Coordination Chemistry Reviews 399 (2019) 213023
Contents lists available at ScienceDirect
Coordination Chemistry Reviews journal homepage: www.elsevier.com/locate/ccr
Review
Linker functionalized metal-organic frameworks Sayed Ali Akbar Razavi, Ali Morsali ⇑ Department of Chemistry, Faculty of Sciences, Tarbiat Modares University, P.O. Box 14117-13116, Tehran, Islamic Republic of Iran
a r t i c l e
i n f o
Article history: Received 26 April 2019 Accepted 8 August 2019
Keywords: Metal-organic frameworks (MOFs) Functional(ized) metal-organic frameworks (FMOFs) Host-guest chemistry of MOFs Function-application properties Function-structure properties
a b s t r a c t Owing to their three dimensionality and high porosity, metal-organic frameworks (MOFs) have attracted the attention of scientists especially chemists and material engineers. The frameworks of this subclass of hybrid materials are made of inorganic metal ions/clusters and organic bridging ligands. Special connections between these two building blocks of MOFs lead to a theoretically unlimited number of frameworks. Unlike other porous materials, MOFs benefit from characteristics such as high crystallinity and regularity, high porosity and surface area, hybrid organic-inorganic nature, moderate to high chemical and thermal stability, and decorable pores with different functional groups. Decorating MOFs with functions is possible through functionalization of organic linkers, inorganic building blocks, and void cavities of the framework. Tunability of MOFs with organic linkers is of particular importance due to the unlimited possibility to design functional or multi-functional organic linkers as well as distinctive chemical properties of organic functional groups (OFGs). The purpose of this review is to gain deeper insight into the effects of organic functional groups on structure, host-guest chemistry and applications of functional metal-organic frameworks (FMOFs) in order to be able to synthesize functional MOFs for specific purposes through an in depth analysis of the literature. Ó 2019 Elsevier B.V. All rights reserved.
Contents 1.
2.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.1. Metal-organic frameworks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2. Functional metal-organic frameworks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.3. Scope of the current review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.4. Classification of OFGs and FMOFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 Functional groups. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 2.1. Carbonyl-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 2.1.1. Urea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 2.1.2. Amide and oxalamide. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 2.1.3. Squaramide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.1.4. Carbonyl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 2.1.5. Imide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.1.6. Carboxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.2. Nitrogen-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.2.1. Heterocyclic azine N-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.2.2. Heterocyclic azole N-Based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.2.3. Non-cyclic N-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 2.2.4. Ionic N-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 2.3. Oxygen-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.3.1. Hydroxyl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.3.2. Ether . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2.3.3. Other oxygen-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
⇑ Corresponding author. E-mail address:
[email protected] (A. Morsali). https://doi.org/10.1016/j.ccr.2019.213023 0010-8545/Ó 2019 Elsevier B.V. All rights reserved.
2
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
2.4.
3.
Sulfur-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.1. Thiol and sulfide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.2. Sulfonate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4.3. Other sulfur-based functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5. Other functional groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.1. Halogen-based functional groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5.2. Phosphonate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusion and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Introduction 1.1. Metal-organic frameworks Since 1995 through the pioneering work of Yaghi, three dimensional porous coordination polymers which are denoted as metalorganic frameworks (MOFs) have received a lot of attention from scientists, especially in the fields of chemistry and material science [1,2]. The great interest in designing novel MOFs is partly due to their unique structural and practical properties. MOFs are an amazing type of crystalline, porous, and hybrid material which can be synthesized by solvothermal, electrochemical, ultrasonic, microwave, and mechanochemical methods in mild conditions through the coordination of metal ions/clusters with organic bridging linkers as inorganic and organic molecular building blocks [3–13]. Unlimited number of MOFs with different structural and practical properties can be developed by changing metal ion/clusters, using various combinations of these inorganic building blocks and infinite types of organic linkers with different lengths, functionalities and geometries. Furthermore, owing to their unlimited structural variation and regular nature, MOFs show very useful properties like highly ordered and crystalline framework, high porosity and accessible pores, noticeable stability against air, humidity, and heat, as well as hybrid inorganic-organic nature and especially tunability in chemical functionalization. As a result of such distinctive characteristics, MOFs are employed for different purposes such as gas storage and separation [14–16], heterogeneous catalysis [17,18] and photocatalysis [19,20], sensing [21,22], removal and separation of hazardous chemicals [23,24], drug delivery, bio sensing and other medical applications [25,26], electrical conductivity and electrochemical applications [27], ion storage and conductivity [28], and designing MOF-based stimuliresponsive and energetic materials [29–31].
topology of MOFs through induced structural changes and different types of secondary interactions. Therefore, it is possible to synthesize FMOFs with desired functionality, stability, and porosity for specific applications through pre-designing of FMOFs by functionalization using chemically ideal organic functional groups. 1.3. Scope of the current review Considering the importance of organic functionalization of MOFs, in this review the effects of organic functional groups on MOF structures, host-guest chemistry and applications are discussed. Although there are some published papers on FMOFs, none of them have exclusively discussed about linker by linker approach to functionalize MOFs [32–36]. The content of these reviews are primarily focused on: (I) linker design by focusing on length, geometry, symmetry, and hapticity of linkers for discovering new MOFs and (II) application-based classification (gas adsorption, catalysis, and . . .) approach for examination of FMOFs through linker, pore, and metal ion functionalization without a focus on any specific functional groups. Therefore, we think that in those reviews the main objective has been to discuss the applications of FMOFs rather explanation of functionalization effects. Given the focus of our research team on designing FMOFs using different organic functional groups and examining the effects of organic functional groups chemistry on the structure and applications of FMOFs, and considering the growing importance of functionalization of porous martials especially MOFs by organic functional groups in past few years, we decided to write the current review with a completely different approach, coherent
1.2. Functional metal-organic frameworks One of the most important features of MOFs is their ability to be decorated with various types of functions which can be incorporated into different parts of the MOF structure including metal ions/clusters, organic bridging ligands, and empty spaces inside the cavities (Fig. 1). Due to the variety of organic functional groups and their rich host-guest chemistry, organic functional groups are abundantly used for construction of functional metal-organicframeworks (FMOFs). The most important reason for introduction of organic functional groups into the structure of MOFs is to enrich the host-guest chemistry between MOFs as host and other small molecules as guests. Tuning and optimizing the host-guest chemistry for MOFs through functionalization with organic functional groups is an excellent, practical, and rational idea for improving the efficiency of MOFs in different applications. On the other hand, functionalization has also crucial effects on the structural properties such as crystallinity, porosity, flexibility, stability, and
42 42 43 44 44 44 45 46 49 49
Fig. 1. Functionalizable parts of MOFs.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
3
Fig. 2. Classification of organic functional groups based on their chemical characteristics and structural properties.
content, and particular focus on decoration of MOFs with organic functional groups and their effects on the structure, host-guest chemistry, and application of FMOFs. Unlike other reviews that categorize and present content based on applications of MOFs, our classification in this review is based on similarity and differences in chemical and structural characteristics of functional
groups. After classification of the general content using function based classification approach, findings on each one of the mentioned organic functional groups are discussed under ‘‘function-a pplications” and ‘‘function-structure” models. Advantages of this classification are: (I) the effects of each organic functional groups on the structure and applications of FMOFs are easily identified
4
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
2. Functional groups In this section, four major categories of FMOFs are discussed. The effects of organic functional group on FMOF behavior are classified into two groups: (I) function-application properties and (II) function-structure properties. For each one of the organic functional groups, after presenting a brief description of chemical and structural properties, the effects of organic functional groups on the structure, host-guest chemistry, and application of the FMOFs are discussed and some examples are presented. 2.1. Carbonyl-based functions
Fig. 3. Classification of organic functional groups based on their role and position in the framework.
and can be easily compared with other organic functional groups and (II) it is possible to efficiently design FMOFs tailored for specific purposes through having such a precise and comprehensive information about the effects organic functional groups on the structure and applications of FMOFs. 1.4. Classification of OFGs and FMOFs Here, we classified organic functional groups by two approaches. In the first approach, we classified organic functional groups based on their chemical and structural properties into four major categories including (I) carbonyl-based functions, (II) nitrogen-based functions, (III) oxygen-based functions, and (IV) sulfur-based functions (Fig. 2). Each of these main groups includes a number of sub-categories or functional groups. Afterward, FMOFs are explained according to this approach. In next approach, we classified organic functional groups in two groups based on their role in MOF structures (Fig. 3): (I) functional groups as coordinating sites and (II) functional groups as guestinteractive sites. Coordinating functional groups are groups of functions which coordinate to metal ions during self-assembly construction MOFs. The functions which are located in this group most often include functions such as carboxylate, sulfonate, phosphonate, and some of heterocyclic groups like azoles, pyridine, and diazines. In addition to coordinating sites, these functions are also applied as guest-interactive sites. Here we focus on guestinteractive characteristics of these functions, but the ability and chemistry of these functions as coordinating sites are also briefly considered. Guest-interactive sites consist of broader functions. This group of functions remains intact in self-assembly process and are free and ready to interact with guest molecules. Much of this review is focused on guest-interactive capability of functional groups. It is necessary to mention that functional groups which are discussed here may be categorized in only one group or in both groups. For example some functions like carboxylate, sulfonate, phosphonate, and some of heterocyclic groups like azoles, pyridine, and diazines are applied as both coordinating and guest-interactive sites, but some other functions like urea, amide, oxalamide, imide, triazine and tetrazine are mostly applied as guest interactive sites. Applications of functional groups as guest-interactive sites inside the structure of MOFs along with their possible host-guest chemistry and mechanism are presented in Table 1.
A considerable number of the functions introduced inside the structure of MOFs are carbonyl-containing functions including urea, amide, oxalamide, squaramide, ketone, imide, carboxyl, and some other groups such as hydrazone [664–666]. Due to the presence of carbonyl group, all of these functions have partially mutual host-guest chemistry. However, the difference in their host-guest chemistry is because of the various groups connected to the carbonyl function. Direct connections between carbonyl and other groups including amine(s) and hydroxy provide different properties for this group of FMOFs. 2.1.1. Urea In field of synthetic organic chemistry, urea provides the most commonly employed moieties as two-point hydrogen bond donor catalysts through N–H bonds and hydrogen bond acceptor site from (C@O) motif. Owing to the potential of urea group in noncovalent interactions, through multiple supramolecular hydrogen bonds, it is a desirable function for interaction with electronegative and (partially) negatively charged species (Fig. 4a) [667]. The concepts of anion binding or binding with negatively (partially) charged heteroatoms is a key to designing new method for organocatalytic transformations, because organocatalysts display interactions with anionic moieties. It means that urea can be applied as anion recognizer and organocatalyst site toward electron rich species [668]. Along with such advantages of urea, there is an important problem in homogeneous urea containing compounds. Urea functions have a significant tendency in self-association and make oligomers through urea(–NH2). . ..(–OC–)urea hydrogen bonds (Fig. 4b). As a result, this tendency for self-aggregation decreases the solubility and reactivity of homogeneous urea. Recently, chemists have found a promising strategy to solve this problem by preventing the self-association of urea functions. Hupp and coworkers reported that through immobilization of urea functions inside the structure of MOFs, their self-association is prevented (Fig. 4c) and thus the wealth of its hydrogen bond based chemistry can be exploited in designing FMOFs [43]. Hupp and coworkers reported the first urea containing MOF with ideas of using organocatalyst hydrogen bond donors and preventing self-association of homogeneous urea catalysts [43]. Since then, urea functionalized MOFs have been applied as hydrogenbond organocatalysts in different reactions (Table 1). NU-601 with formula Zn2(bipy)2(H4L) where H4L is 5,50 -(carbonylbis(azanediyl)) diisophthalic acid (Fig. 5a) has been used as urea functionalized MOF for alkylation between nitro-alkenes and pyrroles. Findings show that NU-601 has a higher activity as the heterogeneous catalyst compared to the diphenyl urea as homogeneous catalyst (90% conv. at 18 h vs 65% conv. at 18 h) and control reaction (21% conv. at 48 h) (Fig. 5b). Urea FMOFs have also been applied for interaction with oxoanions and electron rich oxygen containing organics through (urea) NH. . ..O(partially negative) hydrogen bonding which can be used
5
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 Table 1 Applications of functional MOFs and their corresponding host-guest interaction and mechanisms. Functional group
Application
Details of application
Host-guest interaction and mechanisms
Ref.
Urea
Catalyst
Freidel-Craft Henry Epoxide ring opening Nitroimidazole NO2 SO2 Artificial sweeteners NH3 Phosphorylated peptides Pb2+ Hg2+ SO2 4 Cr2O2 7 Nitroaromatics Phenol TNP
Hydrogen bonding with electron-rich or anionic spices
[37–44] [45] [46] [47] [48] [49] [50] [49] [51] [52] [52] [53–57] [58] [59] [60] [52,58,61]
Removal
Sensing
Amide
Gas adsorption
CO2 C2H2 H2
Catalyst Removal
Oxalamide
Sensing Gas adsorption
Knoevenagel Henry Pb2+ U6+ n-butanol TNP CO2 CH4
Catalyst Photoelectronic Squaramide Carbonyl
Catalyst Sensing Gas adsorption Sensing
Stimuli Responsivity Imide
Gas adsorption Photocatalyst
Sensing
Stimuliresponsivity
Electrochemistry Conductivity Carboxy
CO2 conversion Antenna for energy transfer Friedel-Craft Lactose CO2 Organic Solvents Aniline Phenol Moisture Ratiometric luminescent thermometer Polarized fluorescence CO2 H2 C–H and C–C reduction C–OH and C–NH2 oxidation Organic Solvents Organic Amines Substituted phenyl ring Photochromic Electrochromic Electrical conductivity Energy transfer Li storage OH
Conductivity
Cd2+ Cr3+ Cu2+ Fe3+ Tb3+ TNP Benzaldehyde Pyridine H+
Removal
NH3
Sensing
Analyte(NO2) (NH)urea hydrogen bonding Analyte(SO) (NH)urea hydrogen bonding Ammonia(NH) (NH)urea hydrogen bonding Analyte(PO) (NH)urea hydrogen bonding Urea metal ion coordination interactions Oxoanion(O) (NH)urea hydrogen bonding Analyte(NO2) (NH)urea hydrogen bonding Phenol(OH) (C@O)urea hydrogen bonding TNP(OH) (C@O)urea hydrogen bonding TNP(NO2) (NH)urea hydrogen bonding Amide(O) C(CO2) dipole/quadrupole interaction Amide(NH) O(CO2) hydrogen bonding Amide(NH) p(C2H2) hydrogen bonding Amide(O) p(C2H2) polarization Improving H2-framework interaction by increasing positive charge on metal ion Providing basic environment by amide function Amide metal coordination interaction Amide(O) (HO)n-butanol hydrogen bonding Amide(O) (HO)TNP hydrogen bonding Simultaneous oxalamide(O) C(CO2) and oxalamide(NH) O(CO2) interactions Decreasing CH4-framwork interaction in low pressure and increasing at high pressure CO2 fixation and activation in pores by strong oxalamide-CO2 interaction Adsorption of light photons and transfer to metal ions Hydrogen bonding with electron-rich spices Squaramide(NH) (O)lactose hydrogen bonding Carbonyl(O) C(CO2) Lewis base-acid interaction Dipole-quadrupole interaction Charge transfer between electroactive fluorenone and solvents Carbonyl(O) (H2N)aniline hydrogen bonding Carbonyl(O) (HO) hydrogen bonding
[62–84] [72,85] [86] [77,87,88] [89] [90] [91] [87] [92] [93–96] [97] [93] [98] [99–101] [102] [103–05]
Thermally activated metal to fluorene back-energy transfer
[106] [107] [107] [108] [109]
fluorene to metal energy transfer Reduction of NDI core for incorporation of Li+ ions Dipole (C@O of NDI) gas quadrupole interaction light-induced electron transfer through p. . .p stacked organized PDI arrays
[110] [111–115] [113,116] [117] [117]
Charge transfer through formation of e-rich(analyte) e-poor(NDI) complexes
[118] [118] [119] [120–125] [120,126] [121,122,127,128] [129] [130] [131]
Photoinduced NDI/NDI redox couple formation Photoinduced charge transfer from e-rich spices to e-poor NID core Reversible redox conversion of carbonyl of NDI Cation(EVIm) p(NDI) interaction to provide free hydroxide from [ EVIm] OH ion pair Coordination/chelation interaction between metal ions and carboxy function
Chelating and antenna for sensitizing the visible-emitting Tb(III) cation TNP(OH) (O)free carboxy hydrogen bonding Benzaldyde(C@O) (HOOC)Carboxyl hydrogen bonding Pyridine(N) (HOOC)Carboxy hydrogen bonding Hydrogen bonding between carboxy functions and water molecules in the hydrophilic pores Free(COOH) NH3 acid-base interaction
[132,133] [134] [135–137] [138–140] [141] [134] [142] [143] [144–148] [149] (continued on next page)
6
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Table 1 (continued) Functional group
Application
Separation Catalyst Gas adsorption Pyridine
Sensing
Gas adsorption
Catalyst
Diazines
Gas adsorption
Removal Triazine
Gas adsorption
Details of application
Host-guest interaction and mechanisms
Ref.
NO2 SO2 Cu2+ from Co2+ Epoxide ring opening CO2 H2 Nitrophenol Cu2+ Mn2+ Eu3+ CO2 CH4 C2H2 Esterification Direct arylation CO2 Conversion CO2 hydrogenation CH4 C2H2 CO2 Pb2+ Hg2+ CO2 Benzene C2H6
Chemical reactivity capabilities of hydroxy function
[149] [149] [150] [151] [145,152–156] [156] [157] [158–160] [133] [161] [67,162–166] [166,167] [167] [168] [169] [170] [171] [172–174] [175,176] [177–183] [184] [184] [185-190] [191,192] [192,193]
C2H4 C2H2 Sensing
Tetrazine
Catalyst Sensing
Removal
Gas adsorption
Heptazine
Gas adsorption
Pyrazole
Catalyst Gas adsorption
TNP Nitrophenol Fe3+ Hg2+ Naphthalene Anthracene Phenanthrene N,N’-dimethylaniline pH Knoevenagel Hg2+ Al3+ Phenylhydrazine Phenol CHCl3 Nitrous oxide gas Br2 Acetone Water DMF Ethanol Acetonitrile Nitromethane Rose-Bengal B Pyridine Quinoline CH4 CO2 H2 CO2 C2 hydrocarbons
CO2 hydrogenation CO2
O2 Imidazole
Gas adsorption
CO2
Removal Sensing
Light alcohols Methyl Orange Formaldehyde
Stronger Free(COOH) Cu(II) Coordination/chelation interaction Brønsted acidity of free(COOH) groups Carboxy(O) gas molecules polarization interaction Pyridyl(N) (H)OH of nitrophenol hydrogen bonding Pyridyl(N) metal ion coordination interaction
Pyridyl(N) C(CO2) Lewis base- acid interaction Increased dispersion pyridyl(N) CH4 interaction Pyridyl(N) (H)C2H2 hydrogen bonding Lewis basic catalytic assistant of pyridyl ring Fixation and activation of CO2 by (N)pyridine Increased dispersion N atom CH4 interaction Diazine(N) (H)C2H2 hydrogen bonding Diazine(N) (C)CO2 Lewis base- acid interaction Pyrazine(N) metal ion coordination interaction N(triazine) (C)CO2 base- acid interaction C6H6(p-rich) (p-deficient)triazine interaction Improved van der Waals and electrostatic interaction through guest polarization Improved van der Waals and electrostatic interaction through guest polarization C2(p-rich) (p-deficient)triazine interaction Triazine(N) (HO)TNP hydrogen bonding Triazine(N) metal ion coordination interaction analyte(p-rich) (p-deficient)triazine interaction
Protonation/deprotonation of triazine ring in different pH Lewis basic catalytic assistant by triazine N atoms Tetrazine metal ion coordination interaction Tetrazine(p-deficient) (p-rich)analyte interaction Redox switching between dihydro-tetrazine and tetrazine
Solvatochromic effects through interaction between solvent polarity and active electronic states of tetrazine
Analyte(free O/free COO) (NH)dihydro-tetrazine hydrogen bonding Dihydro-tetrazine(NH)∙∙∙(N)NCC hydrogen bonding Increased CH4 lewis basic tetrazine polarization interaction Weak CO2 Tetrazine interaction N(heptazine) (C)CO2 base-acid interaction Improved van der Waals and electrostatic interaction through increasing polarization C2H2,C2H4(p-rich) (p-poor)heptazine interaction Chemical binding to CO2 and H2 dissociating Pyrazole(NH) (O)CO2 hydrogen bonding Pyrazole(p) (C)CO2 donor–acceptor interaction Pyrazole(C) (O)CO2 donor–acceptor interaction Optimization of CO(II) ligand field by pyrazole for formation of CO(II) superoxo moieties by coordination of O2 to cobalt(II) metal center Imidazole(N) (C)CO2 Lewis base-acid Dipole-quadrupole interaction Imidazole(NH,N) (OH)alcohol hydrogen bonding Imidazole(NH) (SO 3 )dye hydrogen bonding Imidazole(N) (H)HCOH hydrogen bonding Imidazole(CH) (O) HCOH hydrogen bonding
[192,193] [185,192,193] [194–196] [194] [197] [198] [199] [199] [199] [200] [194] [201–203] [204] [205] [206] [207] [208] [209] [209] [210] [210] [210] [210] [210,211] [211] [212] [213] [213] [214] [215–217] [218] [219,220] [193,219]
[221] [222–228]
[229] [156,230–234] [232] [235] [236]
7
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 Table 1 (continued) Functional group
Application
Photoelectronic Catalyst
Triazole
Conductivity Energy storage
Gas adsorption Sensing
Removal Tetrazole
Gas adsorption
Details of application
Host-guest interaction and mechanisms
Ref.
Fe3+ pH
Imidazole(N) metal ion coordination interaction Bufferic behavior of imidazole through readily protonation/deprotonation in different pH Incorporation of imidazole ring in intermolecular proton transfer Carbene formation on C2 atom of imidazole ring
[237] [238]
Originating of conduction pathway from imidazole protons Nitrogen-rich ligands with energetic substituents like nitro and aide stabilized in 3D framework with several energetic N@N, N@C and N–C bonds Triazole(N) (C)CO2 Lewis base-acid interaction Triazole(N) (H)C2H2 hydrogen bonding Triazole(N,p) metal ion coordination interaction
[241] [242–244] [242,244–247]
Sm3+ emission sensitizing Conjugate addition of alcohol to a,b-unsaturated enols H+ Insensitivity Detonation heat CO2 C2H2 Hg2+ Ba2+ Fe3+ Al3+ Neutral red Iodine CO2 H2 H2S Xe O2
Amine
Conductivity
H+
Energy storage
Insensitivity and stability Detonation heat CO2
Gas adsorption
H2O H2S H2 Catalyst
Photocatalyst
Removal
Sensing
CO2 conversion Knoevenagel Henry Aldol Transesterification Biginelli Hantzsch Amine N-formylation Tetrahydro-chromenes synthesis Ugi CO2 conversion Cr(VI) reduction CO2 reduction Organic Transformation a-alkylation of aldehydes Chlorine gas Dichloromethane Trichloromethane Iodine Pyrrole Indole Quinoline Pyridine Phenol NO2 Formaldehyde Benzaldehyde DNA Fe3+ Hg2+ Cr3+ 2,4-dinitrophenol TNP
Deprotonation of triazole ring through strong triazole(NH) Al(III) electrostatic interaction Triazole neutral red hydrogen bonding triazole I2 p-anions interaction Tetrazole(N) (C)CO2 Lewis base-acid interaction Enhanced polarization and strengthened physisorption H2S(acidic) tetrazole (basic) interaction High percentage of open N-donor sites and so localized high charge density for more polarization Coordinate O2 with the resulting Cr(III)–O2 complex though control over ligand field stabilization energy of Cr(II) centers by tetrazole group Hydrogen bond between coordinated water and uncoordinated N (tetrazole) atoms between the layers Nitrogen-rich ligands with energetic substituents stabilized in 3D framework with high density of energetic N@N and C@N bonds Amine(N) (C)CO2 Lewis base-acid interaction (chemisorption for alkylamine and strong physisorption for aromatic amine) Amine(NH) (O)CO2 hydrogen bond Acid(gas molecule) base(amine) interaction Electron-donating group such as amino could enhance the interaction between H2 and framework Providing catalytic Lewis basic site through amine group
Providing adsorptive site for higher performance Antenna effect of photosensitizer amine function (Radical formation by amine group upon irradiation)
Electrophilic aromatic substitution on ortho position of amine group Improved interaction of amine group with dipole moment (low loading) and dispersion (high loading) through halomethane(Cl) (H)amine interaction Amine(NH) (I)iodine hydrogen bonding Amine(N) (NH)analyte hydrogen bonding Amine(NH) (N)analyte hydrogen bonding Amine(N) (HO)Phenol hydrogen bonding Redox reaction through amine conversion to diazonium Amine(NH) (CHO) aldehyde hydrogen bonding Hydrogen bonding between amine group and bases in DNA Coordination interaction between free –NH2 and metal ions
Hydrogen bond interaction and electron transfer between Lewis basic free amine and hydroxyl of explosives
[239] [240]
[248-262] [261] [263] [258] [253] [264] [265] [266,267] [155,268–280] [279–281] [269] [282] [283] [284] [285–298] [285–291,293–300] [162,181,182,186,301356] [306] [338] [339] [337,357–362] [316,317,329,363368] [329,367] [362,369] [370,371] [357] [372] [373] [374] [375] [376] [377,378] [379,380] [381] [382] [383] [384,385] [384,385] [386,387] [388] [388-390] [389,390] [391] [392] [393] [394] [394] [395] [396] [397] [398] [399] [308,398-402] (continued on next page)
8
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Table 1 (continued) Functional group
Application
Details of application 2,4,6-trinitrotoluene 2,4-dinitrotoluene Cr2O2 7 Nitric oxide pH
Separation
Electrochemistry Drug delivery Azine
Gas adsorption Catalyst Removal
Azo
Sensing StimuliResponsivity
Catalyst Gas adsorption
Removal Sensing
Azide Ammonium (Quaternary amine) Imidazolium
Sensing Removal
Catalyst
Gas adsorption Removal
Pyridinium
Conductivity Stimuliresponsivity
Aromatic isomers Propyne from propylene Li-storage Ibuprofen Nimesulide CO2 CH4 Knoevenagel Heavy metal ions Phenol Pyrrole Indole TNP Photomodulation of CO2 adsorption Photomodulation of cargo release Photoswitchability Cyanosilylation Direct amidation CO2 H2 U6+ TNP Fe3+ Al3+ Hg2+ CH3CN H2S Indole Quinoline Acid fuschin Azidation of bromoalkane Thiolation of bromoalkane Benzoic condensation CO2 conversion Cyclotrimerization Nitrobenzene reduction Olefin hydrogenation Heck Suzuki-Miymaura Sonogashira Knoevenagel CO2 hydroboration Isomerization of allylic alcohol CO2 H2 Acid orange II Methyl blue Methyl orange Congo red Orange G Proton Photochromism Thermochromism Hydrochromism Mechanochromic Vapochromism On demand CO2 release
Host-guest interaction and mechanisms
Electron transfer from amine group to dichromate Redox reaction through amine reduction Protonation/deprotonation of amine in different pH and altering the fluorescence emission Different interaction of polar amine groups with guest molecules based on different polarity and shape Multiple hydrogen bonds between amine function and C3H4 molecules Host-guest interaction between Li and nitrogen atoms of amine groups Optimized load and release of drug through amine(H) (COO) ibuprofen Optimized load and release of drug through amine(H) (NO2) nimesulide azine CO2 Lewis base-acid interactions Improved azine CH4 polarization interaction Azine function as Lewis basic catalytic sites Azine metal ions coordination interaction Azine(N) (HO) phenol hydrogen bonding Azine(=N-N@)∙∙∙(NH) neutral NCC hydrogen bonding Azine(N) (HO) TNP hydrogen bonding Through cis–trans conformation transformation of azo function and change in steric or polar effects
Ref. [399] [399] [398] [403] [404] [405] [406] [407] [407] [408] [409-415] [214] [416-418] [419] [420] [213] [213] [421] [422-432] [433]
Conversion between cis–trans conformations of azo function in main or side chain Azo function acts as Lewis basic catalytic site Azo(N) (C)CO2 Coordination interaction Adsorption of H2 molecules near azo function in parallel and perpendicular to azo backbone Azo metal ion coordination interaction Azo(N) (HO)TNP hydrogen bonding Azo metal ion coordination interaction
Acetonitrile(CN) (N@N)azo p-p interaction Redox reaction and conversion of (N3) to (NH2) Quaternary amine(cation) (p)indole interaction Quaternary amine(acid) (base)quinoline interaction Electrostatic between cationic quaternary amine function and anionic dye Cationic framework acts as solid-phase transfer catalyst between organic and aqueous solution Ability to stabilize negative charge in active aldehyde intermediate through acidic carbenic proton of imidazolium ring Imidazolium ring act as fixation and activation site toward CO2 Immobilization of Pd(II) centers on carbenic C2 atom of imidazolium ring
Anchoring of Cu(II) sites on carbenic C2 atom of imidazolium ring Formation of Ir(III)–NHC carbenic heterocyclic complex on imidazolium ring Strong dipole-quadrupole interaction between imidazolium ring and gas molecules Electrostatic interaction between imidazolium ring and anionic dye
Through acidic methylene hydrogen on C2 atom of imidazolium Stimuli-induced (light, pressure, temperature, H2O molecusles) electron transfer from electron rich spices to electron deficient pyridinium ring
Photoinduced elimination of charge guardian of pyridinium ring and weakening the CO2-framework interaction
[434-443] [444] [445] [413,446-451] [452,453] [91,454] [455] [456] [456] [457] [458] [459] [390] [390] [460] [461] [461] [462] [463-466] [467] [468] [468] [468] [468,469] [470] [470] [471] [472] [473-475] [475-477] [478] [478] [478] [468] [479] [480-482] [483-495] [494,495] [483] [496] [493] [497]
9
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 Table 1 (continued) Functional group
Application
Details of application
Host-guest interaction and mechanisms
Ref.
Sensing
Two photon responsivity TNP UV light
Adsorption of two UV photons by pyridinium ring Pyridinium (OH)TNP electrostatic interaction Rapid color change under UV irradiation based on donor(benzoate) (pyridinium)acceptor electron transfer Charge and energy transfer between electron deficient pyridinium ring and electron rich analytes Interaction between anionicCr2O2 7 and cationic pyridinium Through protonation/deprotonation of pendant N atom of the ring Electrostatic and charge transfer interaction between anionic and electron rich analytes and cationic electron poor pyridinium ring
[498] [499] [500]
Removal
Separation Gas adsorption Biosensing
Medical Hydroxy
Gas adsorption
Removal
Sensing
Aniline Phenol Cr(VI) pH I SCN N3 Cl Methyl orange Congo red H2O CH3OH CH3CH2OH NH3 Benzene from cyclohexane CO2 H2 HIV-1 ds-DNA sequences human immunodeficiency virus 1 ds-DNA sudan virus RNA sequences gastric cancer associated microRNAs ebolavirus conserved RNA sequences ebolavirus-encoded miRNAlike fragment Dengue and Zika virus RNA sequences magnetic resonance imaging CO2 SO2 H2 Hg2+ Pharmaceuticals Chloramphenicol Fe3+ Organic Solvents TNP 4-nitrophenol
Ether
Catalyst
Nitrobenzene 1,4-dinitrobenzene 2,4-dinitrochlorobenzene CO2 hydrogenation
Conductivity
H+
Gas adsorption
CO2 H2
Sensing
pH Ln3+ Pyridine
Stimuliresponsivity N-oxide
Thermal expansion
Gas adsorption
Unmasking CO2
Removal Electrochemistry
Cr2O2 7 Li-storage
Strong donor(ammonia) (pyridinium)acceptor interaction p-poor(pyridinium) (benzene)p-rich interaction
[501] [501] [502,503] [487] [492] [492] [492] [492] [504] [505] [506,507] [508] [508] [506,508,509] [510]
Quadrupole-dipole interaction between gas molecule and N-quarterized center Through multiple host-guest interaction like electrostatic interactions, hydrogen bonding and p-p stacking
[155,504] [511] [512,513] [514]
Electrsotatic interaction between anionic dye and cationic pyridinium ring Strong interaction between polarity of Lewis basic guests and cationic pyridine ring
[514,515] [516] [517,518] [518] [519] Contrast agents for in vivo magnetic resonance imaging
[520]
Hydroxy(H) (O)CO2 hydrogen bonding Hydroxy(O) (C)CO2 donor–acceptor interactions Hydroxy(H) (O)CO2 hydrogen bonding Dipole(hydroxy) quadrupole(H2) interaction Hydroxy(O) Hg(II) coordination interaction Multiple host guest interactions like electrostatic and hydrogen bond
[114,301,521-530]
Hydroxy(O) Fe(III) coordination interaction Different binding interaction of hydroxy function with electron donor/ acceptor guest solvents Hydroxy nitroaromatic donor–acceptor interaction Hydroxy(H) (NO2) nitroaromatic hydrogen bonding Hydroxy(OH) (OH)phenolic hydrogen bonding Hydroxy nitroaromatic donor–acceptor interaction Hydroxy(H) (NO2) nitroaromatic hydrogen bonding Providing catalytic site (OCu+) for activation of CO2 through Cu+-H+ exchange of hydroxylic proton Proton migration by water molecules adsorbed through hydrogen bonding in hydroxy rich channels Flexibility and polarity of alkyl ether groups Electrostatic interaction between quadrupole guest and ionic Li+-crown ether centers Hydrogen bonding between H+/OH spices and etheric oxygens Sensitizing of framework to [Ln(H2O)8]3+ in dual emission form through ether(O) (H)H2O hydrogen bonding Tuning the dispersion and exfoliation in aqueous solution by alkyl ether linkers Increased thermal motions (vibrations, rotations) of the alkyl ether side chains with rising temperature Photo- responsivity of 2-nitrobenzyl ether group in side chain N-oxide(O) (C)CO2 donor–acceptor interaction Dipole(N-oxide) quadrupole(CO2) interaction N-oxide(N) (O)dichromate electrostatic interaction Li+ ions (O)N-oxide electrostatic interaction
[530] [529] [531] [532] [533] [534,535] [536] [535] [535] [535] [535] [535] [537] [538] [539-547] [548,549] [550] [551] [552] [553] [554] [555,556] [557,558] [559] (continued on next page)
10
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Table 1 (continued) Functional group
Application
Details of application
Host-guest interaction and mechanisms
Ref.
Azoxy Oxadiazole Thiol
Removal Gas adsorption Removal
Azoxy(O) metal ion electrostatic interaction Oxadiazole(dipole) (quadrupole) CO2 interaction Formation of (SI) spices through interaction between (SH) and I2 Chelation between metal ion and thiol group through soft acid-soft base interaction
Sulfide
Sensing Gas adsorption
Pb2+ CO2 Iodine Hg2+ Cd2+ Pb2+ Hg2+ CO2
[560] [561] [562] [563-569] [570] [570] [571,572] [573,574]
Sensing
Sulfonate
2+
Conductivity Gas adsorption
Hg Pd2+ DMF DMSO Benzaldehyde NMP Biginelli Esterification CO2 conversion H+ CO2
Photocatalyst
H2 Esterification
Catalyst
Removal
Separation
Quinoline Benzothiophene Methylene blue Malachite green Ba2+ Cr3+ Fe3+ H2O TNP C2H2 from C2H6
Thiadazole
Gas adsorption Sensing
CO2 Cd2+
Thiourea Thiocathecole Fluorine
Catalyst Catalyst Gas adsorption
Morita-Baylis-Hillmar C–H functionalization CO2 CH4 H2
Sensing
O2 Kr C2-C3-C4 hydrocarbons
Chlorofluorocarbons Chlorine
Gas adsorption
Bromine
Gas adsorption Sensing Conductivity
Phosphonate
Sensing
Nitro
Gas adsorption Gas adsorption
CO2 CH4 H2 CO2 Hg(II) H+ Nitrobenzene 4-nitrophenol 2,4,6-trinitrophenol 2,4-dinitrophenol 4-nitrotoluene CO2 CO2
Sulfur containing functions provide strong gas-binding site through S atom and C atom of CO2 Soft acid-soft base interaction between metal ions and S atom of sulfide function Donor-acceptor interaction between electron rich sulfide group and electron deficient analytes
Providing catalytic Brønsted acidic sites by (SO3H) function Activation and fixation of CO2 through Sulfonate(O) (C)CO2 interaction Through Brønsted acidic and hydrophilic channels provided by (SO3H) Sulfonate(O) (C)CO2 donor–acceptor interaction Sulfonate CO2 dipole-quadrupole interaction H2(H) (O)sulfonate improved polarization interaction Photo-induced radical generation and electron transfer from sulfonate to metal oxide clusters Acid(SO3H) (analyte S,N atom)base interaction Electrostatic interaction between cationic dyes and anionic (SO 3 ) group Sulfonate(O). . . Barium chelation Interactions coordination interaction between sulfonate(O) and metal ions Hydrogen bonding between hydrogen acceptor sulfonate(O) site and hydrogen donor(OH) analytes + + + P-complexation between (SO 3 Ag ) and ethylene through H -Ag cation exchange Strong interaction between S atom of thiadazole and C atom of CO2 Soft acid-soft base interaction between metal ion and S atom of thiadazole function Thiourea hydrogen bonding site for carbonyl activation Anchoring site for Pd(II) ions High polarity/dipole moment of C–F bond to enhance interaction with gas molecules
Fluorine(F) (H–C)hydrocarbon hydrogen bonding High polarity/dipole moment of C–F bond to enhance interaction with gas molecules Highly polarized and fluorophilic pores and electron deficiency of aromatic core Enhanced polarizability of gases molecules through polar (C–Cl) bond
Enhanced dipole(C–Br) (CO2)quadrupole interaction Hg(II)-Br interaction Proton-transfer through Brønsted acidic pores provided by (PO3H) group Sensitizing of metal centers by acting as light antenna group
Strong phosphonate CO2 interaction Strong NO2 CO2 interaction
for removal and sensing of a large variety of chemicals (Table 1). Janiak and coworkers reported that the urea-functionalized MOF, [Zn2(L1)2(bipy)] where H2L1 = [1,3]diazepine-3,9-dicarboxylic acid, exhibits an uptake of 10.9 mmol.g1 (41 wt%) of SO2 at 293 K and 1 bar. Compounds [Zn2(L1)2(bipy)] and [Zn2(L1)2(bpe)] adsorb 14.3 mmol.g1 (20 wt%) and 17.8 mmol.g1 (23 wt%) NH3, respectively (Fig. 6a-b) [49]. These high uptake values are traced
[575,576] [577] [578] [578] [578] [578] [579] [580,581] [582,583] [581,584-593] [104,154,594-596] [597] [598] [599] [599] [600] [600] [601] [602] [589] [603] [604] [605] [561] [606] [607] [608] [609-634] [635-637] [611,617,618,638640] [641] [642] [612,620,643-647]
[648] [649,650] [651] [651] [649,650] [652] [653-657] [658] [658] [658] [658] [658] [659–661] [333,649,662,663]
to the urea functionality and its hydrogen bonding interactions with the adsorbents (Fig. 6c). In case of anion recognition by FMOFs, it has been reported urea FMOFs have a very special tendency to strongly interact with sulfate oxoanion (Table 1). A Ni- coordination network, [Ni(EDPU)2(H2O)2][(SO4)(H2O)2](H2O)3.5(EtOH)0.4 which is assembeled form EDPU (EDPU = ethylenedi(m-pyridylurea)) as a bis-urea ligand, can selectively recognize and separate Y-shaped
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 4. Schematic chemistry of urea. (a) Hydrogen bonding interaction between urea and electron-rich species. (b) Aggregation and self-quenching of urea groups. (c) Separated urea groups inside the pore-walls of FMOFs.
and highly hydrophilic sulfate oxoanion through multiple hydrogen bonds from aquoes solution (Fig. 7) [55]. This framework can also selectively and efficiently separate sulfate from aquoes solutions containing F, Cl, Br, I, NO 3 , and ClO4 .The topology of the framework and accessibility of urea function inside the MOF pores as well as anion topologies require the construction of directed hydrogen bonds for strategic design of selective anion recognizer urea decorated MOFs. Generally two important factors affect the anion recognition by urea FMOFs: (I) anion topology, accessibility and direction of urea and (II) hydrogen bond strength between urea and oxoanions. Due to the bent nature of urea function and strong structuredirecting hydrogen bonds with other parts of the framework, introduction of urea function into the MOF pores has complicated effects on the structure, topology, stability, and flexibility of the MOFs. In order to examine the effects of introducing urea function as hydrogen bonding site on the topologies and structure of the MOFs, Forgan and coworkers synthesized compounds [Zn2(L1)2(P1)] with pcu topology as parent and [Zn(L1)(P3)] as daughter frameworks, where P1 = 4,40 -bipyridine, H2L1 = [1,10 -biphenyl]-4,40 -dicar boxylic acid and P3 = N,N0 -bis(4-pyridyl)urea (Fig. 8) [669]. The results showed that introduction of urea function lead to the construction of a different framework with four-fold interpenetrated structure and dia topology. The authors explained that the bent nature of the urea function results in a non-linear disposition (calculated angle 160°) of the N-donors of the pyridyl units. Also, hydrogen bonds between the urea NH of each P3 ligand and a carboxylic O of L1 ligand act as two important structure-directing fac-
11
tors that make the solid-state interpenetrated dia topology more favorable. So, direct synthesis of urea functionalized MOFs with isoreticular synthesis strategy can be complicated because of bent nature of the urea and possible structure directing hydrogen bonds of urea containing linkers with each other or other parts of the struture. Urea functionalized MOFs show different behaviors upon activation or solvent exchange: (I) Collapse of the framework and loss of crystallinity (at least noticeably reduced crystallinity); (II) single crystal-single crystal transformation; (III) ability to be retained after activation. Such dynamic behavior of urea FMOFs can be happened mostly because of rotational freedom of urea around the C– N–C bonds and interruption in structure-directing hydrogen bonding of urea with other parts of the structure. But in some cases polycatenated urea decorated frameworks, like TMU-32, save their crystallinity with low-moderate porosity [59]. Here, three isostructure rht-type MOFs have been compared to evaluate the effects of introducing urea function on the stability of framework (Fig. 9) [42,260,670]. Triazole functionalized NTU-105 framework (Cu3(L1)(H2O)3∙11DMF8H2O), which is synthesized by H6L1 linker (5,50 ,500 -(4,40 ,400 -(benzene-1,3,5-triyl)tris(1H-1,2,3-t riazole-4,1-diyl))triisophthalic acid), is highly porous to N2(861 cm3/g, BET (m2.g1) = 3543) and CO2(187 cm3.g1). Replacement of one urea group inside the H6L1 linker, leads to the construction of Cu-UBTA framework (Cu3(L2)(H2O)3∙10DMF9H2O) based on the unsymmetrical H6L2 hexacarboxylic acid linker (5,50 -(4,40 -(5-(3-(3,5-dicarboxyphenyl)ureido)-1,3-pheny lene)bis(1H-1,2,3-triazol-4,1-diyl))diisophthalic acid). Cu-UBTA framework is still porous but with reduced BET surface area (3134 m2.g1) and adsorption capacity (805 cm3.g1 for N2 and 165 cm3.g1 for CO2). In addition, replacement of three urea groups inside the H6L1 linker, leads to the construction of Cu-TUH (Cu3(L3)(H2O)3∙5DMF10H2O) framework based on the symmetrical H6L3 hexacarboxylic acid linker (5,50 ,500 -(((benzene-1,3,5-triyl tris(azanediyl))tris(carbonyl))tris-(azanediyl))-triisophthalic acid). The results reveal that porosity (BET (m2.g1) = 68) and adsorption capacity (40 cm3.g1 for N2 and 37 cm3.g1 for CO2) for Cu-TUH is completely lost. PXRD pattern of Cu-TUH framework after activation indicates the instability and collapse of the framework after removal of guest solvents. The authors attributed this observation to the fact that the ligand H6L3, containing three urea functions, is more flexible than other hexa-carboxylate linkers (H6L1 and H6L2) in rht-MOFs. In another work, upon exchange of DMF molecules with CS2 from [Zn4O(L)3(DMF)2]n (where H2L is N,N0 -bis(4-carboxyphenyl)
Fig. 5. Structure, applied ligands in synthesis of NU-601 (a) Application of NU-601 as hydrogen bond organocatalyst in alkylation between nitro-alkenes and pyrroles (b). Reproduced with permission from Ref. [43]. Copyright 2012 American Chemical Society.
12
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 6. Structural representation of [Zn2(L1)2(bipy)] (a) and [Zn2(L1)2(bpe)] (b). Hydrogen bonding between urea group of H2L1 and SO2 and NH3 guests (c). Reprinted with permission from Ref. [49]. Copyright 2012 American Chemical Society.
Fig. 7. Structure of EDPU ligand (a), crystal structure of [Ni(EDPU)2(H2O)2][(SO4)(H2O)2](H2O)3.5(EtOH)0.4 (b) and Interaction of sulfate via eight hydrogen bonds from four urea groups (c). Reprinted with permission from Ref. [54]. Copyright 2007 Royal Society of Chemistry.
urea) framework, the propensity of [Zn4O(RCO2)6] cluster toward solvation enables the urea linkers to adopt distorted conformations as the MOF breathes, even facilitating rotation from the trans/trans to the trans/cis conformation without compromising the overall topology(Fig. 10) [671]. Moreover, it is observed that some of urea units, which show dramatic configurational rearrangement from the trans/trans to the trans/cis conformation, lose their capability as binding guests for organocatalysis. It seems that the polycatenated chains and hydrogen bond interactions between urea functions with other motives of the framework and/or solvent molecules like DMF have a very crucial role in the stability and direction of the framework [672]. Through activation of the MOFs, such hydrogen bonds can be interrupted leading to the collapse of the structure and reduction in crystallinity. Moreover, with interrupting such structure-directing/sta bilizing hydrogen bonds, freedom of urea function for rotation around C-N can be increased, causing easier trans/trans to trans/cis conformation change in urea function. In these circumstances, the flexibility of the inorganic SBUs and catenation of the structure are of importance. Ability of the ligand to distort, combined with the labile coordination chemistry of the inorganic SBUs, may allow the MOF to undergo structural transformation or negligible loss of crystallinity and porosity, but in case of rigid inorganic SBUs, the framework expect to be collapsed. However, based on the reported PXRD patterns of urea functionalized MOFs, this group of functionalized MOFs did not present good stability and rigidity during activation and gas adsorption measurements in most cases.
In some cases like TMU-32 ([Zn(oba)(L2)]2DMFH2O (H2oba: 4,40 oxidibenzoicacid; L2 = 1,3-di(pyridin-4-yl)urea), the polycatenated framework and strong structure-directing hydrogen bonds are reasons of saved crystallinity of the structure after activation and microporosity measurements (BET (m2.g1) = 432) [59]. 2.1.2. Amide and oxalamide Due to different guest-interactive sites, amide is an attractive function for providing different types of host-guest interactions. Because of different characteristics of amide such as: (I) high polarizing ability and strong electrical field of carbonyl group for dipoledipole/quadrupole interactions, (II) participating in hydrogen bonding through (–NH) site as donor and (–CO–) site as acceptor, and (III) electron donating and Lewis basicity through electron rich oxygen of carbonyl, amide is a candidate functional group for strengthening the host-guest chemistry of amide FMOFs. Based on review of published articles, amide FMOFs have mostly been applied as polar and stable hosts for improvement of gas-framework interaction especially CO2 and C2H2 (Table 1). Amide can improve CO2-framework interactions remarkably and thus it can increase CO2 capacity and selectivity. Improved CO2amide interaction is attributed to amide(C@O) C(CO2) Lewis base-acid and amide(NH) O(CO2) hydrogen bond interactions (Fig. 11a). On the other hand, amide functionalized frameworks show low affinity to non-polar/quadrupolar CH4 and N2 molecules. Therefore, due to the synergic combination of much greater quadruple moment of CO2 compared with nonpolar N2 and CH4
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
13
Fig. 8. Effects of urea function in the main-chain on the structure of MOFs. (a) Structure of H2L1, P1 and P3 ligands. (b) [Zn2(L1)2(P1)] framework as parent MOF with pcu topology. (c) [Zn2(L1)2(P3)] framework with dia topology. (d) Four-fold interpenetrated urea decorated daughter framework. (e) Structure-directing interaction between urea NH of each P3 ligand and a carboxylic O of L1 ligand.
Fig. 9. Representation of urea and triazole functionalized ligands for construction of rht- type MOFs.
along with large dipole moment of amide function, high selectivity of CO2 over CH4 and N2 is provided by amide FMOFs. In case of C2H2, strong interactions including amide(C@O) (H)C2H2, amide (NH) (p)C2H2 hydrogen bonds and amide(C@O) (p)C2H2 polarization interactions are the reasons for improved affinity of amide FMOFs toward C2H2 (Fig. 11b). Schroder and coworkers designed MFM-188 ([Cu4L(H2O)4] ∙12H2O) containing a tetra-amide octacarboxylate linker (H8L, 5,50 ,500 ,5000 -([1,10 -biphenyl]-3,30 ,5,50 -tetracarbonyl)tetrakis (azanediyl)tetraisophthalic acid) with a high surface area (BET = 2568 m2. g1) [84]. The uptake of CO2 and C2H2 by activated MFM-188a was measured up to 1 bar, and in both cases the isotherms showed fully reversible adsorption with a CO2 uptake of 120 cm3.g1 (23.7 wt% or 86.7 v/v) recorded at 298 K and a C2H2 uptake of 232 cm3.g1 (27.0 wt% or 166.7 v/v) at 295 K. The Qst values at zero coverage were 21.0 and 32.5 kJ.mol1 for CO2 and C2H2, respectively. Neutron diffraction and inelastic neutron scattering studies show that C2D2 forms H-bonds as H-acceptor via its p-electrons to the N–H
groups which are pointing into the pores. Likewise, CO2 participates in H-bond interaction as an acceptor from N-H groups. Many other amide functionalized MOFs have been applied in CO2 adsorption and separation (Table 1). All of them clearly show that introduction of amide function inside the MOF structures can be as effective as taking advantage of Lewis basic sites like amine (–NH2) and azine (@N–N@) and polar functions like sulfune (–SO3) and hydroxyl (–OH). Interestingly, another work reported that amide function did not show positive effects on CO2 adsorption and there was no direct binding between the adsorbed CO2 and CH4 molecules and the pendant amide group in the pore [81]. Clearly, it can be understood that not only the presence of functional group, but also accessibility of functions in the cavities and effective orientation of functional groups as guest-interactive sites inside the pores are of special importance for improvement of guest-framework host-guest chemistry. In another work by Space and coworkers, it is reported that presence of polar amide groups inside the structure of Cu-TPBTM
14
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 10. Structural change of [Zn4O(L)3(DMF)2]n after activation by CS2.
Fig. 11. Possible interactions between amide function and gas molecules. (a) Amide-carbon dioxide interaction. (b) Amide-acetylene interactions.
Fig. 12. Effect of introducing amide function in the main-chain of the MOF structure. (a) Representation of H2L1, H2L2, P1 and P2 ligands, (b) [Zn2(L1)2(P1)] framework as parent MOF with pcu topology. (c) [Zn2(L1)2(P2)] showing disposition of the SBUs induced by the amide pillar and [Zn2(L2)2(P1)] daughter frameworks which have a more regular arrangement of SBUs compared to [Zn2(L1)2(P1)].
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
15
Fig. 13. Effects of conformation of amide-function on the constructed compound. (a) Two-fold interpenetrated metal-organic framework based on amide ligands with trans– trans conformation, (b) metal-organic cluster based on amide ligands with cis-cis conformation. Reproduced with permission from Ref. [673]. Copyright 2013Royal Society of Chemistry.
(with formula [Cu3(tbptm6)(H2O)8] where H6tpbtm is N,N0 ,N00 -tris (isophthalyl)-1,3,5-benzenetricarboxamide) indirectly increases the low-pressure physisorption of hydrogen [86]. Hydrogen molecules bind onto the interior coordinatively unsaturated Cu2+ ions at low loading in the structure of Cu-TPBTM and PCN-61(with formula [Cu(H2O)]3(btei)∙5DMF4H2O where H6btei is 5,50 ,500 -ben zene-1,3,5-triyltris(1-ethynyl-2-isophthalate)). Presence of the negatively charged oxygen atom of the amide group in the structure of tpbtm6 hexacarboxylate ligand causes the interior Cu2+ ions to exhibit a higher positive charge through an inductive effect. As a result the H2-Cu2+ interactions are strengthened in Cu-TPBTM framework rather PCN-61. Multi interactive characteristics of amide provide good potential for FMOFs to optimize host-guest interactions for sensing and removal of different types of guests through different interac-
tions like hydrogen bonding, electron donor–acceptor, coordination, and polarization interaction (Table 1). ([Cd(4btapa)2(NO3)2]6H2O2DMF (1), 4-btapa = 1,3,5-benzene tricarboxylic acid tris[N-(4-pyridyl)amide])) in its active form (1a) has been applied as chemical separator for discriminate between nbutanol, n-pentane and n-pentene with the same size and shape [87]. Unique and selective response toward n-butanol was achieved because of hydrogen bonding between amide functions inside the 1a framework and hydroxyl of n-butanol. In another work by our group, it is reported that considering the higher basicity of amide compared to imine, amide functionalized frameworks have higher affinity toward heavy metal ions like Co2+, Cr3+, Cd2+, Cu2+, Fe3+, and especially Pb2+ [90]. As a result of the preferred interaction between amide and oxophil lead(II) cation, selective removal with higher efficiency is achieved.
Fig. 14. (a) Non-functional and amide-functionalized ligands for construction of Nbo-type MOFs. (b) Representation of crystal structure with NbO topology of NJU-Bai-17. Fig. 14b: Reproduced with permission from Ref. [85]. Copyright 2016 Royal Society of Chemistry.
16
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 15. (a) Non-functional and amide-functionalized ligands for construction of rht-type MOFs. (b) Representation of crystal structure with rht topology of Cu-TPBTM. Fig. 15b: Reproduced with permission from Ref. [80]. Copyright 2011 American Chemical Society.
Fig. 16. Structural representation of NOTT-125 and possible interactions between oxalamide and gas molecules and related distances. (a) Structure of H4bdpo ligand (left) and fof topology of NOTT-125 (right). (b) Oxalamide-carbon dioxide interactions which are not possible with single amide. (c) Oxalamide-methane interactions. Reproduced with permission from Ref. [96]. Copyright 2010 Wiley-VCH.
The oxygen atom of carbonyl motif of amide group is electron rich and can act as Lewis basic site. So, it is possible to apply amide functionalized MOFs in Lewis base catalyzed reactions (Table 1). Amide functionalized MOFs, [Cd(4-btapa)2(NO3)2]6H2O2DMF (where 4-btapa is 1,3,5-benzene tricarboxylic acid tris[N-(4pyridyl)amide)) and [Zn2(oba)2(bpta)]∙DMF3 (TMU-22, where H2oba = 4,40 -oxybis(benzoic acid) and bpta = N,N0 -bis(4-pyridyl)terephthalamide)) show good basic catalytic activity in Knoevenagel reaction with high yields [87,88]. Study of the structure of amide functionalized MOFs shows that, the presence of amide function between or instead of phenyl ring of organic linkers does not change the structure and topology of the framework. Although because of the cis-trans conformation exchange of amide function, the MOF framework tolerates some distortions and requires some flexibility, the overall structure pattern is retained. In addition, PXRD patterns of amide functionalized MOFs after activation are similar to synthesized MOFs revealing that amide functionalized MOFs are stable after solvent removal. Because of their polarity and stability, they are widely used in gas adsorption applications.
As mentioned earlier, Forgan and coworkers synthesized compound [Zn2(L1)2(P1)] with pcu topology (Fig. 12). [669]. By replacing non-functionalized ligands with amide functionalized ligands, [Zn(L1)(P2)] and [Zn(L2)(P1)] (where H2L2 = 4-(4carboxybezamido)benzoic acid, P1 = 4,40 -bipyridine and P2 = N-(pyridin-4-yl)-isonicotinamide) were synthesized. Findings indicate that, in contrast to urea, introduction of amide function leads to construction functionalized structures similar to the non-functionalized ones. Although some distortions such as changes in orientation of the SBUs with respect to one another are introduced in the structures, the overall structure is retained. It has been reported that cis-confromation of a tritopic amide containing ligand (H2L2, N,N0 ,N00 -methyl-4,40 ,400 -[1,3,5-benzene triyltris(carbonylimino)]trisbenzoic acid) leads to the formation of metal-organic cluster with formula [Ni14(m3OH)8(L2)6(formate)2(DMF)10(H2O)2] (Fig. 13a); however, applying similar ligands (H3L1, 4,40 ,400 -[1,3,5-benzenetriyltris(carbonyli mino)]trisbenzoic acid) without the methyl group on NH group with trans-conformation of amide functions results in the construction of extended three dimensional metal-organic
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
17
Fig. 17. (a) 3D framework of [Ln2(bdpo)1.5(DMA)3(H2O)]5H2O (b) H4bdpo to Ln metal ions energy transfer and unusually fluent change of luminescent color. Reprinted with permission from Ref. [98]. Copyright 2017 American Chemical Society.
Fig. 18. Significantly different characteristics between urea and squaramide.
frameworks ([Cu3(L1)2(H2O)3]∙14DMF16H2O) (Fig. 13b) [673]. This work shows that flexible amide ligands with conformational degrees of freedom could diversify the dimensionality of metalorganic materials and their net topology [673,674]. Here some examples about the effects of amide function on the structure of MOFs are presented. In well-known NbO topology category of MOFs including NOTT-100 (with formula [Cu2(L1)(H2O)2] where H4L1 is Biphenyl-3,30 ,5,50 -tetracarboxylic acid), NJU-Bai 17 (with formula [Cu2(dbai)(H2O)2]4DMF4H2O where H4dbai is (5-(3,5-dicarboxybenzamido)isophthalic acid), NOTT-101 (with formula [Cu2(L2)(H2O)2] where H4L2 is Terphenyl-3,300 ,5,500 -tetra carboxylic acid), and HNUST-1(with formula [Cu2(bdpt)(H2O)2] where H4bdpt is (bis(3,5-dicarboxyphenyl) terephthalamide) X-ray crystallography clearly shows that introduction of amide function between or instead of phenyl ring of organic linkers does not change the ligand geometry and topology (Fig. 14) [76,85,675]. Same results are achieved by studying a series of rht-topology MOFs, NOTT-112 (with formula [Cu3(L)(H2O)3)]∙8DMSO∙15DMF3 H2O where H6L is 1,3,5-tris(1,3,5-dicarboxy[1,10 -biphenyl]-4-yl)b enzene), PCN-61, and Cu- TPBTM (Fig. 15) [80]. Similar to amide, oxalamide (–NH–CO–CO–NH–) function is also highly polar with two different interactive (–CO–) and (– NH–) sites. Therefore, similar to amide, host-guest chemistry of oxalamide functionalized MOFs are based on hydrogen bonding, donor-acceptor, polarization, and dipole-quadrupole interactions. Schröder and coworkers reported [Cu2(H2O)2(bdpo)]4H2O2DMA, denoted as NOTT-125, where H4bdpo is N,N0 -bis(3,5-dic arboxyphenyl)-oxalamide, as the first oxalamide functionalized MOF with fof topology (Fig. 16a) [96]. Reported BET surface area (m2.g1) and pore volume (cm3.g1) for NOTT-125 are 2471 and 1.1, respectively. CO2 uptake (wt%) at 1 bar, 273 and 298 K are 40.0 and 18.19, respectively [96]. The excellent uptake of CO2
observed in low pressure isotherms observed because of presence of specific CO2–oxalamide interactions. The isosteric heat of CO2 adsorption is estimated to be 25.35 kJ.mol1 at zero coverage. Computational study of NOTT-125 reveals a set of strong interactions between CO2 and the oxalamide motif which are not possible with a single amide in a way that one CO2 molecule interacts with both amide motifs through simultaneous hydrogen bond CO2(O) (NH)amide(1) and dipole-quadrupole CO2(C) (O)amide(2) interactions (Fig. 16b). In contrast to the CO2-oxalamide interactions, the binding identified between NOTT-125 and CH4 is dependent on the presence of only one of the amide groups of the linker (Fig. 16c). Wang and coworkers reported three lanthanide-based oxalamide functionalized MOFs, [Ln2(bdpo)1.5(DMA)3(H2O)]5H2O where Ln = Eu, Gd and Tb and H4bdpo is N,N0 -bis(3,5-dicarboxyphe nyl)-oxalamide (Fig. 16a-Left), showing highly efficient energy transfer and unusually fluent change of luminescent color [98]. The authors claimed such efficient energy transfer to originate from the transfer of energy from tetracarboxylate bdpo4 ligand to Ln metal center (Fig. 17). They have chosen H4bdpo because it contains oxalamide function with a good planarity and p-conjugated system that helps to form a large delocalized p-electron conjugated system inside the framework as a strongly absorbing chromophore. Calculations have confirmed H4bdpo as a sensitizer efficiently absorbing and transferring energy to lanthanide ions. 2.1.3. Squaramide With the strong hydrogen bond donor ability from its two (–NH–) sites and the remarkable hydrogen bond accepting characteristic from its two (–CO–) sites, squaramide is applied in supramolecular chemistry as a new binding site for interaction with negatively (partially) charged atoms as artificial anion receptor in molecular recognition and powerful bifunctional hydrogen bonding catalyst in asymmetric organocatalysis [676–678]. Compared to urea as its closest analogues among the abovementioned hydrogen bonding functional groups, the squaramide functionality differs significantly in various aspects especially in higher Brønsted acidity, structural rigidity, and duality in molecular recognition for anions and cations (Fig. 18) [679]. Successful transfer of such properties of squaramide into the MOFs offers opportunities to simultaneously take advantage of both heterogeneous nature of MOFs and chemical characteristics of squaramide. However, despite considerable progress in the area of urea FMOFs, generation of a squaramide structural motif as part of the MOF catalyst has attracted much less attention. Squaramide functionalized MOFs have been applied as active hydrogen bonding organocatalysts for the Friedel–Crafts reaction and as sensors for detection of lactose. Hupp and coworkers
18
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 19. Application of UiO-67-Squar/bpdc in Friedel–Crafts reaction. Reproduced with permission from Ref. [99]. Copyright 2015 American Chemical Society.
Fig. 20. Representation of non-functional and squaramide functionalized tetratopic ligand which is applied for construction of isostructural Nbo type MOFs.
reported the first squaramide functionalized UiO-67 MOF (Fig. 19) [99]. They have demonstrated that incorporation of this moiety into a porous UiO-67 metal-organic framework derivative leads to dramatic acceleration of the biorelevant Friedel-Crafts reaction between indole and b-nitrostyrene (78% yield for UiO-67-Squar/ bpdc and 38% yield for UiO-67-Urea/bpdc). Since the activity of hydrogen bond donor functions increases as a function of acidity, it is possible to conclude that in order to design FMOFs as efficient organocatalysts and molecular recognizer groups, introduction of squaramide with more acidity requires greater consideration because of the stronger interaction with negatively (partially) charged atoms. In comparison of H4dbda ((5,50 -(3,4-dioxocylcobut-1-ene-1,2-diyl)bis(azanediyl)dii sophthalic acid) and H4tptc (p-terphenyl-3,5,300 ,500 -tetracarboxylic acid) ligands, since the presence of squaramide in main-chain of ligands does not change the orientation of carboxylate coordinating sites relative to each other (Fig. 20), the structure and topology of the related NbO-MOFs, [Cu2(dbda)(CH3OH)2] and [Cu2(tptc) (H2O)2], are retained [100].
2.1.4. Carbonyl Because of dipole moment and high electron density on oxygen atom, carbonyl groups are highly polarize Lewis basic sites and can interact with guest molecules through dipolar interaction. In structural view, carbonyl is completely rigid without any possible conformation rotation. Accordingly, introduction of carbonyl function inside the MOF structures provides a great strategy to synthesize polar and rigid frameworks for improved host-guest interactions and structural permanency. However compared to amide, carbonyl FMOFs have received less attention. Study of carbonyl functionalized MOFs has shown that carbonyl functionalized MOFs are mostly synthesized using fluorenone based ligands which are well known motifs for their responsivity to polar solvents and photon radiation owing to their labile electronic states. Functionalization with polar carbonyl is good for interaction with electron deficient, Lewis acidic, and hydrogen bond donor guests (Table 1). In case of fluorene based MOFs, they have been applied for colorimetric detection of organic sensing (Table 1). Zhou and coworkers reported the first single-lanthanide ratiometric luminescent thermometer MOF (Me2NH2)3[Ln3(fdc)4 (NO3)4]4H2O where (Ln = Eu, Gd, Tb and H2fdc = 9-fluorenone2,7-dicarboxylic acid) [109]. Time-resolved luminescence studies show that in Eu-MOF the energy difference between the H2fdc triplet’s excited state and the 5D0 of Eu3+ level is small enough to allow for a strong thermally activated ion-to-ligand back energy transfer. The Eu3+ to the ligand triplet emission intensity ratio allows measuring the temperature in the 12–320 K range, with a relative thermal sensitivity of up to 2.7% K1 at 170 K (0.33% K1 at 300 K) and a repeatability of up to 96%. Kaskel and coworkers reported a carbonyl decorated MOF with fluorenone motif, [Zr6O4(OH)4(fdc)6)0.1BA7.6H2O] denoted as DUT-122 where BA = benzoic acid and H2fdc = 9-fluorenone-2,7dicarboxylic acid, which shows colorimetric guest-responsive fluorescent properties in presence of saturated vapors of different organic solvents like protic, aprotic, polar and non-polar, aromatic and aliphatic ones (Fig. 21) [106]. This observation of solvent responsive from DUT-122 framework is based on sensitivity of carbonyl functionalized fluorenone motifs to different solvents. In addition to framework polarity improvement, introduction of carbonyl functional group has very distinctive effects on rigidity of
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
19
Fig. 21. Application of DUT-122 in organic solvent sensing. (a) Structure of DUT-122. (b) Colorimetric sensing of guests molecules with related wavelength. Reproduced with permission from Ref. [106]. Copyright 2016 Wiley-VCH.
Fig. 22. Carbonylation of sp3 atoms for construction of rigid and polar carbonyl functionalized MOFs.
the structure. Actually, the rigidity of the ligands are increased by replacing the sp3 methylene moiety with a polar sp2 (–CO–) group. Fig. 22 Represent the carbonylation strategy based on H2fdc (fluorene-2,7-dicarboxylic acid) and H4mdip (5,50 -methylene-di-is ophthalic acid) flexible ligands to gain H2ofdc (9-oxofluorenone-2 ,7-dicarboxylic acid) and H4cdip (5,50 -carbonyldiisophthalic acid) rigid ligands, respectively. Matsuda and coworkers reported two isostructure MOFs, LMOF-201 ([Zn2(ofdc)2(bipy)]) and LMOF-202 ([Zn2(fdc)2(bipy)]); where LMOF-202 contains uncoordinated carbonyl oxygen atoms from ofdc2 dicarboxylate ligands [103]. This slight difference between the two frameworks results in the flexibility and rigidity of LMOF-201 and LMOF-202, respectively. For rigid carbonyl decorated framework, LMOF-202, CO2 adsorption isotherm at 195 K was typical type-I isotherm, whereas for flexible non-functionalized framework, LMOF-201, adsorption isotherm was gate opening type that had a large hysteresis revealing the structural change of LMOF-201 during adsorption measurements. Also, In Situ PXRD pattern of synthesized and activated samples of LMOF-202 are identical, because of LMOF-202’s rigidity. However, in case of LMOF-201, PXRD pattern for guest free sample clearly differs from that of the as-synthesized sample because flexibility of the framework leads to structural change during desolav-
tion. Moreover, a flexible structural change could clearly be observed for LMOF-201 by an in-situ coincident PXRD/adsorption measurement. After discussing urea, amide, ketone, and squaramide we can summarize our conclusions in Fig. 23. Urea is mostly unstable upon activation with strong H bond donor/acceptor sites. Therefore, it has been frequently applied in H-bond organocatalysis of reactions as well as removal and sensing of guests with electron rich motifs like nitro-compounds and oxoanions in liquid phase. However, ketone functionalized MOFs are rigid and only accepting H-bonds. So, ketone functionalized MOFs have been applied for removal and sensing of electron deficient guests and robust and polar platforms for CO2 adsorption. But similar to urea decorated MOFs, amide functionalized MOFs have two (–NH–) and (–CO–) interactive sites and similar to ketone decorated MOFs amide functionalized MOFs are stable upon activation. As a result of the stability and dualinteractive features of amide, amide functionalized MOFs have been frequently applied in CO2 adsorption. Moreover, maybe because of the rigidity of squaramide aromatic ring unlike urea, squaramide functionalized MOFs show good stability after activation and gas adsorption measurements as well as better control over the position of the squaramide group within the pore structure.
20
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 23. Comparison between carbonyl based functions.
Fig. 24. Structure of imide functionalized aromatic cores.
2.1.5. Imide Imide function (–CO–NR–CO–) has particularly been applied in the design of aromatic functionalized molecules with phenyl, naphthalene, and perylene cores (Fig. 24). Desirable electronic and spectroscopic characteristics and better structural properties of naphthalene-diimides (NDI) over pyromellitic-diimides (PyDI) and perylene-diimide (PDI) helps NDI to be considered as an ideal component for architecting functional materials for both covalent and non-covalent systems in supramolecular chemistry. Naphthalene-diimide is neutral, conjugated planar, thermally stable, and chemically robust with a high melting point incorporating various and fully practical chemical features like high electron deficiency and so anion-p, p(deficient)-p(rich) and donor-acceptor interactions, high charge carrier mobility and so high conductivity, special electron transfer and so oxidative properties and active electronic states [680]. NDI containing materials are among the most interdisciplinary fields in material and supramolecular science because of its diverse noncovalent interactions including
hydrogen bonding, p-p stacking, p-ion interactions, and phalogen interactions [681]. In addition to control over host-guest interactions of NDI, this p-deficient characteristic and p-p interactions of NDI can be utilized to implement strong directional interactions for crystal engineering of naphthalene-diimide functionalized materials. Combination of such abilities of NDI core with the arranged and ordered structure of MOFs is beneficial to create optimal properties and new applications through construction of structures which do not suffer from structural defects. Li-storage occurring through electrochemical mechanism, combination of porosity and stability of MOFs along with redox activity of NDI core has led to rational designing FMOF-based lithium batteries. [Cd(ClO4)2(DPNDI)2](DMA)4.5(H2O)2 (2) where DPNDI = N0 , N00 -di(4-pyridyl)-1,4,5,8-naphthalenediimide, has been employed as cathode material for lithium-ion batteries [130]. The results of FT-IR and XPS analyses reveal that, instead of metal ions, only the organic linkers of DPNDI take part in the electrochemical process. The reversible conversion of the C@O bond in lithiation process and the C@C bond in delithiation process confirms the participation of diimide C@O bond of NDI group of DPNDI ligand in the electrochemical redox reaction. Photoactivity of aromatic-diimide functions along with heterogeneity of coordination polymers offer a practical strategy for development of photocatalysts toward the better exploitation of solar energy. Duan and coworkers developed a PDI functionalized coordination polymer, Zn-PDI (where H2PDI is bis(N-carboxyme thyl)peryleneimide), for visible-light-driven reduction of arylhalides through C–H and C–C bonds formation and photooxidation of alcohols to carbonyls and amines to imines (Fig. 25) [117]. NDI and PyDI functionalized MOFs are used is gas adsorption because: (I) p-deficient nature of these groups aid them in selective electrostatic/charge transfer interaction, (II) polarizing and high dipole moment carbonyl motifs are present, and (III) these
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
21
Fig. 25. Illustration of assembling insoluble PDI into organized arrays in Zn-PDI for synthesis of an efficient photocatalyst to reduce aryl halides and oxidize alcohols and amines under visible light. Reprinted with permission from Ref. [117]. Copyright 2016 American Chemical Society.
Fig. 26. (a) Structure of H2NDI ligand. (b) Single network of FJU-66 along a axis. (c) Multiple point interactions of the [EVIm]+ cation with the framework: hydrogen bonding d [O1(–imide) C39([EVIm]+)] = 3.127 Å and lone pair–p interaction d [N([EVIm]+ center of naphthyl)] = 2.838 Å. The minimum distance between OH anion and the [EVIm]+ cation, d [center of imidazolium O7(anion)] = 10.915 Å. Color code: Cu, bright green; C, gray; O, red; N, blue. For the sake of clarity Hydrogen atoms are omitted. Reprinted with permission from Ref. [131]. Copyright 2015 Royal Society of Chemistry.
functions are redox active and reductive doping of alkaline metals can enhance the columbic host-guest interactions. Due to such characteristics of NDI and PyDI, it is possible to improve the gasframework interaction especially for dipole or quadrupole targeted molecules like CO2. Ghosh and coworkers reported that electron lacking characteristic of [Zn(PBDA)(DPNI)]n framework (PBDA: 4,40 -((2-(tert-butyl)-1,4-phenylene)bis(oxy))dibenzoic acid; DPNI: N,N’-di-(4-pyridyl)-1,4,5,8-naphthalenetetracarboxydiimide) owing to presence of NDI function in the DPNI linker, expedite strong interactions with the polarized adsorbate gas CO2 (Qst = 32.1 kj.mol1) consequently leading to high CO2 selective adsorption at 0.2 bar (CO2/N2 = 189.4, CO2/H2 = 256.5 and CO2/CH4 = 12.3) [111]. Hupp and coworkers argued that framework reduction might boost adsorption through increasing the polarizability of the structure and so charge/quadrupole interactions and coulombically displace interwoven frameworks [116]. It is established that in addition to CO2 or H2 uptake capacity, framework reduction leads to improvements in gas adsorption parameters like isosteric heat of H2 and CO2 adsorption, CO2/N2 or CO2/CH4 selectivity [114]. p-Deficiency characteristic of NDI core is the reason of its participation in p(deficiency)-p(rich) interaction, donor acceptor and
anion-p interactions. As a result of such interactions, NDI functionalize MOFs have been applied as chemosensor for electron rich analytes and as a platform for hydroxide conductivity (Table 1). Xiang and coworkers developed a free OH anion-containing MOF with rationally tunable host-guest interactions and high OH conductivity at low OH concentrations by synthesizing NDI functionalized [Cu6(NDI)3]∙2DMF∙6MeOH2H2O (FJU-66 where H2NDI is 2,7-bis(3,5-dimethyl)dipyrazol-1,4,5,8-naphthalene-tetra carboxydiimide) with high affinity to the electron-rich organic amines (Fig. 26) [131]. The host-guest interaction between NDI functions of FJU-66 and imidazolium ring of 1- ethyl-3vinylimidazolium ([EVIm]+) creates free OH anions from the resistance of the cation–anion electrostatic force to enhance the OH conductivity. FJU-66[EVIm]OH with a low OH concentration of 0.34 mmol.g1 is a free OH anion-containing MOF exhibiting high OH conductivity close to 0.1 S.cm1 with the lowest activation energy Ea of 0.11 eV. The high conductivity in FJU-66[EVIm] OH can be attributed to multipoint supramolecular interactions between the robust framework and planar [EVIm]+ cations, which pulls [EVIm]+–OH pairs far apart and frees OH anions. The restraint-free OH anions significantly improve the OH conductivity.
22
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 27. Application of Mg-NDI in colorimetric detection of guests. (a) Solvatochromic behavior of Mg-NDI for organic solvents. (b) Mg-NDI color change in presence of different aromatic-substitutes. Reproduced with permission from Ref. [118]. Copyright 2015 Royal Society of Chemistry.
Mg-NDI framework constructed by H4BINDI ligand (N,N0 -bis(5isophthalic acid)naphthalenediimide) and Zn2(bdc)2(DPNDI) (H2bdc = 1,4-benzenedicarboxylic acid; DPNDI = N,N0 -di(4-pyri dyl)-1,4,5,8-naphthalenediimide) frameworks shows color change in presence of organic amines and benzene substitutes (Fig. 27) [118]. Mg-NDI is capable of sensing a diverse range of solvents in less than 60 s. Moreover, this MOF can selectively sense small sized organic amines by visual color change as well as quenching of its fluorescence efficiency. Electron rich organic amines can form charge transfer complex with the NDI moieties within the framework. NDI behaves as n-type semiconductor and generates stable radical anions upon exposure to external stimuli. Owing to such unique properties of NDI core, NDI functionalized MOFs have been applied in designing photochromic, electrochromic, conductive, and switchable FMOFs. NDI inside the structure of MOFs is sunlight sensitive and after a specific time for each MOF undergoes a color switching through photo-induced electron transfer in photochromic transformation. Photosensitivity and photochromism occurs because of formation of stable NDIradical anions. For almost all reported MOFs, placing a sample in a dark place without sunlight, after certain duration of time which is specific for each MOF, the color is turned to its as-synthesized sample. This phenomenon shows that the photochromic transformation is a reversible process. Incorporating NDI cores inside the extended structure of MOFs prevents the fast decolorization of NDI cores. Speed of reversible photochromic transformation inside the MOFs is related to the optimum and close crystal packing and sufficient orbital overlap of NDI cores through secondary interactions like p-p stacking inside the structure of MOFs. It is found (i) that the short life time of NDIradical anion is a result of short p-stacked NDI cores and (ii) orthogonal orientation of NDI cores leads to radical stability. Banerjee and coworkers reported the application of a series of photochromic NDI functionalized MOFs, Mg-NDI, Ca-NDI, and SrNDI including BINDI ligand (H4BINDI = N,N0 -bis(5-isophthalic acid)naphthalenediimide) for inkless and erasable printing [125]. They mentioned the resultant printing has a good resolution, stability, capability of being read by both human eyes and smart elec-
tronic devices while the paper can be reused for several cycles without any significant loss in intensity. The authors compared the capability of these three MOFs with bare H4BINDI ligand for erasable printing and found that radiated H4BINDI has a poor contrast compared to the non-radiated one, therefore making H4BINDI non suitable for inkless and erasable printing applications. However, all MOFs show good contrast and the capacity for different color printing with similar efficiency by varying the structure of the MOF. Based on high charge mobility and electrical activity of NDI core, Silva and coworkers reported a naphthalene diimide functionalized cobalt based pillared MOF, MOF-CoNDI-py-2, using N, N-bis(4-pyridyl)-1,4,5,8-naphthalene diimide (NDI-py) and Terphthalic acid (H2TpA) ligands, displaying a light induced anisotropic semiconductivity with wavelength dependent photoconductive-photoresistive behavior, with a high responsivity of 2.5 105 A.W1 [128]. They mentioned that conduction mechanism involves a charge transfer from the metal center to the strong p-acceptor NDI-py, promoting hole transport through the Co TpA2 direction while electrons are transported by the NDIpy direction. Excitation at the MLCT band further improves this mechanism by promoting a charge injection from metal to diimide. Such semiconductivity through NDIs is related to the ability of NDIs to undergo light-induced reduction, forming radical anions. Moreover, through high charge mobility and n-type characteristics of NDI, electrons can transport in NDIs direction. In other examples, introduction of aromatic diimide cores helped to improve the physical properties like turn-on fluorescence behavior of MIL-101(Al)–NH2 by post-synthesis modification with PDI groups and formation of excitonically coupled states by highly ordered p-p stacking interaction through parallel arrangement of NDI cores in crystalline structure of Zn-SURMOF-2 [682,683]. Because of direct connection between rigid naphthalene core and two imide functions, NDI functionalized MOFs show stability and permanent porosity upon activation and do not show function-based flexibility behaviors in their structural view. Moreover, XRD and ESR analyses before and after color switching in photo/electrochromic transformation reveal that photoresponsive behavior of NDI functionalized MOFs results from a chemical charge transfer and radical formation rather than a structural transformation in a way that as-synthesized sample shows no ESR peak while sunlight-treated one shows a ESR peak. Both sunlight-treated and as-synthesized samples exhibit the same PXRD pattern. 2.1.6. Carboxy Carboxy function is one of the favorite functions in construction of metal complexes since it can chelate/coordinate to metal ions through deprotonation and formation of metal-carboxylate complexes. On the other hand, carboxy function is hydrogen donor and hydrogen bond donor/acceptor site which can act as Brønsted acid and hydrogen bond participating site. Because of these characteristics, carboxy function is applied as both coordinating and guest-interactive site in the construction of FMOFs. Despite the fairly simple coordination modes of carboxylate and multi-carboxylate family of ligands, carboxylate [–CO 2 ], is one the functions that is widely applied in coordination chemistry to construct materials such as metal-carboxylate complexes, zero dimensional compounds and three-dimensional structures especially MOFs [684]. Carboxylate-based MOFs have remarkably high surface area and uniform pore size distribution, highly ordered crystallinity, well-defined reticular chemistry which is partly due to their prominent metal-carboxylate secondary building units, reversible formation and dissociation of metal-carboxylate bonds, and mild conditions of synthesis. However, their
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 28. Schematic representation of insertion of CO2 into the aryl-COOH bond within UiO-67(dcppy). Reprinted with permission from Ref. [151]. Copyright 2016 Wiley-VCH.
insufficient chemical stability in air and water, and even a kind of low thermal stability, poses a significant problem if they are to be used in industrial or commercial applications [685–689]. This is why a great number of carboxylate-based functions are constructed by non-oxophil metal centers like Zn2+, Cu2+, Co2+, and Ni2+ [690]. On the other hand, moderate pKa for carboxy (–COOH) and delocalized electron density and anionic charge of both oxygen atoms turns carboxylate into a less strongly coordinating site to the mentioned metal ions. However, there are some strategies to overcome the instability of carboxylate-based MOFs including (I) using oxophil metal ions to improve metal–carboxylate bond strength, [685,691] (II) functionalization through pre-synthetic or postsynthetic modification methods by tuning hydrophobicity of the pores, and [692] (III) using another auxiliary coordinating function along with carboxylate by changing metal-ligand coordination geometry and strength [693,694]. As a well-known and high surface area MOF, MOF-5 ([Zn4O (bdc)3]n or IR-MOF-1 where H2bdc is 1,4-benzenedicarboxylic acid) does not present sufficient tolerability in presence of water and humid air [695]. This is mostly owing to the weak Zn-(O)COO interaction because water molecules attack Zn4O metal clusters replacing carboxylate groups and finally making the structure collapse. To prevent such instability, Masel and coworkers substituted 1,4-benzenedicarboxylate ligand with water repellent groups like
23
trifluoromethoxy. [692]. Provided analysis showed that trifluoromethoxy groups play an important role in preventing water molecules from attacking (Zn4O) cores, therefore making the framework more resistant to the moisture in the air. However, the improved stability of the MOFs by such method is always at the price of a reduction in their microporosity and surface area or functionality. Another strategy is replacing Zn2+ metal ions with oxophil metal ion like Zr4+ which results in the construction of UiO-66 (Zr6O4(OH)4(bdc)6 (where H2bdc is benzenedicaroxylic acid) with a different structure and higher stability toward organic solvents, water, humid air and even acidic solutions. This stability is the result of hard-hard carboxylate coordination interaction with highly positive and oxophil Zr4+ centers. Due to the high charge density and bond polarization, there is a strong affinity between Zr4+ and carboxylate O atoms in most carboxylate-based ZrMOFs. For example, activated UMCM-309 (a 2D zirconium-based microporous coordination polymer derived from the tritopic linker 1,3,5-(4-carboxylphenyl)benzene possesses a Zr6(l3-O)4(l3OH)4(RCO2)6(OH)6(H2O)6 cluster with six hexagonal-planarcoordinated linkers) shows extraordinary stability in a way that can save its crystallinity for over four months in highly acidic 1 M HCl solution. However, Zr-based MOFs show less stability in basic aqueous solutions. [685]. The negative charge on O atom of hydroxide is higher than O atom of carboxylate due to charge localization in carboxylate, indicating that OH can form a stronger bond with Zr(IV) than a carboxylate O atom, leading to the decomposition of Zr-MOFs under basic conditions. This is a declaration about insufficient stability of carboxylate-based MOFs despite their crystallinity and porosity. Using auxiliary binding sites like azoles especially pyrazole is another useful strategy to improve chemical stability of MOFs which will be discussed in the section on azolates [693]. In some cases, the carboxy group remains uncoordinated to interact with guest molecules. Custelcean and coworkers reported that the key factors for the formation of this functional MOF appear to be the low pH and the aqueous conditions employed in the synthesis step, which prevent the COOH groups from deprotonation and coordination to metal ions [696]. Moreover, in some cases carboxy groups are introduced in the framework through postsynthesis modification (PSM) and solvent assisted ligand incorporation (SALI) methods [151,697,698]. Ma and coworkers successfully inserted carbon dioxide into aryl C–H bonds of the
Fig. 29. BUT-83 structure: (a) coordination mode of Co(II) atom and the applied ligand; (b) 1D channel in the structure; and (c) 3D framework structure viewed along the c axis (H atoms are omitted for clarity. Color code: Co, magenta; N, blue; C, black; and O, red). Reprinted with permission from Ref. [148]. Copyright 2017 Royal Society of Chemistry.
24
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
backbone of a MOF to generate free carboxylate groups, which serve as Brønsted acid sites for efficiently catalyzing the methanolysis of epoxides (Fig. 28) [151]. FMOFs containing pores decorated with uncoordinated (–COOH) group provide Brønsted acidic and hydrophilic channels which are suitable for proton conductivity and catalytic activity. Also, abilities like coordinating/chelating to metal ions, participating in hydrogen bond, and interaction with dipolar-quadrupolar molecules make carboxy FMOFs suitable for sensing, removal and separation of different type of guest molecules (Table 1). Li and coworkers synthesized BUT-83 ([Co(dcdpp)]∙5H2O where H2dcdpp is 5,15-di(4-carboxylphenyl)-10,20-di(4-pyridyl)por phyrin) as a new porphyrinic metal-organic framework which is surrounded by high-density non-coordinating carboxy groups (Fig. 29) [148]. BUT-83 exhibits a high proton conductivity of 3.9 102 S.cm1 at 80 °C and 97% relative humidity which is the highest among those of carboxy group functionalized MOFs including UiO-66-(COOH)2 (2.3 103 S.cm1 at 90 °C and 95%) and MIL-53(Fe)-(COOH)2 (2.0 106 S.cm1 at 25 °C and 95%). It was found that the carboxyl density of BUT-83 (0.026 mol.cm3) is much higher than those of UiO-66-(COOH)2 (0.003 mol.cm3) and MIL-53(Fe)-(COOH)2 (0.012 mol.cm3). The high density and proper arrangement of uncoordinated carboxy functions should be responsible for creating efficient pathways for proton transfer and the observed high proton conductivity. In order to evaluate the role of the (–COOH) group in enhancing the conductivity of BUT-83, the proton conductivity of an isostructural MOF with (–CH3) instead of (–COOH) group, was checked. At 25 °C and 97% relative humidity, the methylated framework shows a proton conductivity of 4.7 107 S.cm1, which is considerably lower than that of BUT-83 (5.0 103 S.cm1) under the same measurement conditions. Hupp and coworkers reported a novel analogue of UiO-66 modified with oxalic acid (UiO-66-ox) synthesized via solvent assisted ligand incorporation [149]. They mentioned that incorporation of the free carboxylic acid functional groups into UiO-66 allows for the removal of a broad range of toxic chemicals especially ammonia, sulfur dioxide, and nitrogen dioxide through physical and chemical interactions at levels greater than or equal to the base UiO-66. The increased NH3 removal capacity for UiO-66-ox (2.5 mmol.g1) rather than Uio-66 (2.5 mmol.g1) is due to an acid-base interaction with the carboxylic acid group, while the increased capacities for SO2 (0.1 and 0.8 mmol.g1) and NO2 (3.8 and 8.4 mmol.g1) removal are due to the reactive capabilities of carboxylic acidsNH3 can hydrogen bond to the carboxylic acid, increasing the physisorption capacity, but increased capacity of
Fig. 31. Higher accessibility and Lewis basicity of green-circuled N atoms lead to stronger interactions with CO2 inside the pores of functional MOFs.
UiO-66-ox for SO2 is due to a chemisorption interaction between the free carboxylic acid group and SO2. Also, carboxy FMOFs have been applied as anchoring platforms for encapsulation of enantiopure organic amine, Cu2O nanoparticles, and Trypsin catalysts for asymmetric direct aldol, Huisgen 1,3-dipolar cycloaddition, and biocatalyze proteomics analysis reactions, respectively [699–701]. 2.2. Nitrogen-based functions This major group of FMOFs consists of N atom as the only heteroatom in the skeleton of the function. Nitrogen- based functions are mostly employed in the structure of MOFs. Considering the chemical properties and function skeleton of this major group of FMOFs, nitrogen-based functions are classified in four categories: (I) heterocyclic azine N-based functions, (II) heterocyclic-azole Nbased functions, (III) noncyclic N-based functions, and (IV) ionic N-based functions. 2.2.1. Heterocyclic azine N-based functions In this group of nitrogen-containing functional groups, the (–CH) motifs in benzene ring are replaced by nitrogen atoms. This group includes pyridine, diazines (pyrimidine, pyrazine, and pyradazine), triazine (usually 1,3,5-triazine), and tetrazine (1,2,4,5tetrazine) and in some cases heptazine (tri-s-triazine). Considering the different number of electronegative N atoms in the aromatic ring of these functions, they show different chemical properties. Heterocyclic azine members, with low nitrogen number, including pyridine, diazines, and especially pyrazine mostly act as coordinating sites while members with high nitrogen content including
Fig. 30. Comparison of heterocyclic azine N-based fucntions.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
25
Fig. 32. Representation of tetratopic ligands and related Nbo type structures for methane storage applications. UTSA-76 with pyrimidine function shows higher storage and working capacity. (a) Non-functionalized NOTT-101. (b) Pyridine-functionalized ZJU-5. (c) Pyradazine-functionalized UTSA-75. (d) Pyrimidine-functionalized UTSA-76. Reproduced with permission from Ref. [174]. Copyright 2015 Royal Society of Chemistry.
triazine and tetrazine mostly act as guest-interactive sites. By increasing the number of electronegative nitrogen atom, the electron density of p-ring and N atom is increased in a way that the overall ring feels high electron deficiency (Fig. 30) [702]. Generally heterocyclic-azine FMOFs, owing to Lewis basic r-donating and p-accepting nitrogen atom(s), are used as adsorbent and sensor for hydrogen bond donors and metal ions, polarization of CO2, CH4, and C2 hydrocarbons, and Lewis basic catalysis (Table 1). It is good to mention that the nitrogen atoms of N-based functions are located in the aromatic skeleton of the framework rather than in the pores of MOFs. As a result, the porosity, free space, and pore volume of the framework is almost retained which is beneficial for gas adsorption measurements [172]. In case of triazine and tetrazine due to high p-deficiency, they are applied for sensing, removal, and separation of electron rich aromatics (Table 1). Lewis basic functionalized MOFs by pyridine group are applied in storage of gaseous molecules especially CO2 and C2H2 and CO2/ N2, CO2/ CH4 and C2H2/CH4 selective captures. Due to the Lewis basicity, pyridine ring can interact with acetylene molecules through hydrogen bonding and interact with carbon dioxide through the CO2(C) (N)pyridine interactions. Klopper and coworkers theoretically studied the interactions between CO2 and N-containing organic heterocycles, especially pyridine, diazines (pyrazine, pyrimidine and pyradazine), and triazine. [702]. In heterocyclic azine nitrogen-based functions, by increasing the number of nitrogen in the six-member benzene ring, electron deficiency increases on nitrogens and since the CO2(C) (N) heterocy-
cles binding energy is proportional to electron density on nitrogen atoms, the binding energy between the mentioned rings follows the trend: pyridine > pyrimidine, pyrazine, pyradazine > triazine. So, by increasing the number of nitrogen and increasing the electron deficiency on nitrogen atoms, the tendency of the functional group toward carbon dioxide molecules decreases. In another work, it is reported that gas adsorption properties of MOFs decorated with heterocyclic azine nitrogen-based functions depend not only on the number of Lewis basic nitrogen sites but also more importantly on their accessibility in a way that pyridine-N atoms which are pointed to the pores are completely available for occupation by CO2 molecules and can provide much stronger binding sites compared to open metal sites (Fig. 31) [164]. Moreover, bears open nitrogen atoms of bipyridine linkers readily react with a variety of solution- and gas-phase metal sources resulting in the formation of corresponding metalated frameworks like Pd2+ and Cr3+. Through this metal insertion mechanism, Pd2+ and Cr3+ have been used to catalyze Suzuki–Miyaura coupling and C–H borylation reactions, respectively [703,704]. Chen and coworkers synthesized a series of Lewis basic functionalized MOFs including pyridine functionalized ZJU-5 (with formula [Cu2(L2)(H2O)2]. where H2L2 is 5,50 -(pyridine-2,5-diyl) diisophthalic acid), pyradazine functionalized UTSA-75 (with formula [Cu2(L3)(H2O)2]∙3DMF where H4L3 is 5, 50 -(pyridazine-2, 5diyl)diisophthalic acid), and pyrimidine functionalized UTSA-76 (with formula [Cu2(L4)(H2O)2]5DMF3H2O where H4L4 is 5,50 (pyrimidine-2, 5-diyl)diisophthalic acid) frameworks which are isostructure frameworks with NOTT-101 with formula [Cu2(L1)
Fig. 33. Control over function and size of ligands in designing dual pyrimidine functionalized UTSA-110 for high methane storage purposes. Reproduced with permission from Ref. [173]. Copyright 2018 Wiley-VCH.
26
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
(H2O)2] where H4L1 is Terphenyl-3,300 ,5,500 -tetracarboxylic acid (Fig. 32) [174]. They have demonstrated that incorporation of functional groups with Lewis basic nitrogen sites into the MOF pores can significantly improve the total volumetric methane storage capacities. When pyridine groups were incorporated into the framework of NOTT-101, the methane storage capacity of ZJU-5 was increased from 237 (for NOTT-101) to 249 cm3(STP).cm3 at 65 bar. The volumetric methane storage capacities can be further improved in UTSA-75 (251 cm3.cm3) and UTSA-76 (257 cm3. cm3) by respectively incorporating pyridazine and pyrimidine groups into NOTT-101. Thus volumetric working capacities for ZJU-5 (188 cm3.cm3), UTSA-75 (192 cm3.cm3), and UTSA-76 (197 cm3.cm3) are improved compared to NOTT-101 (181 cm3. cm3). These MOFs are isostructures with very close and similar BET, pore volume, pore shape, and lattice parameters. Clearly, the difference in methane storage capacities of these MOFs at room temperature is due to the change of Lewis basic functional groups. Calculations show that CH4 molecules adsorbed next to the pyrimidine sites of UTSA-76 (15.44 kJ.mol1) exhibit rather similar binding energies to those adsorbed on the central phenyl ring of NOTT101 (15.49 kJ.mol1). In functional view, ideal porous MOFs for high methane storage and working capacities are those with suitable functional sites which can significantly enhance methane storage capacities at higher pressures (35 bar) but have minimal effects on methane storage capacities at lower pressures (5 bar). In fact, upon methane adsorption at room temperature and high pressure the central pyrimidine rings within UTSA-76 can be more easily adjusted and oriented to optimize the methane packing at higher pressures than the central benzene rings within NOTT101. Hence, at low pressures, both MOFs show about the same uptake, but at higher loadings of methane, UTSA-76 can adsorb more methane molecules, because of its adjustable pyrimidine ring orientation, yielding higher working capacity compared to NOTT101. The gravimetric uptake of methane for robust MOFs at 65 bars is positively proportional to the porosity. Moreover, high porosity can also contribute in reducing the low-pressure methane adsorption and be beneficial for the higher working capacities. On the other hand, MOFs should have balanced porosities and framework densities as well as high densities of functional sites for high volumetric uptake. These functions should be effective at higher pressures without negative effects at lower pressures. With this idea, the effects of Lewis basic nitrogen sites on methane storage and working capacity have been investigated by synthesizing UTSA110 with optimized porosity and pyrimidine sites through expanding the UTSA-76 framework with linker-extending approach to contain higher density of functional pyrimidine sites (Fig. 33) [173]. As a result of such a creative design, UTSA-110 exhibits very high gravimetric and volumetric working capacities of 317 cm3(STP).g1 and 190 cm3(STP).cm3, respectively. These studies clearly display the effect of functionalization on the performance of methane storage. For example in comparison with MOFs with coordinatively unsaturated metal centers (USMCs), UTSA-110 show much higher working capacities because the USMCs have a very strong affinity to methane molecules at 5 bars. He and coworkers designed three isostructure MOFs with NOTT-101. Replacing the benzene spacer in the organic linker of NOTT-101, 5,50 -(benzene-1,4-diyl)diisophthalate, with the nitrogen containing heterocyclic rings, namely, pyrazine, pyridazine, and pyrimidine results in three MOFs which are respectively denoted as ZJNU-46 (with formula [Cu2L1(H2O)2]3DMF4EtOH3 H2O where H4L1 is 5,50 -(Pyrazine-2,5-diyl)diisophthalic acid), ZJNU-47 (with formula [Cu2L2(H2O)2]3DMF3EtOH2H2O where H4L2 is 5,50 -(pyridazine-3,6-diyl)-diisophthalic acid), and ZJNU48 (with formula [Cu2L3(H2O)2]3DMF4CH3CN3H2O where H4L3 is 5,50 -(Pyrimidine-2,5-diyl)diisophthalic acid) [176]. These three
MOFs show very high acetylene adsorption capacity. At 295 K and 1 atm, the gravimetric acetylene uptakes are 187, 213, and 193 cm3.g1 (STP) for ZJNU-46, ZJNU-47, and ZJNU-48, respectively. When the temperature went down from 295 K to 278 K, the amount of acetylene adsorbed increased to 257, 283, and 265 cm3.g1 (STP), respectively. The authors mentioned that although these MOFs have similar structural properties and close porosity, the difference in the relative position of the nitrogen atoms in the central ring of organic linkers leads to their distinctly different acetylene uptake capacities in a way that pyradazine functionalized ZJNU-47 shows higher affinity and capacity toward acetylene. Similarly in another work, acetylene uptake (from 184 to 216 cm3.g1) and C2H2/CO2 selectivity (from 8 to 9 to 11.5–17) drastically improved by functionalization of central benzene ring of NOTT-101 through pyrazine function leading to construction of pyrazine functionalized ZJU-40 [175]. Such a high C2H2 uptake capacity was mainly attributed to the introduction of functional pyrazine groups with Lewis basic nitrogen sites into the pore surfaces as well as the suitable pore sizes. However, the immobilized Lewis basic nitrogen sites in ZJU-40 (with formula [Cu2(L)(H2O)2] 7DMF4H2O where H4L is 5,50 -(pyrazine-2,5-diyl)diisophthalic acid) show a negligible effect on the CO2 storage capacity, leading to significantly enhanced C2H2/CO2 selectivity for ZJU-40 compared to that of NOTT-101. Triazine functionalized MOFs have been applied in adsorption and separation of light C2 hydrocarbons as well as benzene/cyclohexane separation. Cao and coworkers synthesized FJU-C1 (with formula (Et4N)3[In3(TATB)4] where H3TATB = 4,40 ,400 -s-triazine-2,4, 6-triyltribenzoic acid) showing high adsorption of polar vapors like methanol, water, and ethanol compared to nonpolar molecules and high separation selectivity for benzene over cyclohexane owing to p rich– p deficient interactions between benzene molecules and the s-triazine rings of the porous material [192]. Han and coworkers synthesized MOF(2), [Zn3(TATB)2(H2O)2]n where H3TATB = 4,40 ,400 -s-triazine-2,4,6-triyltribenzoic acid, and applied it as a probe for designing charge transfer (CT) materials with electron-rich DMA (N,N0 -Dimethylaniline) guest molecules [200]. DMA can be successfully included into the triazine incorporated pores of MOF (2) with color change of white for (2) to green for DMA@(2). This color change represents charge transfer between the host triazine units (electron acceptor) and the guest DMA molecules (electron donor). Among heterocyclic-azine rings, tetrazine has most interesting features such as: (I) highly dense ring with accessible nitrogendonor sites for dipole–quadrupole or Lewis acid-base interactions
Fig. 34. Representation of functional pores of TMU MOFs. (a) H,H-s-tetrazine functionalized pores of TMU-34. (b) tetrazine functionalized pores of TMU-34(–2H). Reproduced with permission from Ref. [214]. Copyright 2017 American Chemical Society.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
[705,706], (II) very high p- deficiency due to the presence of four electronegative nitrogen atoms,[707] (III) s-tetrazines are strong oxidizer because of their strong electron-deficient character and can accept electrons and protons simultaneously to and form dihydro-tetrazine (H,H-s-tetrazine) through a redox-switchable reaction,[706] (IV) s-tetrazine derivatives are highly electroactive heterocycles with very HOMO and LUMO close energies,[708] (V) tetrazine derivatives accept an electron in reversible redox reaction to form a stable radical-anion. In the next step, most of the tetrazine derivatives can accept a second electron to form a diracialanion with high basicity which reacts with traces of water or protic impurities and can be easily protonated to produce 1,4-dihydrotetrazine in an irreversible electrochemical reaction [709,710], (VI) due to a low lying p* tetrazine molecular orbitals and weak n-p* electronic transitions in the visible range, they have unique optical, electronic and electronical properties which make them a good candidate in sensory applications [711], (VII) simultaneous low permanent quadrupole moment and high molecular polarizability characteristics turn the tetrazine ring into a good candidate for interaction with both anions and cations [712]. Due to all these appropriate features, the tetrazine functional groups have been widely used in the structure of MOFs to enhance frameworkguest interactions (Table 1). Tetrazine decorated MOFs are nitrogen rich and it seems that because of this characteristic they have been used for adsorption of H2, CH4, and CO2 gases. Overall, contrary to previous mentioned assumptions, this group of functional MOFs does not present high affinity toward carbon dioxide. The binding energies are proportional to the electron density. For the s-tetrazine, the nitrogen atoms are more electron deficient than for the other azine heterocycles. Thus tetrazine has the low binding energy to interact through Lewis base-acid interactions with carbon dioxide. On the other hand, this function shows good affinity toward CH4 molecules. TMU-34(–2H) (with formula [Zn(OBA)(DPT)0.5]nDMF where H2OBA and DPT are 4,40 -oxybis(benzoic acid) and 3,6-di(pyridin-4yl)-1,2,4,5-tetrazine, respectively) as a s-tetrazine and TMU-34 (with formula [Zn(OBA)(H2DPT)0.5]nDMF where H2DPT is 3,6-di(p yridin-4-yl)-1,4-dihydro-1,2,4,5-tetrazine) as a H,H-s-tetrazine functionalized MOF show very high affinity toward CH4 molecules (22 kJ.mol1 for TMU-34(–2H) and 23 kJ.mol1 for TMU-34) (Fig. 34) [214]. With regard to this high affinity of CH4-tetrazine, TMU-34(–2H) has a higher capacity for CH4 molecules even in comparison with high surface area MOFs at room temperature and pressure. [(WS4Cu4)I2(dptz)3] 3DMF (denoted as 1, dptz = 3,6-di-(pyridin4-yl)-1,2,4,5-tetrazine) was synthesized by Zheng and coworkers for sensing small solvent molecules (Fig. 35) [210]. They designed an MOF with solvatochromic organic ligand and Cu+1 was chosen as the metal center because its filled-shell d10 electronic configura-
27
tion offers opportunities to observe other low-lying electronic transitions related to tetrazine ring. After immersion of a synthesized sample in different solvent molecules, compound 1 exhibits different colors depending on the solvent polarity and the band gaps of these solvent@1 complexes are in linear correlation with the polarity of the guest solvents in two different protic and aprotic groups. By increasing the solvent guests’ polarity the absorption bands are blue-shifted showing a negative solvatochromic effect. Although the solvatochromism of 1 and dptz ligand originate from different transitions (MLCT transition for 1 and p ? p* transition for dptz), the results indicate that dptz ligand plays an important role in the solvatochromic response of 1, which should be ascribed to its strong p-acceptor property and labile electronic structure to solvent polarity. Considering the redox-switchable activity of tetrazine ring, dihydro-tetrazine functionalized TMU-34 ([Zn(OBA)(H2DPT)0.5] DMF (H2DPT = where 3,6-di(pyridin-4-yl)-1,4-dihydro-1,2,4,5-tet razine, H2OBA = 4,40 -oxybis(benzoic acid)) have been applied in sensing of chloroform (Fig. 36) [208]. TMU-34 changes color from yellow to pink through reversible dynamic conversion of dihydro-tetrazine into tetrazine upon exposure to chloroform with 2.5 105 concentration in the presence of other VOCs. As a result, TMU-34 can act as a solid-state, naked-eye visual chemosensor for detection of chloroform in liquid and gas phases.
Fig. 36. Application of TMU-34 in colorimetric detection of chloroform. (a) Sensing cyclic and (b) mechanism of detection. Reproduced with permission from Ref. [208]. Copyright 2017 Wiley-VCH.
Fig. 35. Application of [(WS4Cu4)I2(dptz)3] 3DMF in solvatochromism. (a) The coordination environment of theWS4Cu2+ 4 unit. (b) The UV–vis spectra and photograph of the inclusion Compounds solvents@[(WS4Cu4)I2(dptz)3]. Reproduced with permission from Ref. [210]. Copyright 2010 American Chemical Society.
28
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Because of the presence of four p-accepting and r-donating sp2 N atoms, tetrazine ring can behave as an electron donor or acceptor guest-interactive site. Based on this character, a procedure has been designed to sense Hg2+ with a novel two dimensional and highly sensitive and selective method called double solvent sensing method (Fig. 37) [204]. Fluorescence measurements revealed that TMU-34(–2H) with formula [Zn(OBA)(DPT)0.5] where DPT is 3,6-di(pyridin-4-yl)-1,2,4,5-tetrazine, can selectively detect Hg2+ in water (by 243% enhancement) and in acetonitrile (by 90% quenching). This different behavior in water compared with acetonitrile may be attributed to the hydrogen bonding between water molecules and tetrazine moieties. As a result two 1D detection curves of Hg2+ in water and acetonitrile can be combined together for constructing a 2D double solvent sensing curve. The calculated sensitivity factor in the 2D sensing curve is between 0 and 2 for all of the cations other than Hg2+, whereas for Hg2+ it is equal to 41. As mentioned before, pyridine and diazines especially pyrazine are applied in the synthesis of MOFs as binding sites (in bare or carboxylated forms) for constructing the frameworks. Bare pyridine is not applied for construction of MOFs, but it is used as modulator reagent to control the size and morphology of MOF particles [416]. Among the diazines, pyrazine is mostly applied as bare ligand because the two nitrogens of pyrazine occupy para positions which are completely suitable for construction of pillared porous MOFs. In some cases bare pyrazine is used as nitrogen donor ligand along with other ligands with the same geometry but different lengths (such as 4,40 -bipyridine, 1,2-di(pyridin-4-yl)ethyne and 1,4-bis(4-pyridyl)benzene) to form MOFs in order to evaluate the effects of pore size on the structure and applications of MOFs
[713–716]. In contrast to pyrazine, bare pyrimidine and especially pyradazine rings are rarely applied in the structure of MOFs because their N atoms are located in places which are not ideal for construction of porous frameworks. Although introduction of organic functional groups inside the pores of MOFs is a very powerful strategy, introduction of chemically accessible Lewis basic nitrogen sites through pre-synthesis method still remains a big challenge. There are some strategies to overcome this issue like (I) using hard and oxophil metal ions with high charge density and (II) right choice of heteromultitopic carboxylate-heterocyclic azine ligands with special geometry and rational arrays of nitrogen Lewis basic sites. Based on the first strategy, using oxophil metals like Zr4+ and Al3+ is useful to achieve uncoordinated Lewis basic sites. This strategy is specially effective when using ligands with chelating N-based Lewis basic sites like 2,20 -bipyridine-5,50 -dicarboxylic acid. Application of the mentioned ligand with Zn2+, Co2+, and Ni2+ leads to chelation of metal ions to N atoms [717,718]. But combination of 2,2-bipyridine-5,50 -dicarboxylate and Zr4 + results in the construction of a framework with UiO-67 framework containing free N atoms which can be applied as anchoring sites for immobilizing catalytic metals (Fig. 38) [703,704,719]. To discuss the rational design of heteromultitopic carboxylateheterocyclic azine ligands with uncoordinated Lewis basic N-site we classified heteromultitopic ligands into two different groups. In group one, carboxy function(s) are directly connected to pyridine and diazine rings. In the second group, heterocyclic-azine N atoms and carboxylate functions are located on different benzene rings. In group one, it is almost impossible to construct frameworks with free N-sites by pre-synthesis design of the structure. This group
Fig. 37. Application of TMU-34(–2H) in Hg2+ detection. (a) Mechanism of detection and (b) combination of two 1D curves for construction of 2D double solvent sensing curve. Reprinted with permission from Ref. [204]. Copyright 2017 Royal Society of Chemistry.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 38. Structure and coordination modes of 2,20 -bipyridine-5,50 -dicarboxylate. (a) Bare 2,20 -bipyridine-5,50 -dicarboxylate. (b) Free N Lewis basic sites in Zr-based MOF. (c) Coordinated N lewis basic sites in isostructural Zn-Ni and Co based MOFs. 38b: Reproduced with permission from Ref. [720]. Copyright 2015 American Chemical Society. 38c: Reproduced with permission from Ref. [717]. Copyright 2010 Royal Society of Chemistry.
of ligands usually include ligands such as pyridine-4-carboxylate, pyridine-3,5-dicarboxylate, pyridine-2,5-dicarboxylate, pyridine2,4,6-tricarboxylate, pyrazine-2,5-dicarboxylate, pyrazine-2,3dicarboxylate, pyrazine-2-carboxylate, pyrazine-2,3,5,6-tetracar boxylate, pyrimidine-4,6-dicarboxylate, pyrimidine-3-carboxylate. For example, in the structure of [La(pyzdc)1.5(H2O)2]2H2O ,where pyzdc is pyrazine-2,5-dicarboxylate, although just one heteromultitopic ligand is applied, for some pyzdc ligands the heterocyclic-azine N atoms are coordinated to metal ions while in others they do not coordinate (Fig. 39) [721]. A similar situation is observed in other coordination polymers like [Co2(pmc)4(OH)2], [Cd(pmc)2] and [Cu (pmc)2] where pmc is pyrimidine-5-carboxylate, [Gd2Ag6(pzdc)6(H2O)9]8H2O and pzdc is pyrazine-2,3-dicarboxylate [722,723]. Therefore, it is hard to determine freedom or engagement of heterocyclic-azine N atoms before self-assembly process and structural analysis. In group two where heterocyclic-azine N atom(s) and carboxylate are located on different phenyl rings, some critical points in ligand decoration should be considered such as (I) ligand geometry (linear, bent, or multi-chain), (II) position of coordinating site, and (III) strength and type of coordinating sites. Overall, in self-assembly process for the construction of highly crystalline and regular MOF structures, the geometry of ligand and coordinating sites, especially carboxylate, play critical roles through ligand based template effects. This template effect of ligand particularly happens in case of terminal phenyl rings. To achieve free N-Lewis basic sites, the N-atoms should not be placed
Fig. 39. Structure of [La(pyzdc)1.5(H2O)2]2H2O. (a) Coordination modes of the ligand. (b) Tetranuclear SBU. Reprinted with permission from Ref. [721]. Copyright 2014 Wiley-VCH.
29
on positions that are engaged in structure formation through ligand template effect, especially on terminal phenyl rings. In case of series A ligands, the ligand has two terminal phenyl rings (Fig. 40a). Since two phenyl rings are terminal, to develop the MOF structure both carboxylate and heterocyclic-azine N atoms on both phenyl rings can be coordinated to metal ions. Replacement of carboxylate with pyridine in terminal rings results in coordination of nitrogen atom, even in different positions which expresses the effects of location of N atoms in terminal rings with high ligand-based template effects. But by introduction of two N atoms, the ring is pyrimidine-like and both N atoms cannot coordinate simultaneously. This issue (not connecting two N atoms of pyrimidine ring) is the result of reduced Lewis basicity and coordination strength of pyrimidine N atoms. In case of series B and C (Fig. 40b-c), there are two terminal and one central phenyl rings. Replacement of central phenyl ring with pyridine and diazines introduces no changes in the coordination modes of heterocyclic-azine functional ligands because the N atoms are placed in positions which have low ligand-based template effects during structure formation (central phenyl ring). Moreover, position of Lewis basic N atoms and lower coordination strength, compared to carboxylate, helps them to remain free. Therefore, in case of ligands in groups B and C, locating Lewis basic N atoms in central phenyl ring with lower ligand-based template effects along with weaker coordination strength compared to carboxylate lead to freedom of N Lewis basic sites in these ligands [174,178]. However, in case of L23, with strong coordinating carboxylate group on the central phenyl ring and lower ligandbased template effect, carboxylate can coordinate to metal ions and therefore does not remain free. Since in L23 carboxylate groups are located on the phenyl ring with lower ligand-based template effect, the synthesis conditions like solvent and temperature are important to make it free of coordination [137,179,724]. It is clear that in series D (Fig. 40d), by replacing carboxylate with N Lewis basic sites in terminal ring with high ligand-based template effects (L42), N atoms are coordinated. In case of L43, N atoms which are placed in terminal phenyls are coordinated while the others are not [158,725]. In case of series E ligands with bent geometry, central phenyl ring on metha position has high level of ligand-based template effect (Fig. 40e). Some of the N atoms which are on metha positions and have template effect are coordinated to metal ions but the others are not coordinated. In case of L55, with pyrimidine ring, both N atoms are free because of simultaneous low coordination strength of N pyrimidine atoms and low template effects in their positions [160,161,181,726–728]. 2.2.2. Heterocyclic azole N-Based functions Heterocyclic azole N-based functions include pyrazole, imidazole, triazole, and tetrazole. All heterocyclic azoles contain a number of Lewis basic nitrogen atoms which act as both coordinating and guest-interactive sites in supramolecular chemistry and construction of coordination complexes and polymers especially MOFs. Heterocyclic azoles have different nitrogen sites; A Lewis basic pyrrolic N–H group, and Lewis basic pyridinic N-sites. As building blocks in supramolecular and coordination chemistry, heterocyclic azoles usually bind metal ions through pyridinic-N, which leaves the N–H group free [729]. In this condition, free heterocyclic azole N–H group of pyrazole provides two beneficial aspects: (I) coordination chemistry of heterocyclic azoles through N-pyridinic atom may be complemented by the hydrogen-bond-donor ability of the pyrrolic N–H site for the generation of supramolecular polymers with higher dimensionality, (II) free N–H groups can act as supramolecular hosts for anions and other guests. That means heterocyclic azole ring can act as structure-directing/coordinating
30
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
and guest-interactive site inside the structure of coordination polymers. Coordination through just pyridinic atoms leads to construction of metal-azole frameworks (case I) while coordination through pyrrolic nitrogen by deprotonation mechanism leads to construction of metal-azolate frameworks (case II). The most important feature of these functions for acting as guest interactive site is their Lewis basicity which is very helpful for interaction with Lewis acidic guests like CO2, metal ions, hydrogen bond donor guests and providing Lewis basic catalytic sites (Table 1). Additionally, heterocyclic azole functions have distinctive characteristics. Pyrazole has very strong coordination strength
[693,730,731] and because of this characteristic it is used for tuning the ligand field of metal ions for efficient O2 separation [229]. Other than basicity, in case of metal-azole frameworks, they have free (–NH) sites which are good for interacting with guests and providing hydrophilic pores for proton conductivity, especially in case of tetrazole which has pKa near carboxy function (Table 1). Also p-system of these functions is another guest-interactive characteristic of these functions. The nitrogen-rich property of azole ring offers good chemical characteristics such as polarized carbon atoms in the ring which allow the complexation of anions by hydrogen and halogen bonding for ion-pair recognition and p-p interactions (Table 1). Some of heterocyclic azoles have very spe-
Fig. 40. Strategies for ligand design with free N-basic centers inside the structure of MOFs. (a) Locating N Lewis basic atoms in terminal rings with high ligand based template effects leads to coordinating to metal ions. (b) Locating N atoms with weaker coordination strength compared to carboxylate in central phenyl ring with low ligand based template helps them to remain free. (c), (d) and (e) Locating N atoms in positions with high and low template effects helps them to coordinate or remain free.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
cial properties. For example imidazole is used to construct carbene containing MOF based catalysts and triazole and especially tetrazole FMOFs are applied as energetic materials because they have high content of C–N, C@N, and N@N energetic bonds with high level of energy and low sensitivity (Table 1). Heterocyclic azoles including pyrazole, imidazole, triazole, and tetrazole have moderate-high affinity to coordinate to metal ions [693,732]. This is because these functions contain multiple Lewis basic nitrogens as well as their aromaticity. Owing to different number and positions of nitrogen atoms, each of them can bind to different metal ions with different coordination modes (Fig. 41). Coordination modes of azolate are controlled by two parameters: (I) deprotonation of pyrrolic N atom and (II) increasing the pyridinic N number. In both cases azolates hapticity will increase. In case of deprotonation, basicity is increased as well. In heterocyclic azole functions, pKa values alone can reflect the coordination strength of the conjugated bases with metal ions [733]. By increasing the N atom number in the heterocyclic azole rings, negative charge on nitrogen atoms in conjugated bases is reduced and pKa will decrease. Corresponding pKa for tetrazolate, triazolate, imidazolate, and pyrazolate are in ranges of 4–6, 12–
31
13, 16–17, and 18–20 respectively. By increasing pKa values, metal-azolate binding energy will increase as well (Fig. 42). Most of the synthesized MOFs are based on 3d late metals and carboxylate function. Since the carboxylate is a hard coordinating site and 3d late metals are soft Lewis acid centers and have higher affinity toward nitrogen rather oxygen, it is rational to assume that using MOFs with soft heterocyclic azoles and 3d late metals is a good and simple strategy for construction of stable MOFs. 3d late metal-heterocyclic azole(ate) frameworks show high stability in air, humidity, and acidic or basic media compared to 3d late metal-carboxylate frameworks. It is mentioned that Zr4+carboxylate frameworks show low stability in basic solutions, but normally 3d metal-pyrazolate frameworks show higher stability in such solutions. The reason is that late 3d metals especially Zn2+ and Cu2+ are nitrogenphil and soft Lewis acids while pyrazolate is an N donor coordinating site with relatively high negative charge on nitrogens with high pKa which results in very strong Zn2+/Cu–N bonds and ultrastable Zn2+/Cu2+-pyrazolate frameworks. Moreover, most metal-carboxylate frameworks are relatively hydrophilic,[685] but simple metal-azolate frameworks are usually hydrophobic [729]. Some pyrazolate-based MOFs like
Fig. 41. Coordination modes of heterocyclic azole N-based functions.
Fig. 42. Comparison of structural features of more common coordinating azoles.
32
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
[Cu2(Et4BPZ)], [Zn(1,3-BDP)], [Zn3(BDP)2] and [Fe2 (1,4-BDP)3] show high chemical and thermal stabilities [734–737]. As a good example of an ultra-stable MOF, FJU-66 (with formula [Cu6(NDI)3]∙2DMF∙6MeOH2H2O where H2NDI is 2,7-bis(3,5-dime thyl)dipyrazol-1,4,5,8-naphthalene-tetracarboxydiimide)) shows high thermal stability that can keep its structure intact up to 803 K [131]. The chemical stability of FJU-66 was tested by treating its samples under different conditions. FJU-66 remains stable not only in aqueous solutions with pH values in the range of 3–14, but also in 10 M sodium hydroxide solution over 12 h. BUT-31 ([Ni8(OH)4(H2O)2(BDP-X)6] where H2BDP is 1,4-bis(4-pyrazolyl) benzene and X = –CHO, –CN, –COOH) tolerates boiling water and even highly basic aqueous solution (4 m sodium hydroxide) [223]. It is reported that polyazolate-based MOFs are well known for their good stability under humid and basic conditions [145,738–740]. In another work, Kitagawa and coworkers chose 3,30 ,5,50 -tetramethyl-4,40 -bipyrazole as a building unit and successfully prepared Cu(I) framework materials with greater thermal stability (Tdec = 500 °C in nitrogen atmosphere) [741]. The ditopic nature of pyrazole and the coordinative flexibility of pyrazolate mean that the pyrazole ring can be considered as a flexible ligand in coordination chemistry [742–746]. It can also present structure-directing role through hydrogen bonding and p-p stacking interaction [747]. Mahon and coworkers reported seven pyrazole functionalized MOFs with different o-donor carboxylate ligands and nitrogen donor N,N-ditopic ligand di(4-pyridyl)1Hpyrazole (Hdpp). [747]. In compounds 1 (with formula [Zn2(bdc)2(Hdpp)2]2DMF where H2bdc is 1,4-benzenedicarboxylic acid), 2b (with formula [Zn(1,4-ndc)(Hdpp)]DMF where H2(1,4ndc) is 1,4-naphthalene dicarboxylic acid), 3 (with formula [Zn (mbdc)(Hdpp)]DMF where H2mbdc is 1,3-benzenedicarboxylic acid), and 4 (with [Zn2(mbdc-Me)2(Hdpp)2]DMF where H2(mbdc-Me) is 5-methyl-1,3-benzenedicarboxylic acid), the pyrazole NH groups are involved in hydrogen bonding that serves to link either interpenetrated networks or neighboring sheets together. In these structures the presence of pyrazole group provides a means to interlink interpenetrated or interdigitated networks by hydrogen bonding. However, in 2a (with formula [Zn2(1,4-ndc)2(Hdpp)]4DMF), 5a and 5b (with formula [Zn2(2,6ndc)2(Hdpp)]DMF where H2(2,6-ndc) is 2,6-naphthalene dicarboxylic acid) the hydrogen bonding NH groups project into the pores of the framework enabling interactions with guest mole-
cules. Comparing the structures of 2a, 2b and 5a, 5b shows that upon tuning the reaction conditions each of these structures could be selectively obtained. Zhang and coworkers synthesized flexible frameworks which exhibit gate-opening and hysteretic adsorption behaviors in response to light hydrocarbons with high capacity [745]. Combination of coordination modes for heterocyclic azoles and carboxylate indicates that the syn,syn-l-g1,g1-bidentate mode of pyrazolate as the basis of paddle-wheel M2(RCOO)4(HKUST-1 type SBU), trigonal-prismatic M3O(RCOO)6(MIL-101 type SBU), and octahedral M4O(RCOO)6 (MOF-5 type SBU) is similar to carboxylate [729]. In case of octahedral M4O(RCOO)6 building blocks, some of the carboxylate-pyrazolate bifunctional ligands like 3,5-dimethyl4-carboxypyrazolato, 4-(3,5-dimethylpyrazol-4-yl)benzoic acid, and 3,30 ,5,50 -tetramethyl-4,40 -bipyrazolate have been used in the construction of isostructural MOFs with MOF-5 (Fig. 43) [748– 752]. In another case, 1,3,5-tris(1H-pyrazol-4-yl)benzene was applied in the structure of HKUST-1 with M2(RCOO)4 paddle wheel clusters and shows higher thermal and chemical stability compared to HKUST-1 (Fig. 43) [735]. In some cases triazole/azolate and tetrazole/azolate rings form building blocks similar to carboxylate since they can use some coordination modes like pyrazolate [229,251,283,753,754]. As an important subclass of MOFs, zeolite imidazole(ate) frameworks (ZIFs) are developed by bare or substituted imidazolate rings. ZIFs are composed of tetrahedrally coordinated metal ions (Fe, Co, Cu, Zn, and . . .) with metal-imidazolate-metal angle of 140°. These structural properties are very close to inorganic silica with tetrahedral Si(Al)O4 moieties and Si–O–Si angle (145°) [755,756]. So it is clear that ZIFs have zeolite like topologies. ZIFs are resistant to thermal change and chemically stable and are much more restricted to their structure compared to MOFs. This high stability arises from characteristics such as (I) similarity of M–N–M angle to Si–O–Si, (II) interaction between soft coordination sites and soft metal ions, (III) hydrophobic pores provided by imidazolate rings. Synthesis conditions of pyrazolate based MOFs (high temperature and basic solution for deprotonation) are more difficult because of their higher pKa. Synthesis procedure for preparation of MOFs with azolate coordinating sites is relatively insensitive to common species presented in the solution such as anions and solvent molecules which can hardly compete with the targeted
Fig. 43. Representation of constructed HKUST-1 and MOF-5 with pyrazole based functions.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
metal azolate bonds. This is particularly true for those composed of soft metal ions and diazolates [693]. In case of tetrazole, similarity of pKa to carboxylate indicates that similar mild conditions are needed for synthesis. Regarding the coordinating sites in carboxylate and tetrazole functions we can mention the following points: (I) tetrazolatebased frameworks are very sensitive to high temperatures, that can be avoided either by replacing tetrazole with other azole rings or by using mixed ligand systems and combining carboxylate and tetrazolate linkers; (II) tetrazoles are useful in producing robust MOFs with permanent porosity, proper topologies, and gas adsorption properties comparable to those of the pioneering carboxylatebased MOFs; (III) strong coordination bonds between 3d late metal ions and nitrogen sites of tetrazole can result in greater stability of the prepared MOF. However, the relatively weak coordination between 3d late metal ion and oxygen from carboxylate facilitates the growth of big single crystals for structural characterization; (IV) tetrazolate and carboxylate groups have similar planar structures and the same pH, so that similar experimental conditions can be employed to synthesize to synthesize MOFs based on these linkers; and (V) compared to carboxylate (which is a bidentate ligand), tetrazolate linkers are more versatile and can coordinate to metals using one or several nitrogen atoms, increasing the variety of the structures that can be generated. The new approach of designing FMOFs with azole(ate)carboxylate coordinating sites has received a lot of attention because this kind of ligand benefits from advantages of both carboxylate and carboxylate groups. Using azolate groups provides strong metal ligand interaction which is of benefit for improvement of structural stability. Also, using carboxylate ligands leads to construction of frameworks with high crystallinity. A crystalline and stable MOF is easy to characterize and its stability makes it suitable for real-life applications. One important feature of heterocyclic azoles is their ‘‘charge modulation” (Fig. 44). There are three possibilities to modulate charge of heterocyclic azoles: (I) if the azolate ring coordinates to metal ions through deprotonation, the ring is negative (azolate ring) and (II) if azole ring binds to organic groups, the ring is positive (azolium ring). However azolium ring is used in the structure of MOFs by imidazolium (Table 1) and in rare cases the triazolium ring [757]. Construction of azolium ring is a very useful method to synthesize ionic MOFs by pre-synthesis method. 2.2.3. Non-cyclic N-based functions This group of nitrogen-based functions consists of amine, imine [758,759], azo, azine, and azide functions (Table 1). These functions provide similar chemistry based on nitrogen atoms and different host-guest chemistry owing to the different bonding, arrangement and number of nitrogen atoms. 2.2.3.1. Amine. Amine is structurally simple but contains multiple interactive abilities such as (I) hydrogen bond donating and accepting through H and N atoms, (II) Brønsted acidic or Lewis basic characteristics through N atom, and (III) negative charge accumulation
33
on electronegative N atom and polarity. Such simple and multiple chemistry of amine functional group is of benefit in designing functional MOFs for distinctive applications like gas storage and separation, removal, separation and sensing, catalysis and photocatalysis, electrochemistry and drug delivery (Table 1). As a polar and Lewis basic site, amine FMOFs have been extensively applied in development of CO2 adsorptive MOFs because this guest gas molecule is Lewis acidic and quadrupole in a way that combination of chemical features of amine function and CO2 molecules provides high affinity for engaging together through CO2(C) (N)amine and CO2(O) (H)amine interactions. Both arylamine and alkylamine functionalized MOFs are applied as CO2 adsorbents, but different basicity of these groups of amine functional groups leads to different results for CO2 adsorption and separation (Table 1). In case of alkylamines, larger affinity of aliphatic amines leads to chemisorption of CO2 molecules. The use of arylamines could favor strong physisorption (30–50 kJ.mol1) with CO2 rather than chemisorption. Comparing the aromatic and nonaromatic amines, aromatic amines are generally less basic than aliphatic amines because the nitrogen lone-pair electrons are delocalized by interaction with the electrons of aromatic ring and are less available for bonding. As examples of arylamine functional MOFs, CAU-1 exhibits high heat of adsorption (-48 kJ.mol1), high CO2 uptake (7.2 mmol.g1 at 273 K and 1 atm), and remarkable CO2/N2 selectivity (101:1) and Uio-66-NH2 shows improved CO2 adsorption capability compared to the parent MOF (CO2 uptake (mmol.g1) = 8.5 vs 7, zero coverage enthalpy (kJ.mol1) = 32 vs 25.5, selectivity (CO2/N2) (15%/75%)-IAST = 66.5 vs 37.5) [325,333]. Different types of alkylamine ligands like tetraethylenepentamine, N,N0 -dimethylethylene diamine, 1-methylethylenediamine, 1,1-dimethylethylenediamine, ethylenediamine, piperazine, 3 and 4-picolylamine, N,N0 -dimethyle thylenediamine post synthetically grafted in the coordinatively unsaturated metal centers of CO2 adsorbent MOFs and generally show higher Qst and selectivity compared to arylamine functionalized MOFs [302,343,347–356]. As an example, incorporating N,N0 -dimethylethylenediamine (mmen) within CuBTTri and Mg(dobpdc)2 affords a materials with exceptional CO2 affinity and selectivity compared to the parent MOFs (Q0st(kJ.mol1) = -24, –96, –47 and –71 for Cu-BTTri, mmen-Cu-BTTri, Mg(dobpdc)2 and mmen-Mg(dobpdc)2 respectively). For CuBTTri and Mg(dobpdc)2 with Lewis acidic open metal sites, after grafting with mmen, free amine functions appear in the holes as Lewis basic sites [347,349]. By changing the functionality from Lewis acidic to Lewis basic, the affinity of the framework toward CO2 is remarkably improved. The introduction of amine group into the MOFs can enhance efficiency for the catalytic and photocatalytic reaction since it is Lewis basic and acts as antenna group for light sensitization of the framework (Table 1). MIL-68(In) and MIL-68(In)–NH2 are used as catalyst in catalytic synthesis of cyclic carbonated from epoxide (styrene) and CO2 [361]. The adsorption of CO2 occurs on the basic sites to form activated species. Findings show that conversion of styrene oxides is clearly enhanced by –NH2 functionalization. It increased from about 39% for MIL-68(In) to 73% for MIL-68(In)–
Fig. 44. Modulation of charge on imidazolium ring.
34
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
NH2. In another work, the amine functionalized MOF, showed much higher efficiency under visible light irradiations. In addition to direct excitation of Fe3-l3–oxo clusters in MIL-88B(Fe) the amine functionality in MIL-88B(Fe)–NH2 can also be excited to transfer an electron to Fe3-l3–oxo clusters, which is responsible for the enhanced photocatalytic activity for Cr(VI) reduction [378]. The obtained evidence reveals that when NH2-DBC ligand (organic building block of MIL-88B(Fe)–NH2) is irradiated with visible light, an ESR signal with a g value of 2.004 is observed, which is produced by space confined amine group upon irradiation (Fig. 45). Interactive features of amine function like polarity and Lewis basicity and hydrogen bond participating caused the widespread application of amine decorated MOFs in sensing, removal, and separation of hazardous materials (Table 1). Amine decorated compound [Cd(2-NH2bdc)(tib)4H2O0.5DMA]n (2), where the tib = 1,3,5-tris(1-imidazolyl)benzene, acts as a free standing donor for Hg(II). Hg(II)-(NH2) interaction leads to quenching in compound 2 fluorescence peak. In presence of Hg(II) a new N(1 s) peak appears in XPS analysis (406.38 eV) indicating a chelation interaction between N atom of amino group inside the compound 2 framework and Hg(II) [397]. Bio-MOF-1can selectively detect 2,4,6-trinitrophenol (TNP) in presence of other nitro aromatics through fluorescence quenching which is related to the hydrogen bonding interaction between free functional amine and TNP analyte (Fig. 46) [400]. Functionalization of MIL125(Ti) via amine motif has a very significant effect on H2S adsorption from natural gas, because H2S is H donor and NH2 is H acceptor and they can interact mostly through (H2S)H-N(NH2) and partially through (H2S)S-H(NH2) hydrogen bonding [338]. H2S over CH4 selectivity is very high (70:1) because the impact of functionalization on H2S (polar gas) adsorption is more significant compared to CH4 (non-polar gas). 2.2.3.2. Azine. Azine functional group (@N–N@) is Lewis basic and hydrogen bond acceptor site. Since the nonbonding electrons of nitrogen atoms do not participate in aromatic-p resonance, the increased basicity if this group can be identified through lone pair-lone pair electron repulsion and the alpha effect. As a result, activity of azine function in different forms such as coordination site to metal ions, Lewis basicity toward Lewis acidic guests, and hydrogen bond accepting from hydrogen donor guest is increased. These chemical features help azine functionalized MOFs to be constructed and used as basic catalyst platforms for gas capture and separation, sensing, and adsorption of different chemicals (Table 1).
Fig. 46. Crystal structure of bio-MOF-1 showing 1D channels along the c crystallographic direction (above). Highlighted portion is Zn–adeninate chain, in which free amine groups are exposed to the nano space; plausible H-bonding interaction between adenine and TNP (down). Reprinted with permission from Ref. [400]. Copyright 2015 Wiley-VCH.
For instance, azine functionalized TMU-4 (with formula [Zn2(oba)2(4-bpdb)]n∙(DMF)x where 4-bpdb is 1,4-bis(4-pyridyl)2,3-diaza-1,3-butadiene and H2oba is 4,40 -oxybisbenzoic acid), TMU-5 (with formula [Zn(oba)(4-bpdh)0.5]n∙(DMF)y where 4bpdh is 2,5-bis(4-pyridyl)-3,4-diaza-2,4-hexadiene) and imine functionalized TMU-6 (with formula [Zn(oba)(4-bpmb)0.5]n (DMF)z where 4-bpmb is N1,N4-bis-((pyridin-4-yl)methylene)benzene-1,4 -diamine) MOFs have been used as basic catalysts and show 45, 100, and 38% conversion rates in Knoevenagel reaction, respectively [417]. This is because N-donor linkers with various basicities appear in the walls of all three MOFs, which is favorable for the catalytic reaction (Fig. 47). TMU-5 with higher efficiency has the most
Fig. 45. Proposed dual excitation pathways mechanism for photocatalytic reduction of Cr(VI) over NH-MIL-88B(Fe).
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
basic function because the basicity of azine function is improved by presence of electron-donating methyl groups. On the other hand, TMU-6 has the lowest efficiency because phenyl ring is placed between two nitrogen atoms of azine group leading to elimination of alpha effect which results in reduction of TMU-6 framework basicity. In another work, TMU-5 has shown very high affinity toward carbon dioxide and methane molecules [412]. The presence of high basic azine along methyl groups in the structure of TMU-5 leads to high zero coverage enthalpy for both gases (43 kJ.mol1 for CO2 and 28 kJ.mol1 for CH4). The very high affinity of TMU-5 toward CH4 leads to high capacity of TMU-5 in room temperature and pressure equal to 2.54 wt% which is higher that many MOFs with higher surface areas such as Co3(ndc)(HCOO)3 (l3-OH). Moreover, TMU-50 s framework with high basicity is applied in removal of hazardous metal ions (Cd2+, Co2+, Cr3+, Cu2+ and Pb2+) and the results show that for trace amounts of metal ions, the basicity of the N-donor ligands in the TMU-5 determines the adsorption efficiency of the MOFs for the metal ions [419]. Moreover, azine functionalized MOFs are applied as hydrogen bond acceptor sites for removal of phenol and sensing of picric acid [421]. The basic pore walls of TMU-5 increase its affinity toward the acidic pollutant by creating an interaction between (HO–) of phenol and picric acid with the azine group (Fig. 48).
Fig. 47. Structure of azine and imine functionalized linkers of TMU-4, TMU-5, and TMU-6.
35
2.2.3.3. Azo. Azo function contains two nitrogen atoms and like other nitrogen-based functional groups shows some chemical properties like Lewis basicity, hydrogen bond accepting, and polarity. The presence of azo group as photo-active function conjugated with p-system is of benefit for optimization of host-gust interaction and also improvement in luminescence properties of the MOF [760]. Considering these chemical properties, azo functional group is applied in the structure of MOFs and shows diverse properties in different applications like sensing and removal of pollutants, gas adsorption and separation, and catalysis (Table 1). Moreover, azo function contains one double bond and like alkene functional group it consists of two cis and trans conformations which can be transformed to one another by light or heat external source. The responsivity of azo-containing materials to light is one reason for their use in designing artificial switchable catalysts and other kinds of remote-controllable materials [761]. Especially, azobenzene is a functional component used in a wide array of photoresponsive molecules, chelators, and molecular machines. Azobenzene compounds are easily modifiable and cheap to prepare, presenting many opportunities for preparation of new materials. An important group of stimuli-responsive MOFs is based on light-sensitive azo function, especially azobenzene substitutes in the side-chain of the framework. Upon UV or visible irradiation, azobenzene can isomerize from the nonpolar, planar trans form to the nonplanar cis form. By irradiation with visible light or thermal relaxation, the cis azobenzene goes back to its thermodynamically stable trans form. So using organic ligands containing photoresponsive azobenzene side groups, the remote and on-demand control over certain chemical or physical properties of MOFs can be achieved. As a method to control the guest release, modulation in host-guest interactions and control over pore size of cavities of the framework are crucial conditions. In this view, using photoresponsive azobenzene can provide such conditions through light induced trans–cis isomerization. Particularly, photo-responsive azo functionalized MOFs in side chain are highly desirable for cargo release in on-demand drug delivery, modulation of CO2, and light hydrocarbons release. Also, azo functionalized MOFs in main chain are applied for guest uptake. As an example about the effects of azobenzene function as side chain in the structure of an MOF, PCN-123 displays light-triggered capture and release of CO2 (Fig. 49) [425]. After light irradiation, CO2 capacity decreased 26.6% immediately and 53.9% after 5 h exposure to UV light because of trans–cis photoisomerization of azobenzene groups in side chain of framework. CO2 capacity almost fully recovered after heating the sample at 60 °C for 20 h.
Fig. 48. Functionalized pore and host-guest interaction between TMU-5 and TNP. Reproduced with permission from Ref. [421]. Copyright 2016 American Chemical Society.
36
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 49. (Top) Trans-to-cis isomerization of the PCN-123 ligand induced by UV irradiation and the cis-to-trans isomerization induced by heat treatment. (Bottom) Schematic illustration showing the suggested CO2 uptake in MOF-5, PCN-123 trans, and PCN-123 cis. Reprinted with permission from Ref. [425]. Copyright 2012 American Chemical Society.
In another example with azo function in the main-chain of the framework, Hill and coworkers synthesized Zn(AzDC)(4,40 -BPE)0.5 MOF, where azobenzene is converted to 4,40 -dicarboxylate (AzDC) ligand that can be incorporated into MOF architectures along with pillar ligand trans-1,2-bis(4-pyridyl)ethylene (4,40 -BPE) which
Fig. 50. Dynamic photo-switching in the light-responsive MOF Zn-(AzDC)(4,40 BPE)0.5 leads to instantly reversible CO2 uptake. Reprinted with permission from Ref. [428]. Copyright 2013 Wiley-VCH.
shows cis–trans photoisomerizability when coordinated to a metal complex (Fig. 50) [428]. Zn(AzDC)(4,40 -BPE)0.5 can undergo dynamic light-induced structural flexibility, which results in large variations in CO2 uptake and low-energy CO2 capture and release. Zn(AzDC)(4,40 -BPE)0.5 exhibits unprecedented dynamic switching under CO2 adsorption with a 42% desorption capacity under static irradiation conditions and as much as 64%during dynamic measurements. As authors mentioned, light irradiation increased the MOF surface energy, in which intermolecular interactions between CO2 molecules and the surface were weakened, thus triggering instantaneous CO2 release. Wang and coworkers synthesized biocompatible azobenzene functionalized UiO-66-azo framework containing azobenzene function in side-chain as dangling pendant groups as an efficient platform for cargo release in on-demand drug delivery [433]. UiO-66-azo immersed in a Rhodamine B (RhB) aqueous solution and then washed several times with an aqueous b-cyclodextrin (b-CD) solution. Since, b-CD can form complex with the dangling trans-azobenzene groups on the surface, RhB cannot be released. Upon exposure to UV light, azobenzene in trans-conformation switches to cis-conformation which has weaker affinity toward bCD , resulting in the dissociation of the supramolecular complex and the release of the RhB cargo from the cavities. 2.2.4. Ionic N-based functions This group of nitrogen containing functions include ammonium (quaternary amine), imidazolium, and pyridinium. These cationic functions are able to interact with anionic, polar, and quadrupolar species through electrostatic interaction, dipole–dipole and dipolequadrupole interactions. Construction of MOFs with functional ligands containing these functions offers a new method for synthesizing ionic MOFs. 2.2.4.1. Imidazolium. As a proton-carrier nitrogen-containing heterocyclic function, imidazolium ring contains highly polar and positive charge separated nature and acidic hydrogen on the electron deficient carbenic carbon (Fig. 51). So hosts containing imidazolium function and guest molecules interact through different types of interactions such as charge–charge electrostatic, chargedipole/ quadrupole, and (C–H)+ Xn ionic hydrogen bond [762]. It is obvious that guest molecules with more polarity, charge magnitude, and dipole or quadrupole moments are more preferred by imidazolium ring. Also, methylene proton of imidazolium group has the ability to transfer through hydrogen bonding and imidazolium ring reorientation property adds flexibility to the host structure.
Fig. 51. Structure of imidazolium ring, carbene formation, and metalation on C2 atom.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Moreover, due to the organometallic chemistry of imidazolium ring, it is possible to deprotonate the acidic proton at the carbene position by applying a base which makes it readily available for formation of tethered NHC– metal complexes (Fig. 52) [763]. This carbenic site can be stabilized by the two nitrogen atoms near the carbene center. It has been realized that in addition to NHCto-metal r ? d donating bond, when bonded to a metal, NHC fragment is involved in NHC-to-metal p ? d donating and metal-toNHC d ? p* accepting bonds. Such characteristics lead to stabilization of a wide range of electron deficient metals through a p ? d donating bond and electron-rich metal by d ? p* accepting bond (Fig. 53). NHC-metal complexes are applied in catalytic cross coupling and addition reactions as efficient catalyst centers. Such properties of imidazolium ring introduce it as practical functional groups inside MOF structures which greatly influence the hostguest chemistry and structural features of MOFs (Table 1). Charged nature of imidazolium linkers inside MOFs provides suitable conditions for stronger interaction between host MOF and quadruple moment of guest molecules like carbon dioxide and hydrogen. For example, Hupp and coworkers applied NU-301
37
(with formula Zn2(2H-BTBA)0.5(IMTA)]nDMF where 2HBTBA = benzenetetrabenzoicacid and H2IMTA = N,N0 -bis(2,6-dime thyl-3,5-carboxylphenyl)imidazolium) and NU-302 (with formula Zn2(2Br-BTBA)0.5(IMTA) where 2Br-BTBA = dibromo-benzenetetra benzoicacid) for hydrogen adsorption and found that these structures have high affinity toward hydrogen molecules with Qst(H2) equal to 7.0 and 6.9 kj.mol1, respectively [476]. Based on simulation calculations, H2 molecules prefer to be adsorbed in zwitterionic region between two negatively charged Zn2(CO2)5 nodes (net charge: 1.0) and positively charged imidazolium ring (net charge: +0.5). All of the H2 binding locations maximize interactions between the H2 quadrupole and strongly charged atoms on the zwitterionic framework surfaces. Recently, Cao and coworkers synthesized an imidazolium functionalized MOF ((I)Meim-Uio-66) based on ionic ligand 2-(3-Ethyl-imidazol-1-yl)-terephthalic acid [(Etim-H2BDC)+(Br)] [464]. Because of the high affinity of (I) Meim-Uio-66 to CO2 molecules (Qst(CO2) = -44.2 kj.mol1) through charge (imidazolium)-quadruple (CO2) interaction, this MOF is applied in CO2 cycloaddition reaction as an efficient adsorbent for fixation of CO2 guest molecules inside the pores.
Fig. 52. Different methods for metalation of imidazolium ring on carbenic C2 atom.
Fig. 53. Representation of different modes of NHC-metal complex formation bonds. (a) Effects on N atoms on stabilization of carbenic imidazolium ring. (b) NHC to metal rdonation bond. (c) NHC to metal p-donation bond. (d) Metal to NHC p*-backbond.
38
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
It is a useful strategy to optimize host-guest interactions through charge-charge electrostatic interaction of positive imidazolium ring inside the framework and anionic guests. Wu and coworkers reported two imidazolium functionalized MOFs presenting distinct adsorption ability for anionic dyes [468]. The positively charged imidazolium moieties in the channels have high affinity for the anionic congored molecules while the natural and cationic dyes of oil red O, Sudan black B and methylene blue are not the favored guests. Ability of hydrogen transfer is a unique characteristic of a function which can be applied in designing of conductive MOFs. Kitagawa and coworkers designed an imidazolium functionalized Zn-based MOF (with formula [(Zn0.25)8(O)Zn6(L)12(H2O)29 (DMF)69(NO3)2]n where H2L = 1,3-bis(4-carboxyphenyl)imidazolium) with the ability of proton conductivity in fuel cells by applying AC impedance spectroscopy (Fig. 54) [480]. Proton conductivity depends on the amount of water inside the MOF pores. Water molecules in the channels take part in the Grotthuss type conduction of proton originating from the methylene protons of imidazolium groups.
Fig. 54. Diagram along the a axis of imidazolium functionalized Zn-based MOF depicting methylene groups of the imidazolium moieties aligned inside the channels. Reprinted with permission from Ref. [480]. Copyright 2012 American Chemical Society.
So far, carbene site metalation at the imidazolium ring in the heterogeneous catalysis of MOFs has been achieved by three methods: (I) in situ deprotonation or one-pot reaction through MOF synthesis for metals such as Cu; (II) post-synthesis modification for metals such as Pd and Ir; (III) pre-functionalization of the organic linker with metal centers such as Pd, Ir and Cu. Such NHC-metal MOF based catalysts have been applied in different reactions (Table 1). For example, Verpoort and coworkers postsynthetically modified imidazolium functions inside an MOF to NHCs-Pd metal complex for heterogeneous catalysis of sonogashira cross-coupling reaction [470]. The modified MOF presents a high density and uniform distribution of NHC-Pd active sites in the framework and retains its high catalytic activity for at least 4 cycles without losing its structural integrity. Introduction of imidazolium containing linkers, among the useful practical characteristics, confer such attractive structural properties like uncommon inorganic SBUs and structures (like diamond and rotaxane), multi or non-interpenetrated frameworks, flexibility and single-crystal to single crystal transformation [481,764–767]. Bharadwaj and coworkers have reported that, the MOF [Cd2(L)3(DMF)(NO3)](DMF)3(H2O)8 where H2L = 1,3-bis(4carboxyphenyl)imidimidazolium, shows multiple-fold interpenetration. Because of the coordination mode of the metal ion, the angular nature of the bent imidazolium ligand and partial rotation of the imidazolium moiety with respect to the benzene moiety, the grids are highly interpenetrated (Fig. 55) [481]. Any grid forms a wave like 2D sheet and these sheets develop a 3-fold interpenetrated network. In another work, Hupp and coworkers showed that by modifying the number of imidazolium rings inside the structure of MOFs, morphology or the degree of interpenetration can be controlled (Fig. 56) [766]. They concluded that incorporation of different numbers of imidazolium cores (1 or 2) into porous materials can impact the level of catenation or morphology. Although the positive charge of imidazolium ring induces strong electronic interactions with ionic and polar guests, imidazolium functionalized MOFs commonly suffer from two problems: (I) presence of anionic counter ions in the pores of the MOF for charge-neutralization which leads to blocked pores and reduced
Fig. 55. Multi- interpenetrated structure of [Cd2(L)3(DMF)(NO3)](DMF)3(H2O)8. (a) View of 1D channels of the polycatenated framework down the b-axis. (b) Schematic representation of the polycatenated framework viewed down the b-axis. Reproduced with permission from Ref. [481]. Copyright 2014 American Chemical Society.
Fig. 56. Effects of the number of imidazolium rings on the structure. Reprinted with permission from Ref. [766]. Copyright 2011 American Chemical Society.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
porosity; (II) strong interaction between positive charged framework and guest solvents which lead to hard evacuation and activation. As a result, these groups of functionalized MOFs are usually placed among the low or moderate BET surface area MOFs (nonporous to around 1000 m2.g1). 2.2.4.2. Pyridinium. Positively charged pyridinium function has been widely applied as new supramolecular building block in modern chemistry for decades because of several chemical properties including: Lewis acidity, p-deficiency, responsivity to photon, heat, and electricity. Derivative of 4,40 -bipyridine and pyridinium, commonly known as viologen/bipyridinium possesses excellent redox ability and can act as a good electron acceptor because of Lewis acidity and p-deficiency to form charge-transfer (CT) complexes with electron-rich species and in some cases with redox capability to undergo electron transfer via chemical, electrical, or optical stimuli [768,769]. In addition to the mentioned applications, the bipyridinium unit has also been widely used in the field of supramolecular chemistry as a key building block for the assembly of mechanically interlocked molecules and molecular machines. As a result, owing to the synergic effects of highly ordered structure of MOFs and chemical properties of bipyridinium/pyridinium functional groups it is possible to design new functional coordination polymers especially MOFs with predetermined or improved physical and chemical properties. Moreover, combination of these two factors with chemical properties of coordinated metal ions, results in materials possessing intriguing photoelectrochemical functions that have not been observed in a single bipyridinium. Pyridinium functionalized MOFs have been applied in enhancement of CO2 and H2 adsorption (Table 1). Because of the large quadrupole moment and high polarizability of the CO2 molecule, the cationic framework would lead to stronger interactions between the structure and CO2, while H2-framework interaction is enhanced through strong perturbations by free Lewis acidic pyridinium sites with 9.5 kJ.mol1 isosteric heat of adsorption at lower coverage which is comparable with open metals sites [504]. For example, a new UiO-67-type zirconium MOF material UiO-67bpy-Me (bpy = 2,2-bipyridine-4,40 -dicarboxylic acid, Me = methyl) was prepared by N-quaternization of the pyridine sites in UiO67-bpy [504]. After N-quaternization, the pristine neutral framework turned cationic. Cationic pyridinium decorated UiO-67-bpyMe with lower BET surface (1104 m2.g1) area and pore volume (0.526 cm3.g1) compared to UiO-67-bpy framework (2306 m2. g1 and 1.061 cm3.g1) showing 33% enhancement in CO2 uptake (76 cm3.g1 compared to 56 cm3.g1 at 273 K and 850 mmHg). Some chemical features of pyridinium ring include: (I) electrostatic interaction with anionic and highly electronegative species, (II) p-p stacking with p-rich compounds, (III) interacting with dipolar and quadrupolar molecules through its ionic features, and (IV) formation of electron donor-acceptor complex by charge transfer which is the basis for versatile application of pyridinium FMOFs in fields of sensing, removal, and separation (Table 1). MOFs, via electro-active sites, are very useful in construction of donor–acceptor-type smart chromic materials via arranging electron donor and acceptor sites into the regular structure of MOFs. It is known that stacking of the donor and acceptor groups is the main factor that influences the photo-induced electron transfer process. Pyridinium is a well-known electron acceptor unit able to interact with electron donor entities like benzoate/carboxylate, and through its responsiveness to external sources of photon, electricity, heat, and chemicals can show switching ability. Combination of responsiveness to external stimuli of pyridinium function and interaction with electron donor sites causes this group of functionalized MOFs to be photoresponsive through different processes like photochromism, thermochromism, photoluminescenceswitching, and multi-photon adsorption.
39
Wriedt and coworkers constructed a new MOF,[CdBr(L)](ClO4) 2DMF (1) by zwitterionic ligand 1,10 ,100 (benzene-1,3,5-triyl)tris(m ethylene)tris(4-carboxypyridinium)tribromide (H3LBr3) (Fig. 57) [485]. They mentioned that zwitterionic carboxylate-pyridinium ligand has been applied due to the well-separated intramolecular charges. Incorporation of such a ligand into MOFs can create charged organic surfaces within the pore environment as a new means of polarizing guest molecules. This effect has the potential to improve host-guest interactions, and hence to enhance adsorption enthalpies. More importantly, the fact that pyridinium-based zwitterions can be reversibly photo-reduced to radical species may lead to the elimination of any charge gradients, thus representing an unprecedented mechanism for designing on/off switchable adsorption sites. In particular, it is expected that this process does not lead to any major structural changes, as only the electrostatic surface is modified. To test this hypothesis the CO2 adsorption properties before and after UV exposure is characterized [497]. Calculated heat of adsorption exhibits a significant reduction from 40.5 kJ.mol1 to 27.3 kJ.mol1 at zero coverage before UV irradiation and afterward, respectively. Also, a significant decrease of 43.2% in CO2 uptake at 273 K and 1 bar was recorded for the UVirradiated sample (pristine, 18.3 cm3.g1; after UV irradiation, 10.4 cm3.g1) (Fig. 58). In synthetic and structural view, constructing zwitterionic MOFs with dual functionalized pyridinium-carboxylate ligand is a challenge. Since bipyridinium units have strong affinities to interact with carboxylate groups of neighboring ligands through + CO 2 N electrostatic interactions in the solid state, dense coordination polymers are obtained in most cases, however many
Fig. 57. Structure of (H3LBr3) ligand and charge-separated framework of [CdBr(L)] (ClO4)2DMF. Reproduced with permission from Ref. [485]. Copyright 2015 American Chemical Society.
Fig. 58. The ligand used for construction of zwitterionic MOF for CO2 adsorption and the effects of light on host-guest interaction through reversible conversion of pyridinium ring to its radical homolog. Reproduced with permission from Ref. [497]. Copyright 2016 American Chemical Society.
40
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
attempts have been made to synthesize this category of MOFs because of the interesting photochromic properties due to electron transfer from CO 2 groups to viologen units and strong electric field gradient on the MOF backbone due to the well-separated intraligand charged centers. Moreover, Lewis acidic sites inside MOF structures are mainly metal open sites, created when solvent molecules bound to metal ions are removed during the activation process. However, functionalization of MOFs with pyridinium group is a new method to introduce Lewis acidic sites inside MOF structures. In addition to synthesizing MOFs, pyridinium functionalized ligand has been applied in constructing metal-organic rotaxane frameworks (MORFs). In this regard using cationic pyridinium based axle and anionic sulfonate based wheel eliminates the need for independent counter ions in the construction of MORFs. 2.3. Oxygen-based functions Owing to the high electronegativity of oxygen atoms, functional groups included in this group are electron rich and highly polar consisting of hydroxy, ether, enoxide and other functions like azoxy, oxadiazole, and oxopyridine. As a result of high electron density and polarity these function can interact with electron deficient and polar/quadrupolar guest molecules very well. 2.3.1. Hydroxyl Due to high electronegativity of oxygen atom, polarized O–H bond, and acidic H atom, hydroxy function has some chemical properties enabling it to interact with different types of guests. Since hydroxy contains acidic hydrogen, it can be used in designing proton conductive materials with highly hydrophilic channels and can act as hydrogen donor, hydrogen bond donor and catalytically active acidic site. Also, because of its highly negative nature, it is Lewis base, electron donor with high polarity in a way that it can interact with different guests like hydrogen bond donors, Lewis acidic molecules, and molecules containing partially positive atoms. As a result of host-gust interactive ability of both oxygen and hydrogen atoms of hydroxyl function, hydroxyl functionalized MOFs are designed and applied for different purposes such as improvement of CO2 capture and separation, designing efficient proton conductor frameworks, removal and sensing of hazardous chemicals with different mechanisms (Table 1). Similar to amine (–NH2), uncoordinated hydroxyl group (–OH) inside the MOF cavities is another functional group that shows very high affinity toward CO2. Schröder and coworkers synthesized NOTT-300 (with formula [Al2(OH)2(L1)](H2O)6) where H4L1 is biphenyl-3,30 ,5,50 -tetracarboxylic acid) as a hydroxyl containing framework to remove harmful CO2 and SO2 gases [530]. In situ powder X-ray diffraction and inelastic neutron scattering studies combined with modeling reveal that hydroxyl groups bind to CO2 and SO2 through the formation of O = C(S) = O(d). . .H(d+)–O hydrogen bonds, which are reinforced by weak supramolecular interactions with C–H atoms on the aromatic rings of the framework. In some instances it is mentioned that there is higher affinity
Fig. 59. H2BDC based ligand for construction of MOFs based on hard-soft chemistry post-synthesis methods.
between hydroxyl group and CO2 compared to amine through hydroxy(O) (C)CO2 electron donor–acceptor and hydroxyl(H) (O)CO2 hydrogen bond reinforced by dipole-quadrupole interactions. Li and coworkers reported that among the two isostructure hydroxy and amine functionalized MOFs, [Zn(BDC-OH)(TED)0.5] 1.5DMF0.3H2O (2) and [Zn(BDC-NH2)(TED)0.5]xDMFyH2O (1), hydroxy functionalized MOFs show higher affinity toward CO2 (24.2 kJ.mol1 for 2 and 19.8 kJ.mol1 for 1) and H2 (5.5 kJ.mol1 for 2 and 5 kJ.mol1 for 1) at zero coverage [301]. PCN-222 can provide high density of hydroxyl groups and be applied for removal of chloramphenicol drug from wastewater [533]. Investigations indicate that hydrogen bond interaction, electrostatic interaction, and special pore structure of PCN-222 all have important effects on high-efficiency removal of chloramphenicol. PCN-222 exhibits a large adsorption capacity equal to 370 mg.g1 and more importantly, the adsorption equilibrium can be quickly obtained at only 58 s. Furthermore, 99.0% of chloramphenicol in low concentrations can be removed from water. H atom of hydroxy function can be replaced by cationic metal ions for different purposes. For example Li+-H+ exchange is carried out for improvement of CO2/H2-framework interaction through increasing the framework polarity. The protons of the hydroxyl group are exchanged by metal cations via solution methods. This metal-exchange procedure is conducted for some MOFs to improve the interaction between hydroxy functionalized MOFs and guest gas molecules like CO2 and H2. For example, the H atoms of hydroxyl functions of IR-MOF-8-OH and IR-MOF-14-OH are exchanged with Li+ ions to improve the framework-H2 interactions [526]. In another work, hydroxyl-functionalized (Zn2(TCPB)(DPG)) framework, TCPB = 1,2,4,5-tetrakis(4-carboxyphenyl)-benzene and DPG = meso-1,2-bis(4-pyridyl)-1,2-ethanediol, shows improved affinity to CO2 molecules after Li+-H+ cation exchange through enhanced solid–gas interactions [114]. In another work, Cu+-H+ exchange was conducted to provide proper catalytic sites in CO2 hydrogenation reaction [770]. Hard-soft chemistry is a very promising strategy to control structure and functionality in ligands such as the following (Fig. 59): 1,4-benzenedicarboxylic acid (H2BDC), 2,5-dihydroxy-1, 4-benzenedicarboxylic acid (H2BDC-(OH)2) and 2,5-diacetoxy-1,4 -benzenedicarboxylic acid (H2BDC-(OCOCH3)2). Combination of 1,4-benzenedicarboxylic acid and 2,5-dihydroxy-1,4-benzenedicar boxylic acid with Zn2+ ions leads to construction of MOF-5 and MOF-74. Because both hydroxy and carboxylate groups are hard and have high affinity to metal ions, they coordinate to metal ions to form different structures. To achieve hydroxy functionalized MOF-5, hydroxy functions should be protected. After synthesis of the framework with protected acetoxy functions, they are subsequently deprotected [771,772]. So the pore walls are decorated with hydroxy groups (Fig. 60). In another work, it is argued that control over topology and interpenetration can be achieved by introduction of hydroxy function [773]. Cohen and coworkers reported that combination of methoxy and hydroxy functionalized 1,3,5-tris(4-carboxyphenyl)benzene ligands and Zn2+ions leads to development of isostructural and functionalized analogue of MOF-177 (MOF-177–OMe) and a rare interpenetrated hydroxy functionalized pcu-e framework, respectively [774]. 2.3.2. Ether Ether (C–O–C) function as another oxygen containing function applied in the structure of MOFs. Because of the presence of oxygen atom, ether function can interact with electron deficient, Lewis acid, and hydrogen bond donor species. Therefore, functionalization of framework voids with etheric groups provides weakly Lewis basic and polar media for guest molecules. On the other hand, ether functionalized MOFs show remarkable flexibility in main or side chain because of the free rotation in C–O–C motif. As a result
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
41
Fig. 60. Protection of H2BDC-(OH)2 by acetoxy groups and deprotection of H2BDC-(OCOCH3)2 ligand to achieve hydroxy functionalized MOF-5. Reproduced with permission from Ref. [772]. Copyright 2009 American Chemical Society.
Fig. 61. Representation of the interaction between Cd-EDDA and (a) hydrochloric acid or (b) sodium hydroxide computed by molecular force field based calculations. Reprinted with permission from Ref. [550]. Copyright 2016 American Chemical Society.
of these features, ether incorporated MOFs are applied as platforms for breathing and flexible MOFs. In terms of designing functional and flexible MOFs, introduction of some functions like (–O–), (–NH–) and (–S–) groups is a good strategy because of the free
rotational movement around (C–X–C, X = O, S, NH). The flexibility of this group of MOFs can lead to structural transformation upon guest adsorption/removal and change in temperature or pressure. But, there is one important point to consider. O, S, and NH groups have to be connected to at least one alkyl group like –CH2–. Some studies have reported that flexibility and polarity of ether groups in side or main chains lead to selective capture of CO2 over N2 and CH4. Zhao and coworkers reported that by changing the length of alkyl ether chains (–O–(CH2)n–O–CH3), interaction energies between H2 and MOF-5 framework is improved [548]. According to chemical calculations, the distance between H2-linker changes for all the structures and becomes less than 3 Å, indicating the occurrence of strong interactions. The highest binding energy has been observed between hydrogen molecule and linker 2,5-bis (4-methoxybutoxy) benzene dicarboxylic acid, possessing the longest chain of alkyl ether which indicates that the energy of interaction of hydrogen with substituted benzene is slightly enhanced by the addition of electron-donating groups. Fischer and coworkers showed that through functionalization of phenyl ring in 2 and 5 positions with etheric groups inside the rigid framework of MOF-5, an unexpected structural flexibility can be achieved [543]. They demonstrated that structural integrity upon activation and removal of polar solvent guests is related to the chain length of the alkyl ether groups and the substitution pattern of the bdc-type linker. The density of flexible ether groups in the framework is of great significance in a way that flexible and polar groups act as molecular gates at the pore apertures of the porous networks. Due to the polar nature of the ether chains, polar molecules such as CO2 can easily penetrate through the molecular gate,
Fig. 62. Proposed interaction between N-oxide functionalized framework and carbon dioxide. Reprinted with permission from Ref. [555]. Copyright 2013 Royal Society of Chemistry.
42
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
whereas N2 molecules cannot. In another work based on ether functionalized MOFs, Li-crown ether groups were introduced inside the structure of MOFs for improvement of CO2 and H2 affinity toward CO2 molecules through electrostatic interaction of Li centers with CO2 and H2 quadrupole [549]. Wu and coworkers synthesized an electron-rich ether functionalized MOF, Cd-EDDA where H4EDDA is 5,50 -(ethane-1,2-diylbis(o xy))diisophthalic acid, as a Luminescent Probe for the regenerable ratiometric sensor of pH responding to pH in the range from 2.0 to 11.5 in aqueous solutions [550]. The hydrogen bonding of H+/OH species with etheric oxygens of EDDA4 ligands is responsible for the variation in luminescence behavior in different pHs (Fig. 61). Foster and coworkers reported that incorporating alkyl ether functional groups between the layers is of benefit for exfoliation and interacting with solvent molecules to enhance dispersion [552]. Fischer and coworkers reported that modulated MOFs with flexible alkoxy functions leads to extreme thermomechanical properties like temperature responsive breathing and thermal expansion [553]. This behavior originates from increased thermal motion (vibrations, rotations) of the linear side chains with rising temperature in a way that chain length and chemical nature of the pendent side chains govern such particular properties of the thermo-responsive MOF.
2.3.3. Other oxygen-based functions In addition to the hydroxyl and ether functions, there are some other oxygen-based functional groups like azoxy, enoxide, and oxodiazole used in the structure of MOFs, where azoxy and enoxide functions can be obtained by oxidizing the azo and pyridinic nitrogen, respectively. Enoxide functionalized MOFs are applied in high CO2/CH4, CO2/ CO, and CO2/N2 separation and Cr(VI) adsorption as well as electrochemical Li-storage (Table 1). Jiang and coworkers reported that CO2 molecules lying alongside the N-oxide groups with the electron deficient C atom of CO2 form short contacts with the negatively charged O atom of N-oxide, and the electron rich O atoms of CO2 form short contacts with the positively charged N atom of N-oxide (Fig. 62) [555]. Another study reported that enoxide functionalized TMU-30 (with formula [Pb(INO)2]2DMF framework (where INO is isonicotinate N-oxide) showed very high affinity and capacity toward Cr2O2 in a way that the XPS results con7 firmed the formation of a new electrostatic interaction between the nitrogen of N-oxide groups from TMU-30 and oxygen atoms from chromate during the adsorption process [557]. Azoxy functionalized [Zn3L3(BPE)1.5]n MOF (H2L = 4,40 azoxydibenzoic acid, BPE = bis(4-pyridyl)ethylene) is applied in removal of Pb2+ ions showing very selective capture of Pb2+with high capacity (616.64 mg.g1) and efficiency (>99.27%) against background ions (Fig. 63) [560]. XPS analysis reveals that strong electrostatic attraction and coordination interactions between the highly accessible (O) of azoxy groups and Pb2+ ions is responsible for the adsorption process. In comparison with other similar functions such as thiadiazole and selenadiazole, oxadiazole shows higher affinity toward CO2 molecules (Fig. 64) [561]. Since the oxadiazole (6.23 Debye) has higher dipole moment compared to thiadiazole (4.80 Debye) and selenadiazole (2.61 Debye) the oxadiazole functionalized MOF shows higher affinity and capacity toward CO2. 2.4. Sulfur-based functions
Fig. 63. Possible interaction between azoxy functional framework and Pb2+ ions.
This major group of functional MOFs contains less common sulfur atom as a major fragment of the function. This group includes thiol, sulfide, sulfonate, sulfonic acid and some of other functions like thiourea, thiadazole, and thiocathecole. The advantages of this group of functionalized MOFs is that they contain a soft and electron rich sulfur atom which is of benefit for designing electron rich and polar frameworks with soft guest interactive sites.
Fig. 64. Representation of oxadiazole, thiadiazole, and selenadiazole ligands and their related dipole moment. Reprinted with permission from Ref. [561]. Copyright 2016 Royal Society of Chemistry.
2.4.1. Thiol and sulfide S atom containing functional groups including thiol and sulfide show similar chemical properties in comparison to hydroxy and ether, yet they are different in some cases. Regarding the higher electronegativity of S compared to H and C, S-center of thiols and sulfides are electron rich yet soft electron donors. In case of thiols,
Fig. 65. Adsorption of Hg ions on thiol sites of Zr-DM. Reprinted with permission from Ref. [572]. Copyright 2013 American Chemical Society.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
the (–SH) bond is polar with Brønsted acidic hydrogen. In case of sulfides, they can show flexibility around C–S–C bond as well as ethers. However, despite such advantages, S-containing FMOFs are limited because (I) compared to O atom, S atom is heavier leading to lower surface area (cm2.g1) and (II) the (–S–) fragment is partially oxidizable in higher temperatures. However thiol and sulfide functional groups are applied for removal and sensing of softheavy metals especially Hg2+ (Table 1). Through ‘‘hard-soft” strategy for designing functional MOFs, ZrDMBD has been synthesized by reacting ZrCl4 with 2,5-dimer capto-1,4-benzenedicarboxylic acid (H2DMBD) featuring the hard carboxy and soft thiol functions with the carboxy bonded to the hard Zr(IV) center, and the thiol groups decorating the pores (Fig. 65) [572]. About 10 mg of Zr-DMBD captured Hg2+ with an efficiency of over 99% from aqueous solution when the initial Hg2+ concentration was as low as 10 ppm. IR and Raman spectra of the Hg2+@Zr-DMBD sample showed that the characteristic S–H stretching at 2560 cm3 for Zr-DBMD became absent in the Hg2+@Zr-DMBD and a strong band at 355 cm-1 appeared which is consistent with the Hg–S stretching bond. Moreover, mercury vapor sorption experiments show that Zr-DBMD can adsorb mercury vapors when immersed in a sand bath and heated up to 140 °C for 24 h. Additionally, since the C atom of CO2 is electron-deficient, Sdonor ligand can interact strongly with carbon dioxide. Konarand and coworkers reported that thiophene functionalized [Cu(tdc) (bpe)]n2n(H2O)n(MeOH) MOF (H2tdc = 2,5 thiophenedicarboxylic acid; bpe = 1,2-Bis(4-pyridyl)ethane) has very high affinity toward CO2 with a Q0st of approximately 41.79 kJ.mol1 [573].
2.4.2. Sulfonate Sulfonate (–SO 3 )/sulfonic acid (–SO3H) is another S-based function which is extensively applied in the structure of MOFs as coordinating and guest interactive site. Since the sulfonic acid is a strong Brønsted acid, sulfonic acid functionalized MOFs are applied as catalyst in acid catalyzed reactions and since the Brønsted acidity of MOFs is highly correlated with their proton conductivity, sulfonic FMOFs are used as platforms for proton conductivity. Functionalized UiO-66 framework with highly acidic and strongly hydrophilic (–SO3H) functional group shows 0.34 102 S.cm1 conductivity at 303 K and 97% room humidity with the Ea value of 0.27 eV [584]. Molecular simulations show that the functional groups in UiO-66-SO3H form hydrogen bonds with the adsorbed water molecules to originate H-conduction. The high proton conductivity of the functionalized UiO-66 MOF can be mainly attributed to a cooperative effect of the strong hydrophilicity and
43
high acidity by sulfonic which leads to the formation of highdensity hydrogen bond networks. On the other hand, sulfonate group is very polar, and high polar porosity is achieved by decoration of the surface of the cavities inside the structure of MOFs. As a result, the CO2-sulfonate interaction obviously plays an important role in increasing the affinity between CO2 molecules and MOF skeleton and consequently the CO2 capture and conversion (Fig. 66) [583]. This kind of hostguest interaction is confirmed by density functional theory calculations. Similar to CO2, functionalization of MOFs with sulfonate group is an effective way to improve H2 affinity toward the framework. Froudakis and coworkers reported that hydrogen adsorption of RMOF-14 is enhanced by incorporation of negatively charged sulfonate group [597]. As displayed by simulations, a single hydrogen molecule was bonded over the sulfonate group in an ‘‘end-on” configuration. Overall, sulfonate function plays a critical role in adsorption of hydrogen molecules. Jiang and coworkers reported that esterification reaction can be catalyzed by MIL-101-SO3H as a photocatalyst platform (Fig. 67) [598]. As a model reaction, esterification reaction between benzyl alcohol and acetic acid is conducted by MIL-101-SO3H as a catalyst resulting in 83.8% conversion with 100% selectivity in 322 min. Light irradiation can greatly boost the catalytic process over MIL101-SO3H and light-assisted conversion has reached 97% in 160 min while no similar light enhanced phenomenon can be observed in MIL-101. Apparently, light irradiation has a significant effect on the reactivity of MIL-101-SO3H and acidic sites effect inside the MOF should also promote the process. The mechanism of light-enhanced acid-catalytic reactions over MIL-101-SO3H is illustrated by authors in 4 steps: (i) light irradiation; (ii) electron
Fig. 67. Light-enhanced acid-catalytic reaction over MIL-101-SO3H, in which electron is transferred from –SO 3 groups to Cr-oxo clusters facilitating the release of protons from the MOF. Reproduced with permission from Ref. [598]. Copyright 2018 Royal Society of Chemistry.
Fig. 66. Affinity between CO2 and free sulfonate groups. Reproduced with permission from Ref. [583]. Copyright 2016 American Chemical Society.
44
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
transfer from sulfonate group to Cr-oxo clusters; (iii) expediting H+ transfer and acidity enrichment; and (iv) boosting catalyst activity. Based on the calculated results, the HOMO is dominated by –SO 3 group while the LUMO is dominated by Cr-oxo clusters. The release of one proton (H+) leaves one negative charge localized at –SO 3, making the orbital energy level of -SO3 approximate to the HOMO. Thus, the electrons in –SO 3 can be excited to the LUMO if sufficient energy is provided by light. Sulfonate and its conjugated sulfonic acid can show different behaviors in sensing and removal of analytes. As Lewis basic site, Sulfonate (–SO 3 ) can detect metal ions through Lewis basic sites while Sulfoxy group (–SO3H) as an acidic site can bind to basic molecules to improve adsorptive removal capacity of adsorbent MOFs. Also, electrostatic attraction between MOFs containing highly polar and active –SO 3 groups is expected to facilitate the adsorption of cationic guests via electrostatic attraction. As coordinating sites, sulfonate-based ligands have certain advantages in construction of CPs including: [694,775]. (I) Sulfonate groups are considered as poor ligands in coordination chemistry. As a result of such weak metal-sulfonate interactions, sulfonate-based ligands are used to construct soft coordination polymers showing dynamic structural flexibility as well as single crystal to single crystal transformation. It may be argued that sulfonate based frameworks are not sufficiently robust to tolerate permanent porosity. (II) Sulfonate group can be considered as a trioxy anionic unit with local C3v symmetry. This structure offers special ligating directionality compared to planar carboxylate which is responsible for increasing the dimensionality and connectivity of frameworks in order to construct rigid and stable structures. Comparing (I) and (II) indicates that dynamic-flexible behavior of sulfonate- based coordination polymers (CPs) is more dominant than their robustness. (III) Considering the greater possible variations in coordination modes of sulfonate, a lower degree of predictability is anticipated while allowing for a higher degree of polymorphism. This may lead to lower crystallinity and poorly regular inorganic assembly. (IV) Achieving polar pores is easier in sulfonate than carboxylate because normally two oxygen atoms of carboxylate are coordinated to metal ions providing non-polar porosity. But in case of sulfonate, normally the pores contain pendant S–O groups making the pores polar. Carboxylate groups coordinate to different metal sites with different synthesis methods and conditions. However for sulfonate, hydrothermal and solvothermal conditions improve the coordination ability of sulfonate ligands using Ag+, alkaline and alkaline earth, transition and Lanthanide(III) metal ions. Therefore, when comparing carboxylate and sulfonate we can argue that sulfonate-based frameworks are less predictable and crystalline with lower porosity yet higher flexibility, thermal stability, and polarity. To alleviate the disadvantages and combine the advantages of both sulfonate and carboxylate based frameworks, sulfonate-carboxylate ligands have attracted much interest for construction of novel frameworks. The strong crystalline coordination ability of carboxylate groups and the high dimensionally dynamic nature of sulfonate groups in sulfonate-carboxylate ligands expand the number of possible geometrical configurations between the O-donors and the metal ions, allowing for the formation of poly-nuclear structures.
2.4.3. Other sulfur-based functions In addition to thiol, thioether, and sulfonate ligands, there are some of other S-based functions like thiourea, thiadazole, thiocathecole, and sulfenyl iodide which are applied in the structure of MOFs.
Wang and coworkers reported that post-synthesized IRMOF-3– thiourea showed high activity and selectivity in acetalization and Morita–Baylis–Hillman reaction (Fig. 68) [607]. Halogenated aromatic rings such as F and Br, benzaldehyde and 2Hcinnamaldehyde were evaluated with regard to a range of acetalization and Morita–Baylis–Hillman reactions. The catalyst did not suffer from any leaching problems during catalysis and could be recycled several times without loss of activity (yield over 81%) or selectivity. This heterogeneous thiourea incorporation strategy overcomes self-aggregation and solvation issues that exist in a homogeneous thiourea catalyst. 2,3-dimercaptoterephthalate (tcat) ligand has been incorporated in the structure of UiO-66 by post synthetic modification process [776]. Two ortho-thiol functions of tcat ligand have been applied as encapsulation sites for immobilization of soft Pd2+ ions to synthesize Uio-66-tcat-pd which is catalytically active for regioselective functionalization of sp2 C–H bond. Thiadazole function has been applied as guest interactive site toward CO2 and Cd2+ guests (Table 1). In another work by Xu and coworkers, ZrDMBD as a Zr(IV)-based MOF decorated with free-standing thiol (–SH) groups was synthesized and reacted with I2 molecules to form sulfenyl iodide (S-I) units. [608]. The sulfenyl iodide function, generally found to be unstable and elusive in free-flowing solution conditions, has been conveniently generated and stabilized by reacting I2 with the thiol-functionalized solid grid of ZrDMBD. This simple and effective solid host also captures the spatial confinement observed for the complex biomacromolecular scaffolds involved in iodine thyroid chemistry, wherein the spatial isolation and consequent stabilization of sulfenyl/selenenyl iodides are exerted by means of the protein scaffolds. 2.5. Other functional groups In addition to the functional groups discussed in the previous sections, some others have also been applied in the structure of MOFs yet they not discussed as thoroughly as the previous functions. These are halogen-based functions especially fluorine, nitro, and phosphonate functional group. 2.5.1. Halogen-based functional groups Fluorine atom is highly electronegative and lowly polarizable. These two characteristics result in a very polar C–F bond with high density of negative charge on fluorine atom as well as hydropho-
Fig. 68. Proposed acetalization reaction mechanism and determination of firstorder reaction kinetics of the catalyst. Reproduced with permission from Ref. [607]. Copyright 2014 Royal Society of Chemistry.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
bicity of this bond. As a result, fluorinated organic ligands can be applied in the structure of MOFs for improving the gasframework interaction through high polarity of C–F bond and increasing the water stability and hydrophobicity by tuning the pores with fluorine [777–779]. In addition to hydrophobicity, fluorophilicity can be tuned as well to interact with special types of guests. Most halogen functionalized MOFs are based on F atom, but in some cases Cl and Br atoms are applied as substitutes in side chain of MOF aromatic skeleton to improve gas adsorption performance (Table 1). Fluorine containing ligands are classified in two groups: organic and inorganic. Organic-based fluorine containing ligands are in fluoro (–F), trifluoromethyle (–CF3), and perfluoroalkyle (–(CF2)n–) forms while inorganic-based fluorine containing pillars are SiF2 6 , 2 TiF2 , NbOF 6 , BF4 , PF6 , AlF5(H2O) 5 , and CF3SO3 . In both organic and inorganic forms, high electronegativity of fluorine atom with high density of negative charge makes it a preferable adsorption site in the cavities of MOFs. Chen and coworkers reported that non-fluorinated NOTT-101a and fluorinated NOTT-108a (with formula [Cu2(L)(H2O)2]5DMF3H2O where H4L is tetrafluoro-[1,10 :40 , 100 -Terphenyl]-3,300 ,5,500 -tetracarboxylic acid) have very similar porosity but NOTT-108a exhibits a higher volumetric methane storage capacity of 247 cm3(STP).cm3 and a working capacity of 186 cm3(STP).cm3 (at 298 K and 65 bar) compared to 237 cm3(STP).cm3 and 181 cm3(STP).cm3 for NOTT-101a, which is attributed to the higher polarity/dipole moment of C–F bonds compared to that of C–H bonds, resulting in enhanced electrostatic interaction with methane molecules [635]. In another work Space and coworkers showed that when H2 molecules are located near the fluorine atoms, each positively charged H atom of the H2 molecule can interact with an electronegative fluorine atom improving experimental gas adsorption results [640]. But, Rahul Banerjee and coworkers claim that high H2 uptake in fluorinated MOFs is not a universal phenomenon and it is rather system-specific and differs from system to system [611]. Eddaoudi and coworkers reported that using bulky NbOF 5 fluorinated inorganic pillar instead of PF 6 resulted in limited maximum open pore size for KAUST-7 (NiNbOF5(pyrazine)22H2O) framework which is of benefit for selective molecular exclusion of propane from propylene at atmospheric pressure [647]. Zaworotko and coworkers constructed SIFSIX-3 (with formula [M (pyrazine)4(SiF6)], M = Fe or Ni) by pyrazine and SiF2 units for 6 preferential adsorption of Xe over Kr [780]. Results obtained for Xe/Kr separation is attributed to the high polarizability of SiF2 6 anions, which can strongly interact with polarizable gases such as Xe, and the optimally tuned pore size that is commensurate with the size of Xe atom. Fluorine functionalized MOFs show some unique characteristics like hydrophobicity through the presence of alkyl or fluorinated (F, CF3. . .) groups, which is an efficient way to improve the water stability of a given MOF. (–CF3) functionalized FMOF-1 shows superior hydrophobic characteristics [778]. No adsorption step and thus no water uptake were observed for FMOF-1 even near the saturation pressure. As a result of such hydrophobicity, FMOF-1 does not suffer from degradation upon long-term exposure to boiling water which is consistent with the absence of any water adsorption. 2.5.2. Phosphonate Phosphonate-based ligands are effective chelating agents and form stronger bonds with metal ions compared to carboxylate. They can form a large variety of compounds including metalphosphonate molecules and organic-inorganic hybrid materials (Fig. 69) [781]. Phosphonate-based ligands are applied in the synthesis of coordination polymers more than sulfonate yet less than carboxylate.
45
However, in comparing phosphonate and carboxylate coordinating sites some points should be considered: [694,781–783]. (I) Regarding the formation of phosphonate-metal coordination bonds, with increasing the metal ion valance, the solubility of metal-phosphonate will decrease in a way that tetravalent metal-phosphonates are insoluble even in strong acids. Therefore, it is difficult to solve high valance metal-phosphonate to obtain crystalline frameworks. As a result, in contrast to Zr4+–O(carboxylate) bond which is very strong and forms highly crystalline, porous, and stable 3D coordination polymers, coordination bond between phosphonate and highly charged tri- and tetravalent metal ions (like Al3+, Pb4+ and Zr4+) are porous, air and water resistant and thermally stable yet they are not crystalline. (II) Phosphonate group has high affinity to bond maximum metal ions in a way that three O atoms of phosphonate can bind to eight metal centers. Therefore, simple metal-phosphonate can be altered to bind to maximum metal ions and construct dense metal-phosphonate layers. However, in these conditions the porosity of the framework is not maintained which is a challenge for designing porous phosphonate coordination polymers. (III) Similar to sulfonate, higher number of P–O coordinating sites, make the structure predictability hard because of the presence of additional coordinating O atoms. Coordination modes of phosphonate show both similarities and dissimilarities to carboxylate ligand. Considering (I), (II), and (III) we can argue that porous crystalline phosphonate-based MOFs are rare and some of them contain interlayer porosity. On the other hand, permanent porous phosphonate based MOFs are poorly crystalline with high chemical and thermal stability. In case of porous and crystalline coordination polymers based phosphonate ligands; rational design of target material is nearly impossible which means that crystalline phosphonate-based MOFs must be structurally characterized by powder X-ray diffraction (PXRD) instead of single-crystal methods. (IV) Phosphonate-based MOFs are less soluble with strong PO– M bond, increasing their stability against heat, air, and humidity. (V) Similar to sulfonate, phosphonate groups can provide polar porosity through their free (PO) groups in the pore walls. Phosphonate-based MOFs have distinct differences from carboxylate- based MOFs that make them attractive candidates for porous materials. These differences include their thermal stability and extremely low solubility. However, this often makes it difficult to obtain single crystals, which is now routinely used to determine crystal structures. To overcome these challenges of synthesis of phosphonate MOFs and benefit from the advantages of phosphonate MOFs such as formation of stronger metalphosphonate bonds, high thermal stability, and low solubility, phosphonate containing ligands are functionalized with other binding groups like carboxylate and azolate functions since these functions have the advantage of crystalline and directional coordination ability in bridging metal ions. Therefore, there is an increasing number of mixed carboxylate-phosphonate MOFs and these materials can serve to bridge the gap between these two types of MOFs, combining the most desirable attributes of both. Overall, we can say that in comparing carboxylate and phosphonate based MOFs, the latter show higher polarity, thermal and humidity stability coupled with lower solubility, crystallinity, and isoreticularity. Recently phosphonate based MOFs have been applied in proton conductor, gas adsorbent, and sensor-based MOFs (Table 1). Since the rigid phosphonate group can act as sensitizer or antenna group in energy or electron transfer mechanism with other parts of the framework, phosphonate based MOFs can serve as luminescence sensor for detection of different analytes, especially when they are combined with electroactive metal centers. Regarding the high polarity and polarizability of phosphonate groups (–PO3H or –PO 3)
46
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 69. Coordination modes of carboxylate and phosphonate. (a) Similarity between phosphonate and carboxylate ligands in terms of coordination modes. (b) Differences between phosphonate and carboxylate ligands in terms of coordination modes. Reprinted with permission from Ref. [781]. Copyright 2015 American Chemical Society.
they can be applied in designing highly polar channels in MOFs for strong interaction with ionic/polar/quadrupolar guests like CO2. As a good candidate to tune the Brønsted acidity and hydrophilicity of channels, phosphonate incorporated MOFs with uncoordinated phosphonate oxygen atoms provide highly hydrophilic channels for reversible proton transfer through hydrogen bonding with water molecules.
3. Conclusion and outlook Function-Application properties. For further clarification of the relationship between functional groups and FMOF applications, related function-application properties are summarized within some figures. There are two important points about these figures: (I) some functions like amide and oxalamide, pyridine and diazines, triazine and tetrazines, heterocyclic azoles and thiols and sulfide are mentioned together, because they display similar chemical characteristics, (II) others including squaramide, carbonyl, heptazine (tri-s-triazine), azine, hydrazone, imine, N-oxide, azoxy, oxadiazole, thiourea, thiocathecole, thiadazole, halogens, nitro, and ammonium functions are not mentioned because the published articles about these functional groups were not enough to allow for a correct comparison and conclusion. Fig. 62 shows that in some applications almost all functions are used and authors have called them general applications including gas adsorption, catalysis, removal, and sensing. In other applications like antenna, energetic MOFs, stimuli responsiveness and ion storage and conductivity only some of the functions have been
used because of their special characteristics. We refer to them particular applications. In gas adsorption, it is clear that major groups include amine, heterocyclic azoles, amid and oxalamide, and pyridine and diazines. Amine and heterocyclic azole functions show higher performance in CO2 separation through their Lewis basicity, polarity, and hydrogen bonding, while pyridine and especially diazines play unique roles in high performance methane storage. Amide and oxalamide show noticeable improvements in CO2 and especially C2H2 separation/storage through multiple hydrogen bonds and ppolarization (Fig. 70a). FMOFs are applied as heterogeneous catalysts especially as basic catalysts and platforms for CO2 fixation and activation by amines, NHC-heterocyclic-metal catalytic centers by imidazolium ring, hydrogen bond donor sites by urea, and Brønsted acidic by carboxylate and sulfonate. The combination of catalytic activity of functional groups and heterogeneous nature of MOFs provides very unique opportunities for designing FMOF-based catalysts (Fig. 70b). Since sensing of various analytes requires different guestinteractive sites as sensing of analyte can be provided by different host-guest interactions, almost all functions show good results in sensing of chemicals. In case of pyridinium and NDI core, they can provide a wide range of host-guest interactions and these functions are frequently applied for construction of FMOF-based sensors. Interactive features in some functions like hydrogen bonding for urea, chemo-switchability, p-p stacking and Lewis basic interactions for triazine and tetrazine, Lewis basicity and hydrogen bonding characteristics of amine and hydrogen bonding, chelation ability and electron donation properties of carboxylate
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
47
Fig. 70. Representation of the number of published papers on general applications of FMOFs. (a) gas adsorption, (b) catalysis, (c) sensing, and (d) removal.
functions are the reasons why these functions are broadly applied as FMOFs-based sensors (Fig. 70c). In adsorbent MOFs for pollutant analytes, amine functionalized MOFs are applied most extensively because high surface area MOFs can be decorated easily by amine functions in the side-chain of the framework and both high surface area and functionalization of framework along with a multi-character function is provided. However, urea and thiol/sulfide functionalized MOFs are flexibly applied for removal of electron-rich species and soft and heavy metal ions through hydrogen bonding and soft-soft acid-base interactions, respectively (Fig. 70d). Some functions like amine, NDI core, oxalamide, phosphonate, and sulfonate act as antenna which means that these functions are able to adsorb photons and transfer it through energy or electron transfer to other parts of the structure especially metal clusters for photocatalysis applications and sensitizing framework against different analysts (Fig. 71a). Tetrazole and triazole functions are almost only applied for construction of energetic MOFs. However presence of some auxiliary functions or substitutes likes azide, azo, and nitro is of benefit for enhancement of their performance (Fig. 71b). The issue of ion storage and conduction by MOFs mostly focuses on proton conductivity. So most of the functions that are applied in this field are Brønsted acidic like phosphonate, carboxylate, hydroxy, imidazolium, and specially sulfonate. Designing MOFs for Li-storage requires groups that are polar and redox active like amine and NDI core. In addition to Li-storage, NDI functionalized MOFs are applied in OH conductivity through providing free OH (Fig. 71c).
Stimuli responsive functions are active in presence of external stimuli like heat, pressure, light, electricity, and chemicals. Ether and azo functions show photo/thermo switchability behaviors which are very useful for control over structural properties and on-demand release experiments. Stimuli-chromic functions should be electronically active to show color change upon exposure to external stimuli especially photo-responsivity in case of pyridinium and NDI core and chemo-responsivity in case of tetrazine and triazine (Fig. 71d). Function-Structure properties. Overall there are some critical considerations about the effects of functional groups on the structure of FMOFs. If functional groups are applied as coordinating sites, these are some of the points that should be considered: (I) how functional groups affect the crystallinity, porosity, topology, stability, flexibility, and polarity as well as structural- predictability, (II) how synthesis conditions including solvent, temperature, pH, suitable metal ions, etc. can be provided for synthesis of a framework and better characterization and applications. However, if functional groups are applied as guest-interactive sites, then the following points should be considered: (I) How functional groups affect the orientation of coordination sites relative to each other. If an functional groups changes the orientation of coordinating sites relative to each other, the structure may change as well. However, in these cases the multi-chain and geometry of ligand should be considered. For example, if some functions like pyridine, diazines, tetrazine, azo, azine, amide, oxalamide, NDI core, and squaramide are applied in the structure of a ligand with linear chains, the direction of two carboxylate coordinating sites is
48
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
Fig. 71. Representation of the number of published papers for particular applications of FMOFs. (a) antenna effect, (b) energetic MOFs, (c) Ion storage and conductivity, and (d) stimuli responsiveness.
Fig. 72. Effects of functional groups in the main-chain of the ligand on the orientation of carboxylate groups relative to each other. (a) These functions do not change the orientation of carboxylate groups relative to each other in linear ligand. (b) These functions change the orientation of carboxylate groups relative to each other in linear ligand.
Fig. 73. Effects of functional groups in the main-chain of the tri-chain and hexa-topic ligand on the orientation of carboxylate groups relative to each other. In all cases the orientation of carboxylate groups relative to each other is retained.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
not changed relative to each other and based on our observation it is anticipated that the structure of FMOFs will not change relative to the non-functionalized mother framework (Fig. 72a). But in case of functions like urea, ketone, triazole, and imidazolium because of the change in orientation of carboxylate groups relative to each other, it is anticipated that the structure will change (Fig. 72b). On the other hand, in case of functional ligands with three chains, presence of functions like urea and triazole does not change the orientation of carboxylate groups relative to one another and overall structure is the same as non-functionalized mother framework (Fig. 73). (II) How the functional groups affect the structure through structure-directing interactions? Is the structure stable after activation and elimination of secondary interactions? Since the secondary interactions between parts of framework in selfassembly process is very hard to predict as well as ligand directionality, predictability of structure-directing effects of the function is almost impossible. In this review, we have tried to collect information providing a deeper insight into the effects of functional groups on the structure and application of the metal-organic frameworks. To this end, we have gathered lots of papers on different functional groups in which they refer exactly to the impact of functional groups on structure or applications of FMOFs. In order to provide a better understanding of the subject, functional groups have been classified and through considering chemical properties of each functional group, they have been discussed and best examples have been briefly explained. The selected examples in this review capture major advances made in designing functional MOFs from the function-structure to function-application properties indicating the unique and practical accomplishments in strategic synthesis of functional MOFs. Since in this review our strategy was based on the idea that a discussion about each functional group should begin with an explanation about its chemical properties, the readers will easily understand which chemical properties of each function is more investigated and which one is less investigated. Also, the readers will find out which functional groups are suitable for any specific application and which useful properties of a functional group need to be further studied. The most important advantages of MOFs in comparison to other porous materials are as follows: crystallinity and highly ordered structure, high porosity and surface area, inorganic-organic dual character, adjustability in pore size and porosity by isoreticular synthesis, flexibility of framework and structural transformation as well as tunable chemical functionalization which is extensively addressed in this review. We argue that all other factors including crystallinity, porosity, stability, flexibility, topology, and host-guest chemistry of MOFs are influenced by functionalization. As we have seen, an important strategy that can show the effects of functional groups on host-guest chemistry and structure of MOFs is the comparison of functional and non-functional frameworks. In most cases, use of functional groups has improved performance, but in some cases, applying functional groups has shown some positive effects coupled with certain negative ones. To avoid this, chemists should be familiar with the chemical and structural properties of functional groups. A lot of research has been done on some of the functional groups, and for others, more research is needed. In some cases, where the applications of desired functional groups in the structure of MOFs has been limited, it possible to predict the chemical properties of the functionalized MOFs through considering the chemical properties of the simple functional group. In other words, to synthesize a functional MOF, we must consider the effects of functional groups on all of the factors mentioned above in order to obtain the best performance from the desired functional MOFs. This has been our goal in writing this review about functional metal-organic frameworks. We believe that the
49
future of MOFs mainly belongs to functional MOFs and through this review we tried to compile key informations regarding the effects of functionalization on metal-organic framework structure, host-guest chemistry, and applications. We hope this review to help scientists in designing functional MOFs for specific purposes. Acknowledgment Support of this investigation by Tarbiat Modares University is gratefully acknowledged. References [1] O.M. Yaghi, G. Li, H. Li, Nature 378 (1995) 703. [2] O. Yaghi, H. Li, J. Am. Chem. Soc. 117 (1995) 10401–10402. [3] S. Khazalpour, V. Safarifard, A. Morsali, D. Nematollahi, RSC Adv. 5 (2015) 36547–36551. [4] L. Panahi, M.R. Naimi-Jamal, J. Mokhtari, A. Morsali, Micropor. Mesopor. Mat. 244 (2017) 208–217. [5] B. Bhattacharya, D. Ghoshal, CrystEngComm 17 (2015) 8388–8413. [6] J.H. Bang, K.S. Suslick, Adv. Mater. 22 (2010) 1039–1059. [7] V.V.E. Butova, M.A. Soldatov, A.A. Guda, K.A. Lomachenko, C. Lamberti, Russ. Chem. Rev. 85 (2016) 280. [8] N.A. Khan, S.H. Jhung, Coord. Chem. Rev. 285 (2015) 11–23. [9] N. Stock, S. Biswas, Chem. Rev. 112 (2011) 933–969. [10] M.Y. Masoomi, A. Morsali, P.C. Junk, CrystEngComm 17 (2015) 686–692. [11] M.Y. Masoomi, M. Bagheri, A. Morsali, Ultrason. Sonochem. 33 (2016) 54–60. [12] M.Y. Masoomi, A. Morsali, Ultrason. Sonochem. 28 (2016) 240–249. [13] M. Bigdeli, A. Morsali, Ultrason. Sonochem. 27 (2015) 416–422. [14] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.-H. Bae, J.R. Long, Chem. Rev. 112 (2011) 724–781. [15] Y. He, W. Zhou, G. Qian, B. Chen, Chem. Soc. Rev. 43 (2014) 5657–5678. [16] L.J. Murray, M. Dinca˘, J.R. Long, Chem. Soc. Rev. 38 (2009) 1294–1314. [17] J. Liu, L. Chen, H. Cui, J. Zhang, L. Zhang, C.-Y. Su, Chem. Soc. Rev. 43 (2014) 6011–6061. [18] A. Corma, H. García, F. Llabrés i Xamena, Chem. Rev. 110 (2010) 4606–4655. [19] Y. Li, H. Xu, S. Ouyang, J. Ye, Phys. Chem. Chem. Phys. 18 (2016) 7563–7572. [20] T. Zhang, W. Lin, Chem. Soc. Rev. 43 (2014) 5982–5993. [21] L.E. Kreno, K. Leong, O.K. Farha, M. Allendorf, R.P. Van Duyne, J.T. Hupp, Chem. Rev. 112 (2011) 1105–1125. [22] I. Stassen, N. Burtch, A. Talin, P. Falcaro, M. Allendorf, R. Ameloot, Chem. Soc. Rev. 46 (2017) 3185–3241. [23] J.-R. Li, J. Sculley, H.-C. Zhou, Chem. Rev. 112 (2011) 869–932. [24] J. Li, X. Wang, G. Zhao, C. Chen, Z. Chai, A. Alsaedi, T. Hayat, X. Wang, Chem. Soc. Rev. 47 (2018) 2322–2356. [25] I. Imaz, M. Rubio-Martínez, J. An, I. Sole-Font, N.L. Rosi, D. Maspoch, Chem. Commun. 47 (2011) 7287–7302. [26] A.C. McKinlay, R.E. Morris, P. Horcajada, G. Férey, R. Gref, P. Couvreur, C. Serre, Angew. Chem. Int. Ed. 49 (2010) 6260–6266. [27] A. Morozan, F. Jaouen, Energy Environ. Sci. 5 (2012) 9269–9290. [28] P. Ramaswamy, N.E. Wong, G.K. Shimizu, Chem. Soc. Rev. 43 (2014) 5913– 5932. [29] J. Zhang, M.S. Jean’ne, Dalton Trans. 45 (2016) 2363–2368. [30] S.S. Nagarkar, A.V. Desai, S.K. Ghosh, Chem. Asian J. 9 (2014) 2358–2376. [31] J. Zhang, J.N.M. Shreeve, Dalton Trans. 45 (2016) 2363–2368. [32] F.A. Almeida Paz, J. Klinowski, S.M.F. Vilela, J.P.C. Tomé, J.A.S. Cavaleiro, J. Rocha, Chem. Soc. Rev. 41 (2012) 1088–1110. [33] B. Li, H.-M. Wen, Y. Cui, W. Zhou, G. Qian, B. Chen, Adv. Mater. 28 (2016) 8819–8860. [34] M.C. Das, S. Xiang, Z. Zhang, B. Chen, Angew. Chem. Int. Ed. 50 (2011) 10510– 10520. [35] Y. Cui, B. Li, H. He, W. Zhou, B. Chen, G. Qian, Acc. Chem. Res. 49 (2016) 483– 493. [36] A. Kirchon, L. Feng, H.F. Drake, E.A. Joseph, H.-C. Zhou, Chem. Soc. Rev. (2018). [37] E.A. Hall, L.R. Redfern, M.H. Wang, K.A. Scheidt, ACS Catal. 6 (2016) 3248– 3252. [38] C. Zhu, Q. Mao, C. Li, Y. Zhou, X. Wu, Y. Luo, Y. Li, Catal. Commun. 104 (2018) 123–127. [39] X.-W. Dong, T. Liu, Y.-Z. Hu, X.-Y. Liu, C.-M. Che, Chem. Commun. 49 (2013) 7681–7683. [40] P.C. Rao, S. Mandal, Chemcatchem 9 (2017) 1172–1176. [41] Z. Ju, S. Yan, D. Yuan, Chem. Mater. 28 (2016) 2000–2010. [42] X.-J. Wang, J. Li, Q.-Y. Li, P.-Z. Li, H. Lu, Q. Lao, R. Ni, Y. Shi, Y. Zhao, CrystEngComm 17 (2015) 4632–4636. [43] J.M. Roberts, B.M. Fini, A.A. Sarjeant, O.K. Farha, J.T. Hupp, K.A. Scheidt, J. Am. Chem. Soc. 134 (2012) 3334–3337. [44] Y. Uozumi, A. Tazawa, Synfacts 9 (2013) 1251. [45] P.W. Siu, Z.J. Brown, O.K. Farha, J.T. Hupp, K.A. Scheidt, Chem. Commun. 49 (2013) 10920–10922. [46] A.A. Tehrani, S. Abedi, A. Morsali, J. Wang, P.C. Junk, J. Mater. Chem. A 3 (2015) 20408–20415. [47] P.W. Seo, N.A. Khan, S.H. Jhung, Chem Eng J 315 (2017) 92–100.
50
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
[48] A.M. Ebrahim, T.J. Bandosz, Micropor. Mesopor. Mat. 188 (2014) 149–162. [49] S. Glomb, D. Woschko, G. Makhloufi, C. Janiak, A.C.S. Appl, Mater. Interfaces 9 (2017) 37419–37434. [50] P.W. Seo, N.A. Khan, Z. Hasan, S.H. Jhung, A.C.S. Appl, Mater. Interfaces 8 (2016) 29799–29807. [51] X. Yang, Y. Xia, Microchim. Acta 183 (2016) 2235–2240. [52] A. Hakimifar, A. Morsali, Inorg. Chem. (2018). [53] R. Custelcean, B.A. Moyer, B.P. Hay, Chem. Commun. (2005) 5971–5973. [54] B. Wu, J. Liang, J. Yang, C. Jia, X.-J. Yang, H. Zhang, N. Tang, C. Janiak, Chem. Commun. (2008) 1762–1764. [55] R. Custelcean, V. Sellin, B.A. Moyer, Chem. Commun. (2007) 1541–1543. [56] R. Custelcean, B.A. Moyer, V.S. Bryantsev, B.P. Hay, Cryst. Growth Des. 6 (2006) 555–563. [57] D. Krishna Kumar, A. Das, P. Dastidar, New J. Chem. 30 (2006) 1267–1275. [58] W. Liu, X. Huang, C. Xu, C. Chen, L. Yang, W. Dou, W. Chen, H. Yang, W. Liu, Chem. Eur. J. 22 (2016) 18769–18776. [59] A. Azhdari Tehrani, L. Esrafili, S. Abedi, A. Morsali, L. Carlucci, D.M. Proserpio, J. Wang, P.C. Junk, T. Liu, Inorg. Chem. 56 (2017) 1446–1454. [60] L. Esrafili, A. Azhdari Tehrani, A. Morsali, Ultrason. Sonochem. 39 (2017) 307– 312. [61] W. Liu, X. Huang, C. Chen, C. Xu, J. Ma, L. Yang, W. Wang, W. Dou, W. Liu, ChemistryEur. J. (2018). [62] B. Zheng, Z. Yang, J. Bai, Y. Li, S. Li, Chem. Commun. 48 (2012) 7025–7027. [63] X.-T. Liu, Y.-Y. Jia, Y.-H. Zhang, G.-J. Ren, R. Feng, S.-Y. Zhang, M.J. Zaworotko, X.-H. Bu, Inorg. Chem. Front. 3 (2016) 1510–1515. [64] E. Keceli, M. Hemgesberg, R. Grünker, V. Bon, C. Wilhelm, T. Philippi, R. Schoch, Y. Sun, M. Bauer, S. Ernst, S. Kaskel, W.R. Thiel, Micropor. Mesopor. Mat. 194 (2014) 115–125. [65] J. Duan, Z. Yang, J. Bai, B. Zheng, Y. Li, S. Li, Chem. Commun. 48 (2012) 3058– 3060. [66] J. Park, J.-R. Li, Y.-P. Chen, J. Yu, A.A. Yakovenko, Z.U. Wang, L.-B. Sun, P.B. Balbuena, H.-C. Zhou, Chem. Commun. 48 (2012) 9995–9997. [67] P. Deria, S. Li, H. Zhang, R.Q. Snurr, J.T. Hupp, O.K. Farha, Chem. Commun. 51 (2015) 12478–12481. [68] B. Zheng, H. Wang, Z. Wang, N. Ozaki, C. Hang, X. Luo, L. Huang, W. Zeng, M. Yang, J. Duan, Chem. Commun. 52 (2016) 12988–12991. [69] W.-G. Jin, W. Chen, P.-H. Xu, X.-W. Lin, X.-C. Huang, G.-H. Chen, F. Lu, X.-M. Chen, Chem. Eur. J. 23 (2017) 13058–13066. [70] Z. Lu, J. Bai, C. Hang, F. Meng, W. Liu, Y. Pan, X. You, Chem. Eur. J. 22 (2016) 6277–6285. [71] C.-H. Lee, H.-Y. Huang, J.-J. Lee, C.-Y. Huang, Y.-C. Kao, G.-H. Lee, S.-M. Peng, J.C. Jiang, I. Chao, K.-L. Lu, ChemistrySelect 1 (2016) 2923–2929. [72] S. Xiong, Y. He, R. Krishna, B. Chen, Z. Wang, Cryst. Growth Des. 13 (2013) 2670–2674. [73] Y.-Y. Jia, X.-T. Liu, R. Feng, S.-Y. Zhang, P. Zhang, Y.-B. He, Y.-H. Zhang, X.-H. Bu, Cryst. Growth Des. 17 (2017) 2584–2588. [74] V. Safarifard, S. Rodríguez-Hermida, V. Guillerm, I. Imaz, M. Bigdeli, A. Azhdari Tehrani, J. Juanhuix, A. Morsali, M.E. Casco, J. Silvestre-Albero, E.V. RamosFernandez, D. Maspoch, Cryst. Growth Des. 16 (2016) 6016–6023. [75] B. Zheng, X. Lin, Z. Wang, R. Yun, Y. Fan, M. Ding, X. Hu, P. Yi, CrystEngComm 16 (2014) 9586–9589. [76] B. Zheng, H. Liu, Z. Wang, X. Yu, P. Yi, J. Bai, CrystEngComm 15 (2013) 3517– 3520. [77] X.-M. Lin, T.-T. Li, L.-F. Chen, L. Zhang, C.-Y. Su, Dalton Trans. 41 (2012) 10422–10429. [78] C.-H. Lee, H.-Y. Huang, Y.-H. Liu, T.-T. Luo, G.-H. Lee, S.-M. Peng, J.-C. Jiang, I. Chao, K.-L. Lu, Inorg. Chem. 52 (2013) 3962–3968. [79] Z.-H. Xuan, D.-S. Zhang, Z. Chang, T.-L. Hu, X.-H. Bu, Inorg. Chem. 53 (2014) 8985–8990. [80] B. Zheng, J. Bai, J. Duan, L. Wojtas, M.J. Zaworotko, J. Am. Chem. Soc. 133 (2011) 748–751. [81] O. Benson, I. da Silva, S.P. Argent, R. Cabot, M. Savage, H.G.W. Godfrey, Y. Yan, S.F. Parker, P. Manuel, M.J. Lennox, T. Mitra, T.L. Easun, W. Lewis, A.J. Blake, E. Besley, S. Yang, M. Schröder, J. Am. Chem. Soc. 138 (2016) 14828–14831. [82] Z. Chen, K. Adil, Ł.J. Weselin´ski, Y. Belmabkhout, M. Eddaoudi, J. Mater. Chem. A 3 (2015) 6276–6281. [83] Z. Lu, J. Zhang, J. Duan, L. Du, C. Hang, J. Mater. Chem. A 5 (2017) 17287– 17292. [84] F. Moreau, I. da Silva, N.H. Al Smail, T.L. Easun, M. Savage, H.G.W. Godfrey, S.F. Parker, P. Manuel, S. Yang, M. Schröder, Nat. Commun. 8 (2017) 14085. [85] M. Zhang, B. Li, Y. Li, Q. Wang, W. Zhang, B. Chen, S. Li, Y. Pan, X. You, J. Bai, Chem. Commun. 52 (2016) 7241–7244. [86] T. Pham, K.A. Forrest, P. Nugent, Y. Belmabkhout, R. Luebke, M. Eddaoudi, M.J. Zaworotko, B. Space, J. Phys. Chem. C 117 (2013) 9340–9354. [87] S. Hasegawa, S. Horike, R. Matsuda, S. Furukawa, K. Mochizuki, Y. Kinoshita, S. Kitagawa, J. Am. Chem. Soc. 129 (2007) 2607–2614. [88] M. Ghorbanloo, V. Safarifard, A. Morsali, New J. Chem. 41 (2017) 3957–3965. [89] A. Paul, A. Karmakar, M.F.C. Guedes da Silva, A.J.L. Pombeiro, RSC Adv. 5 (2015) 87400–87410. [90] L. Esrafili, V. Safarifard, E. Tahmasebi, M.D. Esrafili, A. Morsali, New J. Chem. 42 (2018) 8864–8873. [91] L.X. Yang, X.F. Feng, W.H. Yin, Y. Tao, H.Q. Wu, J.Q. Li, L.F. Ma, F. Luo, J. Solid State Chem. 265 (2018) 148–154. [92] A. Buragohain, M. Yousufuddin, M. Sarma, S. Biswas, Cryst. Growth Des. 16 (2016) 842–851.
[93] X.-Y. Li, Y.-Z. Li, Y. Yang, L. Hou, Y.-Y. Wang, Z. Zhu, Chem. Commun. 53 (2017) 12970–12973. [94] Z. Wang, B. Zheng, H. Liu, X. Lin, X. Yu, P. Yi, R. Yun, Cryst. Growth Des. 13 (2013) 5001–5006. [95] D. Fu, Y. Xu, M. Zhao, Z. Chang, X. Bu, Sci. Bull. 61 (2016) 1255–1259. [96] N.H. Alsmail, M. Suyetin, Y. Yan, R. Cabot, C.P. Krap, J. Lü, T.L. Easun, E. Bichoutskaia, W. Lewis, A.J. Blake, M. Schröder, Chem. Eur. J. 20 (2014) 7317– 7324. [97] M. Zhang, W. Zhou, T. Pham, K.A. Forrest, W. Liu, Y. He, H. Wu, T. Yildirim, B. Chen, B. Space, Y. Pan, M.J. Zaworotko, J. Bai, Angew. Chem. 129 (2017) 11584–11588. [98] X.-Y. Li, W.-J. Shi, X.-Q. Wang, L.-N. Ma, L. Hou, Y.-Y. Wang, Cryst. Growth Des. 17 (2017) 4217–4224. [99] C.M. McGuirk, M.J. Katz, C.L. Stern, A.A. Sarjeant, J.T. Hupp, O.K. Farha, C.A. Mirkin, J. Am. Chem. Soc. 137 (2015) 919–925. [100] X. Zhang, Z. Zhang, J. Boissonnault, S.M. Cohen, Chem. Commun. 52 (2016) 8585–8588. [101] S.M. Cohen, Z. Zhang, J.A. Boissonnault, Inorg. Chem. 55 (2016) 7281–7290. [102] M. Jiang, P. Li, P. Wu, F. Zhang, X. Tian, C. Deng, J. Wang, Chem. Commun. 54 (2018) 9131–9134. [103] Y. Ma, Y. Harada, A. Hori, Y. Hijikata, L. Li, R. Matsuda, Dalton Trans. 46 (2017) 15200–15203. [104] B. Wang, H. Huang, X.-L. Lv, Y. Xie, M. Li, J.-R. Li, Inorg. Chem. 53 (2014) 9254– 9259. [105] G. Feng, Y. Peng, W. Liu, F. Chang, Y. Dai, W. Huang, Inorg. Chem. 56 (2017) 2363–2366. [106] F. Drache, V. Bon, I. Senkovska, M. Adam, A. Eychmüller, S. Kaskel, Eur. J. Inorg. Chem. 2016 (2016) 4483–4489. [107] Z. Guo, X. Song, H. Lei, H. Wang, S. Su, H. Xu, G. Qian, H. Zhang, B. Chen, Chem. Commun. 51 (2015) 376–379. [108] C.-C. Wang, C.-H. Yang, S.-M. Tseng, G.-H. Lee, H.-S. Sheu, K. Phyu, Inorg. Chim. Acta 357 (2004) 3759–3764. [109] L. Li, Y. Zhu, X. Zhou, C.D.S. Brites, D. Ananias, Z. Lin, F.A.A. Paz, J. Rocha, W. Huang, L.D. Carlos, Adv. Funct. Mater. 26 (2016) 8677–8684. [110] D. Yan, R. Gao, M. Wei, S. Li, J. Lu, D.G. Evans, X. Duan, J. Mater. Chem. C 1 (2013) 997–1004. [111] S. Mukherjee, R. Babarao, A.V. Desai, B. Manna, S.K. Ghosh, Cryst. Growth Des. 17 (2017) 3581–3587. [112] K.L. Mulfort, J.T. Hupp, J. Am. Chem. Soc. 129 (2007) 9604–9605. [113] T.K. Prasad, D.H. Hong, M.P. Suh, Chem. Eur. J. 16 (2010) 14043–14050. [114] Y.-S. Bae, B.G. Hauser, O.K. Farha, J.T. Hupp, R.Q. Snurr, Micropor. Mesopor. Mat. 141 (2011) 231–235. [115] C.F. Leong, T.B. Faust, P. Turner, P.M. Usov, C.J. Kepert, R. Babarao, A.W. Thornton, D.M. D’Alessandro, Dalton Trans. 42 (2013) 9831–9839. [116] K.L. Mulfort, J.T. Hupp, Inorg. Chem. 47 (2008) 7936–7938. [117] L. Zeng, T. Liu, C. He, D. Shi, F. Zhang, C. Duan, J. Am. Chem. Soc. 138 (2016) 3958–3961. [118] A. Mallick, B. Garai, M.A. Addicoat, P.S. Petkov, T. Heine, R. Banerjee, Chem. Sci. 6 (2015) 1420–1425. [119] Y. Takashima, V.M. Martínez, S. Furukawa, M. Kondo, S. Shimomura, H. Uehara, M. Nakahama, K. Sugimoto, S. Kitagawa, Nature Communications 2 (2011) 168. [120] Y.-X. Xie, W.-N. Zhao, G.-C. Li, P.-F. Liu, L. Han, Inorg. Chem. 55 (2016) 549– 551. [121] J.-Z. Liao, J.-F. Chang, L. Meng, H.-L. Zhang, S.-S. Wang, C.-Z. Lu, Chem. Commun. 53 (2017) 9701–9704. [122] J.-J. Liu, Y. Chen, M.-J. Lin, C.-C. Huang, W.-X. Dai, Dalton Trans. 45 (2016) 6339–6342. [123] L. Han, L. Qin, L. Xu, Y. Zhou, J. Sun, X. Zou, Chem. Commun. 49 (2013) 406– 408. [124] F. Wei, Y. Ye, W. Huang, Q. Lin, Z. Li, L. Liu, S. Chen, Z. Zhang, S. Xiang, Inorg. Chem. Commun. 93 (2018) 105–109. [125] B. Garai, A. Mallick, R. Banerjee, Chem. Sci. 7 (2016) 2195–2200. [126] K. AlKaabi, C.R. Wade, M. Dinca˘, Chem 1 (2016) 264–272. [127] C.F. Leong, B. Chan, T.B. Faust, D.M. D’Alessandro, Chem. Sci. 5 (2014) 4724– 4728. [128] E. Castaldelli, K.D.G. Imalka Jayawardena, D.C. Cox, G.J. Clarkson, R.I. Walton, L. Le-Quang, J. Chauvin, S.R.P. Silva, G.J.-F. Demets, Nat. Commun. 8 (2017) 2139. [129] H.J. Park, M.C. So, D. Gosztola, G.P. Wiederrecht, J.D. Emery, A.B.F. Martinson, S. Er, C.E. Wilmer, N.A. Vermeulen, A. Aspuru-Guzik, J.F. Stoddart, O.K. Farha, J.T. Hupp, A.C.S. Appl, Mater. Interfaces 8 (2016) 24983–24988. [130] B. Tian, G.-H. Ning, Q. Gao, L.-M. Tan, W. Tang, Z. Chen, C. Su, K.P. Loh, A.C.S. Appl, Mater. Interfaces 8 (2016) 31067–31075. [131] Z. Li, Z. Zhang, Y. Ye, K. Cai, F. Du, H. Zeng, J. Tao, Q. Lin, Y. Zheng, S. Xiang, J. Mater. Chem. A 5 (2017) 7816–7824. [132] J.-N. Hao, B. Yan, Chem. Commun. 51 (2015) 7737–7740. [133] Y. Yang, L. Chen, F. Jiang, X. Wan, M. Yu, Z. Cao, T. Jing, M. Hong, J. Mater. Chem. C 5 (2017) 4511–4519. [134] X.-X. Jia, R.-X. Yao, F.-Q. Zhang, X.-M. Zhang, Inorg. Chem. 56 (2017) 2690– 2696. [135] T. Gadzikwa, O.K. Farha, K.L. Mulfort, J.T. Hupp, S.T. Nguyen, Chem. Commun. (2009) 3720–3722. [136] W. Wu, B. Li, C. Gu, J. Wang, A. Singh, A. Kumar, J. Mol. Struct. 1148 (2017) 531–536.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 [137] Y. Wu, G.-P. Yang, Y. Zhao, W.-P. Wu, B. Liu, Y.-Y.J.D.T. Wang, Dalton Trans. 44 (2015) 3271–3277. [138] Y.-T. Liang, G.-P. Yang, B. Liu, Y.-T. Yan, Z.-P. Xi, Y.-Y. Wang, Dalton Trans. 44 (2015) 13325–13330. [139] W. Sun, J. Wang, G. Zhang, Z. Liu, RSC Adv. 4 (2014) 55252–55255. [140] M. Arici, Cryst. Growth Des. 17 (2017) 5499–5505. [141] J. Cao, Y. Gao, Y. Wang, C. Du, Z. Liu, Chem. Commun. 49 (2013) 6897–6899. [142] B. Shi, Y. Zhong, L. Guo, G. Li, Dalton Trans. 44 (2015) 4362–4369. [143] Z. Yue, Z. Chen, M. Yao, H. Wang, G. Li, RSC Adv. 4 (2014) 33537–33540. [144] D.D. Borges, S. Devautour-Vinot, H. Jobic, J. Ollivier, F. Nouar, R. Semino, T. Devic, C. Serre, F. Paesani, G. Maurin, Angew. Chem. Int. Ed. 55 (2016) 3919– 3924. [145] T. He, Y.-Z. Zhang, H. Wu, X.-J. Kong, X.-M. Liu, L.-H. Xie, Y. Dou, J.-R. Li, ChemPhysChem 18 (2017) 3245–3252. [146] F. Yang, H. Huang, X. Wang, F. Li, Y. Gong, C. Zhong, J.-R. Li, Cryst. Growth Des. 15 (2015) 5827–5833. [147] A. Shigematsu, T. Yamada, H. Kitagawa, J. Am. Chem. Soc. 133 (2011) 2034– 2036. [148] H. Wu, F. Yang, X.-L. Lv, B. Wang, Y.-Z. Zhang, M.-J. Zhao, J.-R. Li, J. Mater. Chem. A 5 (2017) 14525–14529. [149] J.B. DeCoste, T.J. Demasky, M.J. Katz, O.K. Farha, J.T. Hupp, New J. Chem. 39 (2015) 2396–2399. [150] X. Meng, R.-L. Zhong, X.-Z. Song, S.-Y. Song, Z.-M. Hao, M. Zhu, S.-N. Zhao, H.-J. Zhang, Chem. Commun. 50 (2014) 6406–6408. [151] W.-Y. Gao, H. Wu, K. Leng, Y. Sun, S. Ma, Angew. Chem. 128 (2016) 5562– 5566. [152] H.-H. Wang, L. Hou, Y.-Z. Li, C.-Y. Jiang, Y.-Y. Wang, Z. Zhu, A.C.S. Appl, Mater. Interfaces 9 (2017) 17969–17976. [153] Q. Yang, S. Vaesen, F. Ragon, A.D. Wiersum, D. Wu, A. Lago, T. Devic, C. Martineau, F. Taulelle, P.L. Llewellyn, H. Jobic, C. Zhong, C. Serre, G. De Weireld, G. Maurin, Angew. Chem. Int. Ed. 52 (2013) 10316–10320. [154] Q. Yang, A.D. Wiersum, P.L. Llewellyn, V. Guillerm, C. Serre, G. Maurin, Chem. Commun. 47 (2011) 9603–9605. [155] Y.-P. Zhao, Y. Li, C.-Y. Cui, Y. Xiao, R. Li, S.-H. Wang, F.-K. Zheng, G.-C. Guo, Inorg. Chem. 55 (2016) 7335–7340. [156] K. Jayaramulu, S.K. Reddy, A. Hazra, S. Balasubramanian, T.K. Maji, Inorg. Chem. 51 (2012) 7103–7111. [157] Y.-Q. Wang, Q.-H. Tan, H.-T. Liu, W. Sun, Z.-L. Liu, RSC Adv. 5 (2015) 86614– 86619. [158] B. Liu, L. Hou, W.-P. Wu, A.-N. Dou, Y.-Y. Wang, Dalton Trans. 44 (2015) 4423– 4427. [159] B. Chen, L. Wang, Y. Xiao, F.R. Fronczek, M. Xue, Y. Cui, G. Qian, Angew. Chem. 121 (2009) 508–511. [160] B. Liu, W.-P. Wu, L. Hou, Y.-Y.J.C.C. Wang, Chem. Commun. 50 (2014) 8731– 8734. [161] Y. Gao, X. Zhang, W. Sun, Z. Liu, Dalton Trans. 44 (2015) 1845–1849. [162] D.-M. Chen, N. Xu, X.-H. Qiu, P. Cheng, Cryst. Growth Des. 15 (2015) 961–965. [163] Q.-Q. Dang, Y.-F. Zhan, L.-N. Duan, X.-M. Zhang, Dalton Trans. 44 (2015) 20027–20031. [164] C. Song, J. Hu, Y. Ling, Y. Feng, R. Krishna, D.-L. Chen, Y. He, J. Mater. Chem. A 3 (2015) 19417–19426. [165] J. Qian, Q. Li, L. Liang, T.-T. Li, Y. Hu, S. Huang, Dalton Trans. 46 (2017) 14102– 14106. [166] L. Li, S. Tang, C. Wang, X. Lv, M. Jiang, H. Wu, X. Zhao, Chem. Commun. 50 (2014) 2304–2307. [167] X. Rao, J. Cai, J. Yu, Y. He, C. Wu, W. Zhou, T. Yildirim, B. Chen, G. Qian, Chem. Commun. 49 (2013) 6719–6721. [168] J.S. Seo, D. Whang, H. Lee, S.I. Jun, J. Oh, Y.J. Jeon, K. Kim, Nature 404 (2000) 982. [169] H. Liu, B. Yin, Z. Gao, Y. Li, H. Jiang, Chem. Commun. 48 (2012) 2033–2035. [170] H.-F. Zhou, B. Liu, L. Hou, W.-Y. Zhang, Y.-Y. Wang, Chem. Commun. 54 (2018) 456–459. [171] J. Ye, J.K. Johnson, Catal. Sci. Tech. 6 (2016) 8392–8405. [172] B. Li, H.-M. Wen, H. Wang, H. Wu, M. Tyagi, T. Yildirim, W. Zhou, B. Chen, J. Am. Chem. Soc. 136 (2014) 6207–6210. [173] H.-M. Wen, B. Li, L. Li, R.-B. Lin, W. Zhou, G. Qian, B. Chen, Adv. Mater. 30 (2018) 1704792. [174] B. Li, H.-M. Wen, H. Wang, H. Wu, T. Yildirim, W. Zhou, B. Chen, Energy Environ. Sci. 8 (2015) 2504–2511. [175] H.-M. Wen, H. Wang, B. Li, Y. Cui, H. Wang, G. Qian, B. Chen, Inorg. Chem. 55 (2016) 7214–7218. [176] C. Song, J. Jiao, Q. Lin, H. Liu, Y. He, Dalton Trans. 45 (2016) 4563–4569. [177] Y. Zhu, Y.-M. Wang, S.-Y. Zhao, P. Liu, C. Wei, Y.-L. Wu, C.-K. Xia, J.-M. Xie, Inorg. Chem. 53 (2014) 7692–7699. [178] H.-C. Kim, S. Huh, D.N. Lee, Y. Kim, Dalton Trans. 47 (2018) 4820–4826. [179] L. Du, Z. Lu, K. Zheng, J. Wang, X. Zheng, Y. Pan, X. You, J. Bai, J. Am. Chem. Soc. 135 (2013) 562–565. [180] T. Li, D.-L. Chen, J.E. Sullivan, M.T. Kozlowski, J.K. Johnson, N.L. Rosi, Chem. Sci. 4 (2013) 1746–1755. [181] J. Jiao, L. Dou, H. Liu, F. Chen, D. Bai, Y. Feng, S. Xiong, D.-L. Chen, Y. He, Dalton Trans. 45 (2016) 13373–13382. [182] J. An, S.J. Geib, N.L. Rosi, J. Am. Chem. Soc. 132 (2010) 38–39. [183] L. Du, Z. Lu, M. Ma, F. Su, L. Xu, RSC Adv. 5 (2015) 29505–29508. [184] F. Rouhani, A. Morsali, Chem. Eur. J. 24 (2018) 5529–5537. [185] K. Liu, X. Li, D. Ma, Y. Han, B. Li, Z. Shi, Z. Li, L. Wang, Mater. Chem. Front. 1 (2017) 1982–1988.
51
[186] R. Luebke, J.F. Eubank, A.J. Cairns, Y. Belmabkhout, L. Wojtas, M. Eddaoudi, Chem. Commun. 48 (2012) 1455–1457. [187] S. Pal, A. Bhunia, P.P. Jana, S. Dey, J. Möllmer, C. Janiak, H.P. Nayek, Chem. Eur. J. 21 (2014) 2789–2792. [188] X. Wang, M. Chen, M. Du, Inorg. Chem. 55 (2016) 6352–6354. [189] Y. Chen, H. Wang, J. Li, J.V. Lockard, J. Mater. Chem. A 3 (2015) 4945–4953. [190] J. Kim, S.-T. Yang, S.B. Choi, J. Sim, J. Kim, W.-S. Ahn, J. Mater. Chem. 21 (2011) 3070–3076. [191] B. Manna, S. Mukherjee, A.V. Desai, S. Sharma, R. Krishna, S.K. Ghosh, Chem. Commun. 51 (2015) 15386–15389. [192] Y. Huang, Z. Lin, H. Fu, F. Wang, M. Shen, X. Wang, R. Cao, ChemSusChem 7 (2014) 2647–2653. [193] K. Liu, D. Ma, B. Li, Y. Li, K. Yao, Z. Zhang, Y. Han, Z. Shi, J. Mater. Chem. A 2 (2014) 15823–15828. [194] G.-X. Wen, M.-L. Han, X.-Q. Wu, Y.-P. Wu, W.-W. Dong, J. Zhao, D.-S. Li, L.-F. Ma, Dalton Trans. 45 (2016) 15492–15499. [195] M. Zhu, X.-Z. Song, S.-Y. Song, S.-N. Zhao, X. Meng, L.-L. Wu, C. Wang, H.-J. Zhang, Adv. Sci. 2 (2015) 1500012. [196] H. He, S.-H. Chen, D.-Y. Zhang, E.-C. Yang, X.-J. Zhao, RSC Adv. 7 (2017) 38871–38876. [197] Q. Tang, S. Liu, Y. Liu, J. Miao, S. Li, L. Zhang, Z. Shi, Z. Zheng, Inorg. Chem. 52 (2013) 2799–2801. [198] B.-H. Liu, D.-X. Liu, K.-Q. Yang, S.-J. Dong, W. Li, Y.-J. Wang, Inorg. Chem. Commun. 90 (2018) 61–64. [199] L. Han, J. Zhou, X. Li, C.-Y. Sun, L. Zhao, Y.-T. Zhang, M. Zhu, X.-L. Wang, Z.-M. Su, Inorg. Chem. Commun. 86 (2017) 200–203. [200] L. Han, L. Qin, L.-P. Xu, W.-N. Zhao, Inorg. Chem. 52 (2013) 1667–1669. [201] Y. Tan, Z. Fu, J. Zhang, Inorg. Chem. Commun. 14 (2011) 1966–1970. [202] F. Guo, B. Yuan, W. Shi, Inorg. Chem. Commun. 86 (2017) 285–289. [203] Z. Miao, Y. Luan, C. Qi, D. Ramella, Dalton Trans. 45 (2016) 13917–13924. [204] S.A.A. Razavi, M.Y. Masoomi, A. Morsali, Inorg. Chem. 56 (2017) 9646–9652. [205] S.A.A. Razavi, M.Y. Masoomi, A. Morsali, Ultrason. Sonochem. 41 (2018) 17– 26. [206] S.A.A. Razavi, M.Y. Masoomi, A. Morsali, Ultrason. Sonochem. 37 (2017) 502– 508. [207] P. Rouschmeyer, N. Guillou, C. Serre, G. Clavier, C. Martineau, P. Audebert, E. Elkaïm, C.M. Allain, T. Devic, Inorg. Chem. 56 (2017) 8423–8429. [208] S.A.A. Razavi, M.Y. Masoomi, A. Morsali, Chem. Eur. J. 23 (2017) 12559– 12564. [209] G. Nickerl, I. Senkovska, S. Kaskel, Chem. Commun. 51 (2015) 2280–2282. [210] Z.-Z. Lu, R. Zhang, Y.-Z. Li, Z.-J. Guo, H.-G. Zheng, J. Am. Chem. Soc. 133 (2011) 4172–4174. [211] C.X. Bezuidenhout, C. Esterhuysen, L.J. Barbour, Chem. Commun. 53 (2017) 5618–5621. [212] S.A.A. Razavi, M.Y. Masoomi, A. Morsali, Inorg. Chem. 57 (2018) 11578. [213] S.A.A. Razavi, A. Morsali, ACS Appl. Mater. Interfaces (2019). [214] S.A.A. Razavi, M.Y. Masoomi, T. Islamoglu, A. Morsali, Y. Xu, J.T. Hupp, O.K. Farha, J. Wang, P.C. Junk, Inorg. Chem. 56 (2017) 2581–2588. [215] S. Sen, S. Neogi, A. Aijaz, Q. Xu, P.K. Bharadwaj, Dalton Trans. 43 (2014) 6100– 6107. [216] K. Liu, X. Han, Y. Zou, Y. Peng, G. Li, Z. Shi, S. Feng, Inorg. Chem. Commun. 39 (2014) 131–134. [217] Y. Hijikata, S. Horike, M. Sugimoto, M. Inukai, T. Fukushima, S. Kitagawa, Inorg. Chem. 52 (2013) 3634–3642. [218] Z. Chang, D.-S. Zhang, Q. Chen, R.-F. Li, T.-L. Hu, X.-H. Bu, Inorg. Chem. 50 (2011) 7555–7562. [219] K. Liu, B. Li, Y. Li, X. Li, F. Yang, G. Zeng, Y. Peng, Z. Zhang, G. Li, Z. Shi, Chem. Commun. 50 (2014) 5031–5033. [220] R. Luebke, Ł.J. Weselin´ski, Y. Belmabkhout, Z. Chen, Ł. Wojtas, M. Eddaoudi, Cryst. Growth Des. 14 (2014) 414–418. [221] J. Ye, J.K. Johnson, ACS Catal. 5 (2015) 2921–2928. [222] W.-Y. Gao, T. Pham, K.A. Forrest, B. Space, L. Wojtas, Y.-S. Chen, S. Ma, Chem. Commun. 51 (2015) 9636–9639. [223] T. He, Y.Z. Zhang, B. Wang, X.L. Lv, L.H. Xie, J.R. Li, ChemPlusChem 81 (2016) 864–871. [224] J. Baima, R. Macchieraldo, C. Pettinari, S. Casassa, CrystEngComm 17 (2015) 448–455. [225] G.-P. Li, G. Liu, Y.-Z. Li, L. Hou, Y.-Y. Wang, Z. Zhu, Inorg. Chem. 55 (2016) 3952–3959. [226] H.-H. Wang, L.-N. Jia, L. Hou, W.-J. Shi, Z. Zhu, Y.-Y. Wang, Inorg. Chem. 54 (2015) 1841–1846. [227] K. Tomar, A.K. Gupta, M. Gupta, New J. Chem. 40 (2016) 1953–1956. [228] G. Xu, Y. Peng, Z. Hu, D. Yuan, B. Donnadieu, D. Zhao, H. Cheng, RSC Adv. 5 (2015) 47384–47389. [229] D.J. Xiao, M.I. Gonzalez, L.E. Darago, K.D. Vogiatzis, E. Haldoupis, L. Gagliardi, J.R. Long, J. Am. Chem. Soc. 138 (2016) 7161–7170. [230] C.E. Willans, S. French, L.J. Barbour, J.-A. Gertenbach, P.C. Junk, G.O. Lloyd, J.W. Steed, Dalton Trans. (2009) 6480–6482. [231] S. Nandi, R. Maity, D. Chakraborty, H. Ballav, R. Vaidhyanathan, Inorg. Chem. 57 (2018) 5267–5272. [232] S.-S. Chen, P. Wang, S. Takamizawa, T.-A. Okamura, M. Chen, W.-Y. Sun, Dalton Trans. 43 (2014) 6012–6020. [233] A. Dey, D. Bairagi, K. Biradha, Cryst. Growth Des. 17 (2017) 3885–3892. [234] X. Lv, L. Li, S. Tang, C. Wang, X. Zhao, Chem. Commun. 50 (2014) 6886–6889. [235] Q. Li, D.-X. Xue, Y.-F. Zhang, Z.-H. Zhang, Z. Gao, J. Bai, J. Mater. Chem. A 5 (2017) 14182–14189.
52
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
[236] Y. Yu, X.-M. Zhang, J.-P. Ma, Q.-K. Liu, P. Wang, Y.-B. Dong, Chem. Commun. 50 (2014) 1444–1446. [237] L.-H. Liu, X.-T. Qiu, Y.-J. Wang, Q. Shi, Y.-Q. Sun, Y.-P. Chen, Dalton Trans. 46 (2017) 12106–12113. [238] Y.-H. Han, C.-B. Tian, Q.-H. Li, S.-W. Du, J. Mater. Chem. C 2 (2014) 8065–8070. [239] L. Chen, H. Zhang, M. Pan, Z.W. Wei, H.P. Wang, Y.N. Fan, C.Y. Su, Chem. Asian J. 11 (2016) 1765–1769. [240] M.B. Lalonde, O.K. Farha, K.A. Scheidt, J.T. Hupp, ACS Catal. 2 (2012) 1550– 1554. [241] F.-M. Zhang, L.-Z. Dong, J.-S. Qin, W. Guan, J. Liu, S.-L. Li, M. Lu, Y.-Q. Lan, Z.-M. Su, H.-C. Zhou, J. Am. Chem. Soc. 139 (2017) 6183–6189. [242] S. Li, Y. Wang, C. Qi, X. Zhao, J. Zhang, S. Zhang, S. Pang, Angew. Chem. 125 (2013) 14281–14285. [243] A.A. Dippold, T.M. Klapötke, Chem. Eur. J. 18 (2012) 16742–16753. [244] X. Liu, W. Gao, P. Sun, Z. Su, S. Chen, Q. Wei, G. Xie, S. Gao, Green Chem. 17 (2015) 831–836. [245] J.-G. Xu, C. Sun, M.-J. Zhang, B.-W. Liu, X.-Z. Li, J. Lu, S.-H. Wang, F.-K. Zheng, G.-C. Guo, Chem. Mater. 29 (2017) 9725–9733. [246] Y. Dong, P. Peng, B. Hu, H. Su, S. Li, S. Pang, Molecules 22 (2017) 1068. [247] X. Qu, L. Zhai, B. Wang, Q. Wei, G. Xie, S. Chen, S. Gao, Dalton Trans. 45 (2016) 17304–17311. [248] J.-B. Lin, J.-P. Zhang, X.-M. Chen, J. Am. Chem. Soc. 132 (2010) 6654–6656. [249] P.-Z. Li, X.-J. Wang, R.H.D. Tan, Q. Zhang, R. Zou, Y. Zhao, RSC Adv. 3 (2013) 15566–15570. [250] H.H. Wang, W.J. Shi, L. Hou, G.P. Li, Z. Zhu, Y.Y. Wang, Chem. Eur. J. 21 (2015) 16525–16531. [251] L. Meng, Z. Niu, C. Liang, X. Dong, K. Liu, G. Li, C. Li, Y. Han, Z. Shi, S. Feng, Chem. Eur. J. (2018). [252] B. Liu, S. Yao, C. Shi, G. Li, Q. Huo, Y. Liu, Chem. Commun. 52 (2016) 3223– 3226. [253] J.Q. Liu, G.P. Li, W.C. Liu, Q.L. Li, B.H. Li, R.W. Gable, L. Hou, S.R. Batten, ChemPlusChem 81 (2016) 1299–1304. [254] X. Han, X. Han, R. Li, X.-J. Wang, Y. Zhao, CrystEngComm 18 (2016) 1277– 1281. [255] Y. Ye, S. Xiong, X. Wu, L. Zhang, Z. Li, L. Wang, X. Ma, Q.-H. Chen, Z. Zhang, S. Xiang, Inorg. Chem. 55 (2015) 292–299. [256] Y.-W. Li, K.-H. He, X.-H. Bu, J. Mater. Chem. A 1 (2013) 4186–4189. [257] P.-Z. Li, X.-J. Wang, J. Liu, J.S. Lim, R. Zou, Y. Zhao, J. Am. Chem. Soc. 138 (2016) 2142–2145. [258] Y. Yang, F. Jiang, L. Chen, J. Pang, M. Wu, X. Wan, J. Pan, J. Qian, M. Hong, J. Mater. Chem. A 3 (2015) 13526–13532. [259] K.A. Forrest, T. Pham, B. Space, Phys. Chem. Chem. Phys. 19 (2017) 29204– 29221. [260] X.-J. Wang, P.-Z. Li, Y. Chen, Q. Zhang, H. Zhang, X.X. Chan, R. Ganguly, Y. Li, J. Jiang, Y. Zhao, Sci. Rep. 3 (2013) 1149. [261] Z. Liu, L. Lv, Y. He, Y. Feng, CrystEngComm 19 (2017) 2795–2801. [262] S. Seth, G. Savitha, J.N. Moorthy, Inorg. Chem. 54 (2015) 6829–6835. [263] X. Zhang, T. Xia, K. Jiang, Y. Cui, Y. Yang, G. Qian, J. Solid State Chem. 253 (2017) 277–281. [264] D.K. Singha, P. Mahata, Inorg. Chem. 54 (2015) 6373–6379. [265] Q. Hu, L. Qin, J. Lei, X.-X. Tan, G. Ni, Y.-Q. Wang, J. Li, F.-H. Cao, Polyhedron 151 (2018) 33–36. [266] Q.-K. Liu, J.-P. Ma, Y.-B. Dong, Chem. Commun. 47 (2011) 7185–7187. [267] D.K. Singha, P. Majee, S.K. Mondal, P. Mahata, J. Photochem. Photobiol., A 356 (2018) 389–396. [268] D.-M. Chen, J.-Y. Tian, M. Chen, C.-S. Liu, M. Du, A.C.S. Appl, Mater. Interfaces 8 (2016) 18043–18050. [269] M.I.H. Mohideen, R.S. Pillai, K. Adil, P.M. Bhatt, Y. Belmabkhout, A. Shkurenko, G. Maurin, M. Eddaoudi, Chem 3 (2017) 822–833. [270] F. Wang, H.-R. Fu, Y. Kang, J. Zhang, Chem. Commun. 50 (2014) 12065–12068. [271] B.X. Dong, S.Y. Zhang, W.L. Liu, Y.C. Wu, J. Ge, L. Song, Y.L. Teng, Chem. Commun. 51 (2015) 5691–5694. [272] Q. Li, M.-H. Yu, J. Xu, A.-L. Li, T.-L. Hu, X.-H. Bu, Dalton Trans. 46 (2017) 3223– 3228. [273] L. Hou, L.-N. Jia, W.-J. Shi, Y.-Y. Wang, B. Liu, Q.-Z. Shi, Dalton Trans. 42 (2013) 3653–3659. [274] D. Wang, T. Zhao, G. Li, Q. Huo, Y. Liu, Dalton Trans. 43 (2014) 2365–2368. [275] R. Poloni, B. Smit, J.B. Neaton, J. Am. Chem. Soc. 134 (2012) 6714–6719. [276] P. Cui, Y.-G. Ma, H.-H. Li, B. Zhao, J.-R. Li, P. Cheng, P.B. Balbuena, H.-C. Zhou, J. Am. Chem. Soc. 134 (2012) 18892–18895. [277] Q. Lin, T. Wu, S.-T. Zheng, X. Bu, P. Feng, J. Am. Chem. Soc. 134 (2012) 784– 787. [278] X.-L. Hu, C. Qin, L. Zhao, F.-H. Liu, K.-Z. Shao, Z.-M. Su, RSC Adv. 5 (2015) 49606–49613. [279] P. Pachfule, R. Banerjee, Cryst. Growth Des. 11 (2011) 5176–5181. [280] G. Orcajo, G. Calleja, J.A. Botas, L. Wojtas, M.H. Alkordi, M. Sánchez-Sánchez, Cryst. Growth Des. 14 (2014) 739–746. [281] O. Brea, A. Luna, C. Díaz, I. Corral, ChemPhysChem 19 (2018) 1349–1357. [282] S. Xiong, Q. Liu, Q. Wang, W. Li, Y. Tang, X. Wang, S. Hu, B. Chen, J. Mater. Chem. A 3 (2015) 10747–10752. [283] E.D. Bloch, W.L. Queen, M.R. Hudson, J.A. Mason, D.J. Xiao, L.J. Murray, R. Flacau, C.M. Brown, J.R. Long, Angew. Chem. Int. Ed. 55 (2016) 8605–8609. [284] M.-Y. Li, F. Wang, J. Zhang, Inorg. Chem. Commun. 79 (2017) 37–40. [285] J.-S. Qin, J.-C. Zhang, M. Zhang, D.-Y. Du, J. Li, Z.-M. Su, Y.-Y. Wang, S.-P. Pang, S.-H. Li, Y.-Q. Lan, Adv. Sci. 2 (2015) 1500150. [286] Q. Zhang, J.N.M. Shreeve, Angew. Chem. Int. Ed. 53 (2014) 2540–2542.
[287] Q. Yang, G. Yang, W. Zhang, S. Zhang, Z. Yang, G. Xie, Q. Wei, S. Chen, S. Gao, Chem. Eur. J. 23 (2017) 9149–9155. [288] Y. Feng, X. Liu, L. Duan, Q. Yang, Q. Wei, G. Xie, S. Chen, X. Yang, S. Gao, Dalton Trans. 44 (2015) 2333–2339. [289] Y. Xu, W. Liu, D. Li, H. Chen, M. Lu, Dalton Trans. 46 (2017) 11046–11052. [290] X.-N. Qu, S. Zhang, B.-Z. Wang, Q. Yang, J. Han, Q. Wei, G. Xie, S.-P. Chen, Dalton Trans. 45 (2016) 6968–6973. [291] Y. Shang, B. Jin, R. Peng, Q. Liu, B. Tan, Z. Guo, J. Zhao, Q. Zhang, Dalton Trans. 45 (2016) 13881–13887. [292] S.-H. Wang, F.-K. Zheng, M.-J. Zhang, Z.-F. Liu, J. Chen, Y. Xiao, A.-Q. Wu, G.-C. Guo, J.-S. Huang, Inorg. Chem. 52 (2013) 10096–10104. [293] P. He, J.-G. Zhang, L. Wu, J.-T. Wu, T.-L. Zhang, Inorg. Chim. Acta 455 (2017) 152–157. [294] Y. Feng, Y. Bi, W. Zhao, T. Zhang, J. Mater. Chem. A 4 (2016) 7596–7600. [295] Q. Liu, B. Jin, Q. Zhang, Y. Shang, Z. Guo, B. Tan, R. Peng, Materials 9 (2016). [296] K.-L. Zhang, Z.-C. Pan, Y. Chang, W.-L. Liu, S.W. Ng, Mater Lett 63 (2009) 2136–2138. [297] B.-D. Wu, X.-Y. Li, J. Liang, X.-H. Geng, H.-S. Huang, H. Huang, J.-Y. Wang, W.H. Hu, Polyhedron 139 (2018) 148–154. [298] X. Qu, Q. Yang, J. Han, Q. Wei, G. Xie, S. Chen, S. Gao, RSC Adv. 6 (2016) 46212–46217. [299] S.-H. Wang, F.-K. Zheng, M.-F. Wu, Z.-F. Liu, J. Chen, G.-C. Guo, A.Q. Wu, CrystEngComm 15 (2013) 2616–2623. [300] Y. Friedman, I. Goldberg, Polyhedron 139 (2018) 327–330. [301] Y. Zhao, H. Wu, T.J. Emge, Q. Gong, N. Nijem, Y.J. Chabal, L. Kong, D.C. Langreth, H. Liu, H. Zeng, J. Li, Chem. Eur. J. 17 (2011) 5101–5109. [302] A. Khutia, C. Janiak, Dalton Trans. 43 (2014) 1338–1347. [303] S. Couck, J.F.M. Denayer, G.V. Baron, T. Rémy, J. Gascon, F. Kapteijn, J. Am. Chem. Soc. 131 (2009) 6326–6327. [304] Y. Lin, C. Kong, L. Chen, RSC Adv. 2 (2012) 6417–6419. [305] X. Wang, L. Chen, Q. Guo, Chem Eng J 260 (2015) 573–581. [306] N. Ko, J. Hong, S. Sung, K.E. Cordova, H.J. Park, J.K. Yang, J. Kim, Dalton Trans. 44 (2015) 2047–2051. [307] Y. Lin, Q. Yan, C. Kong, L. Chen, Sci. Rep. 3 (2013) 1859. [308] P. Das, S.K. Mandal, A.C.S. Appl, Mater. Interfaces 10 (2018) 25360–25371. [309] L.A. Darunte, A.D. Oetomo, K.S. Walton, D.S. Sholl, C.W. Jones, ACS Sustain. Chem. Eng. 4 (2016) 5761–5768. [310] B. Arstad, H. Fjellvåg, K.O. Kongshaug, O. Swang, R. Blom, Adsorption 14 (2008) 755–762. [311] Y. Belmabkhout, R. Serna-Guerrero, A. Sayari, Ind. Eng. Chem. Res. 49 (2010) 359–365. [312] R. Vaidhyanathan, S.S. Iremonger, G.K.H. Shimizu, P.G. Boyd, S. Alavi, T.K. Woo, Angew. Chem. 124 (2012) 1862–1865. [313] R. Vaidhyanathan, S.S. Iremonger, K.W. Dawson, G.K.H. Shimizu, Chem. Commun. (2009) 5230–5232. [314] B. Yuan, D. Ma, X. Wang, Z. Li, Y. Li, H. Liu, D. He, Chem. Commun. 48 (2012) 1135–1137. [315] Q. Yan, Y. Lin, C. Kong, L. Chen, Chem. Commun. 49 (2013) 6873–6875. [316] R. Haldar, S.K. Reddy, V.M. Suresh, S. Mohapatra, S. Balasubramanian, T.K. Maji, Chem. Eur. J. 20 (2014) 4347–4356. [317] P. Serra-Crespo, E.V. Ramos-Fernandez, J. Gascon, F. Kapteijn, Chem. Mater. 23 (2011) 2565–2572. [318] Z. Qiao, N. Wang, J. Jiang, J. Zhou, Chem. Commun. 52 (2016) 974–977. [319] X.-L. Hu, Q.-H. Gong, R.-L. Zhong, X.-L. Wang, C. Qin, H. Wang, J. Li, K.-Z. Shao, Z.-M. Su, Chem. Eur. J. 21 (2015) 7238–7244. [320] S. Couck, E. Gobechiya, C.E.A. Kirschhock, P. Serra-Crespo, J. Juan-Alcañiz, A. Martinez Joaristi, E. Stavitski, J. Gascon, F. Kapteijn, G.V. Baron, J.F.M. Denayer, ChemSusChem 5 (2012) 740–750. [321] B. Liu, R. Zhao, G. Yang, L. Hou, Y.-Y. Wang, Q.-Z. Shi, CrystEngComm 15 (2013) 2057–2060. [322] L.A. Darunte, K.S. Walton, D.S. Sholl, C.W. Jones, Curr. Opin. Chem. Eng. 12 (2016) 82–90. [323] J.-W. Zhang, M.-C. Hu, S.-N. Li, Y.-C. Jiang, Q.-G. Zhai, Dalton Trans. 46 (2017) 836–844. [324] B. Liu, R. Zhao, K. Yue, J. Shi, Y. Yu, Y. Wang, Dalton Trans. 42 (2013) 13990– 13996. [325] X. Si, C. Jiao, F. Li, J. Zhang, S. Wang, S. Liu, Z. Li, L. Sun, F. Xu, Z. Gabelica, C. Schick, Energy Environ. Sci. 4 (2011) 4522–4527. [326] I. Erucar, S. Keskin, Ind. Eng. Chem. Res. 52 (2013) 3462–3472. [327] X.-J. Wang, J. Li, P.-Z. Li, L.-B. Xing, H. Lu, H. Wu, Y. Shi, R. Zou, Y. Zhao, Inorg. Chem. Commun. 46 (2014) 13–16. [328] Z.-H. Shi, L. Zhang, H.-Y. Yu, Inorg. Chem. Commun. 54 (2015) 77–80. [329] B. Ugale, S.S. Dhankhar, C.M. Nagaraja, Inorg. Chem. Front. 4 (2017) 348–359. [330] A. Sayari, Y. Belmabkhout, J. Am. Chem. Soc. 132 (2010) 6312–6314. [331] R.W. Flaig, T.M. Osborn Popp, A.M. Fracaroli, E.A. Kapustin, M.J. Kalmutzki, R. M. Altamimi, F. Fathieh, J.A. Reimer, O.M. Yaghi, J. Am. Chem. Soc. 139 (2017) 12125–12128. [332] N. Planas, A.L. Dzubak, R. Poloni, L.-C. Lin, A. McManus, T.M. McDonald, J.B. Neaton, J.R. Long, B. Smit, L. Gagliardi, J. Am. Chem. Soc. 135 (2013) 7402– 7405. [333] G.E. Cmarik, M. Kim, S.M. Cohen, K.S. Walton, Langmuir 28 (2012) 15606– 15613. [334] O. Shekhah, Y. Belmabkhout, Z. Chen, V. Guillerm, A. Cairns, K. Adil, M. Eddaoudi, Nat. Commun. 5 (2014) 4228. [335] J. Cai, H. Wang, H. Wang, X. Duan, Z. Wang, Y. Cui, Y. Yang, B. Chen, G. Qian, RSC Adv. 5 (2015) 77417–77422.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 [336] R. Vaidhyanathan, S.S. Iremonger, G.K.H. Shimizu, P.G. Boyd, S. Alavi, T.K. Woo, Science 330 (2010) 650–653. [337] H.-C. Kim, S. Huh, S.-J. Kim, Y. Kim, Sci. Rep. 7 (2017) 17185. [338] S. Vaesen, V. Guillerm, Q. Yang, A.D. Wiersum, B. Marszalek, B. Gil, A. Vimont, M. Daturi, T. Devic, P.L. Llewellyn, C. Serre, G. Maurin, G. De Weireld, Chem. Commun. 49 (2013) 10082–10084. [339] L. Wu, M. Xue, S.-L. Qiu, G. Chaplais, A. Simon-Masseron, J. Patarin, Micropor. Mesopor. Mat. 157 (2012) 75–81. [340] J.G. Vitillo, M. Savonnet, G. Ricchiardi, S. Bordiga, ChemSusChem 4 (2011) 1281–1290. [341] J. Ethiraj, E. Albanese, B. Civalleri, J.G. Vitillo, F. Bonino, S. Chavan, G.C. Shearer, K.P. Lillerud, S. Bordiga, ChemSusChem 7 (2014) 3382–3388. [342] K.-J. Chen, R.-B. Lin, P.-Q. Liao, C.-T. He, J.-B. Lin, W. Xue, Y.-B. Zhang, J.-P. Zhang, X.-M. Chen, Cryst. Growth Des. 13 (2013) 2118–2123. [343] C.P. Cabello, G. Berlier, G. Magnacca, P. Rumori, G.T. Palomino, CrystEngComm 17 (2015) 430–437. [344] T.M. McDonald, W.R. Lee, J.A. Mason, B.M. Wiers, C.S. Hong, J.R. Long, J. Am. Chem. Soc. 134 (2012) 7056–7065. [345] X. Su, L. Bromberg, V. Martis, F. Simeon, A. Huq, T.A. Hatton, A.C.S. Appl, Mater. Interfaces 9 (2017) 11299–11306. [346] X. Huang, J. Lu, W. Wang, X. Wei, J. Ding, Appl Surf Sci 371 (2016) 307–313. [347] T.M. McDonald, D.M. D’Alessandro, R. Krishna, J.R. Long, Chem. Sci. 2 (2011) 2022–2028. [348] H. Jo, W.R. Lee, N.W. Kim, H. Jung, K.S. Lim, J.E. Kim, D.W. Kang, H. Lee, V. Hiremath, J.G. Seo, H. Jin, D. Moon, S.S. Han, C.S. Hong, ChemSusChem 10 (2016) 541–550. [349] A. Demessence, D.M. D’Alessandro, M.L. Foo, J.R. Long, J. Am. Chem. Soc. 131 (2009) 8784–8786. [350] J.S. Yeon, W.R. Lee, N.W. Kim, H. Jo, H. Lee, J.H. Song, K.S. Lim, D.W. Kang, J.G. Seo, D. Moon, B. Wiers, C.S. Hong, J. Mater. Chem. A 3 (2015) 19177–19185. [351] C. Montoro, E. García, S. Calero, M.A. Pérez-Fernández, A.L. López, E. Barea, J.A. R. Navarro, J. Mater. Chem. 22 (2012) 10155–10158. [352] X. Wang, H. Li, X.-J. Hou, J. Phys. Chem. C 116 (2012) 19814–19821. [353] W.R. Lee, H. Jo, L.-M. Yang, H. Lee, D.W. Ryu, K.S. Lim, J.H. Song, D.Y. Min, S.S. Han, J.G. Seo, Y.K. Park, D. Moon, C.S. Hong, Chem. Sci. 6 (2015) 3697–3705. [354] W.R. Lee, S.Y. Hwang, D.W. Ryu, K.S. Lim, S.S. Han, D. Moon, J. Choi, C.S. Hong, Energy & Environmental Science 7 (2014) 744–751. [355] S. Choi, T. Watanabe, T.-H. Bae, D.S. Sholl, C.W. Jones, The Journal of Physical Chemistry Letters 3 (2012) 1136–1141. [356] W.S. Drisdell, R. Poloni, T.M. McDonald, T.A. Pascal, L.F. Wan, C.D. Pemmaraju, B. Vlaisavljevich, S.O. Odoh, J.B. Neaton, J.R. Long, D. Prendergast, J.B. Kortright, Phys. Chem. Chem. Phys. 17 (2015) 21448–21457. [357] A. Verma, D. De, K. Tomar, P.K. Bharadwaj, Inorg. Chem. 56 (2017) 9765– 9771. [358] A.C. Kathalikkattil, R. Roshan, J. Tharun, R. Babu, G.-S. Jeong, D.-W. Kim, S.J. Cho, D.-W. Park, Chem. Commun. 52 (2016) 280–283. [359] P. Patel, B. Parmar, R.I. Kureshy, N.-U.H. Khan, E. Suresh, Dalton Trans. 47 (2018) 8041–8051. [360] Y.-H. Han, Z.-Y. Zhou, C.-B. Tian, S.-W. Du, Green Chemistry 18 (2016) 4086– 4091. [361] T. Lescouet, C. Chizallet, D. Farrusseng, Chemcatchem 4 (2012) 1725–1728. [362] P. Rani, R. Srivastava, New J. Chem. 41 (2017) 8166–8177. [363] J. Gascon, U. Aktay, M.D. Hernandez-Alonso, G.P.M. van Klink, F. Kapteijn, J Catal 261 (2009) 75–87. [364] Y. Yang, H.-F. Yao, F.-G. Xi, E.-Q. Gao, J. Mol. Catal. A: Chem. 390 (2014) 198– 205. [365] M. Hartmann, M. Fischer, Micropor. Mesopor. Mat. 164 (2012) 38–43. [366] A. Taher, D.-J. Lee, B.-K. Lee, I.-M. Lee, Synlett 27 (2016) 1433–1437. [367] J. Huang, C. Li, L. Tao, H. Zhu, G. Hu, J. Mol. Struct. 1146 (2017) 853–860. [368] P.V. Dau, S.M. Cohen, Inorg. Chem. 54 (2015) 3134–3138. [369] F. Vermoortele, R. Ameloot, A. Vimont, C. Serre, D. De Vos, Chem. Commun. 47 (2011) 1521–1523. [370] M. Savonnet, S. Aguado, U. Ravon, D. Bazer-Bachi, V. Lecocq, N. Bats, C. Pinel, D. Farrusseng, Green Chemistry 11 (2009) 1729–1732. [371] J. Chen, R. Liu, H. Gao, L. Chen, D. Ye, J. Mater. Chem. A 2 (2014) 7205–7213. [372] S. Rostamnia, H. Xin, Appl. Organomet. Chem. 28 (2014) 359–363. [373] S. Rostamnia, Z. Karimi, Inorg. Chim. Acta 428 (2015) 133–137. [374] V. Safarifard, S. Beheshti, A. Morsali, CrystEngComm 17 (2015) 1680–1685. [375] S. Rostamnia, M. Jafari, Appl. Organomet. Chem. 31 (2017) e3584. [376] U.J. Ryu, S.J. Kim, H.-K. Lim, H. Kim, K.M. Choi, J.K. Kang, Sci. Rep. 7 (2017) 612. [377] H. Wang, X. Yuan, Y. Wu, G. Zeng, X. Chen, L. Leng, Z. Wu, L. Jiang, H. Li, J. Hazard. Mater. 286 (2015) 187–194. [378] L. Shi, T. Wang, H. Zhang, K. Chang, X. Meng, H. Liu, J. Ye, Advanced Science 2 (2015) 1500006. [379] Y. Fu, D. Sun, Y. Chen, R. Huang, Z. Ding, X. Fu, Z. Li, Angew. Chem. Int. Ed. 51 (2012) 3364–3367. [380] D. Sun, Y. Fu, W. Liu, L. Ye, D. Wang, L. Yang, X. Fu, Z. Li, Chem. Eur. J. 19 (2013) 14279–14285. [381] J. Long, S. Wang, Z. Ding, S. Wang, Y. Zhou, L. Huang, X. Wang, Chem. Commun. 48 (2012) 11656–11658. [382] P. Wu, C. He, J. Wang, X. Peng, X. Li, Y. An, C. Duan, J. Am. Chem. Soc. 134 (2012) 14991–14999. [383] J.B. DeCoste, M.A. Browe, G.W. Wagner, J.A. Rossin, G.W. Peterson, Chem. Commun. 51 (2015) 12474–12477. [384] Q.-K. Liu, J.-P. Ma, Y.-B. Dong, Chem. Commun. 47 (2011) 12343–12345. [385] F. Tian, X. Zhang, Y. Chen, RSC Adv. 6 (2016) 63895–63904.
53
[386] R.-X. Yao, X. Cui, X.-X. Jia, F.-Q. Zhang, X.-M. Zhang, Inorg. Chem. 55 (2016) 9270–9275. [387] V. Safarifard, A. Morsali, CrystEngComm 16 (2014) 8660–8663. [388] I. Ahmed, S.H. Jhung, J. Hazard. Mater. 283 (2015) 544–550. [389] P.W. Seo, I. Ahmed, S.H. Jhung, Phys. Chem. Chem. Phys. 18 (2016) 14787– 14794. [390] I. Ahmed, N.A. Khan, J.W. Yoon, J.-S. Chang, S.H. Jhung, A.C.S. Appl, Mater. Interfaces 9 (2017) 20938–20946. [391] Z. Hasan, M. Tong, B.K. Jung, I. Ahmed, C. Zhong, S.H. Jhung, J. Phys. Chem. C 118 (2014) 21049–21056. [392] T. Boontongto, K. Siriwong, R. Burakham, Chromatographia 81 (2018) 735– 747. [393] G.W. Peterson, J.J. Mahle, J.B. DeCoste, W.O. Gordon, J.A. Rossin, Angew. Chem. Int. Ed. 55 (2016) 6235–6238. [394] T. Moradpour, A. Abbasi, K. Van Hecke, J. Solid State Chem. 228 (2015) 36–41. [395] H.-T. Zhang, J.-W. Zhang, G. Huang, Z.-Y. Du, H.-L. Jiang, Chem. Commun. 50 (2014) 12069–12072. [396] D. Zhao, X.-H. Liu, Y. Zhao, P. Wang, Y. Liu, M. Azam, S.I. Al-Resayes, Y. Lu, W.Y. Sun, J. Mater. Chem. A 5 (2017) 15797–15807. [397] L. Wen, X. Zheng, K. Lv, C. Wang, X. Xu, Inorg. Chem. 54 (2015) 7133–7135. [398] R. Lv, J. Wang, Y. Zhang, H. Li, L. Yang, S. Liao, W. Gu, X. Liu, J. Mater. Chem. A 4 (2016) 15494–15500. [399] Z. Xu, Y. Wen, L. Tian, G. Li, Inorg. Chem. Commun. 77 (2017) 11–13. [400] B. Joarder, A.V. Desai, P. Samanta, S. Mukherjee, S.K. Ghosh, Chem. Eur. J. 21 (2014) 965–969. [401] S. Mukherjee, A.V. Desai, B. Manna, A.I. Inamdar, S.K. Ghosh, Cryst. Growth Des. 15 (2015) 4627–4634. [402] S.S. Nagarkar, A.V. Desai, P. Samanta, S.K. Ghosh, Dalton Trans. 44 (2015) 15175–15180. [403] A.V. Desai, P. Samanta, B. Manna, S.K. Ghosh, Chem. Commun. 51 (2015) 6111–6114. [404] J. Aguilera-Sigalat, D. Bradshaw, Chem. Commun. 50 (2014) 4711–4713. [405] Q.-K. Liu, J.-P. Ma, Y.-B. Dong, J. Am. Chem. Soc. 132 (2010) 7005–7017. [406] H.-M. Wen, L. Li, R.-B. Lin, B. Li, B. Hu, W. Zhou, J. Hu, B. Chen, J. Mater. Chem. A 6 (2018) 6931–6937. [407] Y. Lin, Q. Zhang, C. Zhao, H. Li, C. Kong, C. Shen, L. Chen, Chem. Commun. 51 (2015) 697–699. [408] I.M.P. Silva, M.A. Carvalho, C.S. Oliveira, D.M. Profirio, R.B. Ferreira, P.P. Corbi, A.L.B. Formiga, Inorg. Chem. Commun. 70 (2016) 47–50. [409] A. Pal, J.B. Lin, S. Chand, M.C. Das, ChemistrySelect 3 (2018) 917–921. [410] O. Alduhaish, H. Wang, B. Li, H.D. Arman, V. Nesterov, K. Alfooty, B. Chen, ChemPlusChem 81 (2016) 764–769. [411] B. Bhattacharya, D.K. Maity, R. Mondal, E. Colacio, D. Ghoshal, Cryst. Growth Des. 15 (2015) 4427–4437. [412] M.Y. Masoomi, K.C. Stylianou, A. Morsali, P. Retailleau, D. Maspoch, Cryst. Growth Des. 14 (2014) 2092–2096. [413] B. Bhattacharya, R. Dey, P. Pachfule, R. Banerjee, D. Ghoshal, Cryst. Growth Des. 13 (2012) 731–739. [414] S. Parshamoni, S. Konar, CrystEngComm 18 (2016) 4395–4404. [415] M.-Y. Sun, D.-M. Chen, Inorg. Chem. Commun. 89 (2018) 18–21. [416] M.Y. Masoomi, S. Beheshti, A. Morsali, Cryst. Growth Des. 15 (2015) 2533– 2538. [417] M.Y. Masoomi, S. Beheshti, A. Morsali, J. Mater. Chem. A 2 (2014) 16863– 16866. [418] B. Ugale, C.M. Nagaraja, RSC Adv. 6 (2016) 28854–28864. [419] E. Tahmasebi, M.Y. Masoomi, Y. Yamini, A. Morsali, Inorg. Chem. 54 (2015) 425–433. [420] M.Y. Masoomi, M. Bagheri, A. Morsali, P.C. Junk, Inorganic Chemistry Frontiers 3 (2016) 944–951. [421] M. Bagheri, M.Y. Masoomi, A. Morsali, A. Schoedel, ACS Appl. Mater. Interfaces 8 (2016) 21472–21479. [422] C.B. Fan, Z.Q. Liu, L.L. Gong, A.M. Zheng, L. Zhang, C.S. Yan, H.Q. Wu, X.F. Feng, F. Luo, Chem. Commun. 53 (2017) 763–766. [423] R. Huang, M.R. Hill, R. Babarao, N.V. Medhekar, J. Phys. Chem. C 120 (2016) 16658–16667. [424] H. Huang, H. Sato, T. Aida, J. Am. Chem. Soc. 139 (2017) 8784–8787. [425] J. Park, D. Yuan, K.T. Pham, J.-R. Li, A. Yakovenko, H.-C. Zhou, J. Am. Chem. Soc. 134 (2012) 99–102. [426] N. Prasetya, B.C. Donose, B.P. Ladewig, J. Mater. Chem. A 6 (2018) 16390– 16402. [427] Z. Wang, A. Knebel, S. Grosjean, D. Wagner, S. Bräse, C. Wöll, J. Caro, L. Heinke, Nature Communications 7 (2016) 13872. [428] R. Lyndon, K. Konstas, B.P. Ladewig, P.D. Southon, P.C.J. Kepert, M.R. Hill, Angew. Chem. 125 (2013) 3783–3786. [429] L.L. Gong, X.F. Feng, F. Luo, Inorg. Chem. 54 (2015) 11587–11589. [430] L.-L. Dang, X.-J. Zhang, L. Zhang, J.-Q. Li, F. Luo, X.-F. Feng, J. Coord. Chem. 69 (2016) 1179–1187. [431] A. Knebel, L. Sundermann, A. Mohmeyer, I. Strauß, S. Friebe, P. Behrens, J. Caro, Chem. Mater. 29 (2017) 3111–3117. [432] W.-C. Song, X.-Z. Cui, Z.-Y. Liu, E.-C. Yang, X.-J. Zhao, Sci. Rep. 6 (2016) 34870. [433] X. Meng, B. Gui, D. Yuan, M. Zeller, C. Wang, Science Advances 2 (2016). [434] J.S. Caddy, T.B. Faust, I.M. Walton, J.M. Cox, J.B. Benedict, M.B. Solomon, P.D. Southon, C.J. Kepert, D.M. D’Alessandro, Aust. J. Chem. 70 (2017) 1171–1179. [435] S. Castellanos, A. Goulet-Hanssens, F. Zhao, A. Dikhtiarenko, A. Pustovarenko, S. Hecht, J. Gascon, F. Kapteijn, D. Bléger, Chem. Eur. J. 22 (2015) 746–752.
54
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
[436] K. Müller, A. Knebel, F. Zhao, D. Bléger, J. Caro, L. Heinke, Chem. Eur. J. 23 (2017) 5434–5438. [437] J.W. Brown, B.L. Henderson, M.D. Kiesz, A.C. Whalley, W. Morris, S. Grunder, H. Deng, H. Furukawa, J.I. Zink, J.F. Stoddart, O.M. Yaghi, Chem. Sci. 4 (2013) 2858–2864. [438] W.-Q. Fu, M. Liu, Z.-G. Gu, S.-M. Chen, J. Zhang, Cryst. Growth Des. 16 (2016) 5487–5492. [439] A. Modrow, D. Zargarani, R. Herges, N. Stock, Dalton Trans. 40 (2011) 4217– 4222. [440] A. Modrow, D. Zargarani, R. Herges, N. Stock, Dalton Trans. 41 (2012) 8690– 8696. [441] S. Bernt, M. Feyand, A. Modrow, J. Wack, J. Senker, N. Stock, Eur. J. Inorg. Chem. 2011 (2011) 5378–5383. [442] D. Hermann, H. Emerich, R. Lepski, D. Schaniel, U. Ruschewitz, Inorg. Chem. 52 (2013) 2744–2749. [443] Z. Wang, L. Heinke, J. Jelic, M. Cakici, M. Dommaschk, R.J. Maurer, H. Oberhofer, S. Grosjean, R. Herges, S. Bräse, K. Reuter, C. Wöll, Phys. Chem. Chem. Phys. 17 (2015) 14582–14587. [444] Y.-P. Li, L.-J. Zhang, W.-J. Ji, J. Mol. Struct. 1133 (2017) 607–614. [445] L.T.M. Hoang, L.H. Ngo, H.L. Nguyen, H.T.H. Nguyen, C.K. Nguyen, B.T. Nguyen, Q.T. Ton, H.K.D. Nguyen, K.E. Cordova, T. Truong, Chem. Commun. 51 (2015) 17132–17135. [446] P. Kanoo, R. Haldar, S.K. Reddy, A. Hazra, S. Bonakala, R. Matsuda, S. Kitagawa, S. Balasubramanian, T.K. Maji, Chem. Eur. J. 22 (2016) 15864–15873. [447] S. Chand, A. Pal, M.C. Das, Chem. Eur. J. 24 (2018) 5982–5986. [448] Q. Yang, V. Guillerm, F. Ragon, A.D. Wiersum, P.L. Llewellyn, C. Zhong, T. Devic, C. Serre, G. Maurin, Chem. Commun. 48 (2012) 9831–9833. [449] R. Singh, M. Ahmad, P.K. Bharadwaj, Cryst. Growth Des. 12 (2012) 5025– 5034. [450] P.P. Meng, L.L. Dang, X.J. Zhang, L.F. Ma, J.Q. Li, L.L. Gong, L. Zhang, L.N. Meng, F. Luo, CrystEngComm 18 (2016) 6336–6340. [451] S. Chand, A. Pal, S.C. Pal, M.C. Das, Eur. J. Inorg. Chem. 2018 (2018) 2785– 2792. [452] J. Jiang, AlChE J. 55 (2009) 2422–2432. [453] X.-S. Wang, S. Ma, K. Rauch, J.M. Simmons, D. Yuan, X. Wang, T. Yildirim, W.C. Cole, J.J. López, A. Meijere, H. Zhou, Chem. Mater. 20 (2008) 3145–3152. [454] S. Liu, M. Luo, J. Li, F. Luo, L. Ke, J. Ma, J. Radioanal. Nucl. Chem. 310 (2016) 353–362. [455] N. Goel, N. Kumar, RSC Adv. 8 (2018) 10746–10755. [456] S. Chand, M. Mondal, S.C. Pal, A. Pal, S. Maji, D. Mandal, M.C. Das, New J. Chem. 42 (2018) 12865–12871. [457] A. Pankajakshan, D. Kuznetsov, S. Mandal, Inorg. Chem. (2019). [458] Z. Jin, H. Zhao, D. Yang, X. Yao, G. Zhu, Inorg. Chem. Commun. 25 (2012) 74– 78. [459] S.S. Nagarkar, T. Saha, A.V. Desai, P. Talukdar, S.K. Ghosh, Sci. Rep. 4 (2014) 7053. [460] H. Hahm, S. Kim, H. Ha, S. Jung, Y. Kim, M. Yoon, M. Kim, CrystEngComm 17 (2015) 8418–8422. [461] J.-C. Wang, J.-P. Ma, Q.-K. Liu, Y.-H. Hu, Y.-B. Dong, Chem. Commun. 52 (2016) 6989–6992. [462] C. Naga Babu, P. Suresh, K. Srinivas, A. Sathyanarayana, N. Sampath, G. Prabusankar, Dalton Trans. 45 (2016) 8164–8173. [463] D. Liu, G. Li, H. Liu, Appl Surf Sci 428 (2018) 218–225. [464] J. Liang, R.-P. Chen, X.-Y. Wang, T.-T. Liu, X.-S. Wang, Y.-B. Huang, R. Cao, Chem. Sci. 8 (2017) 1570–1575. [465] J. Liang, Y.-Q. Xie, X.-S. Wang, Q. Wang, T.-T. Liu, Y.-B. Huang, R. Cao, Chem. Commun. 54 (2018) 342–345. [466] F. Adam, J.N. Appaturi, E.-P. Ng, J. Mol. Catal. A: Chem. 386 (2014) 42–48. [467] Y. Dong, Y. Li, Y.-L. Wei, J.-C. Wang, J.-P. Ma, J. Ji, B.-J. Yao, Y.-B. Dong, Chem. Commun. 52 (2016) 10505–10508. [468] G.-Q. Kong, S. Ou, C. Zou, C.-D. Wu, J. Am. Chem. Soc. 134 (2012) 19851– 19857. [469] G.-Q. Kong, X. Xu, C. Zou, C.-D. Wu, Chem. Commun. 47 (2011) 11005–11007. [470] C.I. Ezugwu, B. Mousavi, M.A. Asraf, Z. Luo, F. Verpoort, J Catal 344 (2016) 445–454. [471] A. Burgun, R.S. Crees, M.L. Cole, C.J. Doonan, C.J. Sumby, Chem. Commun. 50 (2014) 11760–11763. [472] F. Carson, E. Martínez-Castro, R. Marcos, G.G. Miera, K. Jansson, X. Zou, B. Martín-Matute, Chem. Commun. 51 (2015) 10864–10867. [473] J.Y. Lee, J.M. Roberts, O.K. Farha, A.A. Sarjeant, K.A. Scheidt, J.T. Hupp, Inorg. Chem. 48 (2009) 9971–9973. [474] S. Sen, S. Neogi, A. Aijaz, Q. Xu, P.K. Bharadwaj, Inorg. Chem. 53 (2014) 7591– 7598. [475] S. Wang, Q. Yang, J. Zhang, X. Zhang, C. Zhao, L. Jiang, C.-Y. Su, Inorg. Chem. 52 (2013) 4198–4204. [476] M.B. Lalonde, R.B. Getman, J.Y. Lee, J.M. Roberts, A.A. Sarjeant, K.A. Scheidt, P. A. Georgiev, J.P. Embs, J. Eckert, O.K. Farha, R.Q. Snurr, J.T. Hupp, CrystEngComm 15 (2013) 9408–9414. [477] P.K. Capon, A. Burgun, C.J. Coghlan, R.S. Crees, C.J. Doonan, C.J. Sumby, CrystEngComm 18 (2016) 7003–7010. [478] W. He, N. Li, X. Wang, T. Hu, X. Bu, Chin. Chem. Lett. 29 (2018) 857–860. [479] B.-Q. Song, X.-L. Wang, Y.-T. Zhang, X.-S. Wu, H.-S. Liu, K.-Z. Shao, Z.-M. Su, Chem. Commun. 51 (2015) 9515–9518. [480] S. Sen, N.N. Nair, T. Yamada, H. Kitagawa, P.K. Bharadwaj, J. Am. Chem. Soc. 134 (2012) 19432–19437.
[481] S. Sen, T. Yamada, H. Kitagawa, P.K. Bharadwaj, Cryst. Growth Des. 14 (2014) 1240–1244. [482] B.-Q. Song, X.-L. Wang, G.-S. Yang, H.-N. Wang, J. Liang, K.-Z. Shao, Z.-M. Su, CrystEngComm 16 (2014) 6882–6888. [483] C. Zhang, L. Sun, Y. Yan, H. Shi, B. Wang, Z. Liang, J. Li, J. Mater. Chem. C 5 (2017) 8999–9004. [484] L.-X. Cai, C. Chen, B. Tan, Y.-J. Zhang, X.-D. Yang, J. Zhang, CrystEngComm 17 (2015) 2353–2358. [485] D. Aulakh, J.R. Varghese, M. Wriedt, Inorg. Chem. 54 (2015) 1756–1764. [486] J.-K. Sun, L.-X. Cai, Y.-J. Chen, Z.-H. Li, J. Zhang, Chem. Commun. 47 (2011) 6870–6872. [487] H.-Y. Li, Y.-L. Wei, X.-Y. Dong, S.-Q. Zang, T.C.W. Mak, Chem. Mater. 27 (2015) 1327–1331. [488] P.-X. Li, M.-S. Wang, G.-C. Guo, Cryst. Growth Des. 16 (2016) 3709–3715. [489] S. Hu, M. You, S. Chen, Z. Fu, CrystEngComm 18 (2016) 7221–7224. [490] X.-D. Yang, C. Chen, Y.-J. Zhang, L.-X. Cai, B. Tan, J. Zhang, Dalton Trans. 45 (2016) 4522–4527. [491] J.-K. Sun, P. Wang, C. Chen, X.-J. Zhou, L.-M. Wu, Y.-F. Zhang, J. Zhang, Dalton Trans. 41 (2012) 13441–13446. [492] J.-K. Sun, P. Wang, Q.-X. Yao, Y.-J. Chen, Z.-H. Li, Y.-F. Zhang, L.-M. Wu, J. Zhang, J. Mater. Chem. 22 (2012) 12212–12219. [493] C. Zhang, L. Sun, Y. Yan, Y. Liu, Z. Liang, Y. Liu, J. Li, J. Mater. Chem. C 5 (2017) 2084–2089. [494] H. Chen, G. Zheng, M. Li, Y. Wang, Y. Song, C. Han, J. Dai, Z. Fu, Chem. Commun. 50 (2014) 13544–13546. [495] C. Zhang, L. Sun, C. Zhang, S. Wan, Z. Liang, J. Li, Inorganic Chemistry Frontiers 3 (2016) 814–820. [496] J.-K. Sun, C. Chen, L.-X. Cai, C.-X. Ren, B. Tan, J. Zhang, Chem. Commun. 50 (2014) 15956–15959. [497] W. An, D. Aulakh, X. Zhang, W. Verdegaal, K.R. Dunbar, M. Wriedt, Chem. Mater. 28 (2016) 7825–7832. [498] J. Yu, Y. Cui, C.-D. Wu, Y. Yang, B. Chen, G. Qian, J. Am. Chem. Soc. 137 (2015) 4026–4029. [499] K.-M. Wang, L. Du, Y.-L. Ma, Q.-H. Zhao, Inorg. Chem. Commun. 68 (2016) 45– 49. [500] S. Hu, J. Zhang, S. Chen, J. Dai, Z. Fu, A.C.S. Appl, Mater. Interfaces 9 (2017) 39926–39929. [501] T. Gong, P. Li, Q. Sui, J. Chen, J. Xu, E.-Q. Gao, J. Mater. Chem. A 6 (2018) 9236– 9244. [502] S. Rapti, D. Sarma, S.A. Diamantis, E. Skliri, G.S. Armatas, A.C. Tsipis, Y.S. Hassan, M. Alkordi, C.D. Malliakas, M.G. Kanatzidis, T. Lazarides, J.C. Plakatouras, M.J. Manos, J. Mater. Chem. A 5 (2017) 14707–14719. [503] C. Zhang, Y. Liu, L. Sun, H. Shi, C. Shi, Z. Liang, J. Li, Chem. Eur. J. 24 (2017) 2718–2724. [504] L. Xu, Y. Luo, L. Sun, S. Pu, M. Fang, R.-X. Yuan, H.-B. Du, Dalton Trans. 45 (2016) 8614–8621. [505] B. Wu, W.-H. Zhang, Z.-G. Ren, J.-P. Lang, Chem. Commun. 51 (2015) 14893– 14896. [506] M. Leroux, N. Mercier, J.-P. Bellat, G. Weber, I. Bezverkhyy, Cryst. Growth Des. 17 (2017) 2828–2835. [507] Y.-J. Zhang, C. Chen, L.-X. Cai, B. Tan, X.-D. Yang, J. Zhang, M. Ji, Dalton Trans. 46 (2017) 7092–7097. [508] M. Leroux, N. Mercier, M. Allain, M.-C. Dul, J. Dittmer, A.H. Kassiba, J.-P. Bellat, G. Weber, I. Bezverkhyy, Inorg. Chem. 55 (2016) 8587–8594. [509] B. Tan, C. Chen, L.-X. Cai, Y.-J. Zhang, X.-Y. Huang, J. Zhang, Inorg. Chem. 54 (2015) 3456–3461. [510] H. Tan, Q. Chen, T. Chen, H. Liu, ACS Appl. Mater. Interfaces (2018). [511] M. Higuchi, K. Nakamura, S. Horike, Y. Hijikata, N. Yanai, T. Fukushima, J. Kim, K. Kato, M. Takata, D. Watanabe, S. Oshima, S. Kitagawa, Angew. Chem. Int. Ed. 51 (2012) 8369–8372. [512] H.-Q. Zhao, G.-H. Qiu, Z. Liang, M.-M. Li, B. Sun, L. Qin, S.-P. Yang, W.-H. Chen, J.-X. Chen, Anal. Chim. Acta 922 (2016) 55–63. [513] B. Sun, H.-Q. Zhao, B.-P. Xie, L.-P. Bai, Z.-H. Jiang, J.-X. Chen, J. Inorg. Biochem. 176 (2017) 17–23. [514] S.-P. Yang, S.-R. Chen, S.-W. Liu, X.-Y. Tang, L. Qin, G.-H. Qiu, J.-X. Chen, W.-H. Chen, Anal. Chem. 87 (2015) 12206–12214. [515] S.-P. Yang, W. Zhao, P.-P. Hu, K.-Y. Wu, Z.-H. Jiang, L.-P. Bai, M.-M. Li, J.-X. Chen, Inorg. Chem. 56 (2017) 14880–14887. [516] G.-H. Qiu, W.-Z. Lu, P.-P. Hu, Z.-H. Jiang, L.-P. Bai, T.-R. Wang, M.-M. Li, J.-X. Chen, J. Inorg. Biochem. 177 (2017) 138–142. [517] L. Qin, L.-X. Lin, Z.-P. Fang, S.-P. Yang, G.-H. Qiu, J.-X. Chen, W.-H. Chen, Chem. Commun. 52 (2016) 132–135. [518] G.-H. Qiu, Z.-H. Weng, P.-P. Hu, W.-J. Duan, B.-P. Xie, B. Sun, X.-Y. Tang, J.-X. Chen, Talanta 180 (2018) 396–402. [519] B.-P. Xie, G.-H. Qiu, P.-P. Hu, Z. Liang, Y.-M. Liang, B. Sun, L.-P. Bai, Z.-H. Jiang, J.-X. Chen, Sens. Actuators, B 254 (2018) 1133–1140. [520] L. Qin, Z.-Y. Sun, K. Cheng, S.-W. Liu, J.-X. Pang, L.-M. Xia, W.-H. Chen, Z. Cheng, J.-X. Chen, A.C.S. Appl, Mater. Interfaces 9 (2017) 41378–41386. [521] Y. Hu, X. Dong, J. Nan, W. Jin, X. Ren, N. Xu, Y.M. Lee, Chem. Commun. 47 (2011) 737–739. [522] R. Zhong, J. Liu, X. Huang, X. Yu, C. Sun, G. Chen, R. Zou, CrystEngComm 17 (2015) 8221–8225. [523] Y. Ling, F. Yang, M. Deng, Z. Chen, X. Liu, L. Weng, Y. Zhou, Dalton Trans. 41 (2012) 4007–4011. [524] Z. Chen, S. Xiang, H.D. Arman, P. Li, D. Zhao, B. Chen, Eur. J. Inorg. Chem. 2011 (2011) 2227–2231.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 [525] D. Wu, J.J. Gassensmith, D. Gouvêa, S. Ushakov, J.F. Stoddart, A. Navrotsky, J. Am. Chem. Soc. 135 (2013) 6790–6793. [526] E. Klontzas, A. Mavrandonakis, E. Tylianakis, G.E. Froudakis, Nano Lett. 8 (2008) 1572–1576. [527] E. García-Pérez, P. Serra-Crespo, S. Hamad, F. Kapteijn, J. Gascon, Phys. Chem. Chem. Phys. 16 (2014) 16060–16066. [528] J. Yang, X. Yan, T. Xue, Y. Liu, RSC Adv. 6 (2016) 55266–55271. [529] Z. Lu, Y. Xing, L. Du, H. He, J. Zhang, C. Hang, RSC Adv. 7 (2017) 47219–47224. [530] S. Yang, J. Sun, A.J. Ramirez-Cuesta, S.K. Callear, W.I.F. David, D.P. Anderson, R. Newby, A.J. Blake, J.E. Parker, C.C. Tang, M. Schröder, Nature Chemistry 4 (2012) 887. [531] F. Luo, J.L. Chen, L.L. Dang, W.N. Zhou, H.L. Lin, J.Q. Li, S.J. Liu, M.B. Luo, J. Mater. Chem. A 3 (2015) 9616–9620. [532] J.Y. Song, S.H. Jhung, Chem Eng J 322 (2017) 366–374. [533] X. Zhao, H. Zhao, W. Dai, Y. Wei, Y. Wang, Y. Zhang, L. Zhi, H. Huang, Z. Gao, J. Colloid Interface Sci. 526 (2018) 28–34. [534] H. Xu, Y. Dong, Y. Wu, W. Ren, T. Zhao, S. Wang, J. Gao, J. Solid State Chem. 258 (2018) 441–446. [535] S. Pal, P.K. Bharadwaj, Cryst. Growth Des. 16 (2016) 5852–5858. [536] G.-L. Liu, Y.-J. Qin, L. Jing, G.-Y. Wei, H. Li, Chem. Commun. 49 (2013) 1699– 1701. [537] T. Maihom, S. Wannakao, B. Boekfa, J. Limtrakul, J. Phys. Chem. C 117 (2013) 17650–17658. [538] S. Biswas, J. Chakraborty, V. Singh Parmar, S.P. Bera, N. Ganguli, S. Konar, Inorg. Chem. 56 (2017) 4956–4965. [539] P.V. Dau, L.R. Polanco, S.M. Cohen, Dalton Trans. 42 (2013) 4013–4018. [540] T.K. Kim, M.P. Suh, Chem. Commun. 47 (2011) 4258–4260. [541] Y.-Q. Lan, H.-L. Jiang, S.-L. Li, Q. Xu, Adv. Mater. 23 (2011) 5015–5020. [542] L. Bai, B. Tu, Y. Qi, Q. Gao, D. Liu, Z. Liu, L. Zhao, Q. Li, Y. Zhao, Chem. Commun. 52 (2016) 3003–3006. [543] S. Henke, R. Schmid, J.-D. Grunwaldt, R.A. Fischer, Chem. Eur. J. 16 (2010) 14296–14306. [544] S. Henke, D.C. Florian Wieland, M. Meilikhov, M. Paulus, C. Sternemann, K. Yusenko, R.A. Fischer, CrystEngComm 13 (2011) 6399–6404. [545] S. Henke, R.A. Fischer, J. Am. Chem. Soc. 133 (2011) 2064–2067. [546] P.K. Thallapally, J. Tian, M. Radha Kishan, C.A. Fernandez, S.J. Dalgarno, P.B. McGrail, J.E. Warren, J.L. Atwood, J. Am. Chem. Soc. 130 (2008) 16842–16843. [547] S. Henke, A. Schneemann, A. Wütscher, R.A. Fischer, J. Am. Chem. Soc. 134 (2012) 9464–9474. [548] Z.I. Salih, Y.-J. Guo, J.-J. Zheng, X. Zhao, Computational and Theoretical Chemistry 1058 (2015) 28–33. [549] G. Orcajo, H. Montes-Andrés, J.A. Villajos, C. Martos, J.A. Botas, G. Calleja, Int. J. Hydrogen Energy (2018). [550] J. Wang, Y. Li, M. Jiang, Y. Liu, L. Zhang, P. Wu, Chem. Eur. J. 22 (2016) 13023– 13027. [551] Y.-Y. Jiang, S.-K. Ren, J.-P. Ma, Q.-K. Liu, Y.-B. Dong, Chem. Eur. J. 15 (2009) 10742–10746. [552] J.A. Foster, S. Henke, A. Schneemann, R.A. Fischer, A.K. Cheetham, Chem. Commun. 52 (2016) 10474–10477. [553] S. Henke, A. Schneemann, R.A. Fischer, Adv. Funct. Mater. 23 (2013) 5990– 5996. [554] R.K. Deshpande, G.I.N. Waterhouse, G.B. Jameson, S.G. Telfer, Chem. Commun. 48 (2012) 1574–1576. [555] Y. Xiong, Y.-Z. Fan, R. Yang, S. Chen, M. Pan, J.-J. Jiang, C.-Y. Su, Chem. Commun. 50 (2014) 14631–14634. [556] A.S. Munn, S. Amabilino, T.W. Stevens, L.M. Daniels, G.J. Clarkson, F. Millange, M.J. Lennox, T. Düren, S. Bourelly, P.L. Llewellyn, R.I. Walton, Cryst. Growth Des. 15 (2015) 891–899. [557] L. Aboutorabi, A. Morsali, E. Tahmasebi, O. Büyükgüngor, Inorg. Chem. 55 (2016) 5507–5513. [558] N. Shokouhfar, L. Aboutorabi, A. Morsali, Dalton Trans. 47 (2018) 14549– 14555. [559] L. Hu, X.-M. Lin, J. Lin, R.-Q. Zhang, D.-L. Zhang, Y.-P. Cai, CrystEngComm 18 (2016) 9307–9315. [560] C. Yu, Z. Shao, H. Hou, Chem. Sci. 8 (2017) 7611–7619. [561] C. Song, Y. Ling, L. Jin, M. Zhang, D.-L. Chen, Y. He, Dalton Trans. 45 (2016) 190–197. [562] A.S. Munn, F. Millange, M. Frigoli, N. Guillou, C. Falaise, V. Stevenson, C. Volkringer, T. Loiseau, G. Cibin, R.I. Walton, CrystEngComm 18 (2016) 8108– 8114. [563] T. Liu, J.-X. Che, Y.-Z. Hu, X.-W. Dong, X.-Y. Liu, C.-M. Che, Chem. Eur. J. 20 (2014) 14090–14095. [564] L. Liang, Q. Chen, F. Jiang, D. Yuan, J. Qian, G. Lv, H. Xue, L. Liu, H.-L. Jiang, M. Hong, J. Mater. Chem. A 4 (2016) 15370–15374. [565] M.-Q. Li, Y.-L. Wong, T.-S. Lum, K. Sze-Yin Leung, P.K.S. Lam, Z. Xu, J. Mater. Chem. A 6 (2018) 14566–14570. [566] F. Ke, L.-G. Qiu, Y.-P. Yuan, F.-M. Peng, X. Jiang, A.-J. Xie, Y.-H. Shen, J.-F. Zhu, J. Hazard. Mater. 196 (2011) 36–43. [567] L. Huang, M. He, B. Chen, B. Hu, J. Mater. Chem. A 4 (2016) 5159–5166. [568] M.R. Sohrabi, Microchim. Acta 181 (2014) 435–444. [569] L. Ding, X. Luo, P. Shao, J. Yang, D. Sun, ACS Sustainable Chemistry & Engineering (2018). [570] J. Zhang, Z. Xiong, C. Li, C. Wu, J. Mol. Liq. 221 (2016) 43–50. [571] S. Halder, J. Mondal, J. Ortega-Castro, A. Frontera, P. Roy, Dalton Trans. 46 (2017) 1943–1950.
55
[572] K.-K. Yee, N. Reimer, J. Liu, S.-Y. Cheng, S.-M. Yiu, J. Weber, N. Stock, Z. Xu, J. Am. Chem. Soc. 135 (2013) 7795–7798. [573] S. Parshamoni, S. Sanda, H.S. Jena, S. Konar, Dalton Trans. 43 (2014) 7191– 7199. [574] M. Yoon, D. Moon, Micropor. Mesopor. Mat. 215 (2015) 116–122. [575] S.-Y. Jiang, W.-W. He, S.-L. Li, Z.-M. Su, Y.-Q. Lan, Inorg. Chem. 57 (2018) 6118–6123. [576] K. Wang, H. Huang, W. Xue, D. Liu, X. Zhao, Y. Xiao, Z. Li, Q. Yang, L. Wang, C. Zhong, CrystEngComm 17 (2015) 3586–3590. [577] J. He, M. Zha, J. Cui, M. Zeller, A.D. Hunter, S.-M. Yiu, S.-T. Lee, Z. Xu, J. Am. Chem. Soc. 135 (2013) 7807–7810. [578] X.-Y. Yang, S. Yuan, J.-S. Qin, C. Lollar, A. Alsalme, H.-C. Zhou, Materials Chemistry Frontiers 1 (2017) 1764–1767. [579] J.-H. Wang, G.-M. Tang, Y.-T. Wang, Y.-Z. Cui, J.-J. Wang, S.W. Ng, Dalton Trans. 44 (2015) 17829–17840. [580] Y. Zang, J. Shi, F. Zhang, Y. Zhong, W. Zhu, Catal. Sci. Tech. 3 (2013) 2044– 2049. [581] M.G. Goesten, J. Juan-Alcañiz, E.V. Ramos-Fernandez, K.B. Sai Sankar Gupta, E. Stavitski, H. van Bekkum, J. Gascon, F. Kapteijn, J. Catal 281 (2011) 177–187. [582] G. Zhang, H. Yang, H. Fei, ACS Catal. 8 (2018) 2519–2525. [583] G. Zhang, G. Wei, Z. Liu, S.R.J. Oliver, H. Fei, Chem. Mater. 28 (2016) 6276– 6281. [584] W.J. Phang, H. Jo, W.R. Lee, J.H. Song, K. Yoo, B. Kim, C.S. Hong, Angew. Chem. Int. Ed. 54 (2015) 5142–5146. [585] N. Wong, J.A. Hurd, R. Vaidhyanathan, G.K.H. Shimizu, Can. J. Chem. 93 (2015) 988–991. [586] G. Zhang, H. Fei, Chem. Commun. 53 (2017) 4156–4159. [587] X.-Y. Dong, R. Wang, J.-B. Li, S.-Q. Zang, H.-W. Hou, T.C.W. Mak, Chem. Commun. 49 (2013) 10590–10592. [588] X. Meng, S.-Y. Song, X.-Z. Song, M. Zhu, S.-N. Zhao, L.-L. Wu, H.-J. Zhang, Chem. Commun. 51 (2015) 8150–8152. [589] H. selective Fe, Journal of Materials Chemistry A: Materials for Energy and Sustainability, 3 (2015) 641-647. [590] Z. Li, G. He, Y. Zhao, Y. Cao, H. Wu, Y. Li, Z. Jiang, J. Power Sources 262 (2014) 372–379. [591] G.K.H. Shimizu, J.M. Taylor, S. Kim, Science 341 (2013) 354–355. [592] S. Shalini, V.M. Dhavale, K.M. Eldho, S. Kurungot, T.G. Ajithkumar, R. Vaidhyanathan, Sci. Rep. 6 (2016) 32489. [593] N. Sikdar, D. Dutta, R. Haldar, T. Ray, A. Hazra, A.J. Bhattacharyya, T.K. Maji, J. Phys. Chem. C 120 (2016) 13622–13629. [594] E. Neofotistou, C.D. Malliakas, P.N. Trikalitis, Chem. Eur. J. 15 (2009) 4523– 4527. [595] W. Lu, D. Yuan, J. Sculley, D. Zhao, R. Krishna, H.-C. Zhou, J. Am. Chem. Soc. 133 (2011) 18126–18129. [596] N. Zhang, D. Peng, H. Wu, Y. Ren, L. Yang, X. Wu, Y. Wu, Z. Qu, Z. Jiang, X. Cao, Journal of Membrane Science 549 (2018) 670–679. [597] A. Mavrandonakis, E. Klontzas, E. Tylianakis, G.E. Froudakis, J. Am. Chem. Soc. 131 (2009) 13410–13414. [598] C. Xu, K. Sun, Y.-X. Zhou, X. Ma, H.-L. Jiang, Chem. Commun. 54 (2018) 2498– 2501. [599] I. Ahmed, Z. Hasan, N.A. Khan, S.H. Jhung, Applied Catalysis B: Environmental 129 (2013) 123–129. [600] X.-P. Luo, S.-Y. Fu, Y.-M. Du, J.-Z. Guo, B. Li, Micropor. Mesopor. Mat. 237 (2017) 268–274. [601] C. Kang, Y. Peng, Y. Tang, H. Huang, C. Zhong, Ind. Eng. Chem. Res. 56 (2017) 13866–13873. [602] H.-H. Wang, L.-J. Zhou, Y.-L. Wang, Q.-Y. Liu, Inorg. Chem. Commun. 73 (2016) 94–97. [603] A. Weiss, N. Reimer, N. Stock, M. Tiemann, T. Wagner, Phys. Chem. Chem. Phys. 17 (2015) 21634–21642. [604] N. Singh, U.P. Singh, R.J. Butcher, CrystEngComm 19 (2017) 7009–7020. [605] Y. Zhang, B. Li, R. Krishna, Z. Wu, D. Ma, Z. Shi, T. Pham, K. Forrest, B. Space, S. Ma, Chem. Commun. 51 (2015) 2714–2717. [606] N. Wei, Y.-R. Zhang, Z.-B. Han, CrystEngComm 15 (2013) 8883–8886. [607] Y. Luan, N. Zheng, Y. Qi, J. Tang, G. Wang, Catal. Sci. Tech. 4 (2014) 925–929. [608] K.-K. Yee, Y.-L. Wong, Z. Xu, Dalton Trans. 45 (2016) 5334–5338. [609] T.K. Pal, D. De, S. Senthilkumar, S. Neogi, P.K. Bharadwaj, Inorg. Chem. 55 (2016) 7835–7842. [610] D.-S. Zhang, Z. Chang, Y.-F. Li, Z.-Y. Jiang, Z.-H. Xuan, Y.-H. Zhang, J.-R. Li, Q. Chen, T.-L. Hu, X.-H. Bu, Sci. Rep. 3 (2013) 3312. [611] P. Pachfule, Y. Chen, J. Jiang, R. Banerjee, Chem. Eur. J. 18 (2011) 688–694. [612] O. Alduhaish, R.-B. Lin, H. Wang, B. Li, H.D. Arman, T.-L. Hu, B. Chen, Cryst. Growth Des. 18 (2018) 4522–4527. [613] Y. Liu, J.-R. Li, W.M. Verdegaal, T.-F. Liu, H.-C. Zhou, Chem. Eur. J. 19 (2013) 5637–5643. [614] A.K. Gupta, D. De, K. Tomar, P.K. Bharadwaj, Dalton Trans. 47 (2018) 1624– 1634. [615] Y.-S. Bae, O.K. Farha, J.T. Hupp, R.Q. Snurr, J. Mater. Chem. 19 (2009) 2131– 2134. [616] A. Santra, M.S. Lah, P.K. Bharadwaj, Z Anorg Allg Chem 640 (2014) 1134– 1140. [617] S.S. Dhankhar, M. Kaur, C.M. Nagaraja, Eur. J. Inorg. Chem. 2015 (2015) 5669– 5676. [618] P. Pachfule, R. Das, P. Poddar, R. Banerjee, Inorg. Chem. 50 (2011) 3855–3865. [619] P. Deria, J.E. Mondloch, E. Tylianakis, P. Ghosh, W. Bury, R.Q. Snurr, J.T. Hupp, O.K. Farha, J. Am. Chem. Soc. 135 (2013) 16801–16804.
56
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023
[620] K.-J. Chen, Hayley S. Scott, David G. Madden, T. Pham, A. Kumar, A. Bajpai, M. Lusi, Katherine A. Forrest, B. Space, John J. Perry, Michael J. Zaworotko, Chem 1 (2016) 753–765. [621] R. Kotani, A. Kondo, K. Maeda, Chem. Commun. 48 (2012) 11316–11318. [622] A. Kondo, H. Noguchi, S. Ohnishi, H. Kajiro, A. Tohdoh, Y. Hattori, W.-C. Xu, H. Tanaka, H. Kanoh, K. Kaneko, Nano Lett. 6 (2006) 2581–2584. [623] S.-I. Noro, K. Fukuhara, Y. Hijikata, K. Kubo, T. Nakamura, Inorg. Chem. 52 (2013) 5630–5632. [624] A. Kondo, A. Chinen, H. Kajiro, T. Nakagawa, K. Kato, M. Takata, Y. Hattori, F. Okino, T. Ohba, K. Kaneko, H. Kanoh, Chem. Eur. J. 15 (2009) 7549–7553. [625] P.M. Bhatt, Y. Belmabkhout, A. Cadiau, K. Adil, O. Shekhah, A. Shkurenko, L.J. Barbour, M. Eddaoudi, J. Am. Chem. Soc. 138 (2016) 9301–9307. [626] S.-I. Noro, D. Tanaka, H. Sakamoto, S. Shimomura, S. Kitagawa, S. Takeda, K. Uemura, H. Kita, T. Akutagawa, T. Nakamura, Chem. Mater. 21 (2009) 3346– 3355. [627] S.-I. Noro, R. Matsuda, Y. Hijikata, Y. Inubushi, S. Takeda, S. Kitagawa, Y. Takahashi, M. Yoshitake, K. Kubo, T. Nakamura, ChemPlusChem 80 (2015) 1517–1524. [628] S.-I. Noro, Y. Hijikata, M. Inukai, T. Fukushima, S. Horike, M. Higuchi, S. Kitagawa, T. Akutagawa, T. Nakamura, Inorg. Chem. 52 (2013) 280–285. [629] P. Kanoo, S.K. Reddy, G. Kumari, R. Haldar, C. Narayana, S. Balasubramanian, T.K. Maji, Chem. Commun. 48 (2012) 8487–8489. [630] P. Nugent, V. Rhodus, T. Pham, B. Tudor, K. Forrest, L. Wojtas, B. Space, M. Zaworotko, Chem. Commun. 49 (2013) 1606–1608. [631] S.K. Elsaidi, M.H. Mohamed, H.T. Schaef, A. Kumar, M. Lusi, T. Pham, K.A. Forrest, B. Space, W. Xu, G.J. Halder, J. Liu, M.J. Zaworotko, P.K. Thallapally, Chem. Commun. 51 (2015) 15530–15533. [632] S.K. Elsaidi, M.H. Mohamed, T. Pham, T. Hussein, L. Wojtas, M.J. Zaworotko, B. Space, Cryst. Growth Des. 16 (2016) 1071–1080. [633] S.D. Burd, S. Ma, J.A. Perman, B.J. Sikora, R.Q. Snurr, P.K. Thallapally, J. Tian, L. Wojtas, M.J. Zaworotko, J. Am. Chem. Soc. 134 (2012) 3663–3666. [634] P. Nugent, Y. Belmabkhout, S.D. Burd, A.J. Cairns, R. Luebke, K. Forrest, T. Pham, S. Ma, B. Space, L. Wojtas, M. Eddaoudi, M.J. Zaworotko, Nature 495 (2013) 80. [635] G. Chang, H. Wen, B. Li, W. Zhou, H. Wang, K. Alfooty, Z. Bao, B. Chen, Cryst. Growth Des. 16 (2016) 3395–3399. [636] N. Nijem, P. Canepa, U. Kaipa, K. Tan, K. Roodenko, S. Tekarli, J. Halbert, I.W.H. Oswald, R.K. Arvapally, C. Yang, T. Thonhauser, M.A. Omary, Y.J. Chabal, J. Am. Chem. Soc. 135 (2013) 12615–12626. [637] S.I. Noro, S. Kitagawa, M. Kondo, K. Seki, Angew. Chem. Int. Ed. 39 (2000) 2081–2084. [638] Y. Huang, S.H. Ke, Adv. Mater. Res. 716 (2013) 244–247. [639] Z. Hulvey, E.H.L. Falcao, J. Eckert, A.K. Cheetham, J. Mater. Chem. 19 (2009) 4307–4309. [640] K.A. Forrest, T. Pham, P.A. Georgiev, F. Pinzan, C.R. Cioce, T. Unruh, J. Eckert, B. Space, Langmuir 31 (2015) 7328–7336. [641] C.G. Piscopo, F. Trapani, A. Polyzoidis, M. Schwarzer, A. Pace, S. Loebbecke, New J. Chem. 40 (2016) 8220–8224. [642] S.K. Elsaidi, M.H. Mohamed, C.M. Simon, E. Braun, T. Pham, K.A. Forrest, W. Xu, D. Banerjee, B. Space, M.J. Zaworotko, P.K. Thallapally, Chem. Sci. 8 (2017) 2373–2380. [643] W. Fan, X. Liu, X. Wang, Y. Li, C. Xing, Y. Wang, W. Guo, L. Zhang, D. Sun, Inorganic Chemistry, Frontiers (2018). [644] J. Xie, F. Sun, C. Wang, Q. Pan, Materials 9 (2016) 327. [645] C. Yang, U. Kaipa, Q.Z. Mather, X. Wang, V. Nesterov, A.F. Venero, M.A. Omary, J. Am. Chem. Soc. 133 (2011) 18094–18097. [646] X. Cui, K. Chen, H. Xing, Q. Yang, R. Krishna, Z. Bao, H. Wu, W. Zhou, X. Dong, Y. Han, B. Li, Q. Ren, M.J. Zaworotko, B. Chen, Science 353 (2016) 141–144. [647] A. Cadiau, K. Adil, P.M. Bhatt, Y. Belmabkhout, M. Eddaoudi, Science 353 (2016) 137–140. [648] T.-H. Chen, I. Popov, W. Kaveevivitchai, Y.-C. Chuang, Y.-S. Chen, A.J. Jacobson, O.Š. Miljanic´, Angew. Chem. Int. Ed. 54 (2015) 13902–13906. [649] S. Biswas, T. Ahnfeldt, N. Stock, Inorg. Chem. 50 (2011) 9518–9526. [650] T. Devic, F. Salles, S. Bourrelly, B. Moulin, G. Maurin, P. Horcajada, C. Serre, A. Vimont, J.-C. Lavalley, H. Leclerc, G. Clet, M. Daturi, P.L. Llewellyn, Y. Filinchuk, G. Férey, J. Mater. Chem. 22 (2012) 10266–10273. [651] Y. Wang, C. Tan, Z. Sun, Z. Xue, Q. Zhu, C. Shen, Y. Wen, S. Hu, Y. Wang, T. Sheng, X. Wu, Chem. Eur. J. 20 (2014) 1341–1348. [652] Y. Li, D. Ma, C. Chen, M. Chen, Z. Li, Y. Wu, S. Zhu, G. Peng, J. Solid State Chem. (2018). [653] S. Begum, Z. Wang, A. Donnadio, F. Costantino, M. Casciola, R. Valiullin, C. Chmelik, M. Bertmer, J. Kärger, J. Haase, H. Krautscheid, Chem. Eur. J. 20 (2014) 8862–8866. [654] D. Umeyama, S. Horike, M. Inukai, T. Itakura, S. Kitagawa, J. Am. Chem. Soc. 134 (2012) 12780–12785. [655] S.-S. Bao, K. Otsubo, J.M. Taylor, Z. Jiang, L.-M. Zheng, H. Kitagawa, J. Am. Chem. Soc. 136 (2014) 9292–9295. [656] A.U. Ortiz, A. Boutin, K.J. Gagnon, A. Clearfield, F.-X. Coudert, J. Am. Chem. Soc. 136 (2014) 11540–11545. [657] P. Ramaswamy, N.E. Wong, B.S. Gelfand, G.K.H. Shimizu, J. Am. Chem. Soc. 137 (2015) 7640–7643. [658] R. Fu, S. Hu, X. Wu, J. Mater. Chem. A 5 (2017) 1952–1956. [659] X. Zhao, J.G. Bell, S.-F. Tang, L. Li, K.M. Thomas, J. Mater. Chem. A 4 (2016) 1353–1365. [660] F. Zhai, Q. Zheng, Z. Chen, Y. Ling, X. Liu, L. Weng, Y. Zhou, CrystEngComm 15 (2013) 2040–2043.
[661] P.L. Llewellyn, M. Garcia-Rates, L. Gaberová, S.R. Miller, T. Devic, J.-C. Lavalley, S. Bourrelly, E. Bloch, Y. Filinchuk, P.A. Wright, C. Serre, A. Vimont, G. Maurin, J. Phys. Chem. C 119 (2015) 4208–4216. [662] P.V. Dau, S.M. Cohen, CrystEngComm 15 (2013) 9304–9307. [663] M. Zhang, Q. Wang, Z. Lu, H. Liu, W. Liu, J. Bai, CrystEngComm 16 (2014) 6287–6290. [664] E. Dooris, C. McAnally, E. Cussen, A. Kennedy, A. Fletcher, Crystals 6 (2016) 14. [665] M. Servati-Gargari, G. Mahmoudi, S.R. Batten, V. Stilinovic´, D. Butler, L. Beauvais, W.S. Kassel, W.G. Dougherty, D. VanDerveer, Cryst. Growth Des. 15 (2015) 1336–1343. [666] K. Roztocki, I. Senkovska, S. Kaskel, D. Matoga, Eur. J. Inorg. Chem. 2016 (2016) 4450–4456. [667] R. Custelcean, Chem. Commun. (2008) 295–307. [668] Z. Zhang, P.R. Schreiner, Chem. Soc. Rev. 38 (2009) 1187–1198. [669] R.S. Forgan, R.J. Marshall, M. Struckmann, A.B. Bleine, D.-L. Long, M.C. Bernini, D. Fairen-Jimenez, CrystEngComm 17 (2015) 299–306. [670] Q.-Y. Li, Y. Quan, W. Wei, J. Li, H. Lu, R. Ni, X.-J. Wang, Polyhedron 99 (2015) 1–6. [671] R.J. Marshall, J. McGuire, C. Wilson, R.S. Forgan, Supramol. Chem. 30 (2018) 732–741. [672] S. Glomb, G. Makhloufi, I. Gruber, C. Janiak, Inorg. Chim. Acta 475 (2018) 35– 46. [673] L. Rajput, D. Kim, M.S. Lah, CrystEngComm 15 (2013) 259–264. [674] J. Wu, P. Wang, C. Huang, L. Fu, C. Song, H. Hou, J. Chang, Inorg. Chem. Commun. 15 (2012) 301–304. [675] X. Lin, J. Jia, X. Zhao, K.M. Thomas, A.J. Blake, G.S. Walker, N.R. Champness, P. Hubberstey, M. Schröder, Angew. Chem. Int. Ed. 45 (2006) 7358–7364. [676] J.P. Malerich, K. Hagihara, V.H. Rawal, J. Am. Chem. Soc. 130 (2008) 14416– 14417. [677] D. Quiñonero, A. Frontera, G.A. Suñer, J. Morey, A. Costa, P. Ballester, P.M. Deyà, Chem. Phys. Lett. 326 (2000) 247–254. [678] J. Alemán, A. Parra, H. Jiang, K.A. Jørgensen, Chem. Eur. J. 17 (2011) 6890– 6899. [679] V. Amendola, G. Bergamaschi, M. Boiocchi, L. Fabbrizzi, M. Milani, Chem. Eur. J. 16 (2010) 4368–4380. [680] M. Al Kobaisi, S.V. Bhosale, K. Latham, A.M. Raynor, S.V. Bhosale, Chem. Rev. 116 (2016) 11685–11796. [681] S.V. Bhosale, C.H. Jani, S.J. Langford, Chem. Soc. Rev. 37 (2008) 331–342. [682] C. Dietl, H. Hintz, B. Rühle, J. Schmedt auf der Günne, H. Langhals, S. Wuttke, Chem. Eur. J. 21 (2015) 10714–10720. [683] R. Haldar, A. Mazel, R. Joseph, M. Adams, I.A. Howard, B.S. Richards, M. Tsotsalas, E. Redel, S. Diring, F. Odobel, C. Wöll, Chem. Eur. J. 23 (2017) 14316–14322. [684] C.C. Corrêa, F.M. Scaldini, F.C. Machado, C.B. Pinheiro, J. Struct. Chem. 57 (2016) 1235–1242. [685] Y. Bai, Y. Dou, L.-H. Xie, W. Rutledge, J.-R. Li, H.-C. Zhou, Chem. Soc. Rev. 45 (2016) 2327–2367. [686] S.S. Kaye, A. Dailly, O.M. Yaghi, J.R. Long, J. Am. Chem. Soc. 129 (2007) 14176– 14177. [687] J.A. Greathouse, M.D. Allendorf, J. Am. Chem. Soc. 128 (2006) 10678–10679. [688] J.J. Low, A.I. Benin, P. Jakubczak, J.F. Abrahamian, S.A. Faheem, R.R. Willis, J. Am. Chem. Soc. 131 (2009) 15834–15842. [689] K. Schröck, F. Schröder, M. Heyden, R.A. Fischer, M. Havenith, Phys. Chem. Chem. Phys. 10 (2008) 4732–4739. [690] K.P. Kepp, Inorg. Chem. 55 (2016) 9461–9470. [691] H. Reinsch, M. Krüger, J. Wack, J. Senker, F. Salles, G. Maurin, N.J.M. Stock, M. Materials, Micropor. Mesopor. Mat. 157 (2012) 50–55. [692] T. Wu, L. Shen, M. Luebbers, C. Hu, Q. Chen, Z. Ni, R.I. Masel, Chem. Commun. 46 (2010) 6120–6122. [693] C. Pettinari, A. Ta˘ba˘caru, S. Galli, Coord. Chem. Rev. 307 (2016) 1–31. [694] G.K.H. Shimizu, R. Vaidhyanathan, J.M. Taylor, Chem. Soc. Rev. 38 (2009) 1430–1449. [695] J.A. Greathouse, M.D. Allendorf, J. Am. Chem. Soc. 128 (2006) 10678–10679. [696] R. Custelcean, M.G. Gorbunova, J. Am. Chem. Soc. 127 (2005) 16362–16363. [697] P. Deria, W. Bury, I. Hod, C.-W. Kung, O. Karagiaridi, J.T. Hupp, O.K. Farha, Inorg. Chem. 54 (2015) 2185–2192. [698] R.M. Abdelhameed, L.D. Carlos, A.M.S. Silva, J. Rocha, New J. Chem. 39 (2015) 4249–4258. [699] Y. Liu, X. Xi, C. Ye, T. Gong, Z. Yang, Y. Cui, Angew. Chem. Int. Ed. 53 (2014) 13821–13825. [700] K. Jayaramulu, V.M. Suresh, T.K. Maji, Dalton Trans. 44 (2015) 83–86. [701] Y.-H. Shih, S.-H. Lo, N.-S. Yang, B. Singco, Y.-J. Cheng, C.-Y. Wu, I.-H. Chang, H.Y. Huang, C.-H. Lin, ChemPlusChem 77 (2012) 982–986. [702] K.D. Vogiatzis, A. Mavrandonakis, W. Klopper, G.E.J.C. Froudakis, ChemPhysChem 10 (2009) 374–383. [703] H. Fei, S.M.J.C.C. Cohen, 50 (2014) 4810-4812. [704] M.I. Gonzalez, E.D. Bloch, J.A. Mason, S.J. Teat, J.R.J.I.C. Long, 54 (2015) 2995– 3005. [705] N.J.T. Saracoglu, Tetrahedron 20 (2007) 4199–4236. [706] W.J.C.C.R. Kaim, Coord. Chem. Rev. 230 (2002) 127–139. [707] H.T. Chifotides, I.D. Giles, K.R. Dunbar, J. Am. Chem. Soc. 135 (2013) 3039– 3055. [708] S. Samanta, S. Das, P. Biswas, J. Org. Chem. 78 (2013) 11184–11193. [709] S. Samanta, P.J.R.A. Biswas, RSC Adv. 5 (2015) 84328–84333. [710] S. Samanta, S. Ray, A.B. Ghosh, P.J.R.A. Biswas, RSC Adv. 6 (2016) 39356– 39363.
S. Ali Akbar Razavi, A. Morsali / Coordination Chemistry Reviews 399 (2019) 213023 [711] G. Clavier, P.J.C.R. Audebert, Chem. Soc. Rev. 110 (2010) 3299–3314. [712] B.L. Schottel, H.T. Chifotides, K.R. Dunbar, Chem. Soc. Rev. 37 (2008) 68–83. [713] P. Nugent, Y. Belmabkhout, S.D. Burd, A.J. Cairns, R. Luebke, K. Forrest, T. Pham, S. Ma, B. Space, L.J.N. Wojtas, Nature 495 (2013) 80. [714] T. Panda, P. Pachfule, R. Banerjee, Chem. Commun. 47 (2011) 7674–7676. [715] B. Chen, S. Ma, F. Zapata, F.R. Fronczek, E.B. Lobkovsky, H.-C. Zhou, Inorg. Chem. 46 (2007) 1233–1236. [716] R. Wang, Q. Meng, L. Zhang, H. Wang, F. Dai, W. Guo, L. Zhao, D. Sun, Chem. Commun. 50 (2014) 4911–4914. [717] S. Huh, S. Jung, Y. Kim, S.-J. Kim, S.J.D.T. Park, Dalton Trans. 39 (2010) 1261– 1265. [718] P.-F. Shi, H.-C. Hu, Z.-Y. Zhang, G. Xiong, B. Zhao, Chem. Commun. 51 (2015) 3985–3988. [719] E.D. Bloch, D. Britt, C. Lee, C.J. Doonan, F.J. Uribe-Romo, H. Furukawa, J.R. Long, O.M. Yaghi, J. Am. Chem. Soc. 132 (2010) 14382–14384. [720] M.I. Gonzalez, E.D. Bloch, J.A. Mason, S.J. Teat, J.R.J.I.C. Long, Inorg. Chem. 54 (2015) 2995–3005. [721] R. Plessius, R. Kromhout, A.L.D. Ramos, M. Ferbinteanu, M.C. MittelmeijerHazeleger, R. Krishna, G. Rothenberg, S. Tanase, Chem. Eur. J. 20 (2014) 7922– 7925. [722] J. Seo, N. Jin, H.J.I.C. Chun, Inorg. Chem. 49 (2010) 10833–10839. [723] F. Zhang, X. Zou, P. Yan, J. Sun, G. Hou, G.J.C.G. Li, DesignCryst. Growth Des. 15 (2015) 1249–1258. [724] A.G. Wong-Foy, O. Lebel, A.J. Matzger, J. Am. Chem. Soc. 129 (2007) 15740– 15741. [725] P. Yang, M.-S. Wang, J.-J. Shen, M.-X. Li, Z.-X. Wang, M. Shao, X. He, Dalton Trans. 43 (2014) 1460–1470. [726] Y. Liu, J.R. Li, W.M. Verdegaal, T.F. Liu, H.C. Zhou, Chem. Eur. J. 19 (2013) 5637–5643. [727] P. Wang, K. Chen, Q. Liu, H.-W. Wang, M. Azam, S.I. Al-Resayes, Y. Lu, W.-Y. Sun, Dalton Trans. 46 (2017) 11425–11430. [728] Q. Zhai, Q. Lin, T. Wu, S.-T. Zheng, X. Bu, P. Feng, Dalton Trans. 41 (2012) 2866–2868. [729] J.-P. Zhang, Y.-B. Zhang, J.-B. Lin, X.-M.J.C.R. Chen, Chem. Rev. 112 (2011) 1001–1033. [730] S. Trofimenko, Prog. Inorg. Chem. (1986) 115–210. [731] P. Li, J. Chen, J. Zhang, X.J.S. Wang, P. Reviews, Sep. Purif. Rev. 44 (2015) 19– 27. [732] G. Aromí, L.A. Barrios, O. Roubeau, P.J.C.C.R. Gamez, Coord. Chem. Rev. 255 (2011) 485–546. [733] P. Lu, Y. Wu, H. Kang, H. Wei, H. Liu, M. Fang, J. Mater. Chem. A 2 (2014) 16250–16267. [734] J.H. Wang, M. Li, D. Li, Chem. Eur. J. 20 (2014) 12004–12008. [735] V. Colombo, S. Galli, H.J. Choi, G.D. Han, A. Maspero, G. Palmisano, N. Masciocchi, J.R.J.C.S. Long, Chem. Sci. 2 (2011) 1311–1319. [736] H.J. Choi, M. Dinca˘, A. Dailly, J.R.J.E. Long, E. Science, Energy & Environmental Science 3 (2010) 117–123. [737] Z.R. Herm, B.M. Wiers, J.A. Mason, J.M. van Baten, M.R. Hudson, P. Zajdel, C.M. Brown, N. Masciocchi, R. Krishna, J.R.J.S. Long, Science 340 (2013) 960–964. [738] N.M. Padial, E. Quartapelle Procopio, C. Montoro, E. López, J.E. Oltra, V. Colombo, A. Maspero, N. Masciocchi, S. Galli, I. Senkovska, S. Kaskel, E. Barea, J.A.R. Navarro, Angew. Chem. 125 (2013) 8448–8452. [739] J. Hitzbleck, A.Y. O’Brien, C.M. Forsyth, G.B. Deacon, K. Ruhlandt-Senge, Chem. Eur. J. 10 (2004) 3315–3323. [740] H. Yang, X. Zhang, G. Zhang, H. Fei, Chem. Commun. 54 (2018) 4469–4472. [741] J.-P. Zhang, S. Kitagawa, J. Am. Chem. Soc. 130 (2008) 907–917. [742] H.J. Choi, M. Dinca, J.R. Long, J. Am. Chem. Soc. 130 (2008) 7848–7850. [743] V. Colombo, S. Galli, H.J. Choi, G.D. Han, A. Maspero, G. Palmisano, N. Masciocchi, J.R.J.C.S. Long, Chemical Science, Transactions 2 (2011) 1311– 1319. [744] E. Quartapelle Procopio, T. Fukushima, E. Barea, J.A. Navarro, S. Horike, S. Kitagawa, Chem. Eur. J. 18 (2012) 13117–13125.
57
[745] H.R. Fu, J. Zhang, Chem. Eur. J. 21 (2015) 5700–5703. [746] H.-R. Fu, J.J.C.G. Zhang, Design Cryst. Growth Des. 15 (2015) 1210–1213. [747] A.D. Burrows, S. Chan, W.J. Gee, M.F. Mahon, C. Richardson, V.M. Sebestyen, D. Turski, M.R.J.C. Warren, CrystEngComm 19 (2017) 5549–5557. [748] L. Hou, Y.-Y. Lin, X.-M. Chen, Inorg. Chem. 47 (2008) 1346–1351. [749] C. Montoro, F. Linares, E. Quartapelle Procopio, I. Senkovska, S. Kaskel, S. Galli, N. Masciocchi, E. Barea, J.A.R. Navarro, J. Am. Chem. Soc. 133 (2011) 11888– 11891. [750] C. Heering, I. Boldog, V. Vasylyeva, J. Sanchiz, C. Janiak, CrystEngComm 15 (2013) 9757–9768. [751] C.-T. He, J.-Y. Tian, S.-Y. Liu, G. Ouyang, J.-P. Zhang, X.-M. Chen, Chem. Sci. 4 (2013) 351–356. [752] C. Feng, H. Zhao, Z.-Q. Li, J. Solid State Chem. 258 (2018) 841–844. [753] S.-M. Zhang, Z. Chang, T.-L. Hu, X.-H.J.I.C. Bu, Inorg. Chem. 49 (2010) 11581– 11586. [754] R. Poloni, B. Smit, J.B. Neaton, J. Am. Chem. Soc. 134 (2012) 6714–6719. [755] M. Eddaoudi, D.F. Sava, J.F. Eubank, K. Adil, V. Guillerm, Chem. Soc. Rev. 44 (2015) 228–249. [756] J.-P. Zhang, Y.-B. Zhang, J.-B. Lin, X.-M. Chen, Chem. Rev. 112 (2012) 1001– 1033. [757] A. Burgun, C.J. Doonan, C.J. Sumby, Aust. J. Chem. 66 (2013) 409–418. [758] S. Abedi, A. Azhdari Tehrani, H. Ghasempour, A. Morsali, New J. Chem. 40 (2016) 6970–6976. [759] A. Azhdari Tehrani, H. Ghasempour, A. Morsali, G. Makhloufi, C. Janiak, Cryst. Growth Des. 15 (2015) 5543–5547. [760] D.J.M.C. Bléger, Phys. Macromol. Chem. Phys. 217 (2016) 189–198. [761] J. Choudhury, Tetrahedron Lett. 59 (2018) 487–495. [762] J. Yoon, S.K. Kim, N.J. Singh, K.S. Kim, Chem. Soc. Rev. 35 (2006) 355–360. [763] M.N. Hopkinson, C. Richter, M. Schedler, F. Glorius, Nature 510 (2014) 485. [764] H.-Y. Gong, B.M. Rambo, W. Cho, V.M. Lynch, M. Oh, J.L. Sessler, Chem. Commun. 47 (2011) 5973–5975. [765] N. Farahani, K. Zhu, C.A. O’Keefe, R.W. Schurko, S.J. Loeb, ChemPlusChem 81 (2016) 836–841. [766] J.M. Roberts, O.K. Farha, A.A. Sarjeant, J.T. Hupp, K.A. Scheidt, Cryst. Growth Des. 11 (2011) 4747–4750. [767] S. Sen, S. Neogi, K. Rissanen, P.K. Bharadwaj, Chem. Commun. 51 (2015) 3173–3176. [768] C.L. Bird, A.T. Kuhn, Chem. Soc. Rev. 10 (1981) 49–82. [769] J.-K. Sun, J. Zhang, Dalton Trans. 44 (2015) 19041–19055. [770] T. Maihom, S. Wannakao, B. Boekfa, J. Limtrakul, J. Phys. Chem. C 117 (2013) 17650–17658. [771] T. Yamada, H.J.S.C. Kitagawa, Supramol. Chem. 23 (2011) 315–318. [772] T. Yamada, H. Kitagawa, J. Am. Chem. Soc. 131 (2009) 6312–6313. [773] D. Rankine, A. Avellaneda, M.R. Hill, C.J. Doonan, C.J.J.C.C. Sumby, Chem. Commun. 48 (2012) 10328–10330. [774] P.V. Dau, K.K. Tanabe, S.M.J.C.C. Cohen, Chem. Commun. 48 (2012) 9370– 9372. [775] A.P. Côté, G.K.H. Shimizu, Coord. Chem. Rev. 245 (2003) 49–64. [776] H. Fei, S.M. Cohen, J. Am. Chem. Soc. 137 (2015) 2191–2194. [777] T.-H. Chen, I. Popov, O. Zenasni, O. Daugulis, O.Š.J.C.C. Miljanic´, Chem. Commun. 49 (2013) 6846–6848. [778] C. Serre, Angew. Chem. Int. Ed. 51 (2012) 6048–6050. [779] S. Mukherjee, A.M. Kansara, D. Saha, R. Gonnade, D. Mullangi, B. Manna, A.V. Desai, S.H. Thorat, P.S. Singh, A. Mukherjee, Chem. Eur. J. 22 (2016) 10937– 10943. [780] S.K. Elsaidi, M.H. Mohamed, C.M. Simon, E. Braun, T. Pham, K.A. Forrest, W. Xu, D. Banerjee, B. Space, M.J.J.C.S. Zaworotko, Chem. Sci. 8 (2017) 2373– 2380. [781] J. Goura, V.J.C.R. Chandrasekhar, Chem. Rev. 115 (2015) 6854–6965. [782] J.A. Sheikh, H.S. Jena, A. Clearfield, S. Konar, Acc. Chem. Res. 49 (2016) 1093– 1103. [783] K.J. Gagnon, H.P. Perry, A.J.C.R. Clearfield, Chem. Rev. 112 (2011) 1034–1054.