, N. W. Ockwig, A. R. Millward, D. S. Contreras and O. M. Yaghi, Angew. Chem., 2005, 117, 4823–4827 CrossRef .
  45. M. H. Beyzavi, R. C. Klet, S. Tussupbayev, J. Borycz, N. A. Vermeulen, C. J. Cramer, J. F. Stoddart, J. T. Hupp and O. K. Farha, J. Am. Chem. Soc., 2014, 136, 15861–15864 CrossRef CAS PubMed .
  46. H.-Q. Xu, J. Hu, D. Wang, Z. Li, Q. Zhang, Y. Luo, S.-H. Yu and H.-L. Jiang, J. Am. Chem. Soc., 2015, 137, 13440–13443 CrossRef CAS PubMed .
  47. Z. Zhou, C. He, J. Xiu, L. Yang and C. Duan, J. Am. Chem. Soc., 2015, 137, 15066–15069 CrossRef CAS PubMed .
  48. T. E. Synchrotron, Fast nucleation and growth of nanocrystals of a porous coordination polymer, http://www.esrf.eu/UsersAndScience/Publications/Highlights/2011/scm/scm2, accessed 12th Jan, 2017.
  49. M. Zhu, D. Srinivas, S. Bhogeswararao, P. Ratnasamy and M. A. Carreon, Catal. Commun., 2013, 32, 36–40 CrossRef CAS .
  50. U. O. Liverpool, MOF-5 (or IRMOF-1) Metal Organic Framework, http://www.chemtube3d.com/solidstate/MOF-MOF5.html, accessed 19th Jan, 2017.
  51. S.-H. Lo, D. Senthil Raja, C.-W. Chen, Y.-H. Kang, J.-J. Chen and C.-H. Lin, Dalton Trans., 2016, 45, 9565–9573 RSC .
  52. D. Wang, R. Huang, W. Liu, D. Sun and Z. Li, ACS Catal., 2014, 4, 4254–4260 CrossRef CAS .
  53. C. Volkringer, M. Meddouri, T. Loiseau, N. Guillou, J. Marrot, G. Férey, M. Haouas, F. Taulelle, N. Audebrand and M. Latroche, Inorg. Chem., 2008, 47, 11892–11901 CrossRef CAS PubMed .
  54. K. Barthelet, J. Marrot, G. Ferey and D. Riou, Chem. Commun., 2004, 520–521,  10.1039/B312589K .
  55. T. Lescouet, C. Chizallet and D. Farrusseng, ChemCatChem, 2012, 4, 1725–1728 CrossRef CAS .
  56. P.-Z. Li, X.-J. Wang, J. Liu, J. S. Lim, R. Zou and Y. Zhao, J. Am. Chem. Soc., 2016, 138, 2142–2145 CrossRef CAS PubMed .
  57. Y.-H. Han, Z.-Y. Zhou, C.-B. Tian and S.-W. Du, Green Chem., 2016, 4086–4091 RSC .
  58. K. Koh, A. G. Wong-Foy and A. J. Matzger, Angew. Chem., Int. Ed., 2008, 47, 677–680 CrossRef CAS PubMed .
  59. A. Phan, C. J. Doonan, F. J. Uribe-Romo, C. B. Knobler, M. O'Keeffe and O. M. Yaghi, Acc. Chem. Res., 2010, 43, 58–67 CrossRef CAS PubMed .
  60. T. Panda, K. M. Gupta, J. Jiang and R. Banerjee, CrystEngComm, 2014, 16, 4677–4680 RSC .
  61. L. Yang, L. Yu, G. Diao, M. Sun, G. Cheng and S. Chen, J. Mol. Catal. A: Chem., 2014, 392, 278–283 CrossRef CAS .
  62. A. J. Howarth, Y. Liu, P. Li, Z. Li, T. C. Wang, J. T. Hupp and O. K. Farha, Nat. Rev. Mater., 2016, 1, 15018 CrossRef CAS .
  63. J. B. DeCoste, G. W. Peterson, H. Jasuja, T. G. Glover, Y.-g. Huang and K. S. Walton, J. Mater. Chem. A, 2013, 1, 5642–5650 CAS .
  64. N. C. Burtch, H. Jasuja and K. S. Walton, Chem. Rev., 2014, 114, 10575–10612 CrossRef CAS PubMed .
  65. D. J. Tranchemontagne, J. L. Mendoza-Cortés, M. O'Keeffe and O. M. Yaghi, Chem. Soc. Rev., 2009, 38, 1257–1283 RSC .
  66. J. J. Low, A. I. Benin, P. Jakubczak, J. F. Abrahamian, S. A. Faheem and R. R. Willis, J. Am. Chem. Soc., 2009, 131, 15834–15842 CrossRef CAS PubMed .
  67. M. Yoshida and M. Ihara, Chem. – Eur. J., 2004, 10, 2886–2893 CrossRef CAS PubMed .
  68. S. S. Zhang, J. Power Sources, 2006, 162, 1379–1394 CrossRef CAS .
  69. C. Martín, G. Fiorani and A. W. Kleij, ACS Catal., 2015, 5, 1353–1370 CrossRef .
  70. W. Peppel, Ind. Eng. Chem., 1958, 50, 767–770 CrossRef CAS .
  71. W. K. Cline, US Pat., 2667497, 1954 Search PubMed.
  72. X. Jiang, F. Gou, F. Chen and H. Jing, Green Chem., 2016, 3567 RSC .
  73. T. Ema, Y. Miyazaki, J. Shimonishi, C. Maeda and J.-y. Hasegawa, J. Am. Chem. Soc., 2014, 136, 15270–15279 CrossRef CAS PubMed .
  74. K. Motokura, S. Itagaki, Y. Iwasawa, A. Miyaji and T. Baba, Green Chem., 2009, 11, 1876–1880 RSC .
  75. B. Ochiai and T. Endo, J. Polym. Sci., Part A: Polym. Chem., 2007, 45, 5673–5678 CrossRef CAS .
  76. A. R. Hajipour, Y. Heidari and G. Kozehgary, RSC Adv., 2015, 5, 22373–22379 RSC .
  77. R. Babu, A. C. Kathalikkattil, R. Roshan, J. Tharun, D.-W. Kim and D.-W. Park, Green Chem., 2016, 18, 232–242 RSC .
  78. C. M. Miralda, E. E. Macias, M. Zhu, P. Ratnasamy and M. A. Carreon, ACS Catal., 2012, 2, 180–183 CrossRef CAS .
  79. J. Tharun, G. Mathai, A. C. Kathalikkattil, R. Roshan, Y.-S. Won, S. J. Cho, J.-S. Chang and D.-W. Park, ChemPlusChem, 2015, 80, 715–721 CrossRef CAS .
  80. A. C. Kathalikkattil, R. Babu, R. K. Roshan, H. Lee, H. Kim, J. Tharun, E. Suresh and D.-W. Park, J. Mater. Chem. A, 2015, 3, 22636–22647 CAS .
  81. O. V. Zalomaeva, A. M. Chibiryaev, K. A. Kovalenko, O. A. Kholdeeva, B. S. Balzhinimaev and V. P. Fedin, J. Catal., 2013, 298, 179–185 CrossRef CAS .
  82. H.-Y. Cho, D.-A. Yang, J. Kim, S.-Y. Jeong and W.-S. Ahn, Catal. Today, 2012, 185, 35–40 CrossRef CAS .
  83. X. Huang, Y. Chen, Z. Lin, X. Ren, Y. Song, Z. Xu, X. Dong, X. Li, C. Hu and B. Wang, Chem. Commun., 2014, 50, 2624–2627 RSC .
  84. Y. Ren, Y. Shi, J. Chen, S. Yang, C. Qi and H. Jiang, RSC Adv., 2013, 3, 2167–2170 RSC .
  85. L. Liu, S.-M. Wang, Z.-B. Han, M. Ding, D.-Q. Yuan and H.-L. Jiang, Inorg. Chem., 2016, 55, 3558–3565 CrossRef CAS PubMed .
  86. J. Kim, S.-N. Kim, H.-G. Jang, G. Seo and W.-S. Ahn, Appl. Catal., A, 2013, 453, 175–180 CrossRef CAS .
  87. S. Van der Perre, T. Van Assche, B. Bozbiyik, J. Lannoeye, D. E. De Vos, G. V. Baron and J. F. M. Denayer, Langmuir, 2014, 30, 8416–8424 CrossRef CAS PubMed .
  88. D. Feng, W.-C. Chung, Z. Wei, Z.-Y. Gu, H.-L. Jiang, Y.-P. Chen, D. J. Darensbourg and H.-C. Zhou, J. Am. Chem. Soc., 2013, 135, 17105–17110 CrossRef CAS PubMed .
  89. G. Férey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surblé and I. Margiolaki, Science, 2005, 309, 2040–2042 CrossRef PubMed .
  90. G.-Y. Hwang, R. Roshan, H.-S. Ryu, H.-M. Jeong, S. Ravi, M.-I. Kim and D.-W. Park, J. CO2 Util., 2016, 15, 123–130 CrossRef CAS .
  91. B. Mousavi, S. Chaemchuen, B. Moosavi, Z. Luo, N. Gholampour and F. Verpoort, New J. Chem., 2016, 40, 5170–5176 RSC .
  92. W. Kleist, F. Jutz, M. Maciejewski and A. Baiker, Eur. J. Inorg. Chem., 2009, 3552–3561 CrossRef CAS .
  93. X. Zhou, Y. Zhang, X. Yang, L. Zhao and G. Wang, J. Mol. Catal. A: Chem., 2012, 361–362, 12–16 CrossRef CAS .
  94. Y.-J. Kim and D.-W. Park, J. Nanosci. Nanotechnol., 2013, 13, 2307–2312 CrossRef CAS PubMed .
  95. D. S. Sholl and R. P. Lively, J. Phys. Chem. Lett., 2015, 6, 3437–3444 CrossRef CAS PubMed .
  96. Z. Fang, B. Bueken, D. E. De Vos and R. A. Fischer, Angew. Chem., Int. Ed., 2015, 54, 7234–7254 CrossRef CAS PubMed .
  97. C. Chizallet, S. Lazare, D. Bazer-Bachi, F. Bonnier, V. Lecocq, E. Soyer, A.-A. Quoineaud and N. Bats, J. Am. Chem. Soc., 2010, 132, 12365–12377 CrossRef CAS PubMed .
  98. A. Schejn, A. Aboulaich, L. Balan, V. Falk, J. Lalevee, G. Medjahdi, L. Aranda, K. Mozet and R. Schneider, Catal. Sci. Technol., 2015, 5, 1829–1839 CAS .
  99. R. R. Kuruppathparambil, T. Jose, R. Babu, G.-Y. Hwang, A. C. Kathalikkattil, D.-W. Kim and D.-W. Park, Appl. Catal., B, 2016, 182, 562–569 CrossRef CAS .
  100. U. Ravon, M. Savonnet, S. Aguado, M. E. Domine, E. Janneau and D. Farrusseng, Microporous Mesoporous Mater., 2010, 129, 319–329 CrossRef CAS .
  101. W. Lu, Z. Wei, Z.-Y. Gu, T.-F. Liu, J. Park, J. Park, J. Tian, M. Zhang, Q. Zhang, T. Gentle Iii, M. Bosch and H.-C. Zhou, Chem. Soc. Rev., 2014, 43, 5561–5593 RSC .
  102. W.-Y. Gao, L. Wojtas and S. Ma, Chem. Commun., 2014, 50, 5316–5318 RSC .
  103. Z. Zhang, S. Xian, Q. Xia, H. Wang, Z. Li and J. Li, AIChE J., 2013, 59, 2195–2206 CrossRef CAS .
  104. J. Liang, R.-P. Chen, X.-Y. Wang, T.-T. Liu, X.-S. Wang, Y.-B. Huang and R. Cao, Chem. Sci., 2017, 8, 1570–1575 RSC .
  105. M. Taherimehr, B. Van de Voorde, L. H. Wee, J. A. Martens, D. E. De Vos and P. Pescarmona, ChemSusChem, 2016 DOI:10.1002/cssc.201601768, n/a-n/a .
  106. Q. Liu, Z.-X. Low, L. Li, A. Razmjou, K. Wang, J. Yao and H. Wang, J. Mater. Chem. A, 2013, 1, 11563–11569 CAS .
  107. Z. Zhao, J. Fan, J. Wang and R. Li, Catal. Commun., 2012, 21, 32–37 CrossRef CAS .
  108. Y. Horiuchi, T. Toyao, M. Saito, K. Mochizuki, M. Iwata, H. Higashimura, M. Anpo and M. Matsuoka, J. Phys. Chem. C, 2012, 116, 20848–20853 CAS .
  109. C. G. Silva, A. Corma and H. Garcia, J. Mater. Chem., 2010, 20, 3141–3156 RSC .
  110. B. Pattengale, S. Yang, J. Ludwig, Z. Huang, X. Zhang and J. Huang, J. Am. Chem. Soc., 2016, 138, 8072–8075 CrossRef CAS PubMed .
  111. D. K. Kanan and E. A. Carter, J. Phys. Chem. C, 2012, 116, 9876–9887 CAS .
  112. S. Hamad, N. C. Hernandez, A. Aziz, A. R. Ruiz-Salvador, S. Calero and R. Grau-Crespo, J. Mater. Chem. A, 2015, 3, 23458–23465 CAS .
  113. R. Grau-Crespo, A. Aziz, A. W. Collins, R. Crespo-Otero, N. C. Hernández, L. M. Rodriguez-Albelo, A. R. Ruiz-Salvador, S. Calero and S. Hamad, Angew. Chem., Int. Ed., 2016, 16012–16016 CrossRef CAS PubMed .
  114. M. Fuentes-Cabrera, D. M. Nicholson, B. G. Sumpter and M. Widom, J. Chem. Phys., 2005, 123, 124713 CrossRef PubMed .
  115. C. H. Hendon, D. Tiana, M. Fontecave, C. m. Sanchez, L. D'arras, C. Sassoye, L. Rozes, C. Mellot-Draznieks and A. Walsh, J. Am. Chem. Soc., 2013, 135, 10942–10945 CrossRef CAS PubMed .
  116. K. T. Butler, C. H. Hendon and A. Walsh, ACS Appl. Mater. Interfaces, 2014, 6, 22044–22050 CAS .
  117. W. Liang, R. Babarao and D. M. D'Alessandro, Inorg. Chem., 2013, 52, 12878–12880 CrossRef CAS PubMed .
  118. Y. Lee, S. Kim, J. K. Kang and S. M. Cohen, Chem. Commun., 2015, 51, 5735–5738 RSC .
  119. M. A. Nasalevich, C. H. Hendon, J. G. Santaclara, K. Svane, B. Van Der Linden, S. L. Veber, M. V. Fedin, A. J. Houtepen, M. A. Van Der Veen and F. Kapteijn, Sci. Rep., 2016, 6, 23676 CrossRef CAS PubMed .
  120. Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li, Angew. Chem., 2012, 51, 3364–3367 CrossRef CAS PubMed .
  121. Y. Lee, S. Kim, H. Fei, J. K. Kang and S. M. Cohen, Chem. Commun., 2015, 51, 16549–16552 RSC .
  122. L. Li, S. Zhang, L. Xu, J. Wang, L.-X. Shi, Z.-N. Chen, M. Hong and J. Luo, Chem. Sci., 2014, 5, 3808–3813 RSC .
  123. C. Wang, Z. Xie, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2011, 133, 13445–13454 CrossRef CAS PubMed .
  124. Y. Liu, Y. Yang, Q. Sun, Z. Wang, B. Huang, Y. Dai, X. Qin and X. Zhang, ACS Appl. Mater. Interfaces, 2013, 5, 7654–7658 CAS .
  125. J. Fan, E.-z. Liu, L. Tian, X.-y. Hu, Q. He and T. Sun, J. Environ. Eng., 2010, 137, 171–176 CrossRef .
  126. K. Fujihara, S. Izumi, T. Ohno and M. Matsumura, J. Photochem. Photobiol., A, 2000, 132, 99–104 CrossRef CAS .
  127. K. Das, S. N. Sharma, M. Kumar and S. De, J. Phys. Chem. C, 2009, 113, 14783–14792 CAS .
  128. R. Li, J. Hu, M. Deng, H. Wang, X. Wang, Y. Hu, H. L. Jiang, J. Jiang, Q. Zhang, Y. Xie and Y. Xiong, Adv. Mater., 2014, 26, 4783–4788 CrossRef CAS PubMed .
  129. D. Sun, Y. Fu, W. Liu, L. Ye, D. Wang, L. Yang, X. Fu and Z. Li, Chemistry, 2013, 19, 14279–14285 CrossRef CAS PubMed .
  130. M. B. Chambers, X. Wang, N. Elgrishi, C. H. Hendon, A. Walsh, J. Bonnefoy, J. Canivet, E. A. Quadrelli, D. Farrusseng, C. Mellot-Draznieks and M. Fontecave, ChemSusChem, 2015, 8, 603–608 CrossRef CAS PubMed .
  131. S. Wang, W. Yao, J. Lin, Z. Ding and X. Wang, Angew. Chem., Int. Ed., 2014, 53, 1034–1038 CrossRef CAS PubMed .
  132. Q. Li and H. Kim, Fuel Process. Technol., 2012, 100, 43–48 CrossRef CAS .
  133. R. Huang, Y. Peng, C. Wang, Z. Shi and W. Lin, Eur. J. Inorg. Chem., 2016, 4358–4362 CrossRef CAS .
  134. Q. Zhang, C. F. Lin, B. Y. Chen, T. Ouyang and C. T. Chang, Environ. Sci. Technol., 2015, 49, 2405–2417 CrossRef CAS PubMed .
  135. Y. Hori, A. Murata and R. Takahashi, J. Chem. Soc., Faraday Trans. 1, 1989, 85, 2309–2326 RSC .
  136. J. Hu, J. Wang, T. H. Nguyen and N. Zheng, Beilstein J. Org. Chem., 2013, 9, 1977–2001 CrossRef PubMed .
  137. S. Han, Y. Huang, T. Watanabe, S. Nair, K. S. Walton, D. S. Sholl and J. Carson Meredith, Microporous Mesoporous Mater., 2013, 173, 86–91 CrossRef CAS .
  138. W. P. Mounfield III, C. Han, S. H. Pang, U. Tumuluri, Y. Jiao, S. Bhattacharyya, M. R. Dutzer, S. Nair, Z. Wu and R. P. Lively, J. Phys. Chem. C, 2016, 120, 27230–27240 Search PubMed .
  139. S. Bhattacharyya, S. H. Pang, M. R. Dutzer, R. P. Lively, K. S. Walton, D. S. Sholl and S. Nair, J. Phys. Chem. C, 2016, 120, 27221–27229 CAS .
  140. Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Electrochim. Acta, 1994, 39, 1833–1839 CrossRef CAS .
  141. Y. Hori, K. Kikuchi and S. Suzuki, Chem. Lett., 1985, 1695–1698 CrossRef CAS .
  142. K. W. Frese, J. Electrochem. Soc., 1991, 138, 3338–3344 CrossRef CAS .
  143. C. W. Li, J. Ciston and M. W. Kanan, Nature, 2014, 508, 504–507 CrossRef CAS PubMed .
  144. K. J. P. Schouten, Y. Kwon, C. J. M. van der Ham, Z. Qin and M. T. M. Koper, Chem. Sci., 2011, 2, 1902–1909 RSC .
  145. D. Kim, J. Resasco, Y. Yu, A. M. Asiri and P. Yang, Nat. Commun., 2014, 5, 4948 CrossRef CAS PubMed .
  146. J. Albo, M. Alvarez-Guerra, P. Castaño and A. Irabien, Green Chem., 2015, 17, 2304–2324 RSC .
  147. M. Gattrell, N. Gupta and A. Co, J. Electroanal. Chem., 2006, 594, 1–19 CrossRef CAS .
  148. Y. Hori, I. Takahashi, O. Koga and N. Hoshi, J. Phys. Chem. B, 2002, 106, 15–17 CrossRef CAS .
  149. K. P. Kuhl, T. Hatsukade, E. R. Cave, D. N. Abram, J. Kibsgaard and T. F. Jaramillo, J. Am. Chem. Soc., 2014, 136, 14107–14113 CrossRef CAS PubMed .
  150. R. Hinogami, S. Yotsuhashi, M. Deguchi, Y. Zenitani, H. Hashiba and Y. Yamada, ECS Electrochem. Lett., 2012, 1, H17–H19 CrossRef CAS .
  151. M. Jahan, Z. Liu and K. P. Loh, Adv. Funct. Mater., 2013, 23, 5363–5372 CrossRef CAS .
  152. X. Wang, J. Zhou, H. Fu, W. Li, X. Fan, G. Xin, J. Zheng and X. Li, J. Mater. Chem. A, 2014, 2, 14064–14070 CAS .
  153. L. Yang, C. Xu, W. Ye and W. Liu, Sens. Actuators, B, 2015, 215, 489–496 CrossRef CAS .
  154. W. Xia, A. Mahmood, R. Zou and Q. Xu, Energy Environ. Sci., 2015, 8, 1837–1866 CAS .
  155. T. Maihom, S. Wannakao, B. Boekfa and J. Limtrakul, J. Phys. Chem. C, 2013, 117, 17650–17658 CAS .
  156. R. Senthil Kumar, S. Senthil Kumar and M. Anbu Kulandainathan, Electrochem. Commun., 2012, 25, 70–73 CrossRef CAS .
  157. I. Hod, M. D. Sampson, P. Deria, C. P. Kubiak, O. K. Farha and J. T. Hupp, ACS Catal., 2015, 5, 6302–6309 CrossRef CAS .
  158. A. Aijaz, N. Fujiwara and Q. Xu, J. Am. Chem. Soc., 2014, 136, 6790–6793 CrossRef CAS PubMed .
  159. S. Zhao, H. Yin, L. Du, L. He, K. Zhao, L. Chang, G. Yin, H. Zhao, S. Liu and Z. Tang, ACS Nano, 2014, 8, 12660–12668 CrossRef CAS PubMed .
  160. G. Ghadimkhani, N. R. de Tacconi, W. Chanmanee, C. Janaky and K. Rajeshwar, Chem. Commun., 2013, 49, 1297–1299 RSC .
  161. P. Li, H. Jing, J. Xu, C. Wu, H. Peng, J. Lu and F. Lu, Nanoscale, 2014, 6, 11380–11386 RSC .

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6mh00484a

This journal is © The Royal Society of Chemistry 2017