0% found this document useful (0 votes)
10 views11 pages

As Featured in

Uploaded by

mariaclaranakao
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views11 pages

As Featured in

Uploaded by

mariaclaranakao
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 11

Showcasing research from an international collaboration As featured in:

platform: TU-NIMS Joint Research Center at Tianjin University,


China, established by Prof. Jinhua Ye in 2011.

Title: Metal–organic frameworks for photocatalysis

Photocatalysis is an environmental-friendly technology to convert


solar energy into chemical energy in the potential applications
of photodegradation of target pollutants, photocatalytic water
splitting, and CO2 photoreduction into solar fuels. Recently,
Metal–organic frameworks (MOFs) have emerged as novel
photocatalysts due to their catalytically active metal centers/
functional organic linkers, easily tunable photophysical/chemical
properties, in addition to their large surface area/well-ordered
porous structure. This article summarizes various strategies
carried out over MOFs via either modification of the organic See Hua Xu, Jinhua Ye et al.,
linker/metal clusters or incorporation with metal/complex Phys. Chem. Chem. Phys.,
catalysts to enhance the light absorption, charge separation, 2016, 18, 7563.
reactant adsorption/activation of MOF-based photocatalysis
towards the superior photocatalytic performance.

www.rsc.org/pccp
Registered charity number: 207890
PCCP
View Article Online
PERSPECTIVE View Journal | View Issue

Metal–organic frameworks for photocatalysis


Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

Cite this: Phys. Chem. Chem. Phys.,


Ying Li,†ab Hua Xu,†*ab Shuxin Ouyangab and Jinhua Ye*abc
2016, 18, 7563
Photocatalysis is a promising technology to convert solar energy into chemical energy. Recently, metal–organic
frameworks (MOFs) have emerged as novel photocatalysts owing to their inherent structural characteristics
of a large surface area and a well-ordered porous structure. Most importantly, via modulation of the organic
Received 1st October 2015, linker/metal clusters or incorporation with metal/complex catalysts, not only the reactant adsorption and
Accepted 27th October 2015 light absorption but also the charge separation and reactant activation will be largely promoted, leading to
DOI: 10.1039/c5cp05885f superior photocatalytic performance. In this article, we will first introduce the photophysical/chemical
properties of MOFs; then various strategies of modification of MOFs towards better photocatalytic activity
www.rsc.org/pccp will be presented; finally, we will address the challenge and further perspective in MOF-based photocatalysis.

1. Introduction are a class of porous coordination polymers, which have spurred


applications in the fields of gas storage, separation, catalysis,
Photocatalysis is an environmentally-friendly technology to and drug delivery.31–35 Via readily tuning either the constituent
convert solar energy into chemical energy, such as photocatalytic metal clusters or the bridging organic linkers, not only the light
splitting of water into H2,1–4 photodegradation of organic mole- absorption but also the surface functionality of MOFs can be
cules,5–8 and CO2 photoreduction into solar fuels.9–12 Ever since optimized, thus modulation of MOFs to achieve superior visible-
the discovery of water photolysis on a TiO2 electrode reported by light-driven photocatalysis is of utmost importance.
Fujishima and Honda,13 large amounts of effects are focused on In this article, we will first introduce the instructive opinions of
the TiO2 based photocatalysis.14–16 However, the large bandgap the electrical conductivity of MOFs in addition to the proposed
(3.2 eV) of TiO2 largely limits its photocatalytic activity in the UV charge transfer mechanism upon light excitation.36 Then,
light region, which only accounts for 5% of the solar spectrum, several representative MOF photocatalysts (such as MOF-5,37–40
leading to the low photocatalytic efficiency.17 During the past UiO-66(Zr),41,42 MIL-125(Ti),43–46 and MIL-101(Fe)47–50) will be
decade, much effort has been devoted to modifying TiO2 via summarized; a series of strategies, including the modification of
anion/cation doping18–20 or combining with other metal/semi- organic linkers/metal clusters for more visible light absorption
conductors for better light utilization.21–23 Meanwhile, other novel and reactant adsorption48,51 as well as the optimization of supported
visible-light-driven semiconductor photocatalysts such as Ag3PO424,25 metal particles/complex catalysts for better charge separation
and C3N426,27were also developed in the past few years. and reactant activation,52–59 is also demonstrated towards higher
Recently, due to its large surface area/well-ordered porous photocatalytic activity. Finally, we will address the challenge and
structure and tunable organic linkers/metal clusters, which further perspective in MOF-based photocatalysis.
endow promising photophysical/chemical properties, metal–
organic frameworks (MOFs) have received much attention in
the field of photocatalysis.28–30 Particularly, MOF materials 2. MOF based photocatalysis
consisted of metal clusters interconnected with organic linkers,
2.1 History of MOFs as photocatalysts
Inspired from metal carboxylate cluster chemistry, in which the
a
TU-NIMS Joint Research Center, Tianjin Key Laboratory of Composite and
addition of an organic dicarboxylate linker can result in a
Functional Materials, Key Lab of Advanced Ceramics and Machining Technology
of Ministry of Education, School of Materials Science and Engineering, Tianjin
supertetrahedron cluster when capped with monocarboyxlates, it
University, 92 Weijin Road, Nankai District, Tianjin 300072, P. R. China. is excepted that the rigid and divergent character of the attached
E-mail: [email protected], [email protected] organic linker could allow the articulation of clusters into a 3D
b
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), framework, leading to a higher apparent surface area and pore
Tianjin 300072, P. R. China
c
volume.37 In 1999, Yaghi and his co-workers first synthesized the
Environmental Remediation Materials Unit, International Center for Materials
Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS),
materials MOF-5 in a formula of Zn4O(BDC)3(DMF)8(C6H5Cl) via
1-1 Namiki, Tsukuba 305-0044, Japan zinc nitrate and H2BDC (1,4-benzenedicarboxylate) as a precursor.37
† These authors contributed equally to this work. Then, in 2007, Alvaro et al. studied the photocatalytic properties of

This journal is © the Owner Societies 2016 Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 | 7563
View Article Online

Perspective PCCP

MOF-5 in the degradation of phenol in aqueous solution.38 when the temperature is decreased to 153 K, indicating that the
However, Hausdorf and his co-workers found that the structural charge transport in MOFs is affected by the thermal variation.68–70
transformation of zinc carboxylate-based MOFs (such as MOF-5) In contrast to the classical crystalline semiconductor with a
will occur when they are exposed to environments with variable band model, the ‘‘jump’’ of charges through a phonon-assisted
water concentration, in spite of the fact that such a structure can quantum tunneling mechanism occurred in the MOFs with a
be reverted by a thermal treatment.39 localized model.71 This hopping model describes the conduc-
Later, to achieve more stable MOFs as photocatalysts, Cavka tivity with a modified Arrhenius law: s = s0 exp[(T0/T)a], where
et al. first presented the 12 coordinated zirconium-based MOF s0 and T0 are related to the electronic density of the states at the
(UiO-66(Zr): [Zr6O4(OH)4(CO2)12]) in 2008;41 when adopted this Fermi level (the localization length and the phonon frequency),
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

UiO-66 material in photocatalytic H2 evolution, UiO-66 was the exponent a is related to the electronic dimensionality,
suspended in water at 100 1C for 4 h without any structural a = 1/(1 + d), d is the dimensionality. On the basis of the above
change under UV light irradiation.42 Meanwhile, considering the low analysis, it is suggested that 1D MOFs will own better electrical
toxicity and redox activity of titanium, Serre and Sanchez fabricated conductivity than that of 2D- and 3D-MOFs.36
the highly photonic sensitive Ti8O8(OH)4(O2C-C6H4-CO2)6 (MIL-
125(Ti)) via using a mixture solvent of dimethylformamide 2.3 Proposed charge transfer mechanism
(DMF) and methanol as well as titanium tetraisopropoxide as a To date, several MOFs (such as MOF-5, UiO-66(Zr), and MIL-
precursor in 2009, and the results revealed that the reduction of 125(Ti)) have been represented as photocatalysts, in which the
center Ti and oxidation of adsorbed alcohol occur simultaneously organic linkers serve as antennas to absorb light and transfer
over MIL-125(Ti) under UV-visible light irradiation, implying this the photoexcited charge carriers to the central metal clusters
MIL-125(Ti) is a potential stable photocatalyst.43 upon light excitation, namely the linker-to-metal-cluster charge-
As visible-light-driven MOF photocatalysts, in 2013, Larurier transfer (LCCT) mechanism. To confirm this LCCT mechanism,
et al. first reported the iron(III)-based MOFs can photodegrade Li et al. first carried out a photochromic test over NH2-MIL-
Rhodamine 6G in aqueous solution under visible light irradiation,47 125(Ti), in which the color of NH2-MIL-125(Ti) suspension
in which the Fe–O cluster itself in the iron(III)-based MOFs (including turned from original bright yellow to green in the presence of
MIL-101(Fe) and MIL-88(Fe)) can act as a semiconductor to absorb N2, and then the green color changed gradually back to the
visible light and then induce the whole photocatalytic reaction, and original bright yellow when CO2 or O2 was introduced into the
the organic linkers facilitate the charge separation. Now, due to their reaction system (Fig. 1a); the accompanied electron spin resonance
superior topology framework and well-ordered porous structure in (ESR) result revealed that Ti3+ existed in the green-colored but not
addition to the amenable organic linker/metal clusters, MOFs based in the yellow-colored suspension, indicating that the existence of
photocatalysis is emerging as an interesting topic. Ti3+–Ti4+ intervalence charge transfer occurred in the titanium
oxo-clusters. This result is in well-accordance with the LCCT
2.2 Electrical conductivity of MOFs mechanism, as shown in Fig. 1b, in the CO2 photoreduction, the
For inorganic semiconductor photocatalysts, once light is excited, organic ligands ATA as antenna first harvest visible light and be
the negative electrons (e) will move to the conduction band (CB) excited, then transfer electrons to the Ti oxo clusters to reduce
and a positive hole (h+) will be left on the valance band (VB), which Ti(IV) into Ti(III), finally Ti(III) reduces CO2 into formate.44
reacts with the adsorbed reactant separately.60,61 The organic semi- Besides the LCCT mechanism in which the organic linkers
conductor undergoes a similar reaction process but the energy level absorb light then activate the metal clusters, the Fe–O clusters
is described as the highest occupied molecular orbital (HOMO) and in the Fe-containing MOFs (e.g. MIL-101(Fe), MIL-53(Fe), and
the lowest unoccupied molecular orbital (LUMO).62–64 Because of its MIL-88B(Fe)) can be directly excited and induce the electron
optical transition and the visual photocatalytic performance, MOFs transfer from O2 to Fe3+ to form Fe2+ to take part in the
have long been considered as semiconductors.38,65,66 photocatalytic reaction;47–49 Laurier and his coworkers reported
Generally, the electrical properties of a substance can be that the Fe–O clusters in a small size in MIL-100(Fe) and MIL-
distinguished via measuring the electrical conductivity at room 88B(Fe) can be photoexcited to generate charge carriers to
temperature: metal (101–105 S cm1), semiconductor (1010– photodegrade Rhodamine 6G;47 in contrast to the inactive
101 S cm1), and insulator (below ca. 1010 S cm1).36 Kobaya- Fe2O3, the 3D framework of MOFs built with the organic linker
shi et al. reported that Cu[Ni(pdt)2] owns an optical bandgap of can effectively suppress the fast charge recombination.72 More-
B2 eV and a small electrical conductivity of 1  108 S cm1; via over, Wang et al. proposed a dual excitation pathway over NH2-MIL-
further partial oxidation with a stream of I2 vapor at 50 1C, the 101(Fe): (I) excitation of an NH2 functionality followed by an electron
electrical conductivity was dramatically increased by 104-fold transfer to the Fe center (LCCT) and (II) direct excitation of Fe–O
with a value of 1  104 S cm1 after low concentration of iodine clusters.48 In comparison to MIL-101(Fe), the HCOO evolution is
doping, indicating that Cu[Ni(pdt)2] is a p-type semiconductor.67 about 3.0-fold higher for NH2-MIL-101(Fe) (178 mmol in 8 h) under
It is noteworthy that the reliable way to study the electrical visible light irradiation (l 4 420 nm). Additionally, when a 590 nm
behavior of MOFs is to monitor the thermal variation of electrical cutoff filter is used, no HCOO is produced over MIL-101(Fe)
conductivity inside the Ohmic regime.36 Gascon et al. reported that (Fig. 2a) but 5.6 mmol of HCOO is detected over NH2-MIL-101(Fe)
MIL-125(Ti) owns a charge mobility of B105 cm2 V1 s1 upon (Fig. 2b); indicating that besides the direct light excitation of Fe–O
340 nm light illumination; this conductance decreased dramatically clusters similar to that of MIL-101(Fe), the NH2 functionality can also

7564 | Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 This journal is © the Owner Societies 2016
View Article Online

PCCP Perspective
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

Fig. 1 (a) Photos and corresponding ESR spectra of NH2-MIL-125(Ti) under various conditions. (b) Proposed LCCT mechanism for CO2 photoreduction
over NH2-MIL-125(Ti) under visible light irradiation (l 4 420 nm). Reproduced with permission from ref. 44 (Copyright 2012 Wiley-VCH Verlag Gmbh).

absorb light and transfer electron from the organic linker to the composed of Bi3+ and H2mna (2-mercaptonicotinic acid).73
Fe–O cluster via the LCCT mechanism to generate Fe2+ taking part in A strong transient photocurrent response is observed for Bi-mna
the CO2 photoreduction (Fig. 2c).48 under visible light irradiation, indicating that the charge carriers
Recently, a ligand-to-ligand charge transfer (LLCT) process can be photogenerated upon light excitation and then efficiently
is proposed over the bismuth-based MOFs (denoted as Bi-mna) separated over Bi-mna (Fig. 3a). Because of the characteristics of

Fig. 2 (a and b) Dependence of the HCOO evolution on the wavelength of incident light over MIL-101(Fe) and NH2-MIL-101(Fe), respectively. ((I) 440–
455 nm; (II) 490–515 nm; (III) 537–566 nm and (IV) above 590 nm). (c) Proposed dual excitation pathways over NH2-MIL-101(Fe). Reproduced with
permission from ref. 48 (Copyright 2014 American Chemical Society).

This journal is © the Owner Societies 2016 Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 | 7565
View Article Online

Perspective PCCP
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

Fig. 3 (a) Calculated electron localization function plots for Bi-mna. (b) Fukui function for Bi-mna. (The isosurface value is 0.0015 e Å3. Bi: purple; C:
brown; O: red; N: gray; H: pink.) (c) Calculated total density of states and partial density of states of Bi-mna. (d) Transient photocurrent response of Bi-mna
measured with bias values of 0 V (black line) and 0.2 V (gray line). Reproduced with permission from ref. 73 (Copyright 2015 American Chemical Society).

Bi3+ (d10s2),74,75 the theoretical calculations predicted that the linkers of MOFs, the light absorption can be effectively enlarged.
Bi–S, Bi–N and Bi–O bonding in Bi-mna display dominantly Gomes et al. found that by substitution of BDC in UiO-66(Zr)
covalent character (Fig. 3b), and the electrons are highly localized with 2-aminoterephthalic acid (ATA), the amino group led to a
around the organic linker but no distribution in the central Bi sites bathochromic shift in the optical absorption without changing
(Fig. 3c); additionally, the density of states (DOS) plots showed that the photochemistry of NH2-UiO-66(Zr).42 When testing the
the upper valence band is mainly composed of S and the lower photocatalytic properties in H2 evolution using methanol as a
conduction band predominantly consisted of O and N for Bi-mna sacrificial reagent, the activity of NH2-UiO-66(Zr) is 1.17 times
(Fig. 3d). On the basis of the above-mentioned calculation results, it higher than that of UiO-66(Zr) under the irradiation of a 200 W
is proposed that the photoexcited electrons from S atoms of mna xenon-doped mercury lamp, and the apparent quantum yield is
firstly transfer through the central Bi then move to the nonadjacent 3.5% under the monochromatic light irradiation (370 nm) for
pyridine ring, and finally take part in the photocatalytic reaction.73 NH2-UiO-66(Zr).75
After that, Li and his coworkers synthesized the NH2-
functioanlized MIL-125(Ti) (Ti8O8(OH)4(bdc-NH2)6) for CO2 photo-
3. Modification of MOFs for superior reduction in acetonitrile with triethanolamine as a sacrificial agent.
photocatalytic performance In contrast to the parent MIL-125 cannot generate any formate,
the amount of formate reached 8.14 mmol in 10 h can be
Due to their inherent structural characteristics (large surface detected over NH2-MIL-125(Ti) under visible light irradiation.44
area and porous structure) and the tunable organic linkers/metal Besides CO2 photoreduction, Higashimura and Matsuoka also
clusters, MOFs exhibit lots of advantages in comparison with evaluated the H2 evolution efficiency and photoelectrochemical
semiconductors in photocatalysis. However, the photocatalytic properties of NH2-MIL-125(Ti) with the absorption band
efficiency of MOFs still cannot satisfy the practical needs; to extended to B500 nm (Fig. 4a). After visible light irradiation,
further promote the photocatalytic performance, a series of Ti3+ centers in a distorted rhombic oxygen ligand field (gx =
strategies have been adopted via either modulation of their 1.980, gy = 1.953, and gz = 1.889) can be detected over NH2-MIL-
organic linkers/metal clusters or incorporation with metal/ 125(Ti) but not over MIL-125(Ti),76 indicating that the photo-
complex catalysts for enlarged visible light absorption, better excited electrons from an amino functionalized linker could
charge separation, and efficient reactant activation. reduce center Ti(IV) into Ti(III). Furthermore, both the H2
evolution rate and the measured photocurrent are strongly
3.1 Modification of organic linkers dependent on the wavelength in a manner which correlates
As a photocatalyst, the first step is light excitation to generate with the light absorption intensities in the visible light region
charge carriers, thus broadened light absorption will be beneficial up to 500 nm (Fig. 4c and d).45 These findings strongly support
for the photocatalytic reaction. By adjusting the functional organic that the amino functionalization can absorb the incident

7566 | Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 This journal is © the Owner Societies 2016
View Article Online

PCCP Perspective
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

Fig. 4 (a) UV-visible absorption spectra of MIL-125(Ti) and NH2-MIL-125(Ti); inset shows the UV-visible absorption spectra of their related organic
linkers: H2BDC and H2BDC-NH2. (b) ESR spectra of NH2-MIL-125(Ti) observed at 77 K in 0.01 M TEOA solution before and after visible light irradiation
(l 4 420 nm). (c) Dependence of H2 evolution rate of Pt/NH2-MIL-125(Ti) and (d) the photocurrent of NH2-MIL-125(Ti) on the wavelength of incident
light and the UV-visible absorption spectra of NH2-MIL-125(Ti). ((I) 380–420 nm; (II) 420–450 nm; (III) 450–480 nm; (IV) 480–500 nm; (V) 500–550 nm
and (VI) above 550 nm). Reproduced with permission from ref. 45 (Copyright 2012 American Chemical Society).

visible light and then its excited state transfers electrons to the the CO2 photoreduction test revealed that a highest formate
conduction band of metal clusters via the LCCT mechanism. evolution rate (59 mmol in 8 h) is detected over MIL-101(Fe).48
Furthermore, considering the more negative redox potential Since the coordination unsaturated Fe metal sites, existed over
of ZrIV/ZrIII in contrast to TiIV/TiIII, Li et al. also adopted NH2- MIL-101(Fe) but not either on MIL-53(Fe) nor MIL-88B(Fe), could
UiO-66(Zr) in the CO2 photoreduction, the experimental results result in direct CO2 adsorption,82–84 thus MIL-101(Fe) exhibits
showed that NH2-UiO-66(Zr) can produce 13.2 mmol formate higher CO2 adsorption capability of 26.4 g cm3 than that of
under similar reaction conditions, much higher than that of both MIL-53(Fe) (13.5 g cm3) and MIL-88B(Fe) (10.4 g cm3),
NH2-MIL-125(Ti).77 Besides Ti- and Zr-based MOFs, amino leading to the more efficient CO2 photoreduction occurring on
functionalization can also enhance the photocatalytic perfor- the active Fe sites. The better activity of MIL-53(Fe) than MIL-
mance for Fe-based MOFs.48,49 Recently, our group reported that 88B(Fe) might attributed to its 1D framework structure which
amine-functionalized MIL-88B(Fe) exhibited superior photo- owns a better electro conductivity and better CO2 adsorption.48
catalytic activity for Cr(VI) reduction under visible light irradiation, Another theoretical study based on the density functional
in comparison to either the above-mentioned MOFs (NH2-UiO-66(Zr) theory (DFT) carried out by Yang and his coworkers suggested
and NH2-MIL-125(Ti)) or the traditional semiconductors (such as that the band gap engineering of MOF-5 via substitution of
N-doped TiO2, g-C3N4, and commercial CdS). Further contrast Zn4O with X4Y (X = Zn, Cd, Be, Mg, Ca, Sr, Ba; Y = O, S, Se, Te) in
experiments revealed that both the Cr(VI) reduction efficiency the node can tune the band gaps ranging from 1.7 to 3.6 eV,
and the photocurrent density would be promoted with higher which is ascribed to the electronic states of Y atoms in the X4Y
amine group incorporation for MIL-88B(Fe) because of enhanced nodes and C atoms in the organic BDC linkers. From O - S -
light absorption.49 Se - Te, the band gap is decreasing and the absorption edge is
red-shifted.51 Although the above findings are only predicted by
3.2 Modulation of the structure of metal clusters the theory calculation, it gives important hint for exploring
Since the Fe–O clusters itself can be directly photoexcited to drive the band engineered MOF for more visible-light absorption.
visible-light-response photocatalytic reaction, much attention has
been paid to the structure of Fe clusters in the Fe-based MOFs.47,48 3.3 Metals supported on MOFs
In which, MIL-101(Fe) owns a rigid zeotype crystal structure with the For traditional semiconductors, because of the Schottky junction
quasi-spherical cages of two modes (2.9 and 3.4 nm);78 MIL-53(Fe) formed between the metal and the semiconductor, metal particles
consists of chains of –OH corner sharing FeO6 octahedra and a BDC are generally adopted as cocatalysts to drive photoexcited electrons
linker to form 1D pores;79,80 for MIL-88B(Fe), the oxo-centered Fe3O to transfer from the semiconductor and accumulate on metal
trimers interconnect with the BDC linker to build the 3D porous particles for a better charge separation.85 Interestingly, the contribu-
network.81 Wang et al. systematically investigated the structure– tion of various metals in the photocatalytic reaction might be quite
activity relationship of the above-mentioned three Fe-MOFs, and different when loading metal on MOFs.52,53,86

This journal is © the Owner Societies 2016 Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 | 7567
View Article Online

Perspective PCCP

For instance, Li et al. found that the loading of Pt or Au on CQO bond energy of 750 kJ mol1.87–91 Huang and his coworkers
NH2-MIL-125(Ti) did not obviously affect the light absorption introduced the Cu2+ in a porphyrin based Al-MOF and found that
ability (Fig. 5a); however, in comparison to the pure NH2-MIL- the metallization of Cu in Al-MOF (denoted as SCu) can largely
125(Ti) which only produces formate, since both Au and Pt can enhance the methanol evolution rate, about 7 times higher than
trap electrons from Ti–O cluster and provide redox reaction the Al-MOF (denoted as SP). The CO2 adsorption–desorption iso-
sites for H2 evolution, H2 is additionally detected for both therms (Fig. 6a) showed that the amount of CO2 adsorption
Au/NH2-MIL-125(Ti) (40.2 mmol in 8 h) and Pt/NH2-MIL- capacity over SCu (277.4 mg g1 measured at 1 atm, 298 K) is much
125(Ti) (235 mmol in 8 h) (Fig. 5b).53 Based on the LCCT higher than that of SP (153.1 mg g1). Most importantly, in contrast
mechanism, the electrons coming from the excited organic to the fresh SCu, SCu under a CO2 atmosphere exhibits three new
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

linker would become less at the Ti-O clusters to generate Ti3+ peaks which are ascribed to the asymmetric stretching vibrations
for the formate production; experimentally, the formate evolution nas(OCO) of the ‘‘end-on’’ (1780 cm1) and ‘‘C-coordination’’
is 16% decreased over Au/NH2-MIL-125(Ti) but 21% higher over (1656 cm1) coordination states as well as the symmetric stretching
Pt/NH2-MIL-125(Ti) in contrast to that of pure NH2-MIL-125(Ti) vibration of CO2 for both of the above-mentioned two coordination
(Fig. 5c). The ESR revealed that Ti3+ (g = 1.988) is detected only modes (1225 cm1) over SCu (Fig. 6b); whereas no changes could
over Pt/NH2-MIL-125(Ti) but not over either Au/NH2-MIL-125(Ti) be observed in the FT-IR spectra for SP and SP under a CO2
or pure NH2-MIL-125(Ti) (Fig. 5d–f); additionally, the DFT calcula- atmosphere. Those above-mentioned results confirm that CO2
tions confirmed that the migration of H from Pt onto the bridging could be chemically adsorbed on the Cu metallization Al-MOF,
O linked with Ti atoms is much easier than that of Au–O–Ti, and thus the linear CO2 molecules would bend to lower the reaction
such hydrogenation will weaken the O–Ti bond then directly result barrier, leading to a higher CO2 photoreduction efficiency.54
in Ti3+. Therefore, the higher formate evolution rate is ascribed to
the extra Ti3+ caused by the hydrogen spillover for Pt/NH2-MIL- 3.4 MOF-based composite materials
125(Ti).53 Furthermore, Shen et al. also reported that the Pd loaded To date, even light absorption can be enlarged via optimizing
NH2-UiO-66(Zr) could exhibit excellent photocatalytic performance either the organic linker or metal clusters, however, for MOFs
for the simultaneous photodegradation of organic pollutants as photocatalysts, the low efficiency in exciton generation and
(methyl orange (MO) and methylene blue (MB)) and photo- charge separation largely limited its practical application.
reduction of Cr(VI) by individually consuming photogenerated Therefore, another effective solution for an improved photo-
holes and electrons, due to the enlarged light absorption and catalytic efficiency is to combine the gas adsorption in MOFs
better charge separation.52 with exciton generation by other promising catalysts.92–96
On the other hand, for a better CO2 adsorption and activation Lin et al. incorporated various catalytically competent Ir, Re,
in CO2 photoreduction, copper (Cu) is adopted over a porphyrin and Ru complexes with dicarboxylic acid functionalities
based Al-MOF to activate the highly stable CO2 molecule with a (including H2[Cp*IrIII(dcppy)Cl] (H2L1), H2[ReI(CO)3(dcbpy)Cl]

Fig. 5 (a) UV-visible absorption spectra of NH2-MIL-125(Ti), Au/NH2-MIL-125(Ti), and Pt/NH2-MIL-125(Ti). (b) H2 evolution and (c) HCOO evolution
over the prepared samples under visible light irradiation (l 4 420 nm). (d–f) The ESR spectra of NH2-MIL-125(Ti), Au/NH2-MIL-125(Ti), and Pt/NH2-MIL-
125(Ti) measured at various atmospheres in the dark or under visible light irradiation. Reproduced with permission from ref. 53 (Copyright 2014 Wiley-
VCH Verlag Gmbh).

7568 | Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 This journal is © the Owner Societies 2016
View Article Online

PCCP Perspective
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

Fig. 6 (a) CO2 adsorption and desorption isotherms of porphyrin linked MOFs without (SP) or with Cu incorporated (SCu). (b) In situ FT-IR spectra of SP
(black) and SCu (red) measured in a CO2 atmosphere. Reproduced with permission from ref. 54 (Copyright 2013 American Chemical Society).

(H2L4), and H2[RuII(bpy)2(dcbpy)]Cl2) (H2L6) with highly stable photosensitizer [Ru(bpy)3]Cl26H2O and Co-ZIF-9. Via optimiza-
porous-structured UiO-67(Zr) to develop the highly efficient tion of the concentration (1 mg) (for more active redox reaction)
heterogeneous catalysts for solar energy utilization.55 In contrast to and thermal treatment conditions (110 1C for 4 h) (to dredge
the inactivity of the parent UiO-67(Zr), L1-UiO-67(Zr) can effectively the blocking micropores for better CO2 adsorption) of Co-ZIF-9,
oxidize water (cerium ammonium nitrate as an oxidant) with the an apparent quantum yield of 1.48% at monochromatic irra-
turnover frequencies of up to 4.8 h1, a superior activity with a total diation of l = 420 nm can be measured (Fig. 7a). This study
turnover number of 10.9 was achieved in the CO2 photoreduction strongly supports that imidazolium motifs containing MOFs
using trimethylamine as a sacrificial agent over L4-UiO-67(Zr), and play important roles on the CO2 capture and activation.56
L6-UiO-67(Zr) can serve as an efficient photocatalyst for various Meanwhile, a C3N4/Co-ZIF-9 composite photocatalyst was
photocatalytic organic transformations (such as aza-Henry reaction, also developed for CO2 photoreduction, because of better CO2
aerobic amine coupling, and aerobic oxidation of thioanisole).55 adsorption in the framework of Co-ZIF-9 and the promoted
Besides the porous structure which is good for gas capture, when charge separation from C3N4 to Co-ZIF-9, 20.8 mmol of CO and
combined with MOFs with specific functionalized organic linkers, 3.3 mmol of H2 can be detected in 2 h for C3N4/Co-ZIF-9 under
not only the charge separation (from other catalysts to MOFs) but visible light irradiation (l 4 420 nm). As shown in Fig. 7b, the
also the reactant activation (e.g. CO2 activation) will be promoted. It wavelength dependence of the CO evolution demonstrated that
is well known that the CO2 activation process is the rate-limiting step the trend of photocatalytic activity matched well with the light
for CO2 photoreduction;97–99 considering the high CO2 adsorptive absorption characteristics of C3N4 instead of Co-ZIF-9, confirm-
capability and the strong stabilizing effect of CO2 owned by ing that the reaction was indeed induced by the light excitation
the organic imidazolium motifs,100,101 Wang et al. adopted the of C3N4 and the function of Co-ZIF-9 as a cocatalyst.57 Besides
Cobalt-containing benzimidazole MOF Co-ZIF-9 as a cocatalyst to Co-ZIF-9, our group also incorporated the C3N4 nanosheets with
enhance the CO2 photoreduction activity of [Ru(bpy)3]Cl26H2O the zirconium based MOFs UiO-66(Zr) and found that the CO
(bpy = 2 0 2-bipyridine).56 [Ru(bpy)3]Cl26H2O/Co-ZIF-9 compo- evolution yield for C3N4 nanosheets/UiO-66(Zr) (59.4 mmol g1) is
site material exhibits much higher CO (1.4 mmol min1) and 3 times higher than that of bulk C3N4/UiO-66(Zr) (17.1 mmol g1)
H2 (1.0 mmol min1) evolution rates than that of pure in 6 h, ascribing to the more efficient electron transfer across the
[Ru(bpy)3]Cl26H2O (CO: 0.04 mmol min1; H2: 0.06 mmol min1), interface between C3N4 nanosheets and UiO-66(Zr).102
which is ascribed to the better CO2 activation ability over Furthermore, the porous structure and the large surface area
Co-ZIF-9 and the enhanced charge separation between the of MOFs can also facilitate the dispersion of semiconductor

Fig. 7 (a) Effect of the concentration of Co-ZIF-9 in the complexed [Ru(bpy)3]Cl26H2O/Co-ZIF-9 material on the evolution of CO and H2 from CO2
photoreduction. (b) Dependence of the evolution of CO and H2 on the wavelength of incident light over g-C3N4/Co-ZIF-9 composite material.
Reproduced with permission from ref. 56 (Copyright 2014 Wiley-VCH Verlag Gmbh) and reproduced with permission from ref. 57 (Copyright 2014 The
Royal Society of Chemistry).

This journal is © the Owner Societies 2016 Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 | 7569
View Article Online

Perspective PCCP
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

Fig. 8 (a) TEM and (b) HRTEM images of CdS/MIL-101(Cr). (c) H2 evolution over MIL-101(Cr) embedded with different concentrations of CdS nanoparticles
under visible light irradiation. (d) TEM images in low and high magnification (inset) of Co3.9/MIL-101(Cr). (e) EDX elemental mapping results of Co in Co3.9/
MIL-101(Cr). (f) Photocatalytic O2 evolution over various samples under visible light irradiation (l 4 420 nm). Reproduced with permission from ref. 58
(Copyright 2013 The Royal Society of Chemistry) and reproduced with permission from ref. 59 (Copyright 2015 The Royal Society of Chemistry).

nanoparticles, leading to more reaction active centers.103–106 However, the stability of MOFs during the photocatalytic
Regarding this point, He et al. first embedded the CdS nano- reaction is an important issue needed to be further considered.
particles on the framework of MIL-101 (Cr) (Fig. 8a and b) and For instance, the incidental pyrolyzation will destroy the frame-
the H2 evolution test in lactic acid aqueous solution (l 4 work of MOFs under high-intensity light irradiation because of
420 nm) showed that the CdS/MIL-101(Cr)(10) composite can the thermal effect of photons. Generally, the thermal stability of
evolve 75.5 mmol gCdS1 h1 of H2, about 2 times higher than MOFs is determined by the coordination number and local
that of the pure CdS (33.8 mmol gCdS1 h1) (Fig. 8c).58 Then, a coordination environment,107 thus searching MOFs with opti-
similar strategy was adopted to immobilize Co3O4 nano- mized coordination between the metal cluster and the organic
particles (2–3 nm) over the MIL-101(Cr) substrate with good linker for high thermal stability is of great importance. On the
dispersion and narrow size distribution (Fig. 8d and e). Due to other hand, controlling the pyrolysis of MOFs under certain
the promoted charge transfer from Co3O4 to MIL-101(Cr), a conditions to expand their micropores into mesopores or
high turnover frequency of 0.012 s1 per cobalt atom and promote their electrical conductivity while maintaining the
oxygen yield of 88% could be obtained in the [Ru(bpy)3]2+– open diffusion channels and ensuring the monodispersion of
[Na2S2O8] system at pH 9 for the Co3O4/MIL-101(Cr) composite metal centers, is believed to open up a new way to achieve
with a Co content of 3.9%, about 9 times higher activity in higher photocatalytic efficiency.108,109
contrast to the bare Co3O4 (Fig. 8f).59 Another drawback of MOFs is their poor electronic conduc-
tivity, which hampers the charge transfer from the organic
linker to the metal cluster. Previous studies suggested that
doping can largely boost the electrical conductivity of MOFs
4. Conclusions and further perspectives based on the polarons/bipolarons and soliton mechanism110–112
and 1D/2D MOFs exhibited higher electrical conductivity in
In this article, we have briefly introduced the electrical con- contrast to 3D frameworks.36 Therefore, it is expected that
ductivity of MOFs, the charge transfer mechanism over the further electronic band modulation and framework topology
light-excited MOFs, and various strategies to enhance the control of MOFs will bring us more promising photocatalytic
photocatalytic efficiency of MOFs. Because of their unprece- properties in the solar-to-energy conversion.
dented porosity and large surface area, MOFs can not only be
beneficial for the gas adsorption but also act as good substrate Acknowledgements
to confine the size of the loaded catalyst in several nanometers.
Most importantly, since both the metal nodes and organic This work was supported by 973 Program (No. 2014CB239301),
linkers can be easily tuned for specified functionality, thus Natural Science Foundation of Tianjin (15JCQNJC03500,
exhibiting promising photocatalytic performance as compared 15JCYBJC17400), and National Natural Science Foundation of
to the semiconductor, especially in CO2 photoreduction. China (51502198, 21573158, 51502200).

7570 | Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 This journal is © the Owner Societies 2016
View Article Online

PCCP Perspective

References 33 G. Férey, C. Serre, T. Devic, G. Maurin, H. Jobic,


P. L. Llewellyn, G. De Weireld, A. Vimont, M. Daturi and
1 Z. Zou, J. Ye, K. Sayama and H. Arakawa, Nature, 2001, 414, J.-S. Chang, Chem. Soc. Rev., 2011, 40, 550–562.
625–627. 34 M. G. Goesten, F. Kapteijn and J. Gascon, CrystEngComm,
2 J. H. Park, S. Kim and A. J. Bard, Nano Lett., 2006, 6, 24–28. 2013, 15, 9249.
3 A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253–278. 35 G. Sneddon, A. Greenaway and H. H. Yiu, Adv. Energy
4 S. Ouyang, H. Tong, N. Umezawa, J. Cao, P. Li, Y. Bi, Mater., 2014, 4, 1301873.
Y. Zhang and J. Ye, J. Am. Chem. Soc., 2012, 134, 1974–1977. 36 G. Givaja, P. Amo-Ochoa, C. J. Gomez-Garcia and
5 S. Yan, Z. Li and Z. Zou, Langmuir, 2009, 25, 10397–10401. F. Zamora, Chem. Soc. Rev., 2012, 41, 115–147.
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

6 C. Chen, W. Ma and J. Zhao, Chem. Soc. Rev., 2010, 39, 37 H. Li, M. Eddaoudi, M. O’Keeffe and O. M. Yaghi, Nature,
4206–4219. 1999, 402, 276–279.
7 S. Ouyang and J. Ye, J. Am. Chem. Soc., 2011, 133, 7757–7763. 38 M. Alvaro, E. Carbonell, B. Ferrer, F. X. Llabrés i Xamena
8 H. Chen, C. E. Nanayakkara and V. H. Grassian, Chem. and H. Garcia, Chem. – Eur. J., 2007, 13, 5106–5112.
Rev., 2012, 112, 5919–5948. 39 S. Hausdorf, J. Wagler, R. Moßig and F. O. R. L. Mertens,
9 C. Wang, R. L. Thompson, J. Baltrus and C. Matranga, J. Phys. Chem. A, 2008, 112, 7567–7576.
J. Phys. Chem. Lett., 2009, 1, 48–53. 40 T. Tachikawa, J. R. Choi, M. Fujitsuka and T. Majima,
10 T. W. Woolerton, S. Sheard, E. Reisner, E. Pierce, S. W. Ragsdale J. Phys. Chem. C, 2008, 112, 14090–14101.
and F. A. Armstrong, J. Am. Chem. Soc., 2010, 132, 2132–2133. 41 J. H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou,
11 W. Tu, Y. Zhou and Z. Zou, Adv. Mater., 2014, 26, 4607–4626. C. Lamberti, S. Bordiga and K. P. Lillerud, J. Am. Chem.
12 P. Li, Y. Zhou, Z. Zhao, Q. Xu, X. Wang, M. Xiao and Z. Zou, Soc., 2008, 130, 13850–13851.
J. Am. Chem. Soc., 2015, 137, 9547–9550. 42 C. Gomes Silva, I. Luz, F. X. Llabrés i Xamena, A. Corma
13 A. Fujishima and K. Honda, Nature, 1972, 238, 37–38. and H. Garcı́a, Chem. – Eur. J., 2010, 16, 11133–11138.
14 X. Chen and A. Selloni, Chem. Rev., 2014, 114, 9281–9282. 43 M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin,
15 G. Liu, H. G. Yang, J. Pan, Y. Q. Yang, G. Q. Lu and H.-M. C. Sanchez and G. Ferey, J. Am. Chem. Soc., 2009, 131,
Cheng, Chem. Rev., 2014, 114, 9559–9612. 10857–10859.
16 M. Kapilashrami, Y. Zhang, Y.-S. Liu, A. Hagfeldt and 44 Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li,
J. Guo, Chem. Rev., 2014, 114, 9662–9707. Angew. Chem., Int. Ed., 2012, 124, 3420–3423.
17 H. Xu, S. Ouyang, L. Liu, P. Reunchan, N. Umezawa and 45 Y. Horiuchi, T. Toyao, M. Saito, K. Mochizuki, M. Iwata,
J. Ye, J. Mater. Chem. A, 2014, 2, 12642. H. Higashimura, M. Anpo and M. Matsuoka, J. Phys. Chem.
18 R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, C, 2012, 116, 20848–20853.
Science, 2001, 293, 269–271. 46 D. Sun, L. Ye and Z. Li, Appl. Catal., B, 2015, 164, 428–432.
19 X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959. 47 K. G. Laurier, F. Vermoortele, R. Ameloot, D. E. De Vos,
20 X. Chen, L. Liu, Y. Y. Peter and S. S. Mao, Science, 2011, J. Hofkens and M. B. Roeffaers, J. Am. Chem. Soc., 2013,
331, 746–750. 135, 14488–14491.
21 S. Linic, P. Christopher and D. B. Ingram, Nat. Mater., 48 D. Wang, R. Huang, W. Liu, D. Sun and Z. Li, ACS Catal.,
2011, 10, 911–921. 2014, 4, 4254–4260.
22 A. V. Akimov, A. J. Neukirch and O. V. Prezhdo, Chem. Rev., 49 L. Shi, T. Wang, H. Zhang, K. Chang, X. Meng, H. Liu and
2013, 113, 4496–4565. J. Ye, Adv. Sci., 2015, 2, 1500006.
23 J. Y. Park, L. R. Baker and G. A. Somorjai, Chem. Rev., 2015, 50 R. Liang, F. Jing, L. Shen, N. Qin and L. Wu, J. Hazard.
115, 2781–2817. Mater., 2015, 287, 364–372.
24 Z. Yi, J. Ye, N. Kikugawa, T. Kako, S. Ouyang, H. Stuart- 51 L.-M. Yang, G.-Y. Fang, J. Ma, E. Ganz and S. S. Han, Cryst.
Williams, H. Yang, J. Cao, W. Luo and Z. Li, Nat. Mater., Growth Des., 2014, 14, 2532–2541.
2010, 9, 559–564. 52 L. Shen, W. Wu, R. Liang, R. Lin and L. Wu, Nanoscale,
25 Y. Bi, S. Ouyang, N. Umezawa, J. Cao and J. Ye, J. Am. Chem. 2013, 5, 9374–9382.
Soc., 2011, 133, 6490–6492. 53 D. Sun, W. Liu, Y. Fu, Z. Fang, F. Sun, X. Fu, Y. Zhang and
26 X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Z. Li, Chemistry, 2014, 20, 4780–4788.
Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 54 Y. Liu, Y. Yang, Q. Sun, Z. Wang, B. Huang, Y. Dai, X. Qin and
76–80. X. Zhang, ACS Appl. Mater. Interfaces, 2013, 5, 7654–7658.
27 Y. Wang, X. Wang and M. Antonietti, Angew. Chem., Int. 55 C. Wang, Z. Xie, K. E. deKrafft and W. Lin, J. Am. Chem.
Ed., 2012, 51, 68–89. Soc., 2011, 133, 13445–13454.
28 J.-L. Wang, C. Wang and W. Lin, ACS Catal., 2012, 2, 56 S. Wang, W. Yao, J. Lin, Z. Ding and X. Wang, Angew.
2630–2640. Chem., Int. Ed., 2014, 53, 1034–1038.
29 T. Zhang and W. Lin, Chem. Soc. Rev., 2014, 43, 5982–5993. 57 S. Wang, J. Lin and X. Wang, Phys. Chem. Chem. Phys.,
30 S. Wang and X. Wang, Small, 2015, 11, 3097–3112. 2014, 16, 14656–14660.
31 G. Férey, Chem. Soc. Rev., 2008, 37, 191–214. 58 J. He, Z. Yan, J. Wang, J. Xie, L. Jiang, Y. Shi, F. Yuan, F. Yu
32 G. Férey and C. Serre, Chem. Soc. Rev., 2009, 38, 1380–1399. and Y. Sun, Chem. Commun., 2013, 49, 6761–6763.

This journal is © the Owner Societies 2016 Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 | 7571
View Article Online

Perspective PCCP

59 J. Han, D. Wang, Y. Du, S. Xi, J. Hong, S. Yin, Z. Chen, 85 J. Ran, J. Zhang, J. Yu, M. Jaroniec and S. Z. Qiao, Chem.
T. Zhou and R. Xu, J. Mater. Chem. A, 2015, 3, 20607–20613. Soc. Rev., 2014, 43, 7787–7812.
60 X. Chen, S. Shen, L. Guo and S. S. Mao, Chem. Rev., 2010, 86 R. Liang, S. Luo, F. Jing, L. Shen, N. Qin and L. Wu, Appl.
110, 6503–6570. Catal., B, 2015, 176, 240–248.
61 N. Serpone and A. Emeline, J. Phys. Chem. Lett., 2012, 3, 87 K. Xie, N. Umezawa, N. Zhang, P. Reunchan, Y. Zhang and
673–677. J. Ye, Energy Environ. Sci., 2011, 4, 4211–4219.
62 W. Pasveer, J. Cottaar, C. Tanase, R. Coehoorn, P. Bobbert, 88 Y. Wang, A. Lafosse and K. Jacobi, J. Phys. Chem. B, 2002,
P. Blom, D. De Leeuw and M. Michels, Phys. Rev. Lett., 106, 5476–5482.
2005, 94, 206601. 89 R. Caballol, E. Sanchez Marcos and J. C. Barthelat, J. Phys.
Published on 28 October 2015. Downloaded by Universidade de Brasilia on 4/26/2024 2:29:19 AM.

63 V. Coropceanu, J. Cornil, D. A. da Silva Filho, Y. Olivier, Chem., 1987, 91, 1328–1333.


R. Silbey and J.-L. Brédas, Chem. Rev., 2007, 107, 926–952. 90 S. Sakaki, K. Kitaura and K. Morokuma, Inorg. Chem., 1982,
64 L. Kaake, P. F. Barbara and X.-Y. Zhu, J. Phys. Chem. Lett., 21, 760–765.
2010, 1, 628–635. 91 A. Fateeva, P. A. Chater, C. P. Ireland, A. A. Tahir, Y. Z. Khimyak,
65 C. G. Silva, A. Corma and H. Garcı́a, J. Mater. Chem., 2010, P. V. Wiper, J. R. Darwent and M. J. Rosseinsky, Angew. Chem.,
20, 3141–3156. Int. Ed., 2012, 124, 7558–7562.
66 C. H. Hendon, D. Tiana and A. Walsh, Phys. Chem. Chem. 92 J.-R. Li, R. J. Kuppler and H.-C. Zhou, Chem. Soc. Rev., 2009,
Phys., 2012, 14, 13120–13132. 38, 1477–1504.
67 Y. Kobayashi, B. Jacobs, M. D. Allendorf and J. R. Long, 93 R. Li, J. Hu, M. Deng, H. Wang, X. Wang, Y. Hu, H. L. Jiang,
Chem. Mater., 2010, 22, 4120–4122. J. Jiang, Q. Zhang and Y. Xie, Adv. Mater., 2014, 26, 4783–4788.
68 R. A. Marcus, J. Chem. Phys., 1956, 24, 966–978. 94 F. Ke, L. Wang and J. Zhu, Nano Res., 2014, 1–13.
69 M. A. Nasalevich, M. G. Goesten, T. J. Savenije, F. Kapteijn 95 M. Nasalevich, R. Becker, E. Ramos-Fernandez, S. Castellanos,
and J. Gascon, Chem. Commun., 2013, 49, 10575–10577. S. L. Veber, M. V. Fedin, F. Kapteijn, J. Reek, J. I. van der Vlugt
70 T. J. Savenije, A. J. Ferguson, N. Kopidakis and G. Rumbles, and J. Gascon, Energy Environ. Sci., 2015, 8, 364–375.
J. Phys. Chem. C, 2013, 117, 24085–24103. 96 R. Wang, L. Gu, J. Zhou, X. Liu, F. Teng, C. Li, Y. Shen and
71 C. Bellitto, G. Dessy and V. Fares, Inorg. Chem., 1985, 24, Y. Yuan, Adv. Mater. Interfaces, 2015, 2, 1500037.
2815–2820. 97 H. Schwarz and R. Dodson, J. Phys. Chem., 1989, 93, 409–414.
72 C. Aprile, A. Corma and H. Garcia, Phys. Chem. Chem. Phys., 98 B. A. Rosen, A. Salehi-Khojin, M. R. Thorson, W. Zhu, D. T.
2008, 10, 769–783. Whipple, P. J. Kenis and R. I. Masel, Science, 2011, 334, 643–644.
73 G. Wang, Q. Sun, Y. Liu, B. Huang, Y. Dai, X. Zhang and 99 B. A. Rosen, J. L. Haan, P. Mukherjee, B. r. Braunschweig,
X. Qin, Chemistry, 2015, 21, 2364–2367. W. Zhu, A. Salehi-Khojin, D. D. Dlott and R. I. Masel,
74 F. Réal, V. Vallet, J.-P. Flament and J. Schamps, J. Chem. J. Phys. Chem. C, 2012, 116, 15307–15312.
Phys., 2006, 125, 174709. 100 E. D. Bates, R. D. Mayton, I. Ntai and J. H. Davis, J. Am.
75 P. Boutinaud and E. Cavalli, Chem. Phys. Lett., 2011, 503, Chem. Soc., 2002, 124, 926–927.
239–243. 101 C. Wang, H. Luo, D. E. Jiang, H. Li and S. Dai, Angew.
76 M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin, C. Sanchez Chem., Int. Ed., 2010, 122, 6114–6117.
and G. Férey, J. Am. Chem. Soc., 2009, 131, 10857–10859. 102 L. Shi, T. Wang, H. Zhang, K. Chang and J. Ye, Adv. Funct.
77 D. Sun, Y. Fu, W. Liu, L. Ye, D. Wang, L. Yang, X. Fu and Mater., 2015, 25, 5360–5367.
Z. Li, Chemistry, 2013, 19, 14279–14285. 103 D. Farrusseng, S. Aguado and C. Pinel, Angew. Chem., Int.
78 S. Bauer, C. Serre, T. Devic, P. Horcajada, J. Marrot, Ed., 2009, 48, 7502–7513.
G. Férey and N. Stock, Inorg. Chem., 2008, 47, 7568–7576. 104 N. V. Maksimchuk, K. A. Kovalenko, V. P. Fedin and
79 P. Llewellyn, P. Horcajada, G. Maurin, T. Devic, O. A. Kholdeeva, Adv. Synth. Catal., 2010, 352, 2943–2948.
N. Rosenbach, S. Bourrelly, C. Serre, D. Vincent, S. Loera- 105 M. Ranocchiari and J. A. van Bokhoven, Phys. Chem. Chem.
Serna and Y. Filinchuk, J. Am. Chem. Soc., 2009, 131, Phys., 2011, 13, 6388–6396.
13002–13008. 106 J. M. Roberts, B. M. Fini, A. A. Sarjeant, O. K. Farha, J. T. Hupp
80 E. Haque, N. A. Khan, J. H. Park and S. H. Jhung, Chem. – and K. A. Scheidt, J. Am. Chem. Soc., 2012, 134, 3334–3337.
Eur. J., 2010, 16, 1046–1052. 107 B. Mu and K. S. Walton, J. Phys. Chem. C, 2011, 115,
81 A. Dhakshinamoorthy, M. Alvaro, H. Chevreau, 22748–22754.
P. Horcajada, T. Devic, C. Serre and H. Garcia, Catal. Sci. 108 Y. Han, P. Qi, S. Li, X. Feng, J. Zhou, H. Li, S. Su, X. Li and
Technol., 2012, 2, 324–330. B. Wang, Chem. Commun., 2014, 50, 8057–8060.
82 P. Chowdhury, C. Bikkina and S. Gumma, J. Phys. Chem. C, 109 U. Pal, R. Mondal, I. Mondal, S. Bala and A. Goswami,
2009, 113, 6616–6621. J. Mater. Chem. A, 2015, 3, 20288–20296.
83 Y. K. Hwang, D. Y. Hong, J. S. Chang, S. H. Jhung, Y. K. Seo, 110 J. L. Bredas and G. B. Street, Acc. Chem. Res., 1985, 18, 309–315.
J. Kim, A. Vimont, M. Daturi, C. Serre and G. Férey, Angew. 111 M. Nechtschein, F. Devreux, F. Genoud, E. Vieil, J. Pernaut
Chem., Int. Ed., 2008, 47, 4144–4148. and E. Genies, Synth. Met., 1986, 15, 59–78.
84 P. Valvekens, F. Vermoortele and D. De Vos, Catal. Sci. 112 D. Y. Lee, D. V. Shinde, S. J. Yoon, K. N. Cho, W. Lee, N. K.
Technol., 2013, 3, 1435–1445. Shrestha and S.-H. Han, J. Phys. Chem. C, 2013, 118, 16328–16334.

7572 | Phys. Chem. Chem. Phys., 2016, 18, 7563--7572 This journal is © the Owner Societies 2016

You might also like