Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

NJC

View Article Online


PERSPECTIVE View Journal | View Issue

The chemistry of titanium-based


metal–organic frameworks
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Cite this: New J. Chem., 2017,


41, 14030
Ha L. Nguyen

Due to the many potential applications of metal–organic frameworks (MOFs), research focusing on the
design, syntheses, and surface modifications of the new generation of crystalline-based MOF materials
has recently received much attention. Among these materials, MOFs based on Ti–oxo clusters are
potential candidates for practical applications such as photocatalysis, CO2 reduction, and pollutant
degradation because of their high optical responsitivities and photoredox properties. The capabilities of
Ti-MOFs in a wide range of applications are inspiring their use as a replacement for TiO2, which has
recently been classified as a potential carcinogen. However, the number of tianium–organic framework
Received 23rd August 2017, structures is very limited compared to the many structures of other MOFs. In order to emphasize
Accepted 4th October 2017 the importance of discovering new Ti-MOF materials and their related features, this work aims to
DOI: 10.1039/c7nj03153j describe the chemistry of titanium-based MOFs, including their synthetic methods, crystal structures,
and topological analysis, as well as further key novel applications, which are desirable for extension to
rsc.li/njc many industrial fields.

Introduction due to the high reactivity of the metal(IV) ions and carboxylate
linkers, which leads to poorly crystalline or amorphous powders.11
Metal–organic frameworks (MOFs) are constructed by the TiO2 is known as a commercial reagent that is useful as a
combination of rigid covalent bonds of organic linkers and white pigment and in other industrial fields, such as paints,
inorganic metal ion or clusters acting as nodes, designated as plastics, paper, foods, cosmetics, and medicines, because of its
secondary building units (SBUs).1 The organic linkers are also UV blocking features. However, very recently, the European
termed as SBUs from a geometrical standpoint and are used for Chemicals Agency (ECHA) has proposed the classification of
topological analysis. Recently, almost 70 000 structures of P25-TiO2 as a potential carcinogen.12 Titanium–oxo cluster-
MOFs have been reported and studied, showing the tremendous based MOFs are promising materials that may replace TiO2 in
development of these crystalline porous materials.2 Considered many industrial applications due to their extreme stability
as a new class of microporous material possessing promising against harsh working conditions of high temperature and
properties of high porosity, well-defined crystallinity, increased humidity and/or chemical environments.13,14 In this contribution,
number of active sites, catalytic activity, etc., MOFs have been we seek to review the chemistry of MOFs based on titanium–oxo
studied to address many issues, such as environmental clusters; their synthetic methods, crystal structures, topological
problems,3–5 sustainable energy,6 and catalytic transformation.7,8 networks and potential applications will be described and
However, the low stability of MOFs under working conditions discussed in detail.
of high temperature and humidity as well as in fluid solvents
reduces the practical applications of this material. In order to
overcome this challenge, MOFs based on high valence metals Synthetic methods, crystal structures,
and carboxylate linking units have been the subject of increasing and topological analysis
studies. Zr(IV)- and Ti(IV)-based MOFs are candidates that not only
contain strong metal–carboxylate bonds and rigid frameworks, The syntheses of MOFs were carefully presented in early
but also demonstrate good photocatalytic properties.9–11 It should published literature reports,15 in which the formation of MOFs
be noted that the number of these materials is still limited was described through solvothermal, hydrothermal, electro-
compared to MOFs based on di- and tri-valence metal clusters chemical, mechanochemical, and microwave-assisted synthesis
methods. In some cases, slow evaporation was used to obtain
Center for Innovative Materials and Architectures (INOMAR), Vietnam National
large crystals that were utilized for structural elucidation by
University-Ho Chi Minh (VNU-HCM), Ho Chi Minh City 721337, Vietnam. single crystal X-ray diffraction analysis.16,17 In the process of
E-mail: nlha@inomar.edu.vn, hanguyen@manar.edu.vn Ti-MOFs formulation, the strong interaction between highly

14030 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC

reactive titanium sources and organic carboxylate linkers leads X-ray diffraction (PXRD) analysis. Indeed, the ab initio method
to difficult association and dissociation of Ti–carboxylate was applied to elucidate the structure via direct methods using
bonds, which results in the production of poorly crystalline EXPO software. The refinement was performed through Rietveld
products such as sols, gels or amorphous powders.11 Hence, refinement executed by Fullprof. The 3-D structure of MIL-125
highly crystalline Ti-MOFs can only be achieved by choosing revealed that MIL-125 crystallized in a body-centered tetragonal
suitable synthetic conditions, including the organic solvent, lattice with the I4/mmm space group (No. 139). The refined
ratio of the reacted solution mixture, pH environment, and lattice parameters were convergent with a = b = 18.6453(10) Å
temperature. Very recently, Serre and co-workers demonstrated and c = 18.1444(10) Å. Eight (8) oxo atoms were divided by two
the Pourbaix diagrams of titanium calculated for [Ti4+] = parallel planes, in which each plane contained four (4) oxo
103 mol L1 at 25 1C using Hydra and Medusa software.18 atoms. Ti8(OH)4 was located in the middle and parallel with two
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

This calculation can guide the synthesis of Ti-MOFs based on planes holding oxo atoms. The 8-member ring Ti8O8(OH)4(CO2)12
consideration of the pH factor. was built by the connection of 4 carboxylates on the top,
It should be noted that materials containing Ti-based clusters, 4 carboxylates on the bottom and 4 carboxylates in the middle
Ti(BuOH) and MIL-91, produced by reaction of 1,4-butanediol plane, which divided the cluster into two parts (Fig. 2). BDC
(Bu(OH)2) and N,N 0 -piperazinebismethylenephosphonate, respec- moieties acted as 2-connected (2-c) to link 12-connected (12-c)
tively, with a titanium source, were reported as early as 2006. distorted cuboctahedra clusters to each other, resulting in a 3-D
Although Ti(BuOH) dissolves in most organic solvents, which is view of a crystal structure which contains triangular faces and
unusual in MOFs chemistry, its crystal structure reveals the first belongs to the fcu topology. It is noteworthy that the fcu network
3-dimensional (3-D) view of an extended titanium coordination of MIL-125, which is identical to the UiO-66 system,24 was not
complex.19 In the case of MIL-91, this material is the first titanium mentioned in the published literature. The transitivity, defined
diphosphonate with well-defined porosity, showing an internal by [p q r s] (where p is the number of vertices, q is the number of
surface area of 550 m2 g1 based on the Langmuir method.20 In edges, r is the number of faces, and s is the type of tile),25 and the
this context, we focus the discussion on the details of the natural tiling for MIL-125 were found to be [1 1 1 2] and 2[34] +
structural properties and applications of Ti-MOFs based on [38], respectively (Table 1). The linking of these triangular faces
organic carboxylate linker building units. leads to the creation of two types of pores in MIL-125. The

Direct synthesis
The first Ti-MOF constructed from a 1,4-benzenedicarboxylate
(BDC) linker and a Ti–oxo cluster, termed MIL-125, was reported
by Serre and Sanchez and co-workers in 2009.21 MIL-125 was
solvothermally synthesized using a solvent mixture of N,N-
dimethylformamide (DMF)/methanol that dissolved the starting
reagents, BDC and titanium(IV) isopropoxide; the mixture was
heated to 140 1C in a solvothermal oven to yield a crystalline
powder. Undoubtedly, MIL-125 is an exemplary material, possessing
the first octameric Ti–oxo cluster and ditopic linker; it inspired other
Fig. 2 Crystal structure of MIL-125 showing a 3-D view with two kinds of
studies, including novel Ti-MOF synthesis efforts, structural
pores, octahedral and tetrahedral, which are represented by yellow and
modifications, and promising applications (Fig. 1).22,23 orange balls, respectively. 12-c of the 8-member ring of Ti8O8(OH)4(CO2)12
The crystalline powder of MIL-125 was obtained by solvothermal is also displayed on the right side. Color code: O, red; C, black; Ti, blue.
synthesis, and its crystal structure was determined by powder H atoms are omitted for clarity.

Fig. 1 Stages of Ti-based MOFs. Details of the synthetic methods and the observed titanium–oxo clusters are provided along with the materials that
have been synthesized and reported.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14031
View Article Online

NJC Perspective

Table 1 Summary of the topological features of Ti-MOFs

Material Net Node Transitivitya Natural tiling


b
MIL-91 sql 4-c [1 1 1 1] [44]
Ti-BuOHb bct 10-c [1 2 2 1] [34.42]
MIL-125b fcu 12-c [1 1 1 2] 2[34] + [38]
ZTOF-1b bcu 6-c [1 1 1 1] [44]
MIL-125-NH2b fcu 12-c [1 1 1 2] 2[34] + [38]
NTU-9b hcb 3-c [1 1 1 1] [63]
COK-69b acs 3-c [1 1 2 2] [43] + [4362]
Ti-CAT-5b srs 3-c [1 1 1 1] [103]
MIL-167b srsb 3-c [1 1 1 1] [103]
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

MOF-901b hxl 6-c [1 1 1 1] [36]


MOF-902b hxl 6-c [1 1 1 1] [36]
Cd–Ti-MOF-1c,d ctmb 4,6-c [2 3 5 3] [462] + [6282]
Cd–Ti-Silan-1c rtl 3,6-c [2 2 3 2] [462] + [6282]
ZTOF-2c she 6,8-c [2 1 2 2] 3[4482] + [41286]
PCN-22d,e kimb 4,4,7-c [5(10)96] 2[63] + 2[6282] + [4284] + [64122] + [4464122]
MIL-101(Ti) f mtn-e 6,6,6,6-c [4 7 9 5] 17[34] + 2[320512] + [32851264]
a
Transitivity is defined by [p q r s] (where p is the number of vertices, q is the number of edges, r is the number of faces, and s is the type of tile).
b
The structural topology is uninodal. c The structural topology is binodal. d These MOFs were found to have new topologies which were not
named and/or acknowledged by the authors in their publications. e The structural topology is trinodal. f The structural topology is tetranodal.

internal diameters of the octahedral and tetrahedral cages were


found to be ca. 12.55 and 6.13 Å, respectively (Fig. 2).22
Based on an organic linker containing carboxylate and hydroxyl
bifunctionality, 2,5-dihydroxyterephthalic acid (H4DOBDC), which
is the same link used to synthesize the MOF-74 series, Zhang and
co-workers reported the structure of NTU-9 in 2014.26 NTU-9 was
solvothermally produced by the reaction of H4DOBDC and
titanium isopropoxide [Ti(i-OPr)4] in acetic acid. The acidic
medium created by acetic acid prevented seed crystals from
forming too quickly, which increased the possibility of yielding
larger crystals. NTU-9 was obtained as large red crystals after 5 d
of reaction in a Teflon-lined stainless-steel autoclave at 120 1C.
The structure of NTU-9 is a 2-D layer that contains a single
pure Ti atom, in which the connection of 3 carboxylates and
3 hydoxyl groups yields a single octahedral Ti cluster. Each
single cluster is bound by 3 DOBDC linking units, and the
combination of these 3-c points creates an hcb network con-
taining one type of hexagonal face with the signature of [63]
tiling (Table 1). In NTU-9, the layer structure is stacked together
along the c axis, constructing an 11  11 Å 1-D channel. The
distance between the first layer and second layer, calculated
from the crystal structure, is around 6 Å (Fig. 3).
Fig. 3 Single Ti atom cluster formation (a) that is the fundamental unit
In order to reduce the high reactivity and hydrolysis of titanium
used to build the NTU-9 structure (b). The layer distance is also presented
isopropoxide, a Ti-based complex precursor was considered for use on the right side of (a). Color code: O, red; C, black; Ti, blue; second layer,
as a reagent in the synthesis of MOFs. A hydrolytically organo- orange. H atoms are omitted for clarity. Yellow balls are placed in the
metallic Ti precursor, dicyclopentadienyl titanium(IV) dichloride cages to show the pores.
([Cp2TiIVCl2]), was explored to react with a flexible linker,
trans-1,4-cyclohexanedicarboxylic acid (H2CDC), to generate
a breathing Ti-MOF, termed COK-69.27 The microcrystalline material represents the first flexible Ti-based MOF, in which the
powder of COK-69 was solvothermally produced by the reaction linker trans-1,4-cyclohexanedicarboxylate (CDC) can transform its
of [Cp2TiIVCl2] precursor and H2CDC in DMF solvent in the conformation to adopt either the equatorial e,e 0 or the axial a,a 0
presence of acetic acid at 110 1C for 48 h. position. In the as-synthesized sample, COK-69 contains DMF
COK-69 represents a triangular geometrical cluster similar guest molecules inside the pores; the ratio of the a,a 0 /e,e 0 con-
to the exemplary Fe–oxo clusters found in MIL-100, PCN-332, formation was found to be 1/1. Moreover, 100% a,a conformation
and PCN-333.28 In particular, trinuclear Ti(IV)3(mO)O2(CO2)6 is of the CDC linking units, proven by 13C NMR spectroscopy, was
an unprecedented Ti(IV)-based cluster in Ti-MOF chemistry. In observed after the activation step to evacuate the trapped solvents
addition, the crystal structure of COK-69 clearly reveals that this (DMF). The open-pore size was calculated to be 4.9 Å, which is

14032 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Fig. 5 Structural conformation of the triangular Ti(III) cluster containing


Ti–OEt and Ti–solvent linkages (a). A Ti–solvent bond was evacuated after
the activation step conducted under vacuum at 150 1C (b). The Ti(IV)
superoxide (c) and Ti(IV) peroxide (d) formed by the reaction of O2 and
the unsaturated Ti(III) cluster. Color code: O, red; C, black; Ti, blue. Modified
Fig. 4 Crystal structure and topological deconstruction of COK-69. with permission from ref. 29. Copyright 2016 American Chemical Society.
(a) The triangular Ti(IV)3(mO)O2(CO2)6 and CDC linking units are simplified
as 6- and 2-c, respectively. This combination results in the 3-D COK-69
structure (b) corresponding to the acs topology. Color code: O, red; C,
black; Ti, blue polyhedral. H atoms are omitted for clarity. Yellow balls are complexes.31 This observation was proven by O2 adsorption
placed in the cages to show the pores. measurements, in which at 25 1C, MIL-101(Ti) exhibited very
steep isotherms at low pressures and reached a capacity of
0.85 mmol g1 (2.6 wt%) at 0.9 mbar. This phenomenon is due
larger than the closed-pore size of 2.7 Å. The underlying net of to a chemisorption process inside the structure of MIL-101(Ti).
COK-69 was identified after structural simplification, in which The presence of Ti(III) atoms in the structure was proven by
the presence of triangular clusters and ditopic organic linkers electron paramagnetic resonance (EPR) studies, which clearly
were simply reduced to 6-c and 2-c, respectively. The combination showed coincidence between the relevant g values for oxidized
of 2-c and 6-c nodes formed an acs topology in the COK-69 MIL-101(Ti) and literature values.29
structure (Fig. 4), which adopts [1 1 2 2] transitivity and two types Crystalline 3D frameworks of metal-catecholates (M-CAT) are
of tiling abbreviated as [43] + [4362] (Table 1). porous materials constructed from metal ions and catecholate-
At the same time, Long and co-workers reported the first based organic units as a type of linker, similar to MOFs.32 In
Ti-MOF based on a Ti(III)–oxo cluster29 whose structure is 2015, Nguyen et al. determined the synthetic conditions to afford
an analogue of MIL-101.30 The synthesis of MIL-101(Ti) was the first 3-D Me-CAT (Me = V, Ti, or Fe).33 In this study, Ti-CAT-5 was
rigorously maintained under inert conditions using a glovebox constructed from H6THO [THO6 = triphenylene-2,3,6,7,10,11-
and vacuum Schlenk line system. Indeed, the mixture of hexakis(olate)] and a single titanium atom. In particular, Ti-CAT-5
anhydrous DMF/ethanol was transferred to a Schlenk flask that was prepared by heating a solution mixture of DMF/water/
had been charged with TiCl3 salt and 1,4-benzenedicarboxylic methanol, dissolved titanium isopropoxide, H6THO, tetrabutyl-
acid (H2BDC) executed in the glovebox. The reaction solution ammonium nitrate, and amylamine to 180 1C for 48 h.
was then heated at 120 1C under positive N2 pressure with The 3-D crystal structure of Ti-CAT-5 was solved by PXRD
stirring for 18 h to yield MIL-101(Ti) as a dark purple powder. and electron density map (EDM) generation through the
Although it shows clusters with triangular geometry, similar Charge Flipping method.34 In particular, Rietveld refinement
to COK-69, MIL-101(Ti) is the only recent example of an MOF provided the convergence residual factor with Rp = 5.66%, Rwp =
constructed from a Ti(III)–oxo cluster and a tritopic linking 8.05%. Ti-CAT-5 crystallizes in a primitive cubic lattice with a
block (BTC) with high porosity (2970 m2 g1 based on the cell parameter of a = 17.996 Å and the P213 space group, as
BET method).29 The conformation of the Ti(III) cluster in extracted from the indexing results. The SBU of Ti-CAT-5 was
MIL-101(Ti) is presented in Fig. 5, in which all Ti(III) atoms built by the connection of 3 catachol fragments, abbreviated as
are directly capped with solvent molecules (ethanol). After Ti(C2O2)3, which were linked to each other through THO building
solvent exchange and activation under vacuum and at high units. It is noted that the combination of the 3-connected nodes
temperature (150 1C), one guest solvent molecule was evaporated, of the Ti(C2O2)3 cluster and THO in Ti-CAT-5 produces a 2-fold
leaving an unsaturated Ti(III) atom. Due to the strong interaction interpenetrated srs network due to the relatively large pore
of O2 with Ti(III) atom, titanium(IV) superoxo and peroxo species space (Fig. 6). The pore space of this structure, which covered
were generated by reaction of oxygen molecules and titanium(III) two dimethylammonium (DMA) counter ions, was created by

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14033
View Article Online

NJC Perspective
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Fig. 6 Synthetic scheme demonstrating the formation of 3-D Ti-CAT-5. Fig. 7 The 3-c bulding unit of MIL-167 (a) was found to be identical to the
The combination of a single Ti building block and THO unit forms the NTU-9 building block. The linkage of Ti(CO3)3 SBUs via DOBDC building
2-fold interpenetrated framework of Ti-CAT-5. Color code: O, red; C, units forms the 3-D structure of MIL-167, whose topology belongs to the
black; Ti, blue; second framework, grey. The yellow ball indicates the free single frame srs network (b). Color code: O, red; C, black; Ti, blue. H atoms
space in the cage. H atoms and DMA anions are omitted for clarity. are omitted for clarity.

MIL-169, two Ti(IV)(CO3)2 SBUs directly share the m2-O atom,


interlocking two frameworks in which the center-to-center which also acts as the edge that combines two layer chains of
distance between two THO linkers of the framework was found the structure. In contrast, MIL-167 exhibited a 3-D structure in
to be 6.6 Å. The discovery of Ti-CAT-5 contributed greatly to our which the building unit is Ti(CO3)3 (Fig. 7). This SBU is similar to
understanding of the architectures and physical and chemical the SBU of NTU-9. However, the rotated direction of the DOBDC
properties of MOFs as well as the promise of accessing new linking units is different from the structural geometry of NTU-9,
properties in the scope of Ti-MOFs. resulting in a single frame srs topology which contains the
Very recently, Assi et al. presented a study that utilized the transitivity signature of [1 1 1 1] and a natural tiling of [103].
synthetic scheme of reacting 2,5-dihydroxyterephthalic acid
(H4DOBDC) with titanium(IV) precursors to successfully synthe- Heterometallic approach
size three new Ti-based MOFs, namely MIL-167, MIL-168, and Direct synthesis, which involves the reaction of various titanium
MIL-169.35 MIL-167 was solvothermally synthesized by reacting salts and organic linkers, is the typical method used to produce
a solvent mixture of N,N-diethylformamide (DEF)/methanol at Ti-MOFs. However, the high reactivity of the titanium precursor
180 1C for 24 h. An alternative procedure was slightly modified leads to an increase in the reaction speed and a decrease in the
to use glacial acetic acid and diethylamine in DEF solvent. The possibility of obtaining large crystals. In order to form single
temperature was reduced to 150 1C, and the reaction time was crystals of Ti-based MOFs by solvothermal reaction, Cui and
5 d. No explanation was provided for the presence of acetic acid co-workers proposed the bimetalic synthetic strategy, in which a
and diethylamine in the alternative synthetic technique for titanium-based metallosalan organic linking unit was initially
MIL-167. A similar procedure was applied to produce MIL-168, synthesized and used as a starting reagent which was then
which used catechol as the starting modulator. Moreover, MIL-169 combined with biphenyl-4,4 0 -dicarboxylic acid (H2BPDC) and
was obtained at 150 1C in 72 h by the reaction of oxalalatotitanate Cd salt to form a 3-D Cd–Ti–oxo cluster-based MOF. This
precursor, H4DOBDC, and catechol in distilled water. material was the first single crystal Ti-MOF based on a carbox-
In terms of crystallography, MIL-168 and MIL-169 can be ylate linker and chiral salen linking unit (Fig. 8).36
classified as 1-D and 2-D, respectively.35 In particular, MIL-168 Interestingly, the linker H2L first linked to Ti(OBu)4 to
was built from a connection of zig-zag chains containing generate the titanium-based salan linking unit, which is a
Ti(IV)(CO3)2(cat) (cat = catechol) building units. In the case of 6-connected point of extension. The combination of the

14034 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC

the MOF formulation process, such as insoluble metal oxides,


which disrupt crystalline growth in Me(IV)-based MOFs.40,41
The synthetic approach that exploits the presence of
bi-functionalized carboxylate and hydroxyl within the integral
organic building blocks to simultaneously bond to Zn2+– and
Ti4+–oxo clusters was first reported by Chun and co-workers in
2013. Taking advantage of the two metal–oxo clusters, which
are stoichiometrically included in the crystal structures of
ZTOF-1 and ZTOF-2, these frameworks were stabilized and
crystalized in large crystals. In particular, heterometallic MOFs
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

based on 2-hydroxyterephthalic acid (H3OBDC) and Zn6Ti2


clusters were obtained as orange crystals.40 A similar procedure
was used to synthesize ZTOF-2 by a combination of an analogue
linker of H3OBDC, 3-hydroxy-2,7-naphthalenedicarboxylic acid
(H3ONDC), and the bimetallic–oxo cluster Zn6Ti2 in the
presence of 1,4-diazabicyclo-[2.2.2]octane (DABCO), which
acted as the linking pillar of the 3-D framework (Scheme 1).41
Fig. 8 Synthetic scheme for the production of bimetalic Cd–Ti-MOF-1.
Indeed, the SBUs of ZTOF-1 was composed of Zn6Ti2 with
6 Zn atoms linked to each other through m3-oxo linkages.
metallosalan-based building block and hexameric Cd–oxo cluster Moreover, 2 tetrahedral Zn atoms in the center of the Zn6Ti2
via pyridine bonds extended the structure to a 2-D framework. SBU share the edge via O–O bonds from carboxylate, while the
Finally, BPDC acted as a pillar to connect these layers to each octahedral Ti atom is located at the head and terminal of the
other, resulting in a 3-D structure. The hexameric Cd-based Zn6–oxo block (Fig. 10a). Simplification of the heterometallic
building units stacked to create a 0.5 nm  1.5 nm 1-D channel SBUs in ZTOF-1 afforded a uninodal structure constructed by
(Fig. 9). The topological analysis for Cd–Ti-MOF was found to be a the connection of 6-connected nodes through ditopic linking
new net, which had not been published before and does not exist units, resulting in the bcu topology (Fig. 10b). In the case of
in the RCSR or TOPOS databases.37,38 For future reference, we ZTOF-2, the pillar created by DABCO links the Zn3 clusters to
will refer to this net as the ctm topology. The connection of the each other and contributes to the generation of 2 tetrahedral Zn
4,6-c points of extension in ctm leads to a binodal network that building blocks, which contain 3 Zn–O bonds and one Zn–N
possesses 2 different vertices, 3 different edges, 5 different faces, linkage. The last octahedral Zn atom was located in the center
and 3 different tiles. The signature for natural tiling is [462] + [62 of the homometallic SBU. In addition, the heterometallic SBU
82]. The metallosalan-based approach for Ti-MOF single crystal in ZTOF-2 was formulated from 3 tetrahedral Zn atoms located
preparation was successfully applied to synthesize Me–Ti-Silan-1 in the middle of the Zn3Ti2 SBU and 2 octahedral Ti atoms,
(where Me equals Cd or Zn).39 Indeed, the chiral Ti–salan where each single Ti atom acted as a 3-connected node
complex based on carboxylate and a pyridine terminal was (Fig. 10a). In terms of topology, the Zn3Ti2 SBU was simplified
obtained by the designed synthesis. The solvothermal reactions as one node possessing a 6-connected point of extension. This
of Ti-metalloligand and TiL(OBu)2 with CdBr24H2O and point directly linked to the 8-c node of the pinwheel Zn3
Zn(OAc)22H2O afforded Cd–Ti-Silan-1 and Zn–Ti-Silan-1, clusters, generating an seh network (Fig. 10b). The connection
respectively. of Ti atoms in the Zn3Ti2 SBU with the central Zn atoms from
Using heterometallic MOFs that contain metal(II) ion and the pinwheel Zn3 clusters generated a 12-faced polyhedron,
metal(IV) ion within the SBUs can reduce the complications of termed a trapezo-rhombic dodecahedron, covering the pore,
which was found to be ca. 16 Å in internal diameter. It should

Fig. 9 Crystal structure of Cd–Ti-MOF-1 (a) and its topological network


(b), which is unknown in the RCSR and TOPOS databases. Color code: O, Scheme 1 Representations of the syntheses of ZTOF-1 and ZTOF-2,
red; C, black; N, green, Ti, blue; Cd, orange. H atoms are omitted for clarity. which possess heterometalic Zn/Ti–oxo clusters.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14035
View Article Online

NJC Perspective
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Fig. 10 Representative structures of ZTOF-1 and ZTOF-2, which exhibit


heterometalic SBUs, (a) and response topological network (b). Color code:
O, red; C, black; N, green, Ti, blue; Cd, orange. H atoms are omitted for
clarity.

be noted that the surface areas of ZTOF-1 and ZTOF-2 calculated


by the BET method were 1045 and 1878 m2 g1, respectively.
These values are comparable with those of other highly porous
Fig. 11 Crystal strcuture of PCN-22 viewed from the a direction (top) and
Ti-MOF materials, such as MIL-125, MIL-125-NH2, and PCN-22,42 b direction (bottom), comprising a TCPP linking unit and an unprecedented
which will be discussed in the next section. The success of Ti7O6 cluster. Color code: O, red; C, black; Ti, blue. The yellow ball indicates
heterometallic-based MOFs was further investigated by Chun the free space in the cage. H atoms are omitted for clarity.
and co-workers. He combined Co(II)Ti(IV) with terephthalic acid
(BDC), forming CTOF-1, or 2-hydroxyterephthalic acid, producing The topological analysis of PCN-22 was conducted using the
CTOF-2, an isoreticular structure of CTOF-1.43 Both materials TOPOS pro package,38 in which the crystal structure was
were formed from triangular building blocks, Co2TiO(CO2)6 initially simplified to yield a simple net through the edges
SBUs, and linear building units to generate an open acs frame- and nodes. Particularly, the Ti7O6 SBU can be divided into two
work with intriguing dynamic properties. parts that are symmetric to each other by the octahedral Ti
central atom acting as the intersection point of the C2 axis and
Reticular chemistry method Ti6O6 plane. In order to simplify the net, the octahedral Ti
A reversible bond association/dissociation process in Ti-MOF central atom was assigned as the edge linking two Ti3O3 SBUs,
synthesis is extremely difficult due to the strong Ti–O bond, which which were then reduced to two 7-c nodes. The TCPP linking
is a challenge to achieving crystalline Ti-MOFs. The building block units are clearly 4-connected nodes that combine the 7-c nodes
synthesis method uses the metal cluster as a starting reagent to of the Ti–oxo SBUs to generate a binodal 4,4,7-c network,
link to the organic linker. This methodology is also known as showing a new topology (Fig. 12). We will refer to this net as
reticular chemistry,44 meaning that the designed structures were the kim network in future discussions. It is noteworthy that the
precisely predicted by topological principles.38,44 PCN-22 structure can be alternatively simplified by the connection
The linking of terminal ligands to metal–oxo clusters was of 4-c TCPP linking units and 12-c of the Ti7O6 cluster, forming a
subsequently replaced by linking of ditopic organic units to form 4,4,12-c trinodal network with (4-c)2(4-c)(12-c) stoichiometry.
MOF structures. This synthetic method was reported by Guillerm Interestingly, depending on how the augment node is analyzed,
et al., who synthesized UiO-66-type analogues from zirconium 4-c of the point of extension of TCPP could be assigned as the
methacrylate–oxo cluster and the corresponding organic linkers connection of 3 3-c nodes. In this case, the topological network
(trans,trans-muconic acid for Zr-muconate and H2BDC for was found to be a 3,3,12-c net, corresponding to (3-c)8(3-c)4(12-c)
UiO-66).11 The building block approach was also applied to synthe- stoichiometry.
size MIL-125 from Ti8O8(OOCC(CH3)3)16 clusters and H2BDC,22 as
reported by Hendon et al. in 2013. Recently, Yuan et al. reported a In situ cluster generation
similar technique using a hexameric tianium–oxo cluster, Only MIL-125 and PCN-22 were successfully obtained by the
Ti6O6(AB)6(i-OPr)6 (AB = 4-aminobenzoate); this building block building block synthetic approach; finding a new synthetic
reagent was reacted with tetrakis((4-carboxyphenyl)porphyrin) technique to produce Ti-MOFs remains a challenge. Using
(TCPP) to generate a Ti-MOF, PCN-22, based on the porphyrin unit. robust synthetic conditions, which avoid cluster decomposition
Interestingly, after formation of the MOF, the titanium–oxo cluster during MOF growth, in situ cluster generation was described in
was found to be rearranged, yielding an unprecedented heptameric the synthesis of MOF-90145 and MOF-90246 reported by Nguyen
Ti7O6 cluster (Fig. 11).42 et al. In particular, 4-aminobenzoic acid was first reacted with

14036 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Fig. 13 Synthetic strategy combining in situ Ti6O6(AB)6(OMe)6 clusters


and aldehyde fuctionalities (a) to produce MOF-901 and isoreticular MOF-
Fig. 12 Topological deconstruction of PCN-22. (a) Ti7O6 cluster simpli- 902, which are presented as space filling models of their crystal structures
fied by two SBUs of 7-c nodes, linking to 4-c of the TCPP linker to form the (b). Color code: O, red; C, black; N, green; H, pink; Ti, blue; second
unprecedented kim network (b). framework, orange. Capping methoxide moieties have been removed for
clarity. Reprinted with permission from ref. 46. Copyright 2016 American
Chemical Society.

14 and 16 Å, respectively. The 2-D layer hxl topology adopting 36


faces was identified for both MOF-901 and MOF-902.

Post synthesis modification

Scheme 2 Synthetic strategy combining MOF and COF chemistry via


The concept of post synthesis modification (PSM) in MOF
in situ cluster generation and an imine condensation reaction. Color code: materials was first reported by Cohen and Wang,47 who anchored
O, red; C, black, N, green, Ti, blue. an amide functional group into the framework of IR-MOF-3
through a condensation reaction. The PSM technique is useful
for the synthesis of MOFs containing bulky functionalities, which
titanium isopropoxide to generate in situ Ti6O6(AB)6(OMe)6 cannot be obtained through direct synthesis because the geo-
clusters, which were then linked to each other through an metrical hindrance of organic linkers disrupts the MOF structure.
imine condensation reaction by the combination of the amino In addition, metal ion exchange by PSM results in more new
functionality and aldehyde moiety. MOF-902, possessing longer materials that include bimetal–oxo clusters, which cannot be
imine linking units, is an isoreticular structure of MOF-901. synthesized directly due to the highly reactive properties of metal
These are the first two examples exploiting robust synthetic salts (Ti sources or Cr salts), resulting in new attractive properties
conditions to synthesize porous materials based on a combination compared to the host materials.48,49
of MOF and COF chemistry (Scheme 2). The manufacturing of Ti-based MOFs via PSM has been
In particular, MOF-901 is based on the hexameric Ti6O6- demonstrated by many research groups.50–52 Ti atoms were
(AB)6(OMe)6 cluster, which possesses 6 4-aminobenzoate term- exchanged at the positions of host metal atoms in the metal
inal chelates binding to a Ti6O6 triangular prism to form 6-c cluster by immersing the parent MOFs in organic solvent
points of extension. It is clear that the AB chelates are co-planar containing titanium salts such as TiCl3 or TiX4 (X = Br or Cl).
and are then linked to each other through the imine condensation A UiO-66(Zr/Ti)-type structure was reported by Kim et al.,
reaction of the amino terminal and aldehyde functionality to showing successful exchange by an increase of the percentage
yield the 2-D crystal structure. The reaction of AB and 4,4 0 - of Ti atom from 53% to 94% (Fig. 14).53 MOF-5(Zn/Ti) was also
biphenyldicarboxaldehyde (BPDA) linking units formed the synthesized by partial PSM metal exchange of Ti(III) and Zn(II).
imine linker with a distance of 24.10 Å, which is longer than Although the percentage of Ti(III) in MOF-5(Zn/Ti) was only
the 19.97 Å-imine linking blocks in the MOF-901 structure found to be 2%,54 this synthetic scheme provided powerful
(Fig. 13). The staggered frameworks in MOF-901 and MOF-902 evidence that engineering the structural functionality can
generated hexagonal cages, whose pore sizes were found to be afford new materials (Fig. 15).

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14037
View Article Online

NJC Perspective

percentages in PCN-333(Sc/Ti), MIL-100(Sc/Ti), MOF-74(Zn/Ti) and


MOF-74(Mg/Ti) were proven by ICP-MS, which displayed that high
concentrations of Ti replaced Sc, Zn, and Mg atoms, respectively
[88.8%, 48.8%, 94.7%, and 37.9% Ti for PCN-333(Sc/Ti), MIL-
100(Sc/Ti), MOF-74(Zn/Ti) and MOF-74(Mg/Ti), respectively].
Clearly, the success of utilizing the HVMO approach to synthesize
various Ti-based MOFs from MOF platforms not only introduced
Fig. 14 The synthesis procedure of UiO-66(Zn/Ti) obtained by a PSM an effective methodology for the synthesis of Ti-MOFs, but also
method. In this strategy, UiO-66(Zr) was initially synthesized by a sol- opened a pathway to explore the promising features of Ti-based
vothermal method, and the host material was then immersed in a solution
MOF materials, such as photoresponsive properties, which are
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

containing the TiX4 source, resulting in UiO-66(Zn/Ti).


related to photocatalytic applications, and redox activity.

Applications of Ti-MOFs
Photocatalytic applications
Semiconductor photocatalysis provides an environmentally
friendly technology that exploits solar energy to convert carbon
dioxide and water to hydrocarbons and oxygen. The first
photocatalytic study of water splitting was demonstrated by
Fujishima and Honda in 1972, in which titanium dioxide (TiO2)
Fig. 15 The PSM strategy to produce MOF-5(Zn/Me) (where Me = Ti, V, generated hydrogen and oxygen under UV irradiation.56 The
Cr, Mn or Fe). Reprinted with permission from ref. 54. Copyright 2016 discovery of the photocatalytic properties of TiO2 inspired the
American Chemical Society. research field of photocatalysts based on TiO2 and related
materials. However, the large bandgap energy (3.2 eV) of TiO2
not only limits57 its activity in the UV region, which represents
Very recently, Zhou and co-workers published a versatile
only 5% of the solar spectrum, but also makes it unsuitable for
synthetic route for Ti-based MOFs preparation that used the
some photoreactions that require high energy of the catalyst to
high valence metathesis and oxidation (HVMO) approach to
activate the initial reagent. The first study of the photocatalytic
synthesize a series of porous photoactive titanium MOFs, PCN-
properties of a MOFs material demonstrated the degradation
333(Sc)-Ti, MIL-100(Sc)-Ti, MOF-74(Zn)-Ti and MOF-74(Mg)-Ti
of phenol by MOF-558 in aqueous solution. Although the
(Fig. 16).55 Indeed, these materials were initially acquired
structural transformation of MOFs based on Zn–oxo carboxylates
through Sc(III) to Ti(III) or Zn(II)/Mg(II) to Ti(III) metathesis.
was later confirmed by Hausdorf and co-workers,59 the photo-
Subsequently, the metathesized materials underwent an oxidation
activities of MOFs generated more interest in the research fields
step, leading to the oxidization of Ti(III) to Ti(IV). The Ti atomic
of phototransformation and renewable energy.60
MOFs based on Ti–oxo clusters are materials whose redox
and photocatalytic activities may be exploited in future to explore
studies of their photoreactions and photocatalytic applications.
As discussed in the previous section, the exemplary MIL-125
contains 8-member ring Ti–oxo clusters. These clusters are linked
to each other through BDC building units. The photoresponsive
properties of MIL-125 were demonstrated by the light absorption
spectrum, which exhibited broad and intense bands in the visible
light region (400 to 800 nm), as indicated by two maxima located
at 250 and 300 nm.21,22 The band gap energy of MIL-125 was
predicted based on the HSE06-calculated VB and CB energy,
which was found to be 3.60 eV,22,61 which is in good agreement
with the experimental estimation. Interestingly, under UV
irradiation, benzyl alcohol adsorbed inside the MIL-125 frame-
work under inert gas (nitrogen) clearly changed color from
Fig. 16 Synthetic scheme of the stepwise HVMO procedure using the white to purple-gray-blue within a few seconds (Fig. 17).
template MOFs (a) MIL-100(Sc) and PCN-333(Sc) to exchange with Ti(III) via The phenomenon of the fast photochromic effect was
metal metathesis. The oxidation reaction was then followed by metal node
further explained by intervalence electron transfer hopping in
oxidation in air (it should be noted that PCN-333(Sc) and MIL-100(Sc) have
the same topology but different structures). (b) A similar process was
the bivalence-based cluster of MIL-125, which simultaneously
conducted to obtain MOF-74(Zn) and MOF-74(Mg).55 Reprinted with contained Ti(III) and Ti(IV); this was observed and proven by EPR
permission from ref. 55. Copyright 2015 The Royal Society of Chemistry. spectra. The transition valence of MIL-125 was helpful for the

14038 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC

experiments showed that RB and MB were degraded completely


after 80 min and 20 min, respectively.26 In addition, the photo-
catalytic properties and the crystallinity of NTU-9 were retained
after 6 h of experimental testing, as proven by consecutive
recycling results and PXRD analysis, respectively.
Currently, photocatalytic hydrogen production from water
splitting is an important topic that is receiving much consideration
from the chemical research community. Semiconductor-type photo-
catalysts are being developed to conduct this process. However, the
challenge of current photocatalysts is that these materials are only
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

activated under UV irradiation, similar to MIL-125, COK-169 or


MIL-167, which accounts for only 2% to 3% of natural solar
energy.62 In order to overcome these problems, photocatalysts
are being modified to achieve a high optical response that can
Fig. 17 The photochromic effect observed when benzyl alcohol was transform light absorption to the red-shifted region, resulting in
loaded in MIL-125 after UV light irradiation (a). The proposed mechanism low band gap energy materials. In this context, the PSM method
for the reduction of MIL-125 under irradiation (b). Modified with permission was used to prepare materials possessing highly conjugated
from ref. 21. Copyright 2015 American Chemical Society. systems (p–p interaction liking units, using more electron-
donating functional groups in the framework),9,46,63 resulting
in a high photoresponse, which is attributed to ligand-to-metal
photo-oxidation reaction of alcohols to aldehydes under UV charge transfer (LMCT).64 The optical engineering approach for
light excitation. This observation, which demonstrates the reducing the band gap energy of MIL-125 was introduced by
photoredox activity of Ti-based MOFs, additionally suggests Hendon et al. by exploiting the amino functional groups in the
them to be promising candidates as potential photoactive framework.65 The resulting material, MIL-125-NH2, is an iso-
catalysts in practical applications. The alcohol oxidation reaction reticular structure with MIL-125; however, it possesses a lower
of ethanol was similarly conducted using COK-69 under a UV-A band gap energy compared to the host material (2.60 eV for
(l = 315–400 nm) light source, affording a color change from MIL-125-NH2, compared to 3.60 eV for MIL-125). In addition,
white to blue, which indicated the transfer of Ti(IV) to Ti(III).27 the introduction of various functional BDC-X linking units
Although it possesses a triangular Ti(IV)–oxo cluster, COK-69 (X = CH3, Cl, OH) into MIL-125 was reported.
also exhibited an identical bivalence transfer phenomenon to In 2012, Masuoka and co-workers utilized the photocatalytic
MIL-125. activity of MIL-125-NH2 for the photocatalytic hydrogen production
By incorporating the porphyrin-based linking unit into a reaction.66 The LMCT (linker-to-cluster charge transfer (LCCT) was
highly ordered porous material, the resulting MOF, based used in that work) mechanism was illustrated by the results of
on the high efficiency of the porphyrin antenna framework, in situ ESR experiments. After 3 h of visible light irradiation, the
displayed high visible light photo-responsiveness compared to sample was applied to measure ESR at 77 K. It is noted that the
MIL-125 and its derivatives. This desirable PCN-2242 structure characteristic behaviour of paramagnetic Ti3+ atoms in a distorted
results in a low band gap energy of 1.93 eV, which is lower than rhombic oxygen ligand field (gx = 1.980, gy = 1.953, and gz = 1.889)
the values of 3.60 and 2.60 eV for MIL-125 and MIL-125-NH2, was clearly observed. In contrast, no Ti3+ signal was observed on
respectively, and corresponds to a broad range of absorption conducting an identical experiment under the same conditions
(200–640 nm) of visible light. The red-shifted visible light using MIL-125 catalyst. This result confirmed that LMCT in MIL-
absorption of PCN-22 indicates that this material is an active 125-NH2 occurred following the hydrogen production reaction,
photocatalyst that can promote the transformation of photo which transferred an electron from the excited BDC-NH2
reactions. In particular, the oxidation of phenyl alcohol to functionality to the titanium–oxo cluster, resulting in the
benzaldehyde was demonstrated under a PCN-22/TEMPO catalyst presence of Ti3+ (Fig. 18).
system (where TEMPO = 2,2,6,6-tetramethylpiperidinyloxyl) and Taking advantage of the capabilities of a well-defined structure,
various catalysts such as TiO2, TCPP, a mechanical mixture of the surfaces as well as the frameworks of MOFs can be easily
TiO2 and TCPP, and PCN-224 (a Zr-porphyrin-based MOF). The modified through PSM to encapsulate more active molecules.67
experimental results clearly showed that PCN-22 outperformed Recently, using MIL-125-NH2 as a MOF platform to load active
other candidates, as indicated by the observation of a twofold metal complexes for further applications of hydrogen production
higher conversion of benzaldehyde over 2 h of reaction time. led to enhanced photocatalytic properties of the resulting material
Taking advantage of its good photoresponsive properties, compared to the parent MIL-125-NH2.68,69 In particular, MIL-125-
NTU-9 exhibited excellent properties of a p-type semiconductor NH2 exhibited high chemical stability, suitable pore size for large
in promoting the degradation of organic dyes. Indeed, the molecule encapsulation, and fast light-harvesting properties,
photocatalytic activity of NTU-90 was studied for the degradation which is very important due to the necessity of charge separation
of rhodamine B (RB) and methylene blue (MB) in aqueous solution and electron transport inside the material during the reaction
under visible light irradiation (l 4 420 nm). The photocatalytic process. In order to achieve an efficient photocatalyst for hydrogen

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14039
View Article Online

NJC Perspective

It is accepted that capturing and efficient use of CO2 are


currently important issues due to the issue of global warming.
Using MOF photocatalysts to convert CO2 into valuable organic
products has received much attention in both academic and
industrial research. In 2012, Li and co-workers reported for the
first time the use of MIL-125-NH2 for CO2 reduction under
visible light irradiation.70 Acetonitrile (MeCN) and triethanol-
amine (TEOA) were utilized as organic solvent and sacrificial
agent, respectively, to conduct the CO2 reduction experiments.
Under visible light irradiation and the presence of MIL-125-NH2
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Fig. 18 The photocatalytic hydrogen production reaction over Pt-supported catalyst, the formation of formate anion (HCOO) in the liquid
MIL-125-NH2 based on the LMCT mechanism, demonstrating electron phase was detected, which reached 8.14 mmol after 10 h of
transfer from the excited BDC-NH2 linker to the titanium–oxo cluster, reaction time. This study not only opened an avenue for the
generating Ti3+. Reprinted with permission from ref. 66. Copyright 2012
application of CO2 reduction using photoresponsive catalysts
American Chemical Society.
such as MOFs and related porous materials, but also demon-
strated the power of crystal structure modification technique in
production, van der Vlugt and Gascon and co-workers published a crystalline engineering for further applications;71,72 just recently,
study that combined the phototocatalytic features of MIL-125-NH2 Uribe-Romo and co-workers explored the photocatalytic activity
and cobaloxime activity to produce Co@MIL-125-NH2, which of an isoreticular series of titanium-based MIL-125.71 N-Alkyl
was prepared by a stepwise PSM approach (ship-in-a-bottle substituted MIL-125-NHR (where R = methyl, ethyl, isopropyl,
technique).68 Indeed, the ligand was initially loaded into the n-butyl, cyclopentyl, cyclohexyl, or n-heptyl) was synthesized and
pores of MIL-125-NH2, followed by reaction of the Co2+ source fully characterized; it was also applied for CO2 reduction under
with the ligand@MIL-125-NH2 processed inside the pores to blue light. The CO2 conversion using MIL-125-NHR catalyst
create the targeted material. The hydrogen production reaction reached a higher yield compared to pristine MIL-125-NH2, as
was then conducted in Co@MIL-125-NH2 and the host material indicated by the increased excited-state lifetime.
MIL-125-NH2. Interestingly, the catalytically active site of the The framework engineering technique was further studied
Co complex loaded into MIL-125-NH2 exhibited a 20-fold by Gascon and co-workers, who demonstrated the uses of PSM
enhancement in the H2 production reaction compared to the to extend the visible light absorption region of MIL-125-NH2 and
pristine MIL-125-NH2. methyl red MIL-125 (MR-MIL-125) via anchoring of antenna-like
Very recently, Jiang and co-workers introduced the encapsulation moieties.73 Very recently, the photocatalytic performance of MIL-125
of a Co(II) complex, [CoII(TPA)Cl][Cl] (TPA = tris(2-pyridylmethyl)- and analogue MIL-125-NH2 was evaluated with increasing amounts
amine), acting as a molecular photocatalytic site, into MIL-125-NH2 of amino functionality. Mellot-Draznieks and co-workers reported
to prepare a new photocatalyst with efficient spatial separation of the synthesis and characterization of MIL-125-NH2 with various
electrons and holes (Fig. 19).69 The catalytic tests conducted under percentages of BDC-NH2 in the structure (0%, 20%, 46%, 70%, and
visible light showed that the 1.97% Co(II) complex encapsulated into 100%).62 Although 100% BDC-NH2 was able to transform visible
the pores of MIL-125 was able to generate H2 evolution with the absorption to the red-shifted region, corresponding to decreased
highest H2 production rate (553 mmol g1 h1) compared to band gap energy, it did not achieve the best results for the benzyl
17 mmol g1 h1 for pristine MIL-125-NH2. alcohol oxidation reaction. The maximum photocatalytic properties
of MIL-125 with mixed BDC and BDC-NH2 in the structure were
found at B50% BDC-NH2. This finding provides an important
insight to further guide linker exchange and crystal structure
engineering for photocatalytic activity enhancement within MOFs.
Engineering the crystal structure of Ti-based MOFs undoubtedly
leads to targeted structures possessing suitable band gap energies as
well as active sites, which enhances the photoactive properties of the
materials. Taking advantage of the well-defined topological net-
works and stability of the Ti-MOFs, the active sites were decorated
either into the pores of the MOFs through interaction between the
post modification motifs and the MOF framework or on the linking
units via chemical reactions. Another method for post synthesis
modification was reported by Yuan and co-workers, who success-
fully modified and synthesized graphitic carbon nitride as a support
of MIL-1254 to prepare new photocatalytic systems exhibiting high
Fig. 19 The ‘‘ship-in-a-bottle’’ synthetic pathway to produce Co(II)@MIL-
photocatalytic activity for Cr(VI) reduction and dye mineralization in
125-NH2 photocatalyst. Reprinted with permission from ref. 69. Copyright wastewater under visible light irradiation. Recently, Wang et al.
2016 American Chemical Society. reported the technique for in situ synthesis of In2S3@MIL-125

14040 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC

core–shell microparticles. This material acted as an integrated Conclusions and outlook


photocatalytic adsorbent, demonstrating high efficiency for the
removal of antibiotics (i.e., tetracycline) in wastewater.61 Although the discovery of Ti-MOFs faces challenges due to the
high reactivity of the starting titanium sources, the study of the
architectural design and new strategies for the preparation of
Highlighted potential applications new Ti-MOFs are still dramatically developing. In this review,
Based on their high photoresponsitivities, MIL-125 and MIL- we have summarized the progress of Ti-based metal–organic
125-NH2 have been utilized in a wide range of photocatalytic frameworks developed in just one decade, including their
applications. However, the biological activity of Ti-based MOF synthetic methodologies, crystal structures and topological
had not been reported and studied. Very recently, Silva and descriptions, and highlighted their applications. Particularly,
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

co-workers successfully modified fabrics, including viscose, linen, we would like to conclude that the synthetic conditions are an
and 100% cotton, with 3-glycidyloxypropyltrimethoxysilane, important factor that contributes to determining robust synthesis
followed by coating with MIL-125-NH2.74 The coated material methodologies for the preparation of new MOFs. Moreover,
was characterized by PXRD, UV-vis spectroscopy, FT-IR and structural features and topological networks were selectively
scanning electron microscopy (SEM) with energy-dispersive evaluated to provide the chemistry of molecular design, topology,
X-ray spectroscopy, which demonstrated the uniform coating and the power of reticular chemistry. These factors are irreplace-
of the coated fabrics. Interestingly, the modified fabrics able for most structural analyses, opening pathways to synthesize
exhibited antimosquito properties, both attracting and killing crystalline materials. This study contributes to the overall road-
mosquitoes. The proposed mosquito killing mechanism of map for further research based on the crystallographic design of
MIL-125-NH2 is that mosquitoes can easily find a person by novel Ti-MOF chemistry as well as post synthetic engineering for
sensing temperature and carbon dioxide, which was produced novel applications. It is fair to note that most Ti-MOF studies
by the photodegradation of dimethylol dihydroxyethylene urea focus on photocatalytic applications for H2 production, CO2
under the MIL-125-NH2 photocatalyst. Hence, mosquitoes were conversion, organic dye degradation, and alcohol oxidation due
attracted to the modified fabrics. This study provides new to the structural properties of Ti-MOFs, which are correlated to
insight into the use of MOFs materials for novel applications their photoactivities. This may lead to the development of a new
such as biological activity to address public health problems. Ti-based MOF that is not only stable in harsh conditions and
Recently, MIL-125-NH2 has been modified to achieve high provides sufficient power for water oxidation, but can also reduce
crystallinity using various Ti(IV) sources such as Ti(i-PrO)4 and the gap between academic research and practical applications.
Ti(BuO)4 under reflux or solvothermal synthetic methods.
MIL-125-NH2 synthesized from Ti(BuO)4 salt by the solvo- Conflicts of interest
thermal method not only showed the highest internal surface
area (1553 m2 g1) compared to the other published MIL-125- There are no conflicts to declare.
NH2 material, but also exhibited an ideal S-shaped isotherm for
water uptake measurement, with a steep increase at 0.2P/P0 of Acknowledgements
relative pressure at 35 1C. The water uptake capacities of
Ti(BuO)4-derivative-MIL-125-NH2 were found to be 0.550 and H. L. N. would like to pay final respects to his father who just
0.670 g g1 at P/P0 = 0.30 and 0.92, respectively. This result passed away. H. L. N. acknowledges Prof. Davide M. Proserpio
outperformed comparable materials such as A520, MIL-100(Fe), (University of Milan) for advanced discussion on topological
MIL-100(Cr), MIL-101(Cr), and MIL-101-NH2(Cr).75 analysis and Mr Bao N. Truong (KITECH, Korea) for positive
One of the most important applications in industrial fields suggestions and recommendations. H. L. N. also acknowledges the
is polymer chemistry. In particular, the global polymethyl funding from VNU-HCM under grant number VNUB-2017-50-01.
methacrylate (polyMMA) market is anticipated to reach USD
8.16 billion by 2025.76 Hence, developing a strategy based on a Notes and references
new catalyst for uniform and high molecular weight polyMMA
preparation is necessary. In this view, Nguyen et al. developed 1 H. Furukawa, K. E. Cordova, M. O’Keeffe and O. M. Yaghi,
two new Ti-based MOFs materials, MOF-901 and MOF-902, Science, 2013, 341, 1230444.
possessing high optical response and low band gap energies 2 P. Z. Moghadam, A. Li, S. B. Wiggin, A. Tao, A. G. P. Maloney,
(2.65 and 2.50 eV for MOF-901 and MOF-902, respectively).45,46 P. A. Wood, S. C. Ward and D. Fairen-Jimenez, Chem. Mater.,
PolyMMA, which was prepared using MOF-901 photocatalyst 2017, 29, 2618–2625.
under visible light irradiation, reached 26 850 g mol1 and a 3 S.-L. Li and Q. Xu, Energy Environ. Sci., 2013, 6, 1656–1683.
polydispersity index (PDI) of 1.60. MOF-902 was found to be 4 H. Wang, X. Yuan, Y. Wu, G. Zeng, X. Chen, L. Leng and
more active than that of MOF-901, TiO2, and other related H. Li, Appl. Catal., B, 2015, 174, 445–454.
Ti-based MOFs. MOF-902 could promote the reaction of poly- 5 H. Wang, X. Yuan, Y. Wu, X. Chen, L. Leng and G. Zeng, RSC
MMA under identical conditions used for MOF-901; however, it Adv., 2015, 5, 32531–32535.
accounted for a molecular weight of 31 465 g mol1 and a PDI 6 A. Schoedel, Z. Ji and O. M. Yaghi, Nat. Energy, 2016,
value of 1.11. 1, 16034.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14041
View Article Online

NJC Perspective

7 W. Wang, X. Xu, W. Zhou and Z. Shao, Adv. Sci., 2017, 31 Y.-F. Li, U. Aschauer, J. Chen and A. Selloni, Acc. Chem. Res.,
4, 1600371. 2014, 47, 3361–3368.
8 T. V. Tran, H. T. N. Le, H. Q. Ha, X. N. T. Duong, L. H. T. Nguyen, 32 M. Hmadeh, Z. Lu, Z. Liu, F. Gándara, H. Furukawa, S. Wan,
T. L. H. Doan, H. L. Nguyen and T. Truong, Catal. Sci. Technol., V. Augustyn, R. Chang, L. Liao, F. Zhou, E. Perre, V. Ozolins,
2017, 7, 3453–3458. K. Suenaga, X. Duan, B. Dunn, Y. Yamamto, O. Terasaki and
9 Y. Bai, Y. Dou, L.-H. Xie, W. Rutledge, J.-R. Li and O. M. Yaghi, Chem. Mater., 2012, 24, 3511–3513.
H.-C. Zhou, Chem. Soc. Rev., 2016, 45, 2327–2367. 33 N. T. T. Nguyen, H. Furukawa, F. Gándara, C. A. Trickett,
10 T. L. H. Doan, H. L. Nguyen, H. Q. Pham, N.-N. Pham-Tran, H. M. Jeong, K. E. Cordova and O. M. Yaghi, J. Am. Chem.
T. N. Le and K. E. Cordova, Chem. – Asian J., 2015, 10, Soc., 2015, 137, 15394–15397.
2660–2668. 34 F. Gándara, F. J. Uribe-Romo, D. K. Britt, H. Furukawa,
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

11 V. Guillerm, S. Gross, C. Serre, T. Devic, M. Bauer and L. Lei, R. Cheng, X. Duan, M. O’Keeffe and O. M. Yaghi,
G. Férey, Chem. Commun., 2010, 46, 767–769. Chem. – Eur. J., 2012, 18, 10595–10601.
12 The Chemical & Engineering News, http://cen.acs.org/arti 35 H. Assi, L. C. Pardo Pérez, G. Mouchaham, F. Ragon,
cles/95/i25/Europe-call-TiO2-carcinogen. M. Nasalevich, N. Guillou, C. Martineau, H. Chevreau,
html?type=paidArticleContent, accessed June 2017. F. Kapteijn, J. Gascon, P. Fertey, E. Elkaim, C. Serre and
13 M. A. Nasalevich, M. van der Veen, F. Kapteijn and T. Devic, Inorg. Chem., 2016, 55, 7192–7199.
J. Gascon, CrystEngComm, 2014, 16, 4919–4926. 36 W. Xuan, C. Ye, M. Zhang, Z. Chen and Y. Cui, Chem. Sci.,
14 S. Abedia and A. Morsali, New J. Chem., 2015, 39, 931–937. 2013, 4, 3154–3159.
15 N. Stock and S. Biswas, Chem. Rev., 2012, 112, 933–969. 37 K. E. Cordova and O. M. Yaghi, Mater. Chem. Front., 2017, 1,
16 D. L. Murphy, M. R. Malachowski, C. F. Campana and 1304–1309.
S. M. Cohen, Chem. Commun., 2005, 5506–5508. 38 V. A. Blatov, A. P. Shevchenko and D. M. Proserpio, Cryst.
17 A. C. Kathalikkattil, R. Babu, R. K. Roshan, H. Lee, H. Kim, Growth Des., 2014, 14, 3576–3586.
J. Tharun, E. Suresh and D.-W. Park, J. Mater. Chem. A, 2015, 39 C. Zhu, X. Chen, Z. Yang, X. Du, Y. Liu and Y. Cui, Chem.
3, 22636–22647. Commun., 2013, 49, 7120–7122.
18 H. Assi, G. Mouchaham, N. Steunou, T. Devic and C. Serre, 40 K. Hong, W. Bak and H. Chun, Inorg. Chem., 2013, 52, 5645–5647.
Chem. Soc. Rev., 2017, 46, 3431–3452. 41 K. Hong and H. Chun, Chem. Commun., 2013, 49, 10953–10955.
19 C. J. Chuck, M. G. Davidson, M. D. Jones, G. Kociok-Köhn, 42 S. Yuan, T.-F. Liu, D. Feng, J. Tian, K. Wang, J. Qin, Q. Zhang,
M. D. Lunn and S. Wu, Inorg. Chem., 2006, 45, 6595–6597. Y.-P. Chen, M. Bosch, L. Zou, S. J. Teat, S. J. Dalgarno and H.-
20 C. Serre, J. A. Groves, P. Lightfoot, A. M. Z. Slawin, C. Zhou, Chem. Sci., 2015, 6, 3926–3930.
P. A. Wright, N. Stock, T. Bein, M. Haouas, F. Taulelle and 43 K. Hong, W. Bak, D. Moon and H. Chun, Cryst. Growth Des.,
G. Férey, Chem. Mater., 2006, 18, 1451–1457. 2013, 13, 4066–4070.
21 M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin, C. Sanchez 44 O. M. Yaghi, M. O’Keeffe, N. W. Ockwig, H. K. Chae,
and G. Férey, J. Am. Chem. Soc., 2009, 131, 10857–10859. M. Eddaoudi and J. Kim, Nature, 2003, 423, 705–714.
22 C. H. Hendon, D. Tiana, M. Fontecave, C. Sanchez, L. D’arras, 45 H. L. Nguyen, F. Gándara, H. Furukawa, T. L. H. Doan,
C. Sassoye, L. Rozes, C. Mellot-Draznieks and A. Walsh, J. Am. K. E. Cordova and O. M. Yaghi, J. Am. Chem. Soc., 2016, 138,
Chem. Soc., 2013, 135, 10942–10945. 4330–4333.
23 K. Khaletskaya, A. Pougin, R. Medishetty, C. Rösler, C. Wiktor, 46 H. L. Nguyen, T. T. Vu, D. Le, T. L. H. Doan, V. Q. Nguyen
J. Strunk and R. A. Fischer, Chem. Mater., 2015, 27, 7248–7257. and N. T. S. Phan, ACS Catal., 2017, 7, 338–342.
24 J. H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, 47 Z. Wang and S. M. Cohen, J. Am. Chem. Soc., 2007, 129,
S. Bordiga and K. P. Lillerud, J. Am. Chem. Soc., 2008, 130, 12368–12369.
13850–13851. 48 J. Park, D. Feng and H.-C. Zhou, J. Am. Chem. Soc., 2015, 137,
25 M. Li, D. Li, M. O’Keeffe and O. M. Yaghi, Chem. Rev., 2014, 11801–11809.
114, 1343–1370. 49 A. Fracaroli, P. Siman, D. Nagib, M. Suzuki, H. Furukawa,
26 J. Gao, J. Miao, P.-Z. Li, W. Y. Teng, L. Yang, Y. Zhao, B. Liu F. D. Toste and O. M. Yaghi, J. Am. Chem. Soc., 2016, 138,
and Q. Zhang, Chem. Commun., 2014, 50, 3786–3788. 8352–8355.
27 B. Bueken, F. Vermoortele, D. E. P. Vanpoucke, H. Reinsch, 50 A. M. Shultz, O. K. Farha, A. A. Sarjeant, J. T. Hupp and
C.-C. Tsou, P. Valvekens, T. De Baerdemaeker, R. Ameloot, S. T. Nguyen, J. Am. Chem. Soc., 2011, 133, 13252–13255.
C. E. A. Kirschhock, V. Van Speybroeck, J. M. Mayer and 51 T. Grancha, J. Ferrando-Soria, H.-C. Zhou, J. Gascon,
D. De Vos, Angew. Chem., Int. Ed., 2015, 54, 13912–13917. B. Seoane, J. Pasán, O. Fabelo, M. Julve and E. Pardo, Angew.
28 D. Feng, T.-F. Liu, J. Su, M. Bosch, Z. Wei, W. Wan, Y.-P. Chem., Int. Ed., 2015, 54, 6521–6525.
Chen, X. Wang, K. Wang, X. Lian, Z.-Y. Gu, J. Park, D. Yuan, 52 B. Li, B. Gui, G. Hu, D. Yuan and C. Wang, Inorg. Chem.,
X. Zou and H.-C. Zhou, Nat. Commun., 2015, 6, 5979. 2015, 54, 5139–5141.
29 J. A. Mason, L. E. Darago, W. W. Lukens and J. R. Long, 53 M. Kim, J. F. Cahill, H. Fei, K. A. Prather and S. M. Cohen,
Inorg. Chem., 2015, 54, 10096–10104. J. Am. Chem. Soc., 2012, 134, 18082–18088.
30 M. Zhao, K. Yuan, Y. Wang, G. Li, J. Guo, L. Gu, W. Hu, 54 C. K. Brozek and M. Dincă, J. Am. Chem. Soc., 2013, 135,
H. Zhao and Z. Tang, Nature, 2016, 539, 76–80. 12886–12891.

14042 | New J. Chem., 2017, 41, 14030--14043 This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017
View Article Online

Perspective NJC

55 L. Zou, D. Feng, T.-F. Liu, Y.-P. Chen, S. Yuan, K. Wang, X. Wang, 67 B. Rungtaweevoranit, J. Baek, J. R. Araujo, B. S. Archanjo,
S. Fordham and H.-C. Zhou, Chem. Sci., 2016, 7, 1063–1069. K. M. Choi, O. M. Yaghi and G. A. Somorjai, Nano Lett., 2016,
56 A. Fujishima and K. Honda, Nature, 1972, 238, 37–38. 16, 7645–7649.
57 H. Xu, S. Ouyang, L. Liu, P. Reunchan, N. Umezawa and 68 M. A. Nasalevich, R. Becker, E. V. Ramos-Fernandez,
J. Ye, J. Mater. Chem. A, 2014, 2, 12642–12661. S. Castellanos, S. L. Veber, M. V. Fedin, F. Kapteijn, J. N. H.
58 M. Alvaro, E. Carbonell, B. Ferrer, F. X. Llabrés i Xamena Reek, J. I. van der Vlugt and J. Gascon, Energy Environ. Sci.,
and H. Garcia, Chem. – Eur. J., 2007, 13, 5106–5112. 2015, 8, 364–375.
59 S. Hausdorf, J. Wagler, R. Moßig and F. O. R. L. Mertens, 69 Z. Li, J.-D. Xiao and H.-L. Jiang, ACS Catal., 2016, 6,
J. Phys. Chem. A, 2008, 112, 7567–7576. 5359–5365.
60 J. G. Santaclara, F. Kapteijn, J. Gascon and M. A. van der 70 Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu and Z. Li,
Published on 04 October 2017. Downloaded by University of Windsor on 13/12/2017 23:25:30.

Veen, CrystEngComm, 2017, 19, 4118–4125. Angew. Chem., Int. Ed., 2012, 51, 3364–3367.
61 H. Wang, X. Yuan, Y. Wu, G. Zeng, H. Dong, X. Chen, 71 M. W. Logan, S. Ayad, J. D. Adamson, T. Dilbeck, K. Hanson
L. Leng, Z. Wu and L. Peng, Appl. Catal., B, 2016, 186, 19–29. and F. J. Uribe-Romo, J. Mater. Chem. A, 2017, 5, 11854–11863.
62 C. Yu, L. Wei, J. Chen, Y. Xie, W. Zhou and Q. Fan, Ind. Eng. 72 Y. Lee, S. Kim, H. Fei, J. K. Kang and S. M. Cohen, Chem.
Chem. Res., 2014, 53, 5759–5766. Commun., 2015, 51, 16549–16552.
63 J. G. Santaclara, M. A. Nasalevich, S. Castellanos, W. H. Evers, 73 M. A. Nasalevich, M. G. Goesten, T. J. Savenije, F. Kapteijn
F. C. M. Spoor, K. Rock, L. D. A. Siebbeles, F. Kapteijn, and J. Gascon, Chem. Commun., 2013, 49, 10575–10577.
F. Grozema, A. Houtepen, J. Gascon, J. Hunger and M. A. van 74 R. M. Abdelhameed, O. M. H. M. Kamel, A. Amr, J. Rocha and
der Veen, ChemSusChem, 2016, 9, 388–395. A. M. S. Silva, ACS Appl. Mater. Interfaces, 2017, 9, 22112–22120.
64 H. Q. Pham, T. Mai, N.-N. Pham-Tran, Y. Kawazoe, H. Mizuseki 75 M. Sohail, Y.-N. Yun, E. Lee, S. K. Kim, K. Cho, J.-N. Kim,
and D. Nguyen-Manh, J. Phys. Chem. C, 2014, 118, 4567–4577. T. W. Kim, J.-H. Moon and H. Kim, Cryst. Growth Des., 2017,
65 M. B. Chambers, X. Wang, L. Ellezam, O. Ersen, M. Fontecave, 17, 1208–1213.
C. Sanchez, L. Rozes and C. Mellot-Draznieks, J. Am. Chem. Soc., 76 ReportLinker, https://www.reportlinker.com/p04838542/Poly-
2017, 139, 8222–8228. methyl-Methacrylate-PMMA-Market-Analysis-By-Product-Extruded-
66 Y. Horiuchi, T. Toyao, M. Saito, K. Mochizuki, M. Iwata, Sheets-Pellets-And-Acrylic-Beads-By-End-Use-Automotive-
H. Higashimura, M. Anpo and M. Matsuoka, J. Phys. Chem. Construction-Electronics-Signs-Displays-Competitive-Landscape-
C, 2012, 116, 20848–20853. And-Segment-Forecasts, accessed March 2017.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2017 New J. Chem., 2017, 41, 14030--14043 | 14043

You might also like