Chromium terephthalate metal–organic framework MIL-101: synthesis, functionalization, and applications for adsorption and catalysis

Samiran Bhattacharjeea, Chao Chenab and Wha-Seung Ahn*a
aDepartment of Chemistry and Chemical Engineering, Inha University, Incheon, 402-751, Republic of Korea. E-mail: whasahn@inha.ac.kr; Fax: +82 32 872 0959; Tel: +82 32 860 7466
bCollege of Chemistry and Chemical Engineering, Xinyang Normal University, Xinyang 464000, Henan Province, China

Received 27th August 2014 , Accepted 3rd October 2014

First published on 6th October 2014


Abstract

The chromium terephthalate metal–organic framework, MIL-101 (MIL, Matérial Institut Lavoisier), is comprised of trimeric chromium(III) octahedral clusters interconnected by 1,4-benzenedicarboxylates, resulting in a highly porous 3-dimentional structure. The large pores (29 and 34 Å) and high BET surface area (>3000 m2 g−1) with a huge cell volume (≈702[thin space (1/6-em)]000 Å3) together with the coordinatively unsaturated open metal sites that can be subjected to diverse post-synthesis functionalization or guest encapsulation, and excellent hydrothermal/chemical stability, make MIL-101 particularly attractive for applications, such as selective gas adsorption/separation, energy storage and heterogeneous catalysis. This paper reviews the current status of research and development on the synthesis, functionalization and applications of MIL-101 for adsorption/catalytic reactions.


1. Introduction

Metal–organic frameworks (MOFs) are a class of crystalline porous hybrid materials assembled by the bonding of metal ions or clusters linked with polyfunctional organic molecules to form infinite one-, two or three dimensional frameworks. These porous materials are generally constructed from metal ions or clusters of zinc, copper, chromium, aluminium, zirconium and other metals, and organic components of carboxylates or N-containing aromatic groups.1–3 The most interesting properties of MOFs are their well-ordered tunable porous structures with a wide range of pore sizes and exceptional textural properties, such as high surface areas and high pore volumes, which enable a range of applications for gas adsorption/separation,4–9 catalysis,10 and drug delivery.11 In recent years, several extensive review articles dealing with the synthesis, structural description, surface modification, and adsorption/catalytic application of various MOF materials have appeared.12–24

In 2004, Férey et al. prepared MIL-100 (MIL, Matérial Institut Lavoisier) hydrothermally using combined computer simulations and targeted chemistry. The framework of MIL-100 was composed of trimeric chromium(III) interconnected by benzene-1,3,5-tricarboxylate (BTC) anions (Fig. 1), which was the first example of MOFs with crystalline walls having a unique hierarchical pore system and high Langmuir surface area.25 MIL-101 with a chemical composition of {Cr3F(H2O)2O(BDC)3·nH2O} (n ∼ 25; 1,4-benzenedicarboxylate (BDC)) and superior physicochemical properties emerged soon after.26 The robust framework of MIL-101 was comprised of similar trimeric chromium(III) octahedral clusters interconnected by BDC molecules resulting in an augmented MTN zeotype structure (Fig. 1).


image file: c4ra11259h-f1.tif
Fig. 1 Structural representation of MIL-100 and MIL-101.25,26

This structure was comprised of two types of mesoporous cages with diameters of ∼29 and 34 Å accessible through two types of microporous windows (the smaller cages have pentagonal windows with a free opening of ∼12 Å, while the larger cages possess both pentagonal and hexagonal windows with a ∼14.7 Å by 16 Å free aperture) (Fig. 2).


image file: c4ra11259h-f2.tif
Fig. 2 Schematic diagram of the pentagonal and hexagonal windows of MIL-101.26

The material exhibits excellent stability against moisture and other chemicals, and the terminal water molecules in MIL-101 can be removed by heating in air or under vacuum at 423 K, which generates two coordinatively unsaturated open metal sites (CUS) per trimeric Cr(III) octahedral cluster. In addition to its highly porous nature, MIL-101 has attracted considerable attention because functional modifications on MIL-101 can be achieved easily either directly using a functionalized ligand during the synthesis or indirectly via the diverse post-synthesis chemical treatment on the CUS or on the organic linkers.

Among the MOFs known, MIL-101 is one of the most promising porous materials for future energy and environmental applications, surpassing MOF-5 or HKUST-1, owing to its superior physicochemical properties including high hydrothermal/chemical stability and desirable textural properties. This paper reviews the recent progress on the synthesis, functionalization and applications in adsorption/catalysis of the chromium terephthalate MIL-101 since Hong et al. first reported the versatility of MIL-101 in their comprehensive experiments in 2009.27 Readers from engineering disciplines, particularly newcomers to MOFs science, will gain a clearer perspective on the potential of the MOF materials for future sustainability by concentrating on the synthesis and applications of a specific model compound out of the myriad of MOF materials available.

2. Synthesis and characterization

2.1 Hydrothermal synthesis

Férey et al.26 reported the preparation of a porous chromium terephthalate metal–organic framework, MIL-101, and a large number of synthesis studies28–37 were followed afterwards under different synthesis conditions (additive, temperature, time, and with support materials etc.), as listed in Table 1. MIL-101 can also be prepared by microwave heating38–41 or via a drygel conversion method.42
Table 1 MIL-101 synthesis studiesa
Method Synthesis conditions Textural properties Comments Ref.
Synthesis medium Temp. (K) Time (h) SBET (m2 g−1) Vpore (cm3 g−1)
a TMAOH (tetramethylammonium hydroxide), CTAB (cetytrimethylammonium bromide), EG (expanded graphite), 4-NTm (4-nitroimidazole).b Langmuir surface area.
Hydrothermal H2O/HF 493 8 4100 2.0 Synthesis and structure analysis 26
Incorporation of Keggin polyanions
H2O/TMAOH 453 24 3197 1.73 Synthesis in an alkaline medium 28
H2 storage
H2O/HF 493 8 3115 1.58 Synthesis with temperature-programming 29
H2O/CH3COOH 473 8 2736 1.50 Synthesis in acetic acid instead of HF 30
Volatile organic compounds (VOC) adsorption
H2O 491 18 3460 HF-free synthesis and comparison (conventional vs. microwave), phosphotungstic acid composite 31
Baeyer condensation and epoxidation reaction
H2O/stearic acid 453 4 2691 2.95 Synthesis of MIL-101 (19–84 nm in size) using a monocarboxylic acids 32
CO2 adsorption studies
H2O/HF/CTAB 493 8 846   Synthesis of hierarchically mesostructured MIL-101 with different morphologies using cationic surfactant CTAB 33
Dye removal
H2O/HF/EG 493 2 3751 1.8 Synthesis of nano-sized MIL-101 crystals (∼200 nm) using expanded graphite as a structure-directing template 34
H2O/4-NTm 423 24 2542 1.59 Synthesis using 4-nitroimidazole 35
H2O 493 8 Thin films on alumina via in situ seeding of nanoparticles by the use of dimethylacetamide 36
H2O 453 4 Thin films on alumina via dip-coating in the presence of polyethyleneimine with controlled thickness (100–545 μm) 37
Microwave H2O/HF 483 0.67 3900 2.3 Synthesis 38
Benzene sorption
H2O 473 0.02 4200b HF-free synthesis and thin films on silicon wafer by a dip-coating route 39
Water and alcohol sorption properties
H2O/HF 493 1 3054 2.01 Synthesis 40
Studies on adsorption and diffusion of benzene
H2O/NaOH 483 0.25 3223 1.57 Comparison (conventional vs. microwave) 41
Dry-gel conversion H2O/HF 493 12 4164 2.77 Detailed synthesis studies 42
Synthesis of MIL-101-NO2 and incorporation of phosphotungstic acid


MIL-101 can be prepared hydrothermally using a chromium salt and H2BDC with the aid of a small amount of HF in an autoclave under autogenous pressure conditions.26 In a typical synthesis, Cr(NO3)3·9H2O (0.400 g, 1 mmol), H2BDC (0.166 g, 1 mmol) and HF (0.2 mL, 1 mmol) in water (4.8 mL, 265 mmol) are heated at 493 K for 8 h, which forms a crystalline green powder. The solid product contains impurities of unreacted or recrystallized H2BDC present both outside and within the pores of MIL-101. The formation of impurity phases normally result in a low product yield. The removal of impurities is difficult, and rigorous purification steps are normally required. Initially, the green solid is filtered through a large pore fritted glass filter (no. 2) and then through small pore (no. 5) paper filter to remove the large H2BDC crystals. This is followed by multi-step solvent treatments using water, hot ethanol and DMF, and finally aqueous NH4F solutions.27 Several attempts have been made to increase the purity of the MIL-101 product by temperature-programmed synthesis to induce the formation of large residual carboxylic acid impurity crystals that can be isolated more easily using a large pore fritted glass filter29 or the addition of tetramethylammonium hydroxide (TMAOH) to promote the dissolution of H2BDC.28

Thermogravimetric analysis (TGA) in air showed that MIL-101 is stable up to 548 K. The N2 sorption isotherm was of type I with secondary uptakes at P/P0 ∼ 0.1 and at P/P0 ∼ 0.2, which reflects the two types of microporous windows in MIL-101. The BET surface area and pore volume of the sample as high as ∼4100 m2 g−1 and 2.0 cm3 g−1, respectively, were reported.26 On the other hand, a more typical BET surface area of MIL-101 determined by other researchers was in the range, ca. 2500–3200 m2 g−1.

Several new synthetic strategies have been suggested to prepare MIL-101 with higher product purity without the use of toxic and highly corrosive HF. The material was synthesized in an alkali medium with a reaction mixture of Cr(NO3)3·9H2O[thin space (1/6-em)]:[thin space (1/6-em)]H2BDC (1[thin space (1/6-em)]:[thin space (1/6-em)]1) and TMAOH (0.05 mol L−1) at 453 K for 24 h.28 The resulting MIL-101 had a BET surface area of 3197 m2 g−1 after washing with distilled water in high product yield (∼88%). MIL-101 with a BET surface area and pore volume of 2736 m2 g−1 and 1.50 cm3 g−1, respectively, was also prepared using acetic acid instead of HF as an additive at 473 K, followed by a solvent treatment with DMF and ethanol.30 Bromberg et al. synthesized MIL-101 from a reaction mixture of Cr(NO3)3·9H2O and H2BDC in an aqueous medium without acid or alkali at 491 K for 18 h with a 64% product yield.31 The BET surface area of the sample was found to be 3460 m2 g−1 after a DMF treatment at 343 K for 18 h.

Jiang et al. reported the size-controlled synthesis of MIL-101 (19 nm to 84 nm in size) using a variety of monocarboxylic acids, such as stearic acid, benzoic acid, 4-methoxybenzoic acid, 4-nitrobenzoic acid, and perfluorobenzoic acid, as a mediator in a reaction mixture of Cr(NO3)3·9H2O[thin space (1/6-em)]:[thin space (1/6-em)]H2BDC[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1680) at 453 K in ∼49% yield.32 The BET surface area and total pore volume of the samples were found to be 2646–2920 m2 g−1 and 2.33–2.95 cm3 g−1, respectively. Nano-sized MIL-101 crystals (∼200 nm in size) with a BET surface area of 3751 m2 g−1 and a total pore volume of 1.8 cm3 g−1 was also prepared using an expanded graphite-template method at a shorter reaction time of 2 h at 493 K.34 Huang et al. reported the preparation of hierarchically mesostructured MIL-101 using the cationic surfactant, cetyltrimethylammonium bromide (CTAB) as a supramolecular template to produce nanoparticles (120–200 nm) with different morphologies, such as octahedral microcrystals, nanospheres, and nanoflowers.33 The cationic surfactant played a key role in controlling the morphology of the resulting nanoparticles.

MIL-101 in film form was also prepared. Férey et al. immersed a silicon wafer into a nanoparticle suspension of MIL-101 in ethanol to form a thin film with thicknesses up to 160 nm via a dip-coating method at room temperature.39 Jiang et al. reported the synthesis of 2.5 μm thick MIL-101 films on an alumina support by immersing adimethylacetamide (DMA)-wetted alumina plate into a reaction mixture of Cr(NO3)3·9H2O and H2BDC in water at 493 K for 8 h.36 The authors expanded the concept and reported the synthesis of MIL-101 and MIL-101-NH2 films on alumina substrates by immersing alumina into a nanoparticle suspension of MIL-101 in ethanol in the presence of polyethylenimine (PEI) via dip-coating to furnish crack-free films of MIL-101-PEI-n and MIL-101-NH2-PEI-n with a controlled thickness.37 Scanning electron microscopy (SEM) showed that the film thickness was in the range, 100 to 545 nm.

2.2 Microwave synthesis

Microwave-assisted synthesis techniques for the preparation of nanoporous materials under hydrothermal conditions have the advantages of (1) fast crystallization, (2) phase selectivity, (3) narrow particle size distribution, and (4) facile morphology control.43–46

Jhung et al. evaluated the microwave synthesis of MIL-101, and observed a significant decrease in crystal size (40–90 vs. 250–600 nm) and synthesis time compared to hydrothermal electric heating with a corresponding BET surface area and pore volume of 3900 m2 g−1 and 2.3 cm3 g−1, respectively.38 MIL-101 was synthesized within 60 min at 483 K under microwave irradiation at 600 W, in which the molar composition of the substrate mixture was Cr(NO3)3·9H2O[thin space (1/6-em)]:[thin space (1/6-em)]H2BDC[thin space (1/6-em)]:[thin space (1/6-em)]HF[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]280). The material was also synthesized by microwave heating in an aqueous medium without HF at 473 K to form non-aggregated nanoparticles, 22 ± 5 nm in size, with a Langmuir surface area of ∼4200 m2 g−1.39 The nano-sized MIL-101 was synthesized using aqueous NaOH without HF, producing crystals, ∼50 nm in size, with a 37% product yield at 483 K.41

2.3 Dry gel conversion synthesis

Dry gel conversion (DGC) applied to the synthesis of inorganic materials, such as zeolites normally have the following advantages: (1) minimization of waste disposal; (2) reduction in reaction volume; and (3) complete conversion of gel to a uniform crystalline product in high yield. Recently, Ahn et al. reported the drygel conversion synthesis of MIL-101.42 Initially a Cr(NO3)3·9H2O granular precursor was ball-milled, which was then mixed with H2BDC and ground again for 15 min. The finely ground mixture of the substrates (140 mg) was placed on a holed Teflon-plate inside a Teflon-lined stainless steel autoclave. H2O (6.25 mL) and HF (1.0 mg, 47% aq.) were added to the bottom of the autoclave and the entire assembly was heated to 493 K for 12 h. The DGC synthesis produced MIL-101 with a high BET surface area (4164 m2 g−1) after a single wash with distilled water in high product yield (∼90%). SEM, X-ray diffraction (XRD) and Fourier transform infrared (FT-IR) analysis confirmed that no recrystallized H2BDC formed during the DGC process. The authors also conducted the synthesis of a functionalized MIL-101 using 2-nitroterephthalic acid as an organic ligand and the encapsulation of phosphotungstic acid (H3PW) into MIL-101 under DGC reaction conditions without using HF. The method might be useful for the large scale production of MIL-101 for industrial use due to the high product yield and simple washing procedure.

3. Direct and post-synthesis organic functionalization

The functionalization of MOF structures is often necessary to create active sites for a range of catalytic/adsorption applications. Table 2 lists the cases of organic-functionalization carried out on MIL-101.
Table 2 Organic functionalization and post-synthesis modification studies of MIL-101a
Organic linker Reaction conditions Textural property Comments Ref.
Temp. (K) Time (h) SBET (m2 g−1) Vpore (cm3 g−1)
a PSM (post-synthesis modification), TMAOH (tetramethylammonium hydroxide), PEI (polyethyleneimine), NCSs (nitrogen-containing compounds), SCCs (sulphur-containing compounds).
2-Nitroterephthalic acid 453 96 1307–1425 Direct synthesis of MIL-101-NO2 in aqueous medium, PSM of MIL-101 to MIL-101-NO2 using nitrating acid and converted to MIL-101-NH2 and urea derivative 47
Monosodium 2-sulfoterephthalic acid 453 144 1915 Direct synthesis of MIL-101-SO3H in HCl 48
Cellulose hydrolysis
2-Nitroterephthalic acid 453 144 2146 1.19 Direct synthesis of MIL-101-NO2 in HCl 49
Water sorption properties
2-Aminoterephthalic acid 423 12 1675 Direct synthesis of MIL-101-NH2 in aqueous NaOH 50
CO2, CH4 adsorption studies
Substituted terephthalic acid 453 96 991–2282 Synthesis of single- and mixed-linker derivatives using high-throughput methods 51
PSM of MIL-101-Br-NO2, MIL-101-Br-NH2
Terephthalic acid and 2-(diphenylphosphino)terephthalic acid 483 24 3020 Synthesis of MIL-101-PPh2 using mixed-linkers in the presence of TMAOH 52
Monosodium 2-sulfoterephthalic acid 453 24 1754 0.91 Direct synthesis of MIL-101-SO3H 53
Acetalization of aldehydes with diols
2-Aminoterephthalic acid 403 24 2070 2.26 Direct synthesis of MIL-101-NH2 and PSM of MIL-101-NH2 via tandem diazotisation process to produce azo dye-, iodo- and fluoro-functionalized MIL-101 54
CO2 sorption studies
Terephthalic acid 493 9 638 0.18 Grafting NH2 and SO3H groups on organic ligand and CUS by three-step PSM of MIL-101 55
One-pot deacetalization-nitroaldol reaction
Monosodium 2-sulfoterephthalic acid and 2-nitroterephthalic acid 453 144 2190 1.37 Direct synthesis of MIL-101-NO2-SO3H and PSM by the reduction of the NO2 groups to MIL-101-NH2-SO3H 56
One-pot deacetalization-nitroaldol reaction
Terephthalic acid 493 8 3555 2.07 Grafting of amines (ethylenediamine, diethylenetriamine and 3-aminopropyltrialkoxysilane) onto CUS Knoevenagel condensation, Heck coupling reaction 57
Terephthalic acid 493 8 3125–183 1.63–0.10 Incorporation of PEI into MIL-101 58
CO2 sorption studies
2-Aminoterephthalic acid 423 6–16 2297–96 2.12–0.37 Direct synthesis of MIL-101-NH2 and post-synthesis PEI incorporation 59
Separation of CO2 from CO2/CH4 mixture
Terephthalic acid 493 8 3555 1.75 In situ formation of MIL-101 on activated carbon 60
Hydrogen storage studies
Terephthalic acid 493 6 Incorporation of H2SO4 and H3PO4 into as-prepared MIL-101 in an aqueous medium 61
High proton-conducting properties at moderate temperature
Terephthalic acid 493 10 3858 2.09 In situ formation of MIL-101 on graphite oxide 62
Adsorption properties of NCSs and SCCs


Nitro-functionalized MIL-101 (MIL-101-NO2) was synthesized using 2-nitroterephthalic acid (H2BDC-NO2) and a mixture of CrCl3[thin space (1/6-em)]:[thin space (1/6-em)]H2BDC-NO2[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1.6[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]278) at 453 K in 96 h.47 The BET surface area of the sample was 1425 m2 g−1. Akiyama et al. reported the preparation of sulfonic acid functionalized MIL-101 (MIL-101-SO3H) using monosodium 2-sulfoterephthlicacid (H2BDC-SO3Na), in which CrO3, H2BDC-SO3Na, HCl, and water were treated hydrothermally at 453 K for 6 days.48 The same group also reported the synthesis of MIL-101-NO2 under similar reaction conditions, as applied for MIL-101-SO3H.49 The BET surface area of MIL-101-SO3H and MIL-101-NO2 were 1915 and 2146 m2 g−1, respectively. Lin et al. reported the synthesis of amine functionalized MIL-101 (MIL-101-NH2) using 2-aminoterephthalic acid (H2BDC-NH2) in an aqueous sodium hydroxide solution at a lower temperature of 423 K.50 The material showed a mean particle size of ca. 50 nm and a BET surface area of 1675 m2 g−1.

Lammert et al. reported the synthesis of single- and mixed-linker functionalized MIL-101 derivatives using a high-throughput synthesis technique.51 Fig. 3 shows the linkers used in high-throughput screening. They described the effects of three different metal sources (CrO3, CrCl3 and Cr(NO3)3·9H2O); highly crystalline MIL-101 materials were produced using CrCl3 as the metal source and 2-bromoterephthalic or 2-nitroterephthalic acid as the mixed-linker components. Recently, Morel et al. synthesized phosphine-functionalized MIL-101 in an aqueous solution of TMAOH using 2-(diphenylphosphino)terephthalic acid and H2BDC at a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]6 at 483 K with high crystallinity and phase purity.52 The BET surface area of the sample was 3020 m2 g−1.


image file: c4ra11259h-f3.tif
Fig. 3 Structure of the organic linkers used in the high-throughput synthesis technique.51

Table 2 also lists the results of the investigations into the post-synthesis modification (PSM) of MIL-101. Bernt et al. converted MIL-101 to MIL-101-NO2 using an acid mixture (conc. HNO3 and conc. H2SO4) with ice cooling.47 The author reported the PSM of MIL-101-NO2 through the reduction of the NO2 group with SnCl2·2H2O to NH2, which was reacted further with ethyl isocyanate to produce a urea derivative (MIL-101-UR2), as described in Fig. 4.


image file: c4ra11259h-f4.tif
Fig. 4 Schematic representation of PSM of MIL-101 to MIL-101-NO2, MIL-101-NH2 and urea derivative.47

The PSM of MIL-101-NH2 via a tandem diazotization process to furnish azo dye-, iodo- and fluoro-functionalized MIL-101 were also reported (Fig. 5).54


image file: c4ra11259h-f5.tif
Fig. 5 Schematic representation of PSM of MIL-101-NH2 through a tandem diazotization process.54

MIL-101-NH2 was reacted with NaNO2 in the presence of HCl to form an arenediazonium chloride salt as an intermediate, which was then converted to azo dye- and iodo-functionalized MOF (MIL-101-azo and MIL-101-I) by the addition of an aqueous solution of C6H5OH and KI, respectively, and a diazonium tetrafluoroborate salt was formed from a reaction mixture of NaNO2 and HBF4, which was then heated to 373 K to form a fluoro-functionalized MIL-101 (MIL-101-F) by eliminating nitrogen.

A site-isolated acid–base bifunctional MIL-101 was obtained by the three-step PSM of MIL-101: (i) BOC (N-tert-butoxycarbonyl)-protected amino group grafting on CUS, (ii) direct sulfonation of the phenylene backbone with chlorosulfonic acid, and (iii) BOC de-protection of the amino groups by a thermal treatment.55 The material, MIL-101-NH2-SO3H, showed a lower BET surface area (638 m2 g−1) than pure MIL-101 (1333 m2 g−1; suspected of incomplete purification). Recently, Ahn et al. reported the synthesis of MIL-101-NH2-SO3H using a simple two-step process: (i) direct hydrothermal synthesis of MIL-101-NO2–SO3H using H2BDC-SO3Na and H2BDC-NO2 organic linkers and (ii) one-step PSM by reduction of the NO2 groups to NH2 using SnCl2, as described in Fig. 6,56 which achieved significant improvement in the surface area (2190 m2 g−1).


image file: c4ra11259h-f6.tif
Fig. 6 Schematic representation of the synthesis of bifunctional MIL-101-NH2–SO3H.56

Goesten et al. reported the PSM of MIL-101 through the chloromethylation of aromatic rings with methoxyacetal chloride in the presence of aluminium chloride to produce CM-MIL-101, followed by the introduction of diphenylphosphine (PPh2) to form a multifunctional material, PPh2-MIL-101 (Fig. 7).63


image file: c4ra11259h-f7.tif
Fig. 7 Schematic representation of the chloromethylation of the aromatic rings of MIL-101 and the introduction of diphenylphosphine.63

The chloromethylene functional group in CM-MIL-101 was confirmed by 1H magic angle spinning nuclear magnetic resonance (MAS NMR) and FT-IR spectroscopy. The PPh2-MIL-101 was highly porous with a BET surface area of 1800 m2 g−1. The CM-MIL-101 was treated with catalytically active tetra(4-pyridyl)porphyrinatomanganese(III) acetate (Mn(TPyP)OAc) to form a hybrid material, Mn(TPyP)OAc/CM-MIL-101 by the elimination of a chlorine atom (Fig. 8).64 FT-IR, diffuse reflectance UV-Vis spectroscopy (UV-Vis DRS), inductively coupled plasma (ICP) and SEM confirmed the post-synthesis metalation. Nguyen et al.65 reported the binding of vanadyl acetylacetonate to MIL-101 via a two-step PSM route: (i) the incorporation of dopamine (dop) to the unsaturated Cr(III) sites and (ii) reaction with VO (acetylacetonate)2 to form a mono-catecholate vanadium(IV) oxo compound, V(dop)-MIL-101, which had a BET surface area of ∼1770 m2 g−1.


image file: c4ra11259h-f8.tif
Fig. 8 Post-synthetic metalation of CM-MIL-101.64

Several nitrogen-containing organic precursors have been used to prepare amine-grafted MIL-101 composites.57–59,66–68 Hwang et al. reported the grafting of amine molecules onto CUS in MIL-101, as described in Fig. 9.57


image file: c4ra11259h-f9.tif
Fig. 9 Schematic representation of ethylenediamine ligand grafting onto CUS.57

The materials were prepared by the activation of MIL-101 at 423 K for 12 h under vacuum, followed by heating with amine precursors, such as ethylenediamine (ED), diethylenetriamine (DETA) or 3-aminopropyltrialkoxysilane (APS), in toluene under reflux. FT-IR, CO adsorption, BET surface area measurements, and elemental analysis confirmed the successful grafting of the amines onto CUS in MIL-101. Ahn et al. similarly reported the grafting of DETA onto the open metal sites in MIL-101.67 Dialkylaminopyridine ligands (Fig. 10) including 4-dimethylaminopyridine, 4-pyrrolidinopyridine, 4-morpholinopyridine, and 4-(3-aminopropyl)-methylaniline were also incorporated onto the open metal sites in MIL-101 by heating MIL-101 with the precursors in toluene under reflux.68 The polyethyleneimine (PEI)-incorporated MIL-101-NH2 was obtained using a two-step process: (i) direct hydrothermal synthesis of MIL-101-NH2 using 2-aminoterephthalic acid followed by (ii) refluxing it with PEI in methanol at 383 K for 12 h under vacuum conditions.59


image file: c4ra11259h-f10.tif
Fig. 10 Structure of the dialkylaminopyridine ligands.68

4. Guest encapsulation

4.1 Encapsulation of polyoxometalate (POM) and metal phthalocyanine

MIL-101 was subjected to post-synthesis guest encapsulation by making use of its high thermal and chemical stability against water and other common organic solvents, as well as meso-sized cages for the incorporation of catalytically active guest molecules. POM are polyatomic compounds that are composed of bulky clusters of transition metal oxide anions. These bifunctional redox materials have a wide range of applications in material science and catalysis. Metal phthalocyanine (MPc) compounds also show good catalytic activity for the oxidation of organic compounds.

Férey et al. first reported the incorporation of a large Keggin polyanion, PW11O407−, into MIL-101 from an aqueous solution at room temperature, and estimated that approximately five highly charged Keggin moieties can be incorporated within the large cages of MIL-101 (volume of the Keggin unit and large cages of MIL-101 were 2250 and 20[thin space (1/6-em)]600 A3, respectively).26 The encapsulated material showed a significantly lower surface area than pure MIL-101. Moreover, XRD, N2 sorption measurements, TGA, 31P solid state NMR, and FT-IR spectroscopy confirmed the incorporation of Keggin ions within the pores. Several other studies also reported the encapsulation of polyoxotungstates,31,69–74 Ti/Co-monosubstituted Keggin heteropolyanions,75 polyoxovanadate,76 lanthanopolyoxometalates (Ln = Eu3+ and Sm3+),77 zinc-monosubstituted Keggin heteropolyanion,78 and boronperoxotungstate79 into MIL-101 using impregnation or one-pot hydrothermal synthesis routes.

The encapsulation of large metal phthalocyanine (MPc), (MPcF16, M = Fe, Ru) and (FePctBu4)2N complexes into MIL-101 was conducted using a wet infiltration method.80 N2 physisorption measurements and energy-dispersive X-ray spectroscopy (EDS) analysis confirmed the homogeneous distribution of the FePcF16 or RuPcF16 complexes inside the MIL-101 matrix, whereas the bulky (FePctBu4)2N complex was found adsorbed on the outer surfaces. FePcS/MIL-101 (FePcS = irontetrasulfophthalocyanine) was prepared by the adsorption of FePcS in an aqueous medium,81 in which FePcS was deposited on both the internal and external surface of MIL-101 crystals.

Anbia et al. reported the in situ synthesis of MWCNT@MIL-101 (MWCNT = multi-walled carbon nanotube) from a reaction mixture of Cr(NO3)3·9H2O and H2BDC in HF/H2O at 493 K.82 The composite material, AC@MIL-101 (AC = activated carbon) was also synthesized using similar route by adding AC at different loadings in situ during the synthesis of MIL-101.60 The hybrid proton-conducting materials, H2SO4@MIL-101 and H3PO4@MIL-101, were synthesized by introducing the as-prepared MIL-101 to an aqueous solution of H2SO4 (2.7 M) or H3PO4 (2.6 M) at room temperature.61 The composite material, GO/MIL-101 (GO = graphite oxide), was prepared by the in situ formation of MIL-101 on GO by introducing the as-prepared GO into a mixture of Cr(NO3)3·9H2O and H2BDC in HF/H2O at 493 K for 10 h.62,83 The surface area of the GO/MIL-101 was found to be dependent on the amount of GO used. A small amount of GO (ca. 0.25%) exhibited a higher value than MIL-101, but larger amounts of GO (>0.5%) reduced the porosity.62

4.2 Encapsulation of metals nanoparticles (NPs)

The encapsulation of metal NPs into the interior cavities of MOF crystals has attracted significant attention in recent years because of its potential applications in energy storage and catalysis. In most cases, NPs are formed both inside the pores and on the external surface on the MOFs. In metal NPs-based catalysts, it appears that the catalytic activity and selectivity can be controlled by the size of the catalytically active phase and by the nature of the supports. Considerable effort has been made to improve the catalytic performance of MIL-101 through the immobilization of noble metals or non-noble metal NPs, as listed in Table 3.
Table 3 Encapsulation of metal nanoparticles over MIL-101a
Preparation method Synthesis conditions NP size (nm) Comments Ref.
Metal precursor Reducing agent Temp. (K) Time (h)
a Amines (ED, DETA, APS), DMF (dimethyl formamide), acac (acetylacetonate), ac (acetate), AB (ammonia borane).
Anion exchange H2PdCl4, H2PtCl6 NaHB4 273 2 2–4 M-Amine-MIL-101 (M = Pd, Pt, Au) 57
Knoevenagel condensation, Heck reaction
HAuCl4
Incipient wetness impregnation (CHCl3) Pd(acac)2 H2 473 1 ∼1.5 Pd/MIL-101 84
Cyanosilylation reaction, and hydrogenation of styrene and cyclooctene
Impregnation (DMF) Pd(NO3)2 H2 473 2 1.9 ± 0.7 Pd/MIL-101 85
Suzuki–Miyaura and Ullmann coupling reaction
Incipient wetness impregnation (DMF) Pd(NO3)2 H2 473 2 2.5 ± 0.5 Pd@MIL-101 29
4.2 ± 1.1 One-pot methyl isobutyl ketone synthesis
Colloidal deposition HAuCl4 H2 473 2 9.8 ± 3.4 Au/MIL-101 86
2.3 ± 1.1 Aerobic oxidation of alcohol
Impregnation (CHCl3) Pd(acac)2 H2 473 2 2.1 Pd/MIL-101 87
One-pot indole synthesis in water
Solution infiltration Pd(NO3)2 H2 473 4 2.6 ± 0.5 Pd/MIL-101 88
Direct C2-arylation of indoles
Impregnation (H2O) HAuCl4 H2 473 3 2–8 Au–Pd/MIL-101 and Au–Pd/ED-MIL-101 89
H2PdCl4
Dehydrogenation of HCOOH for chemical H2 storage
2–15
Impregnation (H2O) Pd(NO3)2 H2 473 2 ∼3 Pd@MIL-101 90
One-pot synthesis of menthol
Chemical vapour deposition [(C5H5)Pd(C3H5)] [(C5H5)2Ni] H2 343 20 2–3 Ni/Pd@MIL-101 91
Hydrogenation of dialkyl ketones
Impregnation (hexane–H2O) H2PtCl6 H2 473 5 1.8 ± 0.2 Pt@MIL-101 92
Liquid-phase AB hydrolysis, solid-phase AB thermal dehydrogenation, gas-phase CO oxidation
Deposition-precipitation (H2O) HAuCl4 H2 473 2 1.8 ± 0.5 Au/MIL-101 93
Oxidation of cyclohexane
Colloidal deposition (H2O) PdCl2 NaHB4 273 2.5 ± 0.5 Pd/MIL-101 94
Aerobic oxidation of alcohol
Colloidal deposition (H2O) H2PdCl4 NaHB4 273 2–3 Pd/MIL-101 95
Phenol hydrogenation
Incipient wetness impregnation (acetone–DMF) Pd(acac)2 CO + H2 473 1 3–9 Pt/MIL-101, Pd/MIL-101, PtPd/MIL-101, Pt@Pd/MIL-101 96
Pt(acac)2
H2PtCl6 CO oxidation
Impregnation (ethanol) H2PtCl6 HCOONa 368 2 1.5–2.5 Pt/MIL-101 97
Asymmetric hydrogenation of α-ketoesters
Colloidal deposition (H2O) H2PtCl6 NaHB4 273 150–350 Pt@MIL-101 98
Hydrogenation of nitroarenes
Impregnation (hexane–H2O) H2PtCl6, H2PdCl4, HAuCl4 H2 473 5 ∼2.5 M@MIL-101 (M = Pt, Pd, Au or Rh) 99
RhCl3 Reduction of Cr(VI) to Cr(III)
Impregnation (hexane–H2O) HAuCl4 NaHB4 273 1.8 ± 0.2 AuNi@MIL-101 100
NiCl2 AB hydrolysis
Impregnation (H2O) Cu(NO3)2 NH2NH2 2–3 Cu/MIL-101 101
Reduction of aromatic nitro compounds
∼100
Impregnation (acetone–DMF) Pd(ac)2 CH3COCH3 318 24 2–10 Pd/MIL-101 102
Oxidation of alcohol
Hydrogenation of alkenes and aldehyde
Infiltration (H2O) H2PdCl4 NaHB4 273 3 1.8 ± 0.4 Pd@MIL-101 103
Hydrolysis of AB
Incipient wetness impregnation (CHCl3) Pd(acac)2 H2 473 2 2–3 Pd/MIL-101 104
Hydrogenation of 2,3,5-trimethylbenzoquinone
Impregnation (H2O) K2PtCl6 NaHB4 and NH3BH3 273 1.9 ± 0.4 Ni–Pt@MIL-101 105
H2 generation from aqueous alkaline solution of hydrazine
NiCl2
Impregnation (ethanol–H2O) K2PtCl6 H2 373 5 5.0 ± 0.5 Pt-MIL-101 106
Hydrogenation of olefin mixture, benzonitrile and linoleic acid
Impregnation (CH3CN–toluene) C13H19O2Rh 343 2.5 150–250 Rh@MIL-101 107
Hydroformylation of olefin


Férey et al. reported the encapsulation of noble metals, such as Pd, Pt and Au into the amine-grafted MIL-101 by anion exchange through the three-step PSM process (i) treatment of surface amines with aqueous HCl to form positively charged surface ammonium groups, (ii) reaction of the cationic surface NH3+ with [PdCl4]2−, [PtCl6]2−, and [AuCl4], and (iii) reduction of noble metal ions using NaBH4 at low temperature to produce NPs in the range of 2–4 nm. Some NPs (>20 nm) were also found outside the pores.57 Pd NPs supported on MIL-101 can also be prepared from either of the following: (i) chloroform solution of Pd(acac)2 added to an activated MIL-101 and reduction by H2 at 473 K to form Pd particles of ∼1.5 nm,84 (ii) DMF solution of Pd(NO3)2·2H2O and reduction by H2 at 473 K to form Pd particles 1.9 ± 0.7 nm,85 (iii) aqueous solution of PdCl2 using polyvinyl alcohol as a protecting agent and NaBH4 as a reducing agent to produce NPs, 2.5 ± 0.5 nm in size,94 (iv) Pd(CH3COO)2 infused into MIL-101 in an acetonitrile solution and reduction by H2 at 473 K to produce NPs in the range, 2–3 nm,108 or (v) an acetone solution of Pd(CH3COO)2 added to MOF and heated at 318 K for 24 h, followed by a treatment with DMF at 373 K to form NPs in the range, 2–10 nm, dispersed over the external surface and inside the pore system.102

Xu et al. reported the immobilization of Pt NPs into MIL-101 nanopores using a double solvent method.92 In a typical synthesis, metal precursor H2PtCl6 dissolved in a hydrophilic solvent (water) with a volume less than or equal to the pore volume of MIL-101 was added slowly to the suspension of MIL-101 in a large excess of n-hexane as the hydrophobic solvent with stirring, followed by reduction using H2 at 473 K. Transmission electron microscopy (TEM) and electron tomography showed that the encapsulated Pt NPs in Pt@MIL-101 contained very small particles, 1.2–3.0 nm in size, distributed homogeneously throughout the interior cavities of the MIL-101 crystals with no Pt NPs on the external surface. The same group reported the immobilization of AuNi NPs using the combined double solvent method and liquid-phase concentration-controlled reduction route by reduction using NaBH4.100 A high-concentration of NaBH4 produced well dispersed AuNi NPs, 1.8 ± 0.2 nm size, within the pores without any deposition on the external surface of the framework, whereas a low-concentration of reducing agent resulted in NPs aggregation, >5.0 nm in size, on the external surface.

The mono- and bimetallic polyhedral metal nanocrystals supported on MIL-101, Pt/MIL-101, Pd/MIL-101, and PtPd/MIL-101 were obtained by incipient wetness using M(acac)2 salt (M = Pt, Pd or Pt/Pd), followed by reduction using CO/H2 at 473 K.96 The high-angle annular dark field and bright field scanning transmission electron microscopy images showed that cubic Pt, tetrahedral Pd, and octahedral PtPd nanocrystals were deposited on MIL-101. The authors also reported the preparation of core–shell Pt@Pd/MIL-101 (spherical Pt was used as core for the growth of Pd shell) by seed-mediated two-step reduction from H2PtCl6 and Pd(acac)2 salts in an acetone/DMF solvent, followed by the same reduction process.

The encapsulation of Au–Pd, Au–Pt, Ru–Pd, Au, Pd, Pt or Ru NPs into MIL-101, or amine-grafted MIL-101 were obtained by impregnation and subsequent reduction by H2 at 473 K.89 The Au–Pd NPs in Au–Pd/MIL-101 and Au–Pd/ED-MIL-101 were 2–8 and 2–15 nm in size, respectively. Hermannsdörfer et al. reported the preparation of bimetallic Ni/Pd@MIL-101 by the chemical vapour deposition of [(C5H5)Pd(C3H5)] and [(C5H5)2Ni], followed by hydrogenation using H2 at 343 K to produce nanoparticles in the range, 2–3 nm.91 Recently, Cao et al. reported the preparation of bimetallic NiPt@MIL-101 by impregnation using aqueous K2PtCl6 and NiCl2 solutions, followed by reduction using NaBH4 and NH3BH3.105 TEM and EDS showed that the encapsulated NiPt NPs in NiPt@MIL-101 contained particles 1.9 ± 0.4 nm in size, which were well-dispersed within the mesoporous cavities in MIL-101.

Pt (20 wt%)/C doped MIL-101 was obtained by the simple ball mill mixing of MIL-101 and Pt (20 wt%)/C under a 0.2–0.3 MPa argon atmosphere.109 A carbon bridge between MIL-101 and Pt (20 wt%)/C was built by adding sucrose to the mixture, and heating the mixture in a tubular reactor protected by flowing nitrogen at 483 K after ball milling.

5. Applications

5.1 Adsorption

The large surface area and pore volume, high concentration of CUS, excellent thermal stability, as well as high resistance to moisture make MIL-101 an attractive candidate material for the adsorption of gases, water vapour and organic compounds. To enhance its adsorption performance, diverse functional groups or additives have been introduced to the MIL-101 structure to form versatile hybrid materials with supplementary functionalities. Tables 4 and 5 list the reported adsorption studies on MIL-101.
Table 4 Gas adsorption over MIL-101
Adsorbenta Adsorbate Temp. (K) Pressure (bar) Uptake (mmol g−1) ΔHadsorption (kJ mol−1) Comments Ref.
a CB (carbon bridge), MWCNT (multi-walled carbon nanotubes), SWCNT (single-walled carbon nanotubes), AC (activated carbon), PEI (polyethylenimine), DETA (diethylenetriamine), TEPA (tetraethylenepentamine).b mmol cm−3.
MIL-101 H2 77 80 30.5 10–9.3 Adsorption capacity and heat of adsorption 110
MIL-101 H2 77 20 22.5 4.3 Heat of adsorption 111
MIL-101 H2 293 1900 36 H2 storage by ionic clusters-doped MIL-101 112
MIL-101 (pellet) H2 77.3 80 43.5 Adsorption by MIL-101 pellets 113
MIL-101 (pellet) H2 19.5 0.9 52.9 Adsorption by MIL-101 pellets 114
CB-Pt/C-MIL-101 H2 293 50 5.7 Enhance H2 storage by spillover effect 109
CB-Pt/AC-MIL-101 H2 298 100 7.2 21–12 Enhance H2 storage by spillover effect 115
CB-Pd/AC-MIL-101 H2 298 32 2.3 Enhance H2 storage by spillover effect 116
CB-Pt/CMK-3-MIL-101 H2 298 20 6.7 Enhance H2 storage by spillover effect 117
Pd@MIL-101 H2 298 45 1.2 H2 chemisorption 108
Mo6Br8F6-MIL-101 H2 77 60 9.7 H2 storage by [Mo6Br8F6]2− cluster loaded MIL-101 118
Li+-MIL-101 H2 77 1 17 Enhance H2 storage by lithium-doping on MIL-101 119
n-Hexane-MIL-101 H2 298 63 6.3b Enhanced H2 uptake by n-hexane confined in MIL-101 120
SWCNT@MIL-101 H2 298 60 3.2 Enhanced H2 sorption by SWNT incorporated MIL-101 121
AC@MIL-101 H2 77 60 50.5 Enhanced H2 sorption in AC incorporated MIL-101 60
MIL-101 CO2 298 0.1 0.49 Regeneration conditions and influence of flue gas contaminants 122
MIL-101 CO2 303 50 40 44–24 Adsorption capacity and heat of adsorption 123
CH4 60 13.6 18–8
MIL-101 CO2 298 30 22.9 4.0–28.6 Adsorption equilibrium and kinetics 124
MIL-101 CO2 298 15 16.8 31 HF-free-synthesis of MIL-101 for CO2 adsorption; water stability 125
MIL-101 CO2 273 1 3.35 Size-controlled synthesis of MIL-101 nanoparticles, enhanced CO2/N2 selectivity 32
MIL-101 CO2 288 1.1 3.8 38.81 Equilibrium and dynamics gas adsorption studies 126
MIL-101 CO2 295 36 21.3 32.5–21 Effect of open metal sites and adsorbate polarity 127
MIL-101-NH2-PEI film CO2 273 1 3.7 Facile synthesis of MIL-101 and MIL-101-NH2 films; enhanced CO2/N2 selectivity 37
MIL-101-NH2 CO2 289 25 15 51–19 CO2 adsorption by directly synthesized amine-functionalized MIL-101 50
DETA-MIL-101 CO2 298 1 0.7 69–28 CO2 adsorption by amine grafted MIL-101 67
TEPA-MIL-101 CO2 298 0.2 1.4 42.5–24 Selective adsorption of CO2 over CO by amine grafted MIL-101 66
PEI-MIL-101 CO2 298 1 5 Amine-impregnated MIL-101 for selective CO2 capture 58
PEI-MIL-101-NH2 CO2 323 2.1 3.6 Amine-impregnated MIL-101 for selective CO2 capture 59
MIL-101-NH2 CO2 273 0.15 2 43–23 Selective and enhanced CO2 adsorption by NH2 functionalization 128
MWCNT@MIL-101 CO2 298 10 1.35 Enhanced CO2 adsorption by a MSWNT-incorporated MIL-101 82
MIL-101 Ar 283 5.3 0.9 15–12.5 Polarizability and quadrupole moment effect 129
CH4 2.2 19–14.2
CO2 8.0 24–20
SF6 9.1 26–23.5
C3H8 13.4 37–26
MIL-101 C2H2 313 1 6.4 Effect of purification conditions on gas storage and separations 130
CO2 298 1 4.6
MIL-101 SF6 298 18 12.3 High pressure adsorption isotherms 131
CF4 35 7.86
MIL-101 H2S 303.1 20 38.4 High adsorption capacity 132
MIL-101 n-Butane 293 0.774 11.2   Adsorption capacity and heat of adsorption 133


Table 5 Water vapour and organic species adsorption over MIL-101
Adsorbenta Adsorbate Temp. (K) Pressure Uptake (mg g−1) ΔHadsorption (kJ mol−1) Comments Ref.
a GO (graphite oxide), N-SRGO (nitrogen in straight run gas oil), N-LCO (nitrogen in light cycle oil), IL (ionic liquid), H3PW (phosphotungstic acid), HS (hierarchically structured), PED (protonated ethylenediamine), ED (ethylenediamine).b Vwater/Vfilm.
MIL-101 H2O 298 P/P0 = 0.67 428 Adsorption and H2O stability 115
MIL-101 H2O 298 P/P0 = 0.9 1276 45.13 Adsorption and H2O stability 134
MIL-101 H2O 298 P/P0 = 0.92 1010 Adsorption and H2O stability 135
MIL-101 H2O 303 P/P0 = 0.96 1700 78–44 Studies of H2O sorption isotherm, kinetics, and hydrothermal stability 136
MIL-101 H2O 298 P/P0 = 0.85 1360 70–19 Effect of functional groups in MIL-101 on H2O sorption 49
MIL-101 film H2O P/P0 = 1 0.79b Hierarchically-structured thin films of MIL-101 39
MIL-101-NH2 H2O 293 P/P0 = 0.9 1060 43 Cyclic stability of functionalized MIL-101 137
MIL-101 Benzene 303 P/P0 = 0.5 1303 Microwave synthesis of MIL-101; high and rapid adsorption 38
MIL-101 Benzene 298 55 mbar 1172 Adsorption and diffusion of benzene 40
MIL-101 Benzene 305 P/P0 > 0.7 1404 High uptake 27
n-Hexane P/P0 > 0.7 1084
MIL-101 Benzene 313 P/P0 = 0.85 1443 68–34 Co-adsorption of n-hexane and benzene vapour 138
n-Hexane P/P0 = 0.63 1041 61–38
MIL-101 n-Heptane 313 P/P0 = 1 795 40–29 Adsorption of C5–C9 hydrocarbons 139
MIL-101 Acetone 298 P/P0 = 0.55 1291 Adsorption of volatile organic compounds with different molecular size and shape 140
Benzene 1291
Toluene 1096
Ethylbenzene 1105
m-Xylene 727
o-Xylene 758
p-Xylene 1067
MIL-101 p-Xylene 298 5.82 mbar 1009 24.8–44.2 Adsorption equilibrium and kinetics 141
MIL-101 p-Xylene 313 P/P0 = 0.8 1405 67–39 Adsorption and separation of xylene isomer vapours 142
o-Xylene 1315 76–40
m-Xylene 1320 70–39
MIL-101 Ethyl acetate 288 54 mbar 924 36.48–42.25 Adsorption and diffusion 143
MIL-101 n-Butylamine 298 934 75–125 Adsorption of VOC 30
MIL-101 1,2-Dichloroethane 288 8 mbar 1880 42.0–61.6 Adsorption isotherms, kinetics, and desorption 144
GO@MIL-101 Acetone 288 160 mbar 1195 49–58 MOF/graphite oxide composite; high uptake 83
GO@MIL-101 n-Hexane 298 180 mbar 1042 MOF/graphite oxide composite; high uptake 145
MIL-101 N-SRGO 9 Sorption capacities and selectivity towards N-containing compounds 146
N-LCO 20
MIL-101 Benzothiophene 298 0.37 Experiment/computational study; adsorption of organosulfur compounds 147
GO@MIL-101 Benzothiophene 298 28 MOF/graphite oxide composite; adsorption of N-and S-containing compounds 62
Quinoline 549
Indole 319
IL-MIL-101 Benzothiophene 298 87 Ionic liquid supported on MIL-101 for adsorptive desulfurization 148
MIL-101 (H3PW) Quinoline 298 274 H3PW impregnated MIL-101; adsorptive denitrogenation 74
MIL-101 Xylenol orange 298 312 11 Kinetic and thermodynamic studies 149
HS-MIL-101 Methylene blue 23 Hierarchically mesostructured MIL-101 synthesis; accelerated adsorption kinetics 33
PED-MIL-101 Methyl orange 308 219 29.5 Surface modification; adsorption equilibrium and kinetics 150
MIL-101 Uranine 298 127 25.3 Adsorption kinetics, mechanism, and thermodynamics 151
MIL-101 Pyridine 293 950 Experimental and theoretical study; high uptake 152
MIL-101 C70 303 198 18.1 Selective adsorption and extraction of C70 and higher fullerenes 153
MIL-101 Bisphenol A 303 253 Adsorption equilibrium and kinetics 154
MIL-101 Naproxen 298 119 Adsorption equilibrium and kinetics 155
Clofibric acid 312
ED-MIL-101 Naproxen 298 154 Surface modification; adsorption equilibrium and kinetics 156
Clofibric acid 347


5.1.1 H2 adsorption. As a clean and efficient energy carrier, H2 storage has attracted enormous attention. One approach involves the sorption of H2 on the surface of porous solids.157 Latroche et al. reported H2 adsorption on MIL-101, which exhibited approximately 6.1 wt% H2 uptake under 8 MPa and 77 K (Fig. 11) with a heat of adsorption of 10 kJ mol−1 at low surface coverage conditions.110 Small pores in MIL-101 were believed to play a major role in H2 adsorption. H2 adsorption by MIL-101 in pellet form was also reported.113 The specific surface area and micropore volume decreased with increasing pelletizing pressure increases and optimal volumetric adsorption capacity of 40 g L−1 (under 8 MPa, 77.3 K) was obtained.
image file: c4ra11259h-f11.tif
Fig. 11 Adsorption (point-up triangles) and desorption (point-down triangles) pressure–composition isotherm (PCI) curves for MIL-101 at 298 (empty symbols) and 77 K (full symbols).110

MIL-101 has been functionalized via a variety of methods to enhance the H2 adsorption capacity, typically via a spillover effect. Liu et al. studied the H2 storage behaviour of Pt/C doped and carbon bridged Pt/C-modified MIL-101.109 Both greatly enhanced the H2 storage capacity of MIL-101 at ambient temperatures with the latter exhibiting greater H2 uptake. Doping and modification in MIL-101 influenced H2 storage by assisting in H2 dissociation and their effective spillover to the MOF surface. Similarly, the H2 storage capacity was enhanced by a factor of 2.8 for MIL-101 by carbon bridged spillover.115 Szilágyi et al. introduced Pd nanoparticles into MIL-101 without a carbon bridge,108 and also observed an increase in the ambient-temperature H2 uptake, which was attributed to palladium hydride formation rather than H2 spillover. Klyamkin et al. examined the ultra-high-pressure H2 storage performance (up to 1900 bar) over MIL-101 and MIL-101 doped with ionic clusters ([Re4S4F12]4−, [SiW11O39]7−).112 Low temperature (81 K) sorption measurements indicated that doping had no effect, while doping at high temperatures provided an increase in the number of H2 binding sites such that doping of MIL-101 by [Re4S4F12]4− clusters increased the number of H2 binding sites by 2.5 fold at 293 K. Dybtsev et al. studied the influence of [Mo6Br8F6]2− cluster unit inclusion within MIL-101 on H2 storage performance.118 At room temperature and 8 MPa, the H2 storage capacity of the Mo6Br8F6-MIL-101 was over twice that of MIL-101, which may be assigned to higher heat of adsorption for H2 in Mo6Br8F6-MIL-101. Xiang et al. examined the doping of Li+ to the MIL-101 frameworks for H2 storage.119 At 77 K and 1 bar, the H2 uptake in MIL-101 was increased by 43% (from 2.37 to 3.39 wt%) after lithium doping, which was attributed to the strong affinity of Li+ for H2 molecules. Clauzier et al. reported enhanced H2 uptake in hybrid sorbents consisting of n-hexane confined in MIL-101.120 When n-hexane was confined in MIL-101, the H2 solubility increased 22.2 fold compared to that of n-hexane alone. Such enhanced solubility effect was much larger than that observed with either silica aerogel or MCM-41. The important role of surface chemistry and accessible surface area in enhancing the gas solubility was proposed. Prasanth et al. reported H2 adsorption by single-walled carbon nanotube (SWCNT)-incorporated MIL-101 composites.121 The H2 sorption capacities of MIL-101 increased from 6.37 to 9.18 wt% at 77 K up to 60 bar and from 0.23 to 0.64 wt% at 298 K up to 60 bar, which was attributed to the decrease in pore size and increase in micropore volume in MIL-101 by SWCNT incorporation. Similar results were reported with an activated carbon-incorporated MIL-101.60

5.1.2 CO2 adsorption. CO2 is a major anthropogenic greenhouse gas. In recent years, intensive efforts have been made to develop new technologies/processes for effective CO2 capture and storage. Among them, the majority of research activities focused on examining sorption-based technologies and processes, involving solid CO2 adsorbents.158 Llewellyn et al. examined CO2 adsorption on MIL-101, and reported a record capacity of 40 mmol g−1 or 390 cm3STPcm−3 at 5 MPa and 303 K.123 The heat of adsorption of CO2 at low surface coverage (−44 kJ mol−1) was higher than that reported for other MOFs, and was of the same order of magnitude as that measured by zeolites due to the coordination of CO2 molecules onto the chromium open metal sites in MIL-101. Jiang et al. reported higher CO2 adsorption selectivity over N2 by nano-sized MIL-101 (19–84 nm).32 Liu et al. examined the influence of flue gas contaminants on the adsorption of CO2 by MIL-101, as well as the regeneration conditions.122 Three trace flue gas contaminants (H2O, NO, SO2) were each added to a 10 vol% CO2/N2 feed flow but they had minimal impact on the adsorption capacity of CO2 by MIL-101. In addition, complete regeneration could be achieved within 10 min at 328 K in temperature swing adsorption with N2-stripping, or at 348 K for vacuum-temperature swing adsorption at 20 kPa. Almost 99% of the pre-regeneration adsorption capacity was preserved after 5 adsorption–desorption cycles under a gas flow of 10 vol% CO2, 100 ppm SO2, 100 ppm NO, and 10% RH, respectively.

Frequently, amine species have been introduced to MIL-101 to increase its CO2 adsorption capacity and selectivity against other gases. Tetraethylenepentamine (TEPA) grafted on the Cr(III) sites in MIL-101 for the selective adsorption of CO2 over CO was reported by Wang et al.66 The selectivity for CO2 over CO was improved from 1.77 to 70.20 at 298 K and a total pressure of 40 kPa. DETA grafted on the Cr(III) sites of MIL-101 led to better CO2 adsorption capacity at low pressures as well as a higher heat of adsorption compared to the original MIL-101.67 The post-synthetically prepared MIL-101 with amino groups also showed higher selectivity over N2 and CH4, compared to the original MIL-101.128 Chen et al. reported PEI-incorporated MIL-101 adsorbents prepared by a wet impregnation method, which exhibited dramatically enhanced CO2 adsorption capacity at low pressures (4.2 mmol g−1 at 298 K, 0.15 bar with a 100 wt% PEI loading), rapid adsorption kinetics, and ultrahigh selectivity for CO2 over N2 in the flue gas with 0.15 bar CO2 and 0.75 bar N2 (CO2 over N2 selectivity was 770 at 298 K, and 1200 at 323 K).58 In addition, they achieved improved CO2/CH4 selectivity and CO2 adsorption capacity at low pressure by a dual amine functionalized MIL-101 adsorbent prepared by incorporating PEI into the framework of a directly synthesized amine-functionalized MIL-101 (Fig. 12).59 Enhanced CO2 adsorption selectivity over N2 by MIL-101 and MIL-101-NH2 films that were fabricated through a PEI-assisted dip-coating method was also reported.37


image file: c4ra11259h-f12.tif
Fig. 12 CO2 and CH4 adsorption isotherms for amine-MIL-101 (circles), 50PEI-amine-MIL-101 (triangles), 75PEI-amine-MIL-101 (squares) and 100PEI-amine-MIL-101 (stars) at 298 K. Symbols: filled, CO2 adsorption; hollow, CH4 adsorption.59

Zhang et al. published the results of a multi-scale modelling study to examine the effects of functional groups (–CH3, –Cl, –NO2, –CN and –NH2) on the adsorption and separation of CO2/N2 over MIL-101.159 The CO2 uptake in the low-pressure regime increased in the order of MIL-101 < MIL-101-CN < MIL-101-NO2 < MIL-101-Cl < MIL-101-CH3 < MIL-101-NH2, which followed the sequence of strength of the binding energies between CO2 and the functional groups. The CO2/N2 selectivity was similarly enhanced after functionalization, and the predicted breakthrough time was increased with the longest breakthrough time in MIL-101-NH2 being approximately 2 times longer that in MIL-101.

Anbia et al. reported CO2 adsorption on a hybrid composite of acid-treated multi-walled carbon nanotubes (MWCNTs) and MIL-101.82 The composite had the same crystal structure and morphology as that of virgin MIL-101, but the CO2 adsorption capacity was increased by approximately 60% (from 0.84 to 1.35 mmol g−1) at 298 K and 10 bar. The increase in CO2 adsorption capacity by MWCNT@MIL-101 was attributed to the increase in micropore volume of MIL-101 by MWCNT incorporation.

5.1.3 H2O vapour adsorption. Water sorption technologies are widely used commercially in industrial or indoor desiccant applications, fresh water production and adsorption heat transformation. Silica gel and zeolites currently have commercial applications, and tests using MOFs are also being carried out. Li et al. reported H2O vapour adsorption on MIL-101.115 A H2O adsorption amount of 42.8 wt% was measured at 298 K and P/P0 = 0.67, which was much higher than that by HKUST-1 or COF-1. This larger amount was attributed to the high pore volume and metal-oxide clusters present in the framework of MIL-101. In addition, MIL-101 was stable in moist air after exposing the sample to ambient air for 7 days with a relative humidity ∼40% at 298 K. Ehrenmann et al. reported a high H2O uptake of 1.01 g g−1 (at 298 K, P/P0 = 0.92) by MIL-101 with an S-shaped isotherm.135 Only a slight decrease in performance was detected with a H2O uptake capacity of 98.1% and 96.8% after 20 and 40 cycles, respectively, compared to the initial load, indicating the excellent stability of MIL-101. Seo et al. reported the exhaustive analysis of the H2O sorption properties on MIL-101.136 High equilibrium uptakes of 1.50–1.70 g g−1 in dehydrated MIL-101 at 303–313 K and P/P0 > 0.5 were achieved. MIL-101 showed both high adsorption and desorption rates less than 353 K, and excellent stability under cyclic operation of H2O adsorption–desorption. Highly porous thin films of MIL-101 on a silicon wafer with a dual hierarchical porous structure showed double-step H2O adsorption associated with both different sizes of pores in MIL-101 coupled with the inter-particle mesoporosity.39 Akiyama et al. examined the H2O sorption properties of MIL-101 containing different substituents (–H, –NO2, –NH2, –SO3H) in the ligand (Fig. 13).49
image file: c4ra11259h-f13.tif
Fig. 13 Water adsorption isotherms at 298 K for desolvated MIL-101 (a), MIL-101-NH2 (b), MIL-101-SO3H (c), and MIL-101-NO2 (d). P/P0 is the relative pressure of water, where the saturation vapour pressure (P0) is 3.16 kPa. A is the amount adsorbed (g g−1).49

All the materials exhibited a H2O uptake of ca. 0.8–1.2 g g−1, and the adsorbed H2O could be released at ca. 353 K without high evacuation. The sorption profile could be tuned by changing the substituents in the ligand, such that the isotherm lines of MIL-101-NH2 and MIL-101-SO3H moved to lower P/P0 values compared to those of MIL-101, which was attributed to the highly hydrophilic groups on their pore surfaces. Khutia et al. reported H2O sorption cycle measurements on NH2- or NO2-functionalized MIL-101 synthesized by post-synthetic modification.137 The MIL-101-NH2 showed excellent H2O loading (1.06 g g−1 at 293 K, P/P0 = 0.9) as well as H2O stability over the 40 adsorption–desorption cycles.

5.1.4 Other adsorption for energy/environmental applications. As listed in Tables 4 and 5, other adsorption processes by MIL-101 for energy/environmental applications, such as methane storage, adsorptive removal of VOC, dyes, and nitrogen-(or sulphur) containing compounds have also been reported. Llewellyn et al. reported a CH4 uptake of 13.6 mmol g−1 or a volume capacity of 135 cm3STPcm−3 by MIL-101 at 6 MPa and 303 K, which was below the DOE target for methane storage on a volume basis (180 cm3STPcm−3).123 For CH4 storage, MIL-101 itself is unsatisfactory, whereas several other MOF materials meeting the DOE target have emerged recently.160 A high n-butane adsorption capacity (11.2 mmol g−1 at 293 K, 0.774 bar) on MIL-101 was reported by Klein et al., which was approximately three times higher than that on than Cu3(BTC)2, and twice as much as that with activated carbon.134 The CO uptakes of 1.13 mmol g−1 (at 850 mm Hg, 288 K) and 6.5 mmol g−1 (at 66 bar, 295 K) by MIL-101 were reported with a high heat of adsorption (ca. 42 kJ mol−1).126,127 Tanh Jeazet et al. reported mixed-matrix membranes made from water-stable MIL-101 and polysulfone (PSF) for O2/N2 separation.161 MIL-101 microcrystals could adhere well to PSF and yield a very robust mixed-matrix MIL-101-PSF membrane. At 24 wt% MIL-101, the permeability of the membrane increased almost four fold with respect to the pure polymer tested under the same conditions, while keeping the high selectivity for O2 over N2 of 5–6. Hamon et al. reported a high H2S adsorption capacity of 38.4 mmol g−1 on MIL-101 at 2 MPa and 303 K, which was significantly higher than that of MIL-53 (Al, Cr, Fe), MIL-47 (V) and MIL-100 (Cr).132

Jhung et al. reported benzene adsorption by MIL-101 synthesized under microwave irradiation.38 The MIL-101 exhibited a significantly larger sorption capacity for benzene (16.7 mmol g−1 at 303 K and P/P0 = 0.5) than those by SBA-15, HZSM-5, or commercial active carbon (Fig. 14) with rapid sorption kinetics.


image file: c4ra11259h-f14.tif
Fig. 14 Sorption isotherms for benzene at 303 K on MIL-101, a commercial active carbon, HZSM-5 zeolite, and mesoporous silica SBA-15.38

Hong et al. also reported a large sorption uptake of benzene by MIL-101, i.e., 19.5 mmol g−1 at P/P0 > 0.7 and 305 K,27 and observed that MIL-101 also has a high uptake of n-hexane; the equilibrium sorption capacity of n-hexane at 303 K was 12.6 mmol g−1 at P/P0 > 0.7. Trens et al. reported the co-adsorption of n-hexane and benzene vapours onto MIL-101,138 which showed that MIL-101 had a better affinity for benzene than n-hexane; benzene was adsorbed preferentially at low coverage at the expense of n-hexane. Yang et al. revealed the adsorption performance of MIL-101 towards several VOCs, such as acetone, benzene, toluene, ethylbenzene, m-xylene, o-xylene, and p-xylene, with different molecular properties.140 MIL-101 was found to exhibit higher adsorption capacities than zeolites, activated carbons, or other reported adsorbents. The adsorption of VOCs on MIL-101 takes place by a pore filling mechanism, and the size- and shape-selectivity of VOCs molecules were observed. Huang et al. studied the adsorption characteristics of six VOCs with different functional groups and polarities (n-hexane, toluene, methanol, butanone, dichloromethane, and n-butylamine) on MIL-101.30 Again, MIL-101 had significantly higher affinity and higher adsorption capacity towards VOCs than activated carbons, which showed higher affinity towards the VOCs containing a heteroatom or aromatic ring, particularly amines; MIL-101 exhibited the strongest affinity to n-butylamine with an adsorption capacity of 12.8 mmol g−1. CUS in MIL-101 are vital in the sorption process.

Kim et al. reported a large pyridine adsorption capacity (950 mg g−1 at 293 K) on MIL-101.152 Pyridine could be adsorbed either on the carbon atoms of the carboxylic groups of the BDC (benzene-1,4-dicarboxylate) units or on the CUS in MIL-101. The retention of pyridine under evacuation at 423 K, as well as the calculated stabilization energy suggested the strong adsorption of pyridine on the CUS of MIL-101. Ahmed et al. reported the adsorptive removal of nitrogen-containing compounds (NCCs) and sulfur-containing compounds (SCCs) in model fuels by a graphite oxide/MIL-101 composite (GO/MIL-101).62 The composite exhibited the highest adsorption capacity for NCCs among the adsorbents reported, which was attributed to the higher surface area of GO/MIL-101 over MIL-101 (Fig. 15). The adsorbent could be used several times after regeneration without any detectable degradation in performance. The same type of GO/MIL-101 was also shown to have a high acetone adsorption capacity (20.1 mmol g−1 at 288 K, 161.8 mbar), which was much higher than that of bare MIL-101.83


image file: c4ra11259h-f15.tif
Fig. 15 Adsorption isotherms for (a) quinoline and (b) indole on MIL-101 and 0.25% GO/MIL-101.62

Khan et al. produced acidic ionic liquids (ILs)-supported MIL-101 for adsorptive desulfurization.148 A remarkable improvement in the benzothiophene adsorption capacity was observed for ILs supported on MIL-101 compared to bare MIL-101, which was explained by the acid–base interactions between the acidic IL and basic benzothiophene molecules. Similarly, adsorptive denitrogenation with MIL-101 impregnated with H3PW was reported;74 a 1% loading of H3PW on MIL-101 increased the adsorption of quinoline by approximately 20%, owing to the favourable interactions between the acid and base sites.

Haque et al. reported the adsorptive removal of methyl orange (MO) in an aqueous solution by MIL-101.150 The porosity and pore size in MIL-101 were found to play important roles in the adsorption, and a specific electrostatic interaction between MO and the adsorbent was responsible for the rapid and high uptake of the dye. The adsorption of MO was a spontaneous and endothermic process. Similar results regarding xylenol orange149 and uranine adsorption151 have been reported. Huang et al. reported remarkably accelerated adsorption kinetics for methylene blue removal compared to the bulk MIL-101 by a hierarchically-structured MIL-101 prepared using CTAB as a supramolecular template.33 Huo et al. reported the use of MIL-101 for the rapid magnetic solid phase extraction of polycyclic aromatic hydrocarbons (PAHs) from environmental water samples.162 The simple mixing of MIL-101 with silica-coated Fe3O4 micro-particles in solution with sonication was applied to the adsorbent preparation. Hydrophobic and π–π interactions between the PAHs and framework terephthalic acid molecules, and π-complexation between PAHs and the Lewis acid Cr(III) sites in the pores of MIL-101 played a significant role in the adsorption of PAHs. Hasan et al. investigated the liquid-phase adsorption of naproxen and clofibric acid over MIL-101.155 The adsorption was attributed to an electrostatic interaction between the chemicals and adsorbent. The adsorption rate and adsorption capacity were improved using MIL-101 modified with basic NH2 groups, which can adsorb naproxen and clofibric acid through an acid–base interaction.159 Yang et al. examined the selective adsorption and extraction of C70 and higher fullerenes on MIL-101.153 MIL-101 showed the preferential adsorption of C70 and higher fullerenes over C60 with high selectivity (αC70/C60 = 24). Therefore, the selective extraction of C70 and higher fullerenes from crude carbon soot could be achieved easily via a simple adsorption–desorption process. The used MIL-101 could be regenerated by washing with o-dichlorobenzene under ultra-sonication.

5.2 Heterogeneous catalysis

MIL-101 and its organic-functionalized or guest-incorporated forms have been applied as efficient catalysts for a range of reactions involving different active sites. In this section, the performance of the MIL-101 in the Lewis acid, Lewis base, acid–base concerted, tandem, oxidation, hydrogenation, C–C coupling, and other reactions will be presented.
5.2.1 Cyanosilylation of benzaldehyde. MIL-101 was applied to Lewis acid-catalysed cyanosilylation reaction (Scheme 1), and benzaldehyde and trimethylsilylcyanide were converted to the corresponding product in 98.5% yield at 313 K after 3 h, which showed higher catalytic activity of MIL-101 than other MOFs, such as Cu3(BTC)2.84 The heterogeneity of the catalysts was confirmed in a hot filtration test and recycle runs. XRD confirmed that the structure of MIL-101 had been maintained after the recycle runs.
image file: c4ra11259h-s1.tif
Scheme 1 Cyanosilylation of benzaldehyde.
5.2.2 Knoevenagel condensation. Amine-functionalized MIL-101 was applied to the Knoevenagel condensation reaction as a base- and acid-catalysed reaction (Scheme 2).57,67 Férey et al. reported the Knoevenagel condensation of benzaldehyde with ethyl cyanoacetate using amine-grafted MIL-101, which showed good activity (96.3–97.7% conversion and ∼99% product selectivity) in cyclohexane at 353 K.57 ED-MIL-101 exhibited superior activity (10 times higher) than APS-grafted SBA-15 mesoporous silica. No catalytic activity was observed in the Knoevenagel condensation of benzophenone and malonitrile over ED-MIL-101. This dependence of the catalytic activity on the reactant/product size suggests that the reaction occurred mostly inside of the pores of MIL-101, where the active amine groups were available. Juan-Alcañiz et al. also reported the Knoevenagel condensation reaction of benzaldehyde with ethyl cyanoacetate using H3PW/MIL-101 prepared by the direct encapsulation of phosphotungstic acid into MIL-101, and ∼99% conversion was obtained in toluene at 313 K after 200 min.70 The high catalytic performance originated from the partial substitution of tungsten by Cr3+ to create the so-called lacunary structure. The heterogeneity of the catalysis in the liquid phase reaction was confirmed by a hot filtration experiment and recycle runs, and no leaching of the guest was observed in the filtrate after recycling.
image file: c4ra11259h-s2.tif
Scheme 2 Knoevenagel condensation reaction.
5.2.3 CO2 cycloaddition of epoxide. MIL-101 was applied to the Lewis acid/base-catalysed cycloaddition of CO2 to alkene oxides under high pressure CO2 (8–100 bar) conditions over the temperature range, 298–393 K (Scheme 3)163,164 in the presence of tetrabutylammonium bromide (TBAB) as a co-catalyst under solvent-free conditions.163 Styrene oxide and propylene oxide converted to a cyclic carbonate using CO2 (8 bar) in the presence of TBAB at 298 K with an 82% yield (after 24 h) and 95% yield (after 48 h), respectively, whereas cyclohexene oxide exhibited only a 5% yield after 64 h. The recycled catalyst showed significantly lower activity than the fresh one and the BET surface area decreased substantially, suggesting pore blocking by the carbonaceous deposits formed during the reaction. Recently, Ahn et al. reported the cycloaddition reaction of CO2 and styrene oxides in chlorobenzene using different MOFs with different acid–base functionalities.164 MIL-101 showed a 63% yield of styrene carbonate using CO2 (20 bar) at 373 K after 4 h. On the other hand, UiO-66-NH2 exhibited a significantly higher yield of styrene carbonate (∼94%) than MIL-101 under identical reaction conditions. The authors suggested that the high catalytic activity of UiO-66-NH2 originated from high populations of both Lewis acid and base sites, which suggests that the amine-functionalized MIL-101, MIL-101-NH2, would be equally effective in the same reaction. The heterogeneity of the catalyst was confirmed in a hot filtration test and recycle runs. XRD confirmed that the structure of MIL-101 had been retained after the recycle runs.
image file: c4ra11259h-s3.tif
Scheme 3 CO2 cycloaddition of epoxide.
5.2.4 One pot deacetalization-nitroaldol reaction. A cascade or tandem reaction is a consecutive series of intramolecular organic reactions, which often proceed in the presence of highly reactive chemical intermediates. The main advantages of the reactions are (i) the reaction rate can be accelerated because of its intramolecular nature, (ii) displays high atomic utilization efficiency, and (iii) does not involve workup and isolation of many intermediates. Li et al. reported the acid–base-catalysed one-pot tandem reaction using MIL-101-NH2-SO3H possessing both Bronsted acid and base sites,55 which showed good activity in the hydrolysis of benzaldehyde dimethyl acetal and the subsequent Henry reaction with nitromethane to form 2-nitrovinylbenzene (Scheme 4). Ahn et al. synthesized bifunctional MIL-101-NH2-SO3H in a different but much simpler route, as described earlier,56 and the material exhibited excellent catalytic activity in the deacetalisation–nitroaldol reaction with 100% conversion and 99.3% selectivity to 2-nitrovinylbenzene at 323 K after 6 h. The catalytic performance was superior to other bifunctional catalysts, such as periodic mesoporous organosilicas containing organic amines and sulfonic acid groups or heteropoly acid and amine-grafted SBA-15. The catalyst could be reused several times without any loss of its initial catalytic activity. XRD and FT-IR confirmed that the crystallinity of the catalyst had been retained during the reaction with no leaching observed.
image file: c4ra11259h-s4.tif
Scheme 4 One-pot deacetalization-nitroaldol reaction.
5.2.5 Organophosphorous ester degradation. Dialkylaminopyridine (DAAP)-incorporated MIL-101 was applied to the Lewis acid–base catalysed hydrolysis degradation of diethyl 4-nitrophenyl phosphate, which is a biocide applied widely for crop protection and as a structural analogue for chemical warfare agents (Scheme 5).68 The material induced the efficient degradation of organophosphorous ester paraoxon in a water–acetonitrile mixture at room temperature to form diethyl phosphate and p-nitrophenol (79–100% conversion after 24 h at pH 10), showing 7 times and 47 times higher activity than the parent MIL-101 and free dialkylaminopyridine, respectively, because of the synergistic effects of Lewis acidic Cr(III) and DAAP as a Lewis base.
image file: c4ra11259h-s5.tif
Scheme 5 Organophosphorous ester degradation.
5.2.6 Baeyer condensation of benzaldehyde. Bromberg et al. reported the Baeyer condensation of benzaldehyde and 2-naphthol, and in the three-component condensation of benzaldehyde, 2-naphthol and acetamide under solvent-free microwave irradiation using phosphotungstic acid-incorporated MIL-101, H3PW/MIL-101.31 Benzaldehyde and 2-naphthol were converted to dibenzoxanthene in high yield (96%) within a short reaction time (2 min) at 363 K (Scheme 6), and benzaldehyde, 2-naphthol and acetamide formed 1-amidoalkyl-2-naphthol in high yield (95%) at 403 K after 5 min (Scheme 7). The heterogeneous nature of the catalysis was confirmed. H3PW/MIL-101 showed higher catalytic activities in both condensation reactions compared to those by the parent MIL-101.
image file: c4ra11259h-s6.tif
Scheme 6 Baeyer condensation of benzaldehyde and 2-naphthol.

image file: c4ra11259h-s7.tif
Scheme 7 Baeyer condensation of benzaldehyde, 2-naphthol and acetamide.
5.2.7 Oxidation of alkanes. Ahn et al. reported the liquid-phase oxidation of tetralin using tert-butyl hydroperoxide (t-BuOOH) or O2 and trimethylacetaldehyde as an oxidant,165 making direct use of Cr(III) in MIL-101 for oxidation. MIL-101 exhibited higher catalytic activity (68% conversion and 86% selectivity to 1-tetralone) using t-BuOOH at 353 K after 8 h than a chromium-containing aluminophosphate, CrAPO-5 (47% conversion and 82% selectivity) under identical reaction conditions. The conversion and selectivity were strongly dependent on the reaction temperature, time, catalyst amount, and concentration of t-BuOOH. This study also demonstrated that the nature of the solvents and the sources of the oxidant play important roles. The heterogeneity of the catalyst was confirmed in a hot filtration test and recycling runs. XRD confirmed that the structure of the catalyst was unchanged during the reaction.

The hybrid materials, FePcF16@MIL-101, RuPcF16@MIL-101 and N(FePctBu4)2@MIL-101 were similarly applied to the aerobic oxidation of tetralin using molecular oxygen at 383 K.80 FePcF16@MIL-101 and RuPcF16@MIL-101 showed high activity with a turnover number (TON) of 48[thin space (1/6-em)]200 and 46[thin space (1/6-em)]300, respectively, after 24 h. MPcF16 (M = Fe or Ru) species were located inside the MIL-101 matrix and high catalytic performance was observed. On the other hand, dimeric N(FePctBu4)2 species deposited on the external surface of MIL101 were formed as the reaction proceeded and the catalytic activity declined. Cyclohexane oxidation over MIL-101 using t-BuOOH at 343 K produced cyclohexanone and cyclohexanol with 36% conversion and 75% selectivity to cyclohexanone.166 XRD and FT-IR spectroscopy confirmed that the catalyst was reusable and stable under the reaction conditions. The post-synthetic covalently attached tetra(4-pyridyl)porphyrinatomanganese(III) acetate to CM-MIL-101, Mn(TPyP)OAc/CM-MIL-101, as shown in Fig. 8, was found to be an efficient heterogeneous catalyst for the hydroxylation of a variety of alkanes, such as tetralin, ethylbenzene, propylbenzene, cyclohexane, and cyclooctane, using sodium periodate as the oxidant and imidazole as a co-catalyst.64 The hybrid material exhibited high activity (52–72% conversion) and selectivity (58–100%) to the corresponding alcohols and ketones at 298 K.

5.2.8 Oxidation of alkenes and sulfoxidation of aryl sulphide. The POM encapsulated in MIL-101 was applied to the epoxidation of alkenes using H2O2 as the oxidant.31,69,75 Bromberg et al. reported the epoxidation of p-caryophyllene using H3PW/MIL-101.31 The material exhibited high activity (95% yield to caryophyllene oxide) using 30% aqueous H2O2 as the oxidant in acetonitrile at 328 K after 5 min under microwave-assisted heating. In this case, deprotonation of the H3PW anion immobilized within the MIL-101 cages prevented epoxide ring-opening during the reaction. The heterogeneous nature of the catalyst was confirmed by a hot filtering experiment. XRD confirmed that the structure of MIL-101 had been retained after the recycle runs. Titanium-and cobalt-monosubstituted Keggin heteropolyanions in MIL-101 exhibited high activity (88% conversion) and selectivity (100% to oxide) in the epoxidation of caryophyllene using H2O2 (0.2 M) at 323 K after 4 h, whereas α-pinene and cyclohexene were converted to allyllic products (verbenol/verbenone or cyclohexene-2-ol/cyclohexene-2-one) with good activity (∼40% conversion).75 The same group reported H2O2-based alkene epoxidation using [(PW4O24)3−]/MIL-101 and [(PW12O40)3−]/MIL-101.69 The materials were found to be efficient heterogeneous catalysts for the epoxidation of a variety of alkenes (3-carene, limonene, α-pinene, cyclohexene, cyclooctene, 1-octene) using aqueous H2O2 in acetonitrile at 323 K. The heterogeneity of the catalyst in the liquid phase reaction was confirmed. They also revealed the oxidation of cyclohexene and α-pinene using molecular oxygen (1 bar) in the presence of t-BuOOH over MIL-101 under solvent-free conditions, which produced allyllic products.167 The lanthanide-containing Keggin-type POM and zinc-monosubstituted Keggin heteropolyanions encapsulated into MIL-101 were used as a heterogeneous catalyst for the oxidation of alkenes, such as styrene, cyclohexene, cyclopentene, cyclooctene, indene, 1-octene, 1-decene, and 1-dodecene using H2O2 as an oxidant in acetonitrile with good activity.77,78 The epoxidation of a variety of alkenes (cyclohexene, cyclooctene, styrene, 1-octene and 1-decene) over the Mn-complex-immobilized MIL-101, Mn(TPyP)OAc/CM-MIL-101, were also reported.64 The material exhibited good activity in the epoxidation of alkenes with high yield (77–96%). The heterogeneous nature of the catalyst was provided by XRD and UV-Vis DRS, but the leaching of Mn (0.22%) was observed in the filtrate after the first run.

Hwang et al. reported the catalytic activity of MIL-101 for the liquid-phase sulfoxidation of aryl sulfides to sulfoxides using 35% H2O2 in acetonitrile with high conversion (88–99%) and selectivity (100%) at 298 K after 12 h.168 The oxidation of thioanisole using t-BuOOH at 298 K over the vanadyl(monocatecholate)-decorated MIL-101 catalyst was reported.65 The material showed good activity in the oxidation of thioanisole with 57% and 26% yield to sulfoxide and sulfone, respectively.

5.2.9 Oxidation of aromatic alcohol. The selective oxidation of an alcohol to an aldehyde, particularly benzaldehyde from benzyl alcohol, is one of the most versatile and important classes of organic reactions for laboratories and industries. The development of ecologically acceptable catalytic systems using molecular oxygen or air as an oxidant, which only produces water as a by-product, has attracted considerable attention. The Pd nanoparticle-supported on MIL-101 were found to be an efficient heterogeneous catalyst for the oxidation of a range of alcohols, such as benzyl, allylic, aliphatic, and heterocyclic alcohol to aldehydes using air or O2.94,102 Chen et al. reported the aerobic oxidation of alcohols using pure O2 at 353 K, and the catalyst showed high conversion (95–99%) and selectivity (96–99%) in toluene.94 The authors also reported the oxidation of benzyl alcohol under solvent-free conditions and showed high product selectivity (>99%) with a high turnover frequency (16[thin space (1/6-em)]900 h−1). Recycling runs over the spent catalysts, measurements of the Pd content after the catalytic reaction by atomic absorption spectroscopy (AAS) analysis, and XRD before and after the reaction confirmed the heterogeneity of the reaction. Recently, Ahn et al. reported the oxidation of benzyl alcohol to benzaldehyde over Pd-MIL-101 with 85% conversion and 97% selectivity in o-xylene in air at 358 K after 14 h.102 The oxidation of various alcohols using molecular oxygen over Au MPs supported on MIL-101 was also reported.86 The catalyst exhibited high activity (∼99% conversion and ∼99% product selectivity) in toluene at 353 K, and showed high turnover frequency (TOF) (29[thin space (1/6-em)]300 h−1) due to the well dispersed Au NPs and the electron donation effect of the aryl rings to the Au NPs within the cages of MIL-101. Au NPs supported on PMA-MIL-101 [PMA = phenylene-mono(oxamate)] were also applied to the oxidation of benzyl alcohol using air. In addition, benzyl alcohol was converted to benzaldehyde with TOF ≈ 7 min−1 with high selectivity (100%) in toluene at 353 K, in which Au NPs of ∼2 nm in size were encapsulated inside the MOF cages.169
5.2.10 Hydrogenation reaction. Pd, Pt and Ni NPs supported on MIL-101 were employed in the hydrogenation of alkenes, aldehydes, ketones, phenol, ketoester, nitroarene, benzonitrile, and linoleic acid.84,91,95,97,98,102,104,106 The Pd NPs supported on MIL-101 exhibited good catalytic activity in the hydrogenation of the C[double bond, length as m-dash]C bond in alkenes to the corresponding alkanes and benzaldehyde to benzyl alcohol.84,102 The catalyst was reusable several times without loss of initial high catalytic activity, as confirmed by the repeated reaction cycles and a hot filtration experiment. TEM indicated that the original spherical Pd nanoparticles had a similar morphology to that of the fresh catalyst and the good crystalline MIL-101 structure was retained after the recycle runs.102

The Pd nanoparticles-supported MIL-101 was used in the hydrogenation of phenol in an aqueous solution using molecular H2 (5 bar) at 323 K, and showed high activity (85% conversion and 98.8% selectivity to cyclohexanone).95 The Pd NPs were located inside the MIL-101 cavities. Zhao et al. reported the hydrogenation of 2,3,5-trimethylbenzoquinone in the liquid phase using Pd/MIL-101 (2 wt% Pd), and 99% conversion was obtained at 353 K after 100 min using H2 (6.0 bar), in which the Pd nanoparticles, 2–3 nm in size, were located in the mesoporous cages of MIL-101.104 Pd/MIL-101 (2 wt% Pd) showed better activity (TOF of 675 min−1) than 2.0 wt% Pd on activated carbon (TOF = 519 min−1). The heterogeneity of the catalytic reaction was confirmed. Hermannsdörfer et al. reported the hydrogenation of 3-heptanone using a PdNi NPs-supported MIL-101 catalyst.91 The catalytic performance of the bimetallic PdNi@MIL-101 was superior to the monometallic Pd@MIL-101 or Ni@MIL-101, supporting the bimetallic synergistic effect of the Pd–Ni NPs. Pan et al. reported the asymmetric hydrogenation of ethyl pyruvate and ethyl-2-oxo-4-phenylbutyrate using a Pt nanoparticle-supported MIL-101 catalyst.97 The catalyst exhibited high activity (92–97% conversion and ∼77% ee) after modification with cinchonidine. The Pt nanoparticles-supported on MIL-101 were also found to be an efficient heterogeneous catalyst for the hydrogenation of nitroarenes, 1-octene, benzonitrile, and linoleic acid.98,106

5.2.11 Suzuki–Miyaura, Ullmann and Heck coupling reaction. Pd NPs-supported MIL-101 were applied to the Suzuki–Miyaura, Ullmann and Heck coupling reactions.57,67,85,169 Yuan et al. conducted rigorous catalytic tests to evaluate the catalytic properties of Pd/MIL-101 for the Suzuki–Miyaura and Ullmann coupling reaction of aryl chlorides in aqueous media.85 The product yield was dependent on the reaction time, molar ratio of the reactants, and catalyst concentration. NaOMe was an excellent base for the Suzuki–Miyaura coupling reaction (Scheme 8) in the presence of TBAB, and the product yield ranged from 81 to 97%. The efficiency of the Suzuki–Miyaura coupling reaction was dependent on the reaction atmosphere such that the reaction under nitrogen gave a much higher yield (∼96%) of the cross-coupling product in the coupling reaction of 4-chloroanisole with phenylboronic acid than the reaction in open air (∼33%). In the Ullmann homo-coupling reaction (Scheme 9), 4-chloroanisole was converted to 4,4-dimethoxy biphenyl (97% yield) under a nitrogen atmosphere using TBAB and NaOMe as a base. In open air, the reaction also proceeded to give a similar product yield under identical reaction conditions. The catalyst was reusable five times without any loss of its catalytic efficiency, but a small amount of Pd (<0.2% of the total) was observed in the solution at the end of the reaction. The Pd-loaded onto amine-grafted MIL-101 was active in the Heck coupling reaction of acrylic acid and iodobenzene in N,N-dimethylacetamide (DMA) in the presence of triethylamine at 393 K (Scheme 10).57,67 The hot filtration test and recycling experiments supported its heterogeneous nature.67
image file: c4ra11259h-s8.tif
Scheme 8 Pd NPs supported MIL-101 catalyzed Suzuki–Miyaura coupling reaction.

image file: c4ra11259h-s9.tif
Scheme 9 Pd NPs supported MIL-101 catalyzed Ullmann coupling reaction.

image file: c4ra11259h-s10.tif
Scheme 10 Amine grafted Pd NPs supported MIL-101 catalyzed Heck coupling reaction.
5.2.12 One-step synthesis of methyl isobutyl ketone. Pd@MIL-101 was applied to a multifunctional catalytic reaction involving condensation, dehydration and hydrogenation steps for the liquid-phase synthesis of methyl isobutyl ketone (MIBK), as described in Scheme 11.29 The material showed better performance (60% conversion and 90.2% selectivity) in the one-pot synthesis of MIBK in acetone and H2 (7.5 bar) at 423 K than palladium on Zn–Cr mixed oxide or on the MCM-22 catalysts under the same reaction conditions. The catalyst could be reused several times without any significant loss of its initial catalytic activity, and XRD and TEM confirmed that the crystallinity of Pd@MIL-101 had been retained during the reaction with no metal leaching or agglomeration.
image file: c4ra11259h-s11.tif
Scheme 11 One-step synthesis of methyl isobutyl ketone.
5.2.13 One-pot indole synthesis. Pd/MIL-101 (1.0 mol% Pd) was applied as a heterogeneous catalyst for the synthesis of indole in an aqueous medium, and 2-iodoaniline and phenylacetylene were converted to the product with ∼95% yield at 363 K after 15 h (Scheme 12).87 The catalyst showed higher catalytic efficiency than the Pd/MCM-41 catalyst, owing to the enhanced surface hydrophobicity and the presence of Lewis acid sites on MIL-101. In addition, the catalyst showed higher activity than commercial Pd/C, which might be due to the highly dispersed Pd active sites and the confinement effect of Pd NPs inside the cages of MIL-101. The catalyst could be reused ten times without any loss of its initial catalytic activity, and XRD and TEM confirmed that the crystallinity of Pd/MIL-101 had been retained during the reaction with no metal leaching or agglomeration.
image file: c4ra11259h-s12.tif
Scheme 12 One-pot synthesis of indole.
5.2.14 Direct C2 arylation of indoles. The Pd nanoparticles supported on MIL-101 (0.1 mol% Pd) was an efficient heterogeneous catalyst for the direct C2 arylation of the substituted indoles using caesium acetate in DMF at 393 K for 24 h (Scheme 13).88 Pd/MIL-101 exhibited higher activity (∼85% yield) than commercial Pd/C or Pd-MIL-53 (Al)–NH2 (∼25% yield). The higher catalytic performance of Pd/MIL-101 originated from the homogeneously-distributed Pd NPs within the accessible mesoporous cages of MIL-101. The heterogeneity of the Pd/MIL-101 catalyst in the liquid phase reaction was confirmed by a hot filtration experiment, but the minor leaching of Pd (0.4 ppm) was observed in the filtrate solution after recycling.
image file: c4ra11259h-s13.tif
Scheme 13 C2-arylation of substituted indoles.
5.2.15 Chemical H2 storage reaction. The hydrogen-rich material, such as ammonia borane (NH3BH3), hydrous hydrazine (N2H4·H2O) and formic acid (HCOOH), are commonly used as chemical hydrogen storage materials that release H2 through the hydrolysis, dehydrogenation and pyrolysis reactions. Au, Pd, Pt, Ru and Ni nanoparticles supported on MIL-101 were tested for chemical H2 storage.89,92,100,103,105 Gu et al. reported the dehydrogenation of formic acid using a range of bimetallic and monometallic Au, Pd, Pt, and Ru nanoparticles immobilized on MIL-101 and ED-MIL-101 (Scheme 14).89 Both Au–Pd/ED-MIL-101 and Au–Pd/MIL-101 exhibited higher activity in water at 363 K for the generation of high quality H2 than monometallic Pd/ED-MIL-101, owing to the bimetallic synergistic effect of Au–Pd NPs. No activity was observed when Au/ED-MIL-101 or Ru/MIL-101 was used under identical reaction conditions. The stability/durability was confirmed by recycle runs. Recently, Cao et al. reported hydrogen generation from an aqueous alkaline solution (0.5 M NaOH) of hydrazine using NiPt@MIL-101 at 323 K (Scheme 15).105 The material Ni88Pt12@MIL-101 exhibited excellent catalytic activity with a high TOF of 350 mol H2 molmetal−1 h−1. On the other hand, no activity was observed in the dehydrogenation of the aqueous solution of hydrazine using the parent MIL-101. The catalyst exhibited higher catalytic activity when compared to other NiPt NPs on ZIF-8 or Al2O3, Ce2O3, or dendrimer under identical reaction conditions.
image file: c4ra11259h-s14.tif
Scheme 14 Hydrogen storage from formic acid.

image file: c4ra11259h-s15.tif
Scheme 15 Hydrogen storage from aqueous alkaline solution of hydrazine.

Pt@MIL-101 was tested in liquid-phase NH3BH3 hydrolysis and solid-phase NH3BH3 thermal dehydrogenation.92 In the liquid-phase using Pt/NH3BH3 (molar ratio of 0.0029), the material Pt@MIL-101 (2 wt% Pt) showed two times higher activity to release H2 from NH3BH3 at room temperature than 2 wt% Pt/γ-Al2O3, indicating the small size of the Pt NPs (1.8 ± 0.2 nm) within MIL-101, where the active sites located inside the MOF matrix led to higher activity (Scheme 16). When NH3BH3 loaded either into MIL-101 or Pt@MIL-101 was applied to solid phase thermal dehydrogenation (Scheme 17), the latter released pure H2, whereas the former evolved H2 and NH3 but no borazine was formed in either case. XRD and TEM clearly confirmed that the crystallinity and Pt particle size of Pt@MIL-101 were retained after the completion of the reaction. The Pd nanoparticles supported on MIL-101 (4 wt% Pd) were applied to the chemical storage of H2 and exhibited high activity in the hydrolysis of NH3BH3 to release H2 at room temperature with a TOF of 45 mol H2 molPd−1 min−1, which was higher than that reported for other Pd-based nano-catalysts.103 The stability of the catalyst was confirmed by performing recycle runs. The bimetallic catalyst, AuNi@MIL-101, with an Au/Ni atomic ratio of 7[thin space (1/6-em)]:[thin space (1/6-em)]93 was also an efficient catalyst for the generation of H2 in liquid-phase NH3BH3 hydrolysis at room temperature with a TOF of 66.2 mol H2 molcat−1 min−1.100 The catalyst could be reused five times without significant loss of the initial catalytic activity, and XRD and TEM confirmed that the crystallinity of AuNi@MIL-101 had been retained during the reaction with no metal leaching or agglomeration.


image file: c4ra11259h-s16.tif
Scheme 16 Hydrogen release from ammonia borane.

image file: c4ra11259h-s17.tif
Scheme 17 Hydrogen release from ammonia borane loaded on Pt@MIL-101.
5.2.16 Other catalytic reactions. H3PW/MIL-101 was tested for two acid-catalysed reactions of esterification of n-butanol with acetic acid in the liquid phase to butyl acetate and the gas phase dehydration of methanol to dimethyl ether in a fixed bed in good yield.70 The sulfonic acid functionalized MIL-101 was an efficient catalyst for the liquid-phase acetalization of benzaldehyde and glycol to form benzaldehyde glycol acetate with high conversion (97.3%) and selectivity (100%) at 353 K, which are higher in activity than USY (50.3% conversion).53 Pd@MIL-101 was applied to the multifunctional catalytic reaction involving the isomerization and hydrogenation steps for the liquid-phase synthesis of menthol.90 The material exhibited high activity (>99% conversion, 86% selectivity to menthol and 81% diasteroselectivity to (−)-menthol) in the one-pot synthesis of menthol from citronellal at 453 K for 18 h.

Yadav et al. reported the reduction of highly water-soluble and toxic Cr(VI) to the much less toxic Cr(III) using Pt or Pd immobilized MIL-101 in the presence of excess formic acid.99 Pt (2 wt%)@MIL-101 exhibited higher activity (∼100% after 40 min) in the reduction of an aqueous solution of potassium dichromate at 323 K than Pd (2 wt%)@MIL-101 (∼100% after 210 min), as measured by UV-Vis spectroscopy. MIL-101, Au@MIL-101, and Rh@MIL-101 were catalytically inactive for this reaction under identical reaction conditions.

Copper NPs immobilized onto MIL-101 were applied to the catalytic reduction of 4-nitrophenol and 4-nitroaniline using NaBH4 in the aqueous phase, which were converted to 4-aminophenol and p-phenylenediamine, respectively, in good yield, as measured by UV-Vis spectroscopy.101 The catalytic performance of the Cu nanoparticles on MIL-101 was superior to that of the pure CuNPs. Pd- and Pt-containing MIL-101 were found to be efficient catalysts for the one-pot nitroarene reduction and reductive amination of carbonyl compounds.170 The catalysts exhibited better activity than the commercially available Pd and Pt metal catalysts under identical reaction conditions.

6. Other applications

6.1 Drug delivery and biomedical analysis

Horcajada et al. examined the adsorption and delivery of an anti-inflammatory drug, ibuprofen, by MIL-101.171 MIL-101 allowed a high dose of the drug and long controlled delivery, providing advantages for holding large pharmacological molecules; MIL-101 showed an adsorption capacity of 1.4 g of ibuprofen per gram of dehydrated MIL-101 (Fig. 16) and the delivery rate was significantly slower than that with MCM-41. Čendak et al. examined the interactions of indomethacin and tetrahydrofuran solvent molecules within MIL-101.172 Loading a MIL-101 with indomethacin proved to be very efficient (0.9–1.1 g of indomethacin per gram of MIL-101). Solid-state NMR measurements, however, showed that indomethacin did not attach to the MOF, whereas the THF solvent molecules attached to the framework metallic trimeric units through hydrogen bonding and were not removed sufficiently during drying.
image file: c4ra11259h-f16.tif
Fig. 16 Ibuprofen delivery (mg IBU per g dehydrated material vs. t) from MIL-101 in comparison to MIL-100 and MCM-41.171

Gu et al. reported the effective enrichment of peptides with the simultaneous exclusion of proteins from complex biological samples in MIL-101, which was attributed to the molecular sieving effect by the highly ordered micropores in MIL-101 for the selective inclusion and exclusion of large biomolecules.173 MIL-101 was stable throughout the enrichment procedure, and the peptides were unaffected.

6.2 Gas/liquid chromatographic separation

Yan et al. reported a series of studies on the utilization of MIL-101 as a stationary phase for gas/liquid chromatographic separation.174–177 They achieved the baseline separation of p-xylene, o-xylene, m-xylene, and ethylbenzene (EB) on the MIL-101 coated capillary column by GC within 1.6 min without the need for temperature programming (Fig. 17).176
image file: c4ra11259h-f17.tif
Fig. 17 GC separation of xylene isomers and EB on a MIL-101 coated capillary column (15 m long × 0.53 mm i.d.) at 160 °C under a N2 flow rate of 3 mL min−1. The mass of each isomer was 350 ng.176

The excellent selectivity of the MIL-101 coated capillary column for the separation of xylene isomers and EB was attributed to the host–guest interactions and the CUS and suitable polarity of MIL-101. In addition, columns packed with MIL-101 allowed the excellent separation of not only C60 and C70 (achieved within 3 min with a selectivity of αC70/C60 = 17), but also other high fullerenes, such as C76, C78, C82, C84, C86, and C96,174 which was attributed to the differences in fullerene solubility in the mobile phase, and the π–π and van der Waals interactions between the fullerenes and the MIL-101 frameworks. The MIL-101 packed column also gave baseline separation of dichlorobenzene (DCB) and chlorotoluene isomers, and EB and styrene with a high column efficiency and good precision.175 MIL-101 offered high affinity for the ortho-isomers, allowing rapid and selective separation from other isomers within 3 min using dichloromethane as the mobile phase. They also attempted the high performance liquid chromatographic separation of nitroaniline, aminophenol and naphthol isomers as well as a sulfadimidine/sulfanilamide mixture on the MIL-101 packed column.177 The interaction between the open metal sites of MIL-101 and the polar analytes was adjusted by adding the appropriate amount of MeOH to the mobile phase to achieve the effective separation of the polar analytes by making use of the competition of MeOH with the analytes for open metal sites. Huang et al. reported MOF-organic polymer monoliths prepared by the microwave-assisted polymerization of ethylene dimethacrylate (EDMA), butyl methacrylate (BMA) and 2-acrylamido-2-methylpropane sulfonic acid (AMPS) with MIL-101 as the stationary phases for capillary electrochromatography (CEC) and nano-liquid chromatography (nano-LC).178 The hybrid column showed high permeability and efficient separation of various analytes (xylene, chlorotoluene, cymene, aromatic acids, polycyclic aromatic hydrocarbons, and trypsin digested BSA peptides) either in CEC or nano-LC. These merits were attributed to the inclusion of MIL-101 in the column material, which provided high surface areas, nano-sized pores, and aromatic terephthalate moieties in the MOF–polymer monolith. Hu et al. used a MIL-101 packed micro-column for the on-line sorptive extraction and direct analysis of naproxen and its metabolites from urine samples.179 Through the MIL-101 based in-tube sorptive extraction method coupled with high-performance liquid chromatography (HPLC) and fluorescence detection, naproxen and 6-O-desmethylnaproxen in urine samples could be analysed directly without an additional sample pre-treatment. Hu et al. examined the separation of two liquid mixtures (arginine, phenylalanine, and tryptophan in water, and three xylene isomers in hexane) using MIL-101 as the stationary phase by molecular simulation.180 For the first mixtures, the elution order was found to be arginine > phenylalanine > tryptophan, whereas the elution order was p-xylene > m-xylene > o-xylene for the second mixture. A separation mechanism based on the cooperative solute–solvent and solute–framework interactions was proposed.

6.3 Proton conduction

Ponomareva et al. impregnated MIL-101 with H2SO4 or H3PO4 to afford solid materials with proton-conducting properties at moderate temperatures.61 The proton conductivities of the H2SO4@MIL-101 and H3PO4@MIL-101 at 423 K under low humidity conditions outperformed those by other MOF-based materials, and were comparable to Nafion®. The acid molecules residing inside the pores of MIL-101 were not removed by either heating or by hydration in a humid atmosphere. Recently, Dybtsev et al. reported the incorporation of triflic and toluene sulphinic acids into MIL-101 to produce hybrid proton-conducting solid electrolytes.181 The hybrid material, triflic acid@MIL-101 showed high proton conductivity (0.08 S cm−1) at 15% relative humidity and a temperature of 333 K. The material was found to be stable during the multiple measurements or prolonged heating.

6.4 Sensors

Hu et al. reported the surface-enhanced Raman scattering (SERS) detection by Au nanoparticles-embedded MIL-101.182 The as-synthesized AuNPs/MIL-101 nanocomposites were highly sensitive SERS substrates by effectively pre-concentrating the analytes in close proximity to the electromagnetic fields at the SERS-active metal surface owing to a combination of the localized surface plasmon resonance properties of the AuNPs and the high adsorption ability of MIL-101.

7. Conclusions

MIL-101 can be synthesized hydrothermally using acid additives as well as in water in acid-free synthesis under high pressure conditions. The textural properties of MIL-101, however, differ widely among research groups, and a more reproducible synthesis with more effective purification protocols should be established. The MIL-101 materials produced in aqueous reaction media without acid or alkali additives are more environmentally benign, but slight deterioration in textural properties occurs. Alternative synthesis methods demanding less energy and practical synthesis scale-up will be needed in the future. In this regard, the dry gel conversion synthesis of MIL-101 allowing a high product yield with a simple purification/activation procedure can be promising. Organic-functionalization or the encapsulation of various inorganic/organic guest molecules in MIL-101 have been investigated widely for adsorption/catalytic applications, and the scope of research will definitely grow for practical applications.

MIL-101 has high thermal stability and strong resistance to moisture and organic molecules, which makes it a good candidate material for the adsorption of gases, water and volatile organic compounds in an industrial perspective. Large amounts of H2 and CO2 can be adsorbed by MIL-101, both with a high heat of adsorption under high pressure conditions. H2 storage capacity can be enhanced through a variety of methods, such as noble metal spillover and cation doping. The enhanced CO2 adsorption performance by amine-functionalized MIL-101 is also well-established, which has already shown the efficient separation of CO2 and high selectivity over CH4, N2 and CO. The functionalization of MIL-101, particularly through post synthetic methods, normally leads to a hybrid material with a lower surface area, resulting in decreased CO2 uptake under higher pressure conditions. MIL-101 shows high equilibrium uptake, rapid sorption/desorption rates at moderate temperatures and excellent hydrothermal stability for H2O adsorption, making it a promising material for heat transformation applications or energy-efficient H2O sorption applications. MIL-101 also shows high adsorption performance for volatile organic compounds, nitrogen- or sulfur-containing compounds, and dyes. Specific functional groups have been introduced depending on the specific adsorbates to improve the adsorption performance.

Coordinatively unsaturated Cr sites in MIL-101 are active sites for the catalysis of Lewis acid-catalysed and sometimes oxidation reactions. MIL-101 can also be synthesized or modified as a bifunctional catalyst bearing both Brønsted acid and base functionalities. In addition, the immobilization of noble metal nanoparticles on MIL-101 can also produce a wide range of nanoparticle-supported MIL-101, which are useful for cascade or tandem reactions that display high atomic utilization efficiency and do not involve workup and the isolation of many intermediates. Heteropolyacid-impregnated MIL-101 is also promising for a range of liquid phase acid/base and oxidation reactions. In some cases, Cr or impregnated metals have been detected in the filtrate of the liquid phase reactions, and stability of the catalytic system needs to be monitored closely in some cases. The noble metal or non-noble metal nanoparticles deposited on MIL-101 also show promising catalytic activities for chemical hydrogen storage. The number of such hybrid materials is expected to grow rapidly in the near future, which will not only be applicable to clean energy technology, but also to the preparation of materials for the release of drugs and pharmaceutical chemicals.

MIL-101 in chromatography/membrane separation, proton conduction, and surface-enhanced Raman scattering detection is also under investigation.

There is a large potential of MIL-101 for sustainability, and further research on the introduction of new functional groups to the MIL-101 will be useful to produce versatile hybrid materials for future industrial applications.

Acknowledgements

This study was supported a NRF grant (2013005862) funded by Korean government (MEST).

Notes and references

  1. M. Eddaoudi, D. B. Moler, H. Li, B. Chen, T. M. Reineke, M. O'Keeffe and O. M. Yaghi, Acc. Chem. Res., 2001, 34, 319 CrossRef CAS PubMed.
  2. S. L. James, Chem. Soc. Rev., 2003, 32, 276 RSC.
  3. N. Stock and S. Biswas, Chem. Rev., 2012, 112, 933 CrossRef CAS PubMed.
  4. A. R. Millward and O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 17998 CrossRef CAS PubMed.
  5. A. Ö. Yazaydın, R. Q. Snurr, T. H. Park, K. Koh, J. Liu, M. D. LeVan, A. I. Benin, P. Jakubczak, M. Lanuza, D. B. Galloway, J. J. Low and R. R. Willis, J. Am. Chem. Soc., 2009, 131, 18198 CrossRef PubMed.
  6. J. R. Li, R. J. Kuppler and H. C. Zhou, Chem. Soc. Rev., 2009, 38, 1477 RSC.
  7. J. Kim, S. T. Yang, S. B. Choi, J. Sim, J. Kim and W. S. Ahn, J. Mater. Chem., 2011, 21, 3070 RSC.
  8. D. A. Yang, H. Y. Cho, J. Kim, S. T. Yang and W. S. Ahn, Energy Environ. Sci., 2012, 5, 6465 CAS.
  9. 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 CrossRef CAS PubMed.
  10. A. Corma, H. García and F. X. Llabrés i Xamena, Chem. Rev., 2010, 110, 4606 CrossRef CAS PubMed.
  11. P. Horcajada, T. Chalati, C. Serre, B. Gillet, C. Sebrie, T. Baati, J. F. Eubank, D. Heurtaux, P. Clayette, C. Kreuz, J. S. Chang, Y. K. Hwang, V. Marsaud, P. N. Bories, L. Cynober, S. Gil, G. Férey, P. Couvreur and R. Gref, Nat. Mater., 2010, 9, 172 CrossRef CAS PubMed.
  12. G. Férey, Chem. Soc. Rev., 2008, 37, 191 RSC.
  13. D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int. Ed., 2009, 48, 7502 CrossRef CAS PubMed.
  14. J. Y. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen and J. T. Hupp, Chem. Soc. Rev., 2009, 38, 1450 RSC.
  15. A. U. Czaja, N. Trukhan and U. Müller, Chem. Soc. Rev., 2009, 38, 1284 RSC.
  16. Z. Wang and S. M. Cohen, Chem. Soc. Rev., 2009, 38, 1315 RSC.
  17. S. M. Cohen, Chem. Sci., 2010, 1, 32 RSC.
  18. S. T. Meek, J. A. Greathouse and M. D. Allendorf, Adv. Mater., 2011, 23, 249 CrossRef CAS PubMed.
  19. J. Kim, H. Y. Cho and W. S. Ahn, Catal. Surv. Asia, 2012, 16, 106 CrossRef CAS.
  20. A. Dhakshinamoorthy and H. Garcia, Chem. Soc. Rev., 2012, 41, 5262 RSC.
  21. S. M. Cohen, Chem. Rev., 2012, 112, 970 CrossRef CAS PubMed.
  22. M. Kim and S. M. Cohen, CrystEngComm, 2012, 14, 4096 RSC.
  23. Z. Zhang, Z. Z. Yao, S. Xiang and B. Chen, Energy Environ. Sci., 2014, 7, 2868 CAS.
  24. Q. L. Zhu and Q. Xu, Chem. Soc. Rev., 2014, 43, 5468 RSC.
  25. G. Férey, C. Serre, C. Mellot Draznieks, F. Millange, S. Surblé, J. Dutour and I. Margiolaki, Angew. Chem., Int. Ed., 2004, 43, 6296 CrossRef PubMed.
  26. G. Férey, C. Mellot Draznieks, C. Serre, F. Millange, J. Dutour, S. Surblé and I. Margiolaki, Science, 2005, 309, 2040 CrossRef PubMed.
  27. D. Y. Hong, Y. K. Hwang, C. Serre, G. Férey and J. S. Chang, Adv. Funct. Mater., 2009, 19, 1537 CrossRef CAS.
  28. J. Yang, Q. Zhao, J. Li and J. Dong, Microporous Mesoporous Mater., 2010, 130, 174 CrossRef CAS PubMed.
  29. Y. Pan, B. Yuan, Y. Li and D. He, Chem. Commun., 2010, 46, 2280 RSC.
  30. C. Y. Huang, M. Song, Z. Y. Gu, H. F. Wang and X. P. Yan, Environ. Sci. Technol., 2011, 45, 4490 CrossRef CAS PubMed.
  31. L. Bromberg, Y. Diao, H. Wu, S. A. Speakman and T. A. Hatton, Chem. Mater., 2012, 24, 1664 CrossRef CAS.
  32. D. Jiang, A. D. Burrows and K. J. Edler, CrystEngComm, 2011, 13, 6916 RSC.
  33. X. X. Huang, L. G. Qiu, W. Zhang, Y. P. Yuan, X. Jiang, A. J. Xie, Y. H. Shen and J. F. Zhu, CrystEngComm, 2012, 14, 1613 RSC.
  34. L. T. Yang, L. G. Qiu, S. M. Hu, X. Jiang, A. J. Xie and Y. H. Shen, Inorg. Chem. Commun., 2013, 35, 265 CrossRef CAS PubMed.
  35. H. Chen, S. Chen, X. Yuan and Y. Zhang, Mater. Lett., 2013, 100, 230 CrossRef CAS PubMed.
  36. D. Jiang, A. D. Burrows, R. Jaber and K. J. Edler, Chem. Commun., 2012, 48, 4965 RSC.
  37. D. Jiang, A. D. Burrows, Y. Xiong and K. J. Edler, J. Mater. Chem. A, 2013, 1, 5497 CAS.
  38. S. H. Jhung, J. H. Lee, J. W. Yoon, C. Serre, G. Férey and J. S. Chang, Adv. Mater., 2007, 19, 121 CrossRef CAS.
  39. A. Demessence, P. Horcajada, C. Serre, C. Boissière, D. Grosso, C. Sanchez and G. Férey, Chem. Commun., 2009, 7149 RSC.
  40. Z. Zhao, X. Li, S. Huang, Q. Xia and Z. Li, Ind. Eng. Chem. Res., 2011, 50, 2254 CrossRef CAS.
  41. N. A. Khan, I. J. Kang, H. Y. Seok and S. H. Jhung, Chem. Eng. J., 2011, 166, 1152 CrossRef CAS PubMed.
  42. J. Kim, Y. R. Lee and W. S. Ahn, Chem. Commun., 2013, 49, 7647 RSC.
  43. S. H. Jhung, J. S. Chang, J. S. Hwang and S. E. Park, Microporous Mesoporous Mater., 2003, 64, 33 CrossRef CAS.
  44. S. H. Jhung, J. H. Lee, J. W. Yoon, J. S. Hwang, S. E. Park and J. S. Chang, Microporous Mesoporous Mater., 2005, 80, 147 CrossRef CAS PubMed.
  45. K. K. Kang, C. H. Park and W. S. Ahn, Catal. Lett., 1999, 59, 45 CrossRef CAS.
  46. Y. K. Hwang, J. S. Chang, S. E. Park, D. S. Kim, Y. U. Kwon, S. H. Jhung, J. S. Hwang and M. S. Park, Angew. Chem., Int. Ed., 2005, 44, 556 CrossRef CAS PubMed.
  47. S. Bernt, V. Guillerm, C. Serre and N. Stock, Chem. Commun., 2011, 47, 2838 RSC.
  48. G. Akiyama, R. Matsuda, H. Sato, M. Takata and S. Kitagawa, Adv. Mater., 2011, 23, 3294 CrossRef CAS PubMed.
  49. G. Akiyama, R. Matsuda, H. Sato, A. Hori, M. Takata and S. Kitagawa, Microporous Mesoporous Mater., 2012, 157, 89 CrossRef CAS PubMed.
  50. Y. Lin, C. Kong and L. Chen, RSC Adv., 2012, 2, 6417 RSC.
  51. M. Lammert, S. Bernt, F. Vermoortele, D. E. De Vos and N. Stock, Inorg. Chem., 2013, 52, 8521 CrossRef CAS PubMed.
  52. F. L. Morel, M. Ranocchiari and J. A. van Bokhoven, Ind. Eng. Chem. Res., 2014, 53, 9120 CrossRef CAS.
  53. Y. Jin, J. Shi, F. Zhang, Y. Zhong and W. Zhu, J. Mol. Catal. A: Chem., 2014, 383–384, 167 CrossRef CAS PubMed.
  54. D. Jiang, L. L. Keenan, A. D. Burrows and K. J. Edler, Chem. Commun., 2012, 48, 12053 RSC.
  55. B. Li, Y. Zhang, D. Ma, L. Li, G. Li, G. Li, Z. Shi and S. Feng, Chem. Commun., 2012, 48, 6151 RSC.
  56. Y. R. Lee, Y. M. Chung and W. S. Ahn, RSC Adv., 2014, 4, 23064 RSC.
  57. Y. K. Hwang, D. Y. Hong, J. S. Chang, S. H. Jhung, Y. K. Seo, J. Kim, A. Vimont, M. Daturi, C. Serre and G. Férey, Angew. Chem., Int. Ed., 2008, 47, 4144 CrossRef CAS PubMed.
  58. Y. Lin, Q. Yan, C. Kong and L. Chen, Sci. Rep., 2013, 3, 1859 Search PubMed.
  59. Q. Yan, Y. Lin, C. Kong and L. Chen, Chem. Commun., 2013, 49, 6873 RSC.
  60. P. B. S. Rallapalli, M. C. Raj, D. V. Patil, K. P. Prasanth, R. S. Somani and H. C. Bajaj, Int. J. Energy Res., 2013, 37, 746 CrossRef.
  61. V. G. Ponomareva, K. A. Kovalenko, A. P. Chupakhin, D. N. Dybtsev, E. S. Shutova and V. P. Fedin, J. Am. Chem. Soc., 2012, 134, 15640 CrossRef CAS PubMed.
  62. I. Ahmed, N. A. Khan and S. H. Jhung, Inorg. Chem., 2013, 52, 14155 CrossRef CAS PubMed.
  63. M. G. Goesten, K. B. S. S. Gupta, E. V. Ramos-Fernandez, H. Khajavi, J. Gascon and F. Kapteijn, CrystEngComm, 2012, 14, 4109 RSC.
  64. F. Zadehahmadi, S. Tangestaninejad, M. Moghadam, V. Mirkhani, I. Mohammadpoor-Baltork, A. R. Khosropour and R. Kardanpour, Appl. Catal., A, 2014, 477, 34 CrossRef CAS PubMed.
  65. H. G. T. Nguyen, M. H. Weston, O. K. Farha, J. T. Hupp and S. T. Nguyen, CrystEngComm, 2012, 14, 4115 RSC.
  66. X. Wang, H. Li and X. J. Hou, J. Phys. Chem. C, 2012, 116, 19814 CAS.
  67. S. N. Kim, S. T. Yang, J. Kim, J. E. Park and W. S. Ahn, CrystEngComm, 2012, 14, 4142 RSC.
  68. S. Wang, L. Bromberg, H. Schreuder-Gibson and T. A. Hatton, ACS Appl. Mater. Interfaces, 2013, 5, 1269 CAS.
  69. N. V. Maksimchuk, K. A. Kovalenko, S. S. Arzumanov, Y. A. Chesalov, M. S. Melgunov, A. G. Stepanov, V. P. Fedin and O. A. Kholdeeva, Inorg. Chem., 2010, 49, 2920 CrossRef CAS PubMed.
  70. J. Juan-Alcañiz, E. V. Ramos-Fernandez, U. Lafont, J. Gascon and F. Kapteijn, J. Catal., 2010, 269, 229 CrossRef PubMed.
  71. L. Bromberg and T. A. Hatton, ACS Appl. Mater. Interfaces, 2011, 3, 4756 CAS.
  72. S. Ribeiro, A. D. S. Barbosa, A. C. Gomes, M. Pillinger, I. S. Gonçalves, L. Cunha-Silva and S. Balula, Fuel Process. Technol., 2013, 116, 350 CrossRef CAS PubMed.
  73. L. H. Wee, F. Bonino, C. Lamberti, S. Bordiga and J. A. Martens, Green Chem., 2014, 16, 1351 RSC.
  74. I. Ahmed, N. A. Khan, Z. Hasan and S. H. Jhung, J. Hazard. Mater., 2013, 250–251, 37 CrossRef CAS PubMed.
  75. N. V. Maksimchuk, M. N. Timofeeva, M. S. Melgunov, A. N. Shmakov, Y. A. Chesalov, D. N. Dybtsev, V. P. Fedin and O. A. Kholdeeva, J. Catal., 2008, 257, 315 CrossRef CAS PubMed.
  76. D. M. Fernandes, A. D. S. Barbosa, J. Pires, S. S. Balula, L. Cunha-Silva and C. Freire, ACS Appl. Mater. Interfaces, 2013, 5, 13382 CAS.
  77. C. M. Granadeiro, P. Silva, V. K. Saini, F. A. A. Paz, J. Pires, L. Cunha-Silva and S. S. Balula, Catal. Today, 2013, 218–219, 35 CrossRef CAS PubMed.
  78. Z. Saedi, S. Tangestaninejad, M. Moghadam, V. Mirkhani and I. Mohammadpoor-baltork, J. Coord. Chem., 2012, 65, 463 CrossRef CAS.
  79. I. C. M. S. Santos, S. S. Balula, M. M. Q. Simões, L. Cunha-Silva, M. G. P. M. S. Neves, B. de Castro, A. M. V. Cavaleiro and J. A. S. Cavaleiro, Catal. Today, 2013, 203, 87 CrossRef CAS PubMed.
  80. E. Kockrick, T. Lescouet, E. V. Kudrik, A. B. Sorokin and D. Farrusseng, Chem. Commun., 2011, 47, 1562 RSC.
  81. O. V. Zalomaeva, K. A. Kovalenko, Y. A. Chesalov, M. S. Mel'gunov, V. I. Zaikovskii, V. V. Kaichev, A. B. Sorokin, O. A. Kholdeeva and V. P. Fedin, Dalton Trans., 2011, 40, 1441 RSC.
  82. M. Anbia and V. Hoseini, Chem. Eng. J., 2012, 191, 326 CrossRef CAS PubMed.
  83. X. Zhou, W. Huang, J. Shi, Z. Zhao, Q. Xia, Y. Li, H. Wang and Z. Li, J. Mater. Chem. A, 2014, 2, 4722 CAS.
  84. A. Henschel, K. Gedrich, R. Kraehnert and S. Kaskel, Chem. Commun., 2008, 4192 RSC.
  85. B. Yuan, Y. Pan, Y. Li, B. Yin and H. Jiang, Angew. Chem., Int. Ed., 2010, 49, 4054 CrossRef CAS PubMed.
  86. H. Liu, Y. Liu, Y. Li, Z. Tang and H. Jiang, J. Phys. Chem. C, 2010, 114, 13362 CAS.
  87. H. Li, Z. Zhu, F. Zhang, S. Xie, H. Li, P. Li and X. Zhou, ACS Catal., 2011, 1, 1604 CrossRef CAS.
  88. Y. Huang, Z. Lin and R. Cao, Chem.–Eur. J., 2011, 17, 12706 CrossRef CAS PubMed.
  89. X. Gu, Z. H. Lu, H. L. Jiang, T. Akita and Q. Xu, J. Am. Chem. Soc., 2011, 133, 11822 CrossRef CAS PubMed.
  90. F. G. Cirujano, F. X. Llabrés i Xamena and A. Corma, Dalton Trans., 2012, 41, 4249 RSC.
  91. J. Hermannsdörfer, M. Friedrich, N. Miyajima, R. Q. Albuquerque, S. Kümmel and R. Kempe, Angew. Chem., Int. Ed., 2012, 51, 11473 CrossRef PubMed.
  92. A. Aijaz, A. Karkamkar, Y. J. Choi, N. Tsumori, E. Rönnebro, T. Autrey, H. Shioyama and Q. Xu, J. Am. Chem. Soc., 2012, 134, 13926 CrossRef CAS PubMed.
  93. Z. Sun, G. Li, L. Liu and H. Liu, Catal. Commun., 2012, 27, 200 CrossRef CAS PubMed.
  94. G. Chen, S. Wu, H. Liu, H. Jiang and Y. Li, Green Chem., 2013, 15, 230 RSC.
  95. D. Zhang, Y. Guan, E. J. M. Hensen, L. Chen and Y. Wang, Catal. Commun., 2013, 41, 47 CrossRef CAS PubMed.
  96. A. Aijaz, T. Akita, N. Tsumori and Q. Xu, J. Am. Chem. Soc., 2013, 135, 16356 CrossRef CAS PubMed.
  97. H. Pan, X. Li, D. Zhang, Y. Guan and P. Wu, J. Mol. Catal. A: Chem., 2013, 377, 108 CrossRef CAS PubMed.
  98. W. Du, G. Chen, R. Nie, Y. Li and Z. Hou, Catal. Commun., 2013, 41, 56 CrossRef CAS PubMed.
  99. M. Yadav and Q. Xu, Chem. Commun., 2013, 49, 3327 RSC.
  100. Q. L. Zhu, J. Li and Q. Xu, J. Am. Chem. Soc., 2013, 135, 10210 CrossRef CAS PubMed.
  101. F. Wu, L. G. Qiu, F. Ke and X. Jiang, Inorg. Chem. Commun., 2013, 32, 5 CrossRef CAS PubMed.
  102. S. Bhattacharjee, J. Kim and W. S. Ahn, J. Nanosci. Nanotechnol., 2014, 14, 2546 CrossRef CAS PubMed.
  103. H. Dai, J. Su, K. Hu, W. Luo and G. Cheng, Int. J. Hydrogen Energy, 2014, 39, 4947 CrossRef CAS PubMed.
  104. X. Zhao, Y. Jin, F. Zhang, Y. Zhong and W. Zhu, Chem. Eng. J., 2014, 239, 33 CrossRef CAS PubMed.
  105. N. Cao, J. Su, W. Luo and G. Cheng, Int. J. Hydrogen Energy, 2014, 39, 9726 CrossRef CAS PubMed.
  106. H. Khajavi, H. A. Stil, H. P. C. E. Kuipers, J. Gascon and F. Kapteijn, ACS Catal., 2013, 3, 2617 CrossRef CAS.
  107. T. V. Vu, H. Kosslick, A. Schulz, J. Harloff, E. Paetzold, M. Schneider, J. Radnik, N. Steinfeldt, G. Fulda and U. Kragl, Appl. Catal., A, 2013, 468, 410 CrossRef CAS PubMed.
  108. P. Á. Szilágyi, E. Callini, A. Anastasopol, C. Kwakernaak, S. Sachdeva, R. van de Krol, H. Geerlings, A. Borgschulte, A. Züttel and B. Dam, Phys. Chem. Chem. Phys., 2014, 16, 5803 RSC.
  109. Y. Y. Liu, J. L. Zeng, J. Zhang, F. Xua and L. X. Sun, Int. J. Hydrogen Energy, 2007, 32, 4005 CrossRef CAS PubMed.
  110. M. Latroche, S. Surble, C. Serre, C. Mellot-Draznieks, P. L. Llewellyn, J. H. Lee, J. S. Chang, S. H. Jhung and G. Ferey, Angew. Chem., Int. Ed., 2006, 45, 8227 CrossRef CAS PubMed.
  111. B. Schmitz, U. Muller, N. Trukhan, M. Schubert, G. Ferey and M. Hirscher, ChemPhysChem, 2008, 9, 2181 CrossRef CAS PubMed.
  112. S. N. Klyamkin, E. A. Berdonosova, E. V. Kogan, K. A. Kovalenko, D. N. Dybtsev and V. P. Fedin, Chem.–Asian J., 2011, 6, 1854 CrossRef CAS PubMed.
  113. O. Ardelean, G. Blanita, G. Borodi, M. D. Lazar, I. Misan, I. Coldea and D. Lupu, Int. J. Hydrogen Energy, 2013, 38, 7046 CrossRef CAS PubMed.
  114. H. Oh, D. Lupu, G. Blanita and M. Hirscher, RSC Adv., 2014, 4, 2648 RSC.
  115. Y. Li and R. T. Yang, AIChE J., 2008, 54, 269 CrossRef CAS.
  116. K. S. Lin, A. K. Adhikari, Y. H. Su, C. W. Shu and H. Y. Chan, Adsorption, 2012, 18, 483 CrossRef CAS.
  117. M. Anbia and S. Mandegarzad, J. Alloys Compd., 2012, 532, 61 CrossRef CAS PubMed.
  118. D. Dybtsev, C. Serre, B. Schmitz, B. Panella, M. Hirscher, M. Latroche, P. L. Llewellyn, S. Cordier, Y. Molard, M. Haouas, F. Taulelle and G. Férey, Langmuir, 2010, 26(13), 11283 CrossRef CAS PubMed.
  119. Z. Xiang, Z. Hu, W. Yang and D. Cao, Int. J. Hydrogen Energy, 2012, 37, 946 CrossRef CAS PubMed.
  120. S. Clauzier, L. N. Ho, M. Pera-Titus, B. Coasne and D. Farrusseng, J. Am. Chem. Soc., 2012, 134, 17369 CrossRef CAS PubMed.
  121. K. P. Prasanth, P. Rallapalli, M. C. Raj, H. C. Bajaj and R. Vir Jasra, Int. J. Hydrogen Energy, 2011, 36, 7594 CrossRef CAS PubMed.
  122. Q. Liu, L. Ning, S. Zheng, M. Tao, Y. Shi and Y. He, Sci. Rep., 2013, 3, 2916 Search PubMed.
  123. P. L. Llewellyn, S. Bourrelly, C. Serre, A. Vimont, M. Daturi, L. Hamon, G. D. Weireld, J. S. Chang, D. Y. Hong, Y. K. Hwang, S. H. Jhung and G. Férey, Langmuir, 2008, 24, 7245 CrossRef CAS PubMed.
  124. Z. Zhang, S. Huang, S. Xian, H. Xi and Z. Li, Energy Fuels, 2011, 25, 835 CrossRef CAS.
  125. Z. Liang, M. Marshall, C. H. Ng and A. L. Chaffee, Energy Fuels, 2013, 27, 7612 CrossRef CAS.
  126. K. Munusamy, G. Sethia, D. V. Patil, P. B. S. Rallapalli, R. S. Somani and H. C. Bajaj, Chem. Eng. J., 2012, 195–196, 359 CrossRef CAS PubMed.
  127. P. Chowdhury, S. Mekala, F. Dreisbach and S. Gumma, Microporous Mesoporous Mater., 2012, 152, 246 CrossRef CAS PubMed.
  128. A. Khutia and C. Janiak, Dalton Trans., 2014, 43, 1338 RSC.
  129. P. Chowdhury, C. Bikkina and S. Gumma, J. Phys. Chem. C, 2009, 113, 6616 CAS.
  130. S. J. Lee, J. W. Yoon, Y. K. Seo, M. B. Kim, S. K. Lee, U. H. Lee, Y. K. Hwang, Y. S. Bae and J. S. Chang, Microporous Mesoporous Mater., 2014, 193, 160 CrossRef CAS PubMed.
  131. I. Senkovska, E. Barea, J. A. R. Navarro and S. Kaskel, Microporous Mesoporous Mater., 2012, 156, 115 CrossRef CAS PubMed.
  132. L. Hamon, C. Serre, T. Devic, T. Loiseau, F. Millange, G. Férey and G. D. Weireld, J. Am. Chem. Soc., 2009, 131, 8775 CrossRef CAS PubMed.
  133. N. Klein, A. Henschel and S. Kaskel, Microporous Mesoporous Mater., 2010, 129, 238 CrossRef CAS PubMed.
  134. P. Küsgens, M. Rose, I. Senkovska, H. Frode, A. Henschel, S. Siegle and S. Kaskel, Microporous Mesoporous Mater., 2009, 120, 325 CrossRef PubMed.
  135. J. Ehrenmann, S. K. Henninger and C. Janiak, Eur. J. Inorg. Chem., 2011, 2011, 471 CrossRef.
  136. Y. K. Seo, J. W. Yoon, J. S. Lee, Y. K. Hwang, C. H. Jun, J. S. Chang, S. Wuttke, P. Bazin, A. Vimont, M. Daturi, S. Bourrelly, P. L. Llewellyn, P. Horcajada, C. Serre and G. Férey, Adv. Mater., 2012, 24, 806 CrossRef CAS PubMed.
  137. A. Khutia, H. U. Rammelberg, T. Schmidt, S. Henninger and C. Janiak, Chem. Mater., 2013, 25, 790 CrossRef CAS.
  138. P. Trens, H. Belarbi, C. Shepherd, P. Gonzalez, N. A. Ramsahye, U. H. Lee, Y. K. Seo and J. S. Chang, J. Phys. Chem. C, 2012, 116, 25824 CAS.
  139. T. K. Trung, N. A. Ramsahye, P. Trens, N. Tanchoux, C. Serre, F. Fajula and G. Férey, Microporous Mesoporous Mater., 2010, 134, 134 CrossRef CAS PubMed.
  140. K. Yang, Q. Sun, F. Xue and D. Lin, J. Hazard. Mater., 2011, 195, 124 CrossRef CAS PubMed.
  141. Z. Zhao, X. Li and Z. Li, Chem. Eng. J., 2011, 173, 150 CrossRef CAS PubMed.
  142. P. Trens, H. Belarbi, C. Shepherd, P. Gonzalez, N. A. Ramsahye, U. H. Lee, Y. K. Seo and J. S. Chang, Microporous Mesoporous Mater., 2014, 183, 17 CrossRef CAS PubMed.
  143. J. Shi, Z. Zhao, Q. Xia, Y. Li and Z. Li, J. Chem. Eng. Data, 2011, 56, 3419 CrossRef CAS.
  144. S. Xian, X. Li, F. Xu, Q. Xia and Z. Li, Sep. Sci. Technol., 2013, 48, 1479 CrossRef CAS.
  145. X. Sun, Q. Xia, Z. Zhao, Y. Li and Z. Li, Chem. Eng. J., 2014, 239, 226 CrossRef CAS PubMed.
  146. A. L. Nuzhdin, K. A. Kovalenko, D. N. Dybtsev and G. A. Bukhtiyarova, Mendeleev Commun., 2010, 20, 57 CrossRef CAS PubMed.
  147. L. Wu, J. Xiao, Y. Wu, S. Xian, G. Miao, H. Wang and Z. Li, Langmuir, 2014, 30, 1080 CrossRef CAS PubMed.
  148. N. A. Khan, Z. Hasan and S. H. Jhung, Chem.–Eur. J., 2014, 20, 376 CrossRef CAS PubMed.
  149. C. Chen, M. Zhang, Q. Guan and W. Li, Chem. Eng. J., 2012, 183, 60 CrossRef CAS PubMed.
  150. E. Haque, J. E. Lee, I. T. Jang, Y. K. Hwang, J. S. Chang, J. Jegal and S. H. Jhung, J. Hazard. Mater., 2010, 181, 535 CrossRef CAS PubMed.
  151. F. Leng, W. Wang, X. J. Zhao, X. L. Hu and Y. F. Li, Colloids Surf., A, 2014, 441, 164 CrossRef CAS PubMed.
  152. M. J. Kim, S. M. Park, S. J. Song, J. Won, J. Y. Lee, M. Yoon, K. Kim and G. Seo, J. Colloid Interface Sci., 2011, 361, 612 CrossRef CAS PubMed.
  153. C. X. Yang and X. P. Yan, J. Mater. Chem., 2012, 22, 17833 RSC.
  154. F. X. Qin, S. Y. Jia, Y. Liu, H. Y. Li and S. H. Wu, Desalin. Water Treat., 2014, 1 CrossRef.
  155. Z. Hasan, J. Jeon and S. H. Jhung, J. Hazard. Mater., 2012, 209–210, 151 CrossRef CAS PubMed.
  156. Z. Hasan, E. J. Choi and S. H. Jhung, Chem. Eng. J., 2013, 219, 537 CrossRef CAS PubMed.
  157. A. C. Dillon, K. M. Jones, T. A. Bekkedahl, C. H. Kiang, D. S. Bethune and M. J. Heben, Nature, 1997, 386, 377 CrossRef CAS.
  158. Q. Wang, J. Luo, Z. Zhong and A. Borgna, Energy Environ. Sci., 2011, 4, 42 CAS.
  159. K. Zhang, Y. Chen, A. Nalaparaju and J. Jiang, CrystEngComm, 2013, 15, 10358 RSC.
  160. J. A. Mason, M. Veenstra and J. R. Long, Chem. Sci., 2014, 5, 32 RSC.
  161. H. B. Tanh Jeazet, C. Staudt and C. Janiak, Chem. Commun., 2012, 48, 2140 RSC.
  162. S. H. Huo and X. P. Yan, Analyst, 2012, 137, 3445 RSC.
  163. O. V. Zalomaeva, A. M. Chibiryaev, K. A. Kovalenko, O. A. Kholdeeva, B. S. Balzhinimaev and V. P. Fedin, J. Catal., 2013, 298, 179 CrossRef CAS PubMed.
  164. J. Kim, S. N. Kim, H. G. Jang, G. Seo and W. S. Ahn, Appl. Catal., A, 2013, 453, 175 CrossRef CAS PubMed.
  165. J. Kim, S. Bhattacharjee, K. E. Jeong, S. Y. Jeong and W. S. Ahn, Chem. Commun., 2009, 3904 RSC.
  166. N. V. Maksimchuk, K. A. Kovalenko, V. P. Fedin and O. A. Kholdeeva, Chem. Commun., 2012, 48, 6812 RSC.
  167. I. Y. Skobelev, A. B. Sorokin, K. A. Kovalenko, V. P. Fedin and O. A. Kholdeeva, J. Catal., 2013, 298, 61 CrossRef CAS PubMed.
  168. Y. K. Hwang, D. Y. Hong, J. S. Chang, H. Seo, M. Yoon, J. Kim, S. H. Jhung, C. Serre and G. Férey, Appl. Catal., A, 2009, 358, 249 CrossRef CAS PubMed.
  169. J. Juan-Alcañiz, J. Ferrando-Soria, I. Luz, P. Serra-Cresp, E. Skupien, V. P. Santos, E. Pardo, F. X. Llabrés i Xamena, F. Kapteijn and J. Gascon, J. Catal., 2013, 307, 295 CrossRef PubMed.
  170. F. G. Cirujano, A. Leyva-Pérez, A. Corma and F. X. Llabrés i Xamena, ChemCatChem, 2013, 5, 538 CrossRef CAS.
  171. P. Horcajada, C. Serre, M. Vallet-Regi, M. Sebban, F. Taulelle and G. Férey, Angew. Chem., Int. Ed., 2006, 45, 5974 CrossRef CAS PubMed.
  172. T. Čendak, E. Žunkovič, T. U. Godec, M. Mazaj, N. Z. Logar and G. Mali, J. Phys. Chem. C, 2014, 118, 6140 Search PubMed.
  173. Z. Y. Gu, Y. J. Chen, J. Q. Jiang and X. P. Yan, Chem. Commun., 2011, 47, 4787 RSC.
  174. C. X. Yang, Y. J. Chen, H. F. Wang and X. P. Yan, Chem.–Eur. J., 2011, 17, 11734 CrossRef CAS PubMed.
  175. C. X. Yang and X. P. Yan, Anal. Chem., 2011, 83, 7144 CrossRef CAS PubMed.
  176. Z. Y. Gu and X. P. Yan, Angew. Chem., Int. Ed., 2010, 49, 1477 CrossRef CAS PubMed.
  177. Y. Y. Fu, C. X. Yang and X. P. Yan, Langmuir, 2012, 28, 6794 CrossRef CAS PubMed.
  178. H. Y. Huang, C. L. Lin, C. Y. Wu, Y. J. Cheng and C. H. Lin, Anal. Chim. Acta, 2013, 779, 96 CrossRef CAS PubMed.
  179. Y. Hu, C. Song, J. Liao, Z. Huang and G. Li, J. Chromatogr. A, 2013, 1294, 17 CrossRef CAS PubMed.
  180. Z. Hu, Y. Chen and J. Jiang, Langmuir, 2013, 29, 1650 CrossRef CAS PubMed.
  181. D. N. Dybtsev, V. G. Ponomareva, S. B. Aliev, A. P. Chupakhin, M. R. Gallyamov, N. K. Moroz, B. A. Kolesov, K. A. Kovalenko, E. S. Shutova and V. P. Fedin, ACS Appl. Mater. Interfaces, 2014, 6, 5161 CAS.
  182. Y. Hu, J. Liao, D. Wang and G. Li, Anal. Chem., 2014, 86, 3955 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.