Iodine capture in porous organic polymers and metal–organic frameworks materials

Wei Xie a, Di Cui a, Shu-Ran Zhang a, Yan-Hong Xu *ab and Dong-Lin Jiang *c
aKey Laboratory of Preparation and Applications of Environmental Friendly Materials (Jilin Normal University), Ministry of Education, Changchun, 130103, China. E-mail: xuyh198@163.com
bSchool of Chemistry and Environmental Engineering, Changchun University of Science and Technology, Changchun 130022, China
cDepartment of Chemistry, Faculty of Science, National University of Singapore, 3 Science Drive 3, Singapore 117543, Singapore. E-mail: chmjd@nus.edu.sg

Received 26th December 2018 , Accepted 15th April 2019

First published on 15th April 2019


Abstract

The enrichment of radioactive iodine in the waste of nuclear industries threatens human health, and thus the efficient capture of iodine has attracted a great deal of attention in recent years. Porous organic polymers (POPs) and metal–organic frameworks (MOFs), new classes of porous materials, act as outstanding candidate adsorbent materials in this field due to their high surface areas, permanent tunable porosities, controllable structures, high thermal/chemical stabilities, versatility in molecular design and potential for post-synthetic modification. Herein, this review focuses on the research progress of these two types of porous materials for highly efficient iodine capture. We analyze and discuss some valid strategies for enhancing their iodine uptake, including increasing their surface area and pore volume, using organic building units with unique configurations and functions, introducing chemical functional groups to provide high-enthalpy binding sites, and further processing of POP and MOF materials. Indeed, there are many special structural and functional features found in porous POP and MOF materials, which make them unique and merit further exploration. Thus, we expect to see their usage grow as this field progresses.


image file: c8mh01656a-p1.tif

Wei Xie

Wei Xie is a Lecturer at the Key Laboratory of Preparation and Applications of Environmental Friendly Materials in Jilin Normal University. She received her Bachelor's degree in Chemical Engineering of Changchun University of Science and Technology in 2012. In the same year, she joined in Professor Zhongmin Su's group and studied in the Institute of Functional Material Chemistry of Northeast Normal University. She obtained her PhD degree in 2017. Her current scientific interests are focused on the design, synthesis and functions of metal organic frameworks (MOFs) and MOFs-based composite materials.

image file: c8mh01656a-p2.tif

Di Cui

Di Cui is a Master's student at the School of Jilin Normal University. She received her Bachelor's degree in Chemical Education in Jilin Normal University in 2016. She is now pursuing a Master's degree in Professor Yanhong Xu's group. Her current scientific interests are focused on the design, synthesis and functions of CMPs and organic porous materials.

image file: c8mh01656a-p3.tif

Shu-Ran Zhang

Shuran Zhang is a Teacher at the Key Laboratory of Preparation and Applications of Environmental Friendly Materials in Jilin Normal University. She received her Bachelor's degree in Chemical Education of Harbin Normal University in 2011. In the same year, she began her studies at the Institute of Functional Material Chemistry of Northeast Normal University and received her PhD degree under the supervision of Professor Zhongmin Su in 2016. Her current scientific interests are focused on the design, synthesis and functions of MOFs.

image file: c8mh01656a-p4.tif

Yan-Hong Xu

Yanhong Xu received her Bachelor's degree in Harbin Normal University in 2005 and Master's degree in Northeast Normal University in 2009. In 2012, she obtained her PhD degree from Donglin Jiang's group. She then was a postdoctoral researcher until 2014 at the Institute for Molecular Science for Advanced Studies (Japan) under the supervision of Professor Donglin Jiang. In 2014, She joined Jilin Normal University as a Professor and set up an independent laboratory. She was selected as a Changbai Mountain scholar. Her current scientific interests are focused on the design, synthesis and functions of porous organic frameworks materials.

image file: c8mh01656a-p5.tif

Dong-Lin Jiang

Donglin Jiang received his PhD degree from the University of Tokyo in 1998. In 2005, he took a concurrent associate professorship at the Graduate University for Advanced Studies, Japan, where he started his independent laboratory and research in covalent organic frameworks. In 2016, he was promoted to a full professor at Japan Advanced Institute of Science and Technology. From February 2018, he was appointed the strategic hiring professor at National University of Singapore. His career work has been recognized by awards including the 34th Chemical Society of Japan Award for Creative Works (2017), the Young Scientists' Prize, and so on.


1. Introduction

With the rapid increase in energy demand, nuclear power, as a green and reliable energy source, has long drawn increasing attention around the world.1,2 However, radioiodine in the waste of nuclear industries prevents its promotion and poses an exceptional challenge because it is a highly flowing gas, can easily dissolve in water, and has negative impacts on organisms.3,4 Among the radioisotopes of iodine, 129I has a long half-life (1.57 × 107 years)5 and 131I isotope is short-lived (half-life of 8.02 days),6 but they both directly affect human metabolic processes due to bioaccumulation through the food chain.7,8 As is known, the Chernobyl and Fukushima nuclear power disasters involved in radioisotopes of 131I and 129I, which caused thyroid injury and were carcinogenic, highlighting the need for materials and technology to capture radioactive iodine.9,10 Therefore, dealing with iodine promptly and effectively has become an urgent problem to be solved.

Accordingly, some adsorbents for iodine have been studied in the last few decades, including silica,11 chalcogenide aerogels,12 activated carbon, zeolites,13,14 and microporous polymers.15,16 Presently, the traditional industrial method of capturing iodine is to impregnate an organic amine (such as triethyl diamine) on a porous material (such as activated carbon).17 The resulting adsorbent has a high affinity for radioactive iodine and purification effect, and the process to capture iodine is relatively simple; however, it has the following shortcomings: (i) sublimation of organic amine occurs easily, leading to a significant reduction in the iodine absorption efficiency of impregnated carbon materials; (ii) iodine adsorption is an exothermic process, and organic amines reduce the ignition point of impregnated carbon, exacerbating the potential safety hazard; and (iii) impregnated carbon is prone to aging and weathering, has poor regeneration capacity, is difficult to re-use, and easily causes the captured iodine to desorb and cause secondary pollution. Therefore, the exploration and development of adsorbent materials with iodine binding ability, high iodine adsorption, and excellent thermal stability have important academic value and application prospects.

Recently, solid porous materials as iodine sorbents have proven to be superior due to their easy operation, low energy consumption, fast speed and recycling. Porous organic polymers (POPs)18–20 and metal–organic frameworks (MOFs)21–23 have been studied as excellent candidates for iodine capture. POPs are constructed from various organic building blocks by strong covalent bonds. In contrast, MOFs are comprised of inorganic building units (metal ions or clusters) and organic linkers connected by coordination bonds of moderate strength. POPs and MOFs are new classes within the versatile family of porous materials, which can be synthesized from a wide range of organic monomers using many different chemistries and can incorporate a variety of chemical functionalities. POP and MOF materials possess versatility in molecular design, high surface areas, tunable pore sizes and volumes, adjustable skeletons, good physicochemical robustness, potential for post-synthetic modification (PSM), etc. Therefore, POPs and MOFs provide outstanding platforms for iodine uptake due to their advantages.

This review mainly focuses on the recent advances in the implementation of iodine capture on two types of solid adsorbents, -POP and MOF materials. Although there have been several reviews about the adsorption of iodine on porous solid sorbents,24–27 POPs and MOFs were simply summarized as two types of adsorbents. However, numerous POPs and MOFs with high iodine uptake have been reported. Based on this, in this tutorial review, we will summarize in detail the current progress of porous POP and MOF materials for iodine adsorption, discuss the effective strategies for increasing their iodine capture capacity, and analyze issues and challenges of porous materials for their application in iodine capture. To some extent, we hope our this work will help researchers design and develop porous adsorbent materials.

2. Iodine capture in porous organic polymers (POPs) materials

As a class of functional porous materials, from the view of structure, POPs can be divided into five categories, including conjugated microporous polymers (CMPs),28–30 polymers of intrinsic microporous (PIMs),31–33 hyper-crosslinked polymers (HCPs),34–37 covalent organic frameworks (COFs),38–42 and porous aromatic frameworks (PAFs).43,44 The unique nature of POPs gives them potential applications in the fields of guest adsorption and separation, heterogeneous catalysis, sensing, photo-electricity, energy storage, etc.45–48 More recently, plenty of research efforts have been devoted to designing POPs for ultrahigh iodine capture. In contrast to the traditional inorganic porous materials, besides large specific surface areas and high porosities, POP materials possess the following unique advantages: (i) wide diversity of POP networks, where the structures and properties of the organic building blocks used to construct POPs are diverse, and their synthetic conditions and other options can be regulated, which easily facilitate the realization of pre- or post-functionalization of POPs; (ii) through the formation of strong covalent bonds, POPs have good thermal/chemical stability; and (iii) POPs are constructed using light elements (such as C, H, N, O, and B), and thus their skeletons densities are very low. To the best of our knowledge, to date, HCPs and PIMs have not been reported for iodine capture. In this section, we discuss in detail the performances of different types of POPs (CMPs, COFs, and PAFs) for iodine adsorption. The relevant structural characteristics, pore properties, and iodine adsorption capacities of selected POPs are summarized and tabulated in Table 1.
Table 1 Iodine adsorption capacities in selected POP and MOF materials
Materials BET surface area (m2 g−1) Pore size (Å) Pore volume (cm3 g−1) Iodine uptake (wt%) Ref.
Note: test conditions:a Iodine vapor adsorption, 75–80 °C, ambient pressure.b Iodine vapor adsorption, 358.15 K, 1.0 bar.c Iodine vapor adsorption, 60 °C, ambient pressure.d Iodine vapor adsorption, 343.3 K, ambient pressure.e Iodine vapor adsorption, room temperature, ambient pressure.f Iodine solution of cyclohexane adsorption, room temperature, ambient pressure.
TPB-DMTP 1927 33 1.28 620a 98
TPT-BD-COF 543a 99
TTA-TTB 1733 22 1.01 500a 98
TTPPA 512.39 19.1 0.2997 490a 80
SIOC-COF-7 618 11.8 0.41 481a 96
CalPOF-1 303 477a 70
COF-DL229 1762 14 0.64 470a 97
ETTA-TPA 1822 14, 27 0.95 470a 98
TPT-DHBD25-COF 465a 99
TTPB 222 19.4 0.127 443a 15
NDB-H 116.93 74.6 0.13 443a 79
TPT-DHBD50-COF 430a 99
NDB-S 56.45 75.5 0.11 425a 79
TPT-DHBD75COF 412a 99
CalPOF-2 154 406a 70
TPT-DHBD-COF 388a 99
POP-2 41 20, 400 382a 73
COP01 380c 90
TFBCz-PDA 1441 15 0.74 370a 98
POP-1 12 20, 400 357a 73
CalPOF-3 91 353a 70
ADB-HS 148.20 62.8 0.14 345a 79
SCMP-II 119.76 20 345a 86
ADB-S 41.53 62.9 0.07 342a 79
HCMP-3 82 <10 0.08 336b 72
TTDAB 1.643 17.6–27.0 0.004 313a 80
CalP4_Li 445 0.588 312a 91
Tm-MTDAB 2.778 27.9 0.007 304a 80
HCMP-1 430 <10 0.22 291b 72
AzoPPN 400 5.8–12.7 0.68 290a 69
HCOF-1 290a 100
BDP-CPP-1 635 49 0.78 283a 89
HCMP-2 153 <10 0.06 281b 72
Zr6O4(OH)4(peb)6 2650 14.2 1.16 279e 159
COP02 277c 90
PAF-24 276a 104
PAF-23 271a 104
TTA-TFB 1163 16 0.55 270a 98
NAPOP-4 626 12.7 0.15, 1.17 265a 76
PAF-25 260a 104
COP2++ 258c 90
CalP3_Li 308 0.558 248a 91
NAPOP-3 702 <4.3, 5.8 0.18, 1.01 241a 76
NAPOP-2 458 <4.2, 5.9 0.10, 0.78 239a 76
Azo-Trip 510.4 6.5 0.47 238a 71
BDP-CPP-2 235 30 0.18 223a 89
SCMP-2 855 1.50 222a 16
HCMP-4 <10 222b 72
NRPP-2 1028 7.0 0.81 222a 77
CalP4 759 1.08 220a 91
COP1++ 212c 90
COP2•+ 211c 90
FCMP-600@2 780 17.4 0.665 209a 88
CMPN-3 1368 ∼20 2.36 208d 92
NAPOP-1 657 <3.9, 7.2 0.25, 1.49 206a 76
NiP-CMP 2630 7.0–12.5 2.288 202a 67
HKUST-1 1850 5.0–13.5 0.74 175a 118
[(ZnI2)3(tpt)2] 173e 205
MFM-300(Sc) 1250 8.1 0.50 154a 164
HMTI-1 18 × 21 150f 153
MFM-300(Fe) 1192 7.8 0.46 129a 164
ZIF-8 1630 3.4–11.6 0.66 125a 115
UiO-66-PYDC 1030 0.43 125f 170
MFM-300(In) 1050 7.6 0.41 116a 164
Zr6O4(OH)4(sdc)6 2900 11.9 1.33 107e 159
Zn3(DL-lac)2(pybz)2 762.5 10.5 0.40 101f 123
HMTI-2 4 × 18 100f 154


2.1 CMPs for iodine capture

Since Copper's group first reported CMPs in 2007,49 a variety of CMPs with different functionalities have been published. CMPs are generally formed by the interconnection of multiple carbon–carbon bonds or aromatic benzene rings and often have large conjugated structures. Due to the rigidity of their conjugated units at the molecular level and the expansion of their conjugated structures, the CMP network skeletons can effectively support porous channels and form inherent pores.50–53 CMP materials are commonly prepared via a series of polycondensation reactions, including the Suzuki cross-coupling reaction,54,55 Yamamoto cross-coupling reaction,56,57 Sonogashira–Hagihara cross-coupling reaction,58,59 oxidative coupling reaction,60–62 Schiff-base reaction,63–65 Friedel–Crafts reaction,50 and phenazine ring fusion reaction.66 The inherent porosity, extended π-conjugated skeletons, higher surface areas and larger pore volumes make CMPs a great platform as iodine absorbents. Herein, we analyzed the structures of CMPs and illustrate their effect on the capacities of iodine adsorption of the diverse CMPs.

The building blocks and linking units directly affect the structures, compositions, and porosities of CMPs. From the perspective of the core of the skeleton structure, porphyrin-based CMPs facilitate strong interactions between the host adsorbent and guest. For example, Liu and co-workers first reported a metalloporphyrin-based NiP-CMP for the capture of iodine,67 which was constructed by utilizing nickel(II) 5,10,15,20-tetrakis(4-bromophenyl) porphyrin as a building block via Yamamoto cross-coupling polymerization (Fig. 1a). NiP-CMP possessed a high BET (Brunauer Emmett Teller) surface area of 2630 m2 g−1 and large pore volume of 2.288 cm3 g−1. It exhibited a wide pore size distribution, with pore widths centered at around 0.70, 0.85, 0.95 and 1.25 nm. NiP-CMP could capture 202 wt% iodine vapor, and also capable of trapping iodine in solution. In addition, the iodine loaded NiP-CMP could be recovered by immersing the material in ethanol to remove the adsorbed iodine. Subsequently, another Por-Py-CMP was prepared, in which the porphyrin acted the as core and pyrene units as linkers (Fig. 1b).68 Por-Py-CMP had a BET surface area of 1014 m2 g−1 and pore volume of 0.81 cm3 g−1, and showed capacity for iodine capture (130 wt%). Thus, the results indicated that CMPs with porphyrin units can effectively absorb iodine molecules in vapor and solution conditions.


image file: c8mh01656a-f1.tif
Fig. 1 Synthesis of (a) NiP-CMP and (b) Por-Py-CMP.

Inspired by the structural design of strategic molecule manipulation, Han and co-workers designed and prepared a porous azo-bridged CMP containing a large number of porphyrin groups and phthalocyanine units (AzoPPN) (Fig. 2a).69 The unique π-conjugated systems of porphyrin and phthalocyanine allowed the enrichment of π-electrons clouds, which make them suitable hosts for the adsorption of iodine guests. AzoPPN had a BET specific surface area of 400 m2 g−1 with a wide pore-size distribution of 0.58–1.27 nm and pore volume of 0.68 cm3 g−1. Compared with NiP-CMP (SBET = 2630 m2 g−1 and pore volume = 2.288 cm3 g−1), AzoPPN had a much lower surface area and pore volume; however, it exhibited a prominent adsorption capacity of 290 wt% towards iodine vapor. The high I2 capture of AzoPPN may be attributed to its azo-linkage and π-electron conjugated aromatic framework with plenty of porphyrin and phthalocyanine units, which could afford high affinity towards iodine. Additionally, X-ray photoelectron spectroscopy measurements indicated the process of adsorbing iodine by AzoPPN is a hybrid of physisorption and chemisorption processes, that is, it captured both iodine and triiodide ions. The above results suggest that incorporating macrocyclic molecules in CMPs including porphyrin and phthalocyanine provides a new approach for designing effective absorbents for volatile iodine.


image file: c8mh01656a-f2.tif
Fig. 2 Synthesis of (a) AzoPPN, (b) CalPOFs and (c) Azo-Trip.

Considering the performance of AzoPPN, azo-linked CMPs may show potential applications for the capture of iodine. For example, by changing the building units and the reaction conditions, azo-bridged calix[4]resorcinarene and triptycene-based CMPs, CalPOFs (Fig. 2b)70 and Azo-Trip71 (Fig. 2c) have been reported. The porous properties of CalPOFs could be adjusted by varying the alkyl chain lengths. With an increase in the alkyl chain length from methyl and ethyl to propyl, the BET surface area varied from 303 and 154 to 91 m2 g−1, respectively. Among these CalPOFs polymers, CalPOFs with a methyl chain exhibited the highest iodine vapor uptake of up to 477 wt%. In addition, CalPOFs with ethyl and propyl moieties captured iodine with a capacity of 406 wt% and 353 wt%. On the other hand, the Azo-Trip polymer network showed a BET surface area, pore size, and pore volume of 510.4 m2 g−1, 0.65–5.0 nm, and 0.47 cm3 g−1, respectively. Also, Azo-Trip displayed an efficient guest uptake of 238 wt% iodine vapor. The high capacity of azo-bridged CMPs for iodine may be due to their numerous effective sorption sites including azo (–N[double bond, length as m-dash]N–) groups, macrocyclic π-rich cavities and high affinity of I2 to the host network, which favour the enhancement of their interactions with iodine and possible exchange of electrons to a stable state.

Apart from azo-linkers, nitrogen-containing electron-donating –NH– functional groups have been explored for the assembly of CMPs and iodine uptake. Liao et al. designed a series of hexaphenylbenzene-based CMPs, that is, HCMP-1, 2, 3, and 4 (Fig. 3a), which were synthesized via Buchwald–Hartwig cross-coupling of a hexakis(4-bromophenyl) benzene core and aryl diamine linkers.72 Although these three-dimensional (3D) imino-linked HCMPs possessed low BET surface areas (SHCMP-1 = 430 m2 g−1, SHCMP-2 = 153 m2 g−1, SHCMP-3 = 82 m2 g−1, and SHCMP-4 = na) and pore volumes (VHCMP-1 = 0.22 cm3 g−1, VHCMP-2 = 0.06 cm3 g−1, VHCMP-3 = 0.08 cm3 g−1, and VHCMP-4 = na), they exhibited high iodine affinity with uptake capacities of up to 291, 281, 336, 222 wt%, respectively. Interestingly, the iodine uptake capacities of these HCMPs increased with the increase in number of –NH– groups, which may be attributed to the stronger charge transfer interactions between the –NH– units and the oxidizing iodine. In addition, the two other –NH-containing CMPs, POP-1 and POP-2 reported via the same Buchwald–Hartwig cross-coupling reaction captured iodine vapor of 357 and 382 wt% (Fig. 3b), respectively.73


image file: c8mh01656a-f3.tif
Fig. 3 Synthesis of –NH-containing HCMP-1,2,3,4, POP-1 and POP-2.

Thomas and Faul74 successfully constructed a series of polyaniline-like porous materials, NCMP1, NCMP2 and NCMP3 (Fig. 4), via the facile FeCl3-catalyzed oxidation polymerization of multi-connected aniline precursors. The NCMPs with high N contents (7.39–11.84 wt%) showed BET surface areas and iodine vapor uptake of 58 m2 g−1 and 215 wt% for NCMP1, 280 m2 g−1 and 186 wt% for NCMP2, 485 m2 g−1 and 161 wt% for NCMP3, respectively. Compared to NCMP2 and NCMP3, NCMP1 with the highest nitrogen content had the highest iodine capture capacity among them. Additionally, our group constructed a series of –NH-containing pore-tunable porous materials (NT-POPs),75 and then NT-POPs were further pyrolyzed under nitrogen conditions at 800 °C to form porous materials (NT-POP@800). NT-POP@800 possessed moderate BET surface areas in the range of 475–736 m2 g−1 and displayed relatively rapid guest uptake of 56–192 wt% iodine vapor.


image file: c8mh01656a-f4.tif
Fig. 4 Synthesis of NCMP1–3.

Zhang and co-workers76 reported four aminal-linked CMPs (NAPOP-1, 2, 3 and 4), which were synthesized through the Schiff base condensation of 1,4-bis(4,6-diamino-s-triazin-2-yl) benzene with four types of aldehydes substituted with different N-heterocyclic groups (Fig. 5a). The BET surface areas and pore volumes of NAPOP-1, 2, 3 and 4 were 657 m2 g−1 and 1.74 cm3 g−1, 458 m2 g−1 and 0.88 cm3 g−1, 702 m2 g−1 and 1.19 cm3 g−1, 626 m2 g−1 and 1.32 cm3 g−1, respectively. Also, their pore size distributions ranged from 0.39 to 20 nm and their capture iodine uptake was about 206, 239, 241, and 265 wt%, respectively. The NAPOP polymers possessed moderate BET surface areas, but they showed a higher iodine capture capacity than that of CMPs with larger surface areas and pore volumes, such as NiP-CMP (SBET = 2630 and 202 wt%)67 and Por-Py-CMP (SBET = 1014 and 130 wt%).68 Based on the same core building block (1,4-bis-(2,4-diamino-1,3,5-triazine)-benzene), the other two polymers NRPP-1 and NRPP-277 were produced. The BET surface areas and total pore volumes were 1579 m2 g−1 and 0.91 cm3 g−1 for NRPP-1 and 1028 m2 g−1 and 0.81 cm3 g−1 for NRPP-2, respectively. NRPP-1 and NRPP-2 had an adsorption capacity for iodine of 192 wt% and 222 wt%, respectively.


image file: c8mh01656a-f5.tif
Fig. 5 Synthesis of (a) NAPOP-1, 2, 3, and 4, and (b) pha-HCOP-1.

Furthermore, pha-HcOP-178 (Fig. 5b) was synthesized via a mixed-linker approach, which made it possible to accommodate the –CO–NH– functionality to enhance the binding affinity between the adsorbents and iodine molecules. Its BET surface area, total pore volume and pore size were 217 m2 g−1, 1.048 cm3 g−1 and 2.45 nm, respectively. It could capture iodine of 131 wt% at 353 K and ambient pressure. Compared with Por-Py-CMP (SBET = 1014 m2 g−1 and Vtotal = 0.81 cm3 g−1),68 pha-HcOP-1 possessed a lower BET surface area and similar pore volume, but it showed an equal adsorption capacity. This result further demonstrates that the specific surface area is not the only factor determining the iodine absorption capacity. Additionally, four other amorphous POPs (NDB-H, NDB-S, ADB-HS and ADB-S) based on Schiff-base chemistry were obtained under different solvothermal conditions.79 It is noteworthy that excellent efficiency for removing iodine vapor was observed for NDB-S (∼425 wt%), NDB-H (∼443 wt%), ADB-HS (∼345 wt%) and ADB-S (∼342 wt%).

Geng and co-workers also designed a series of nitrogen-rich CMPs (TTPPA, TTPB, TTDAB, and Tm-MTDAB) (Fig. 6),15,80 which were synthesized via Friedel–Crafts polymerization reaction of 2,4,6-trichloro-1,3,5-triazine (TCT) and derivatives of phenylamine (N,N,N′,N′-tetraphenylbenzidine (TPB), N,N,N′,N′-tetraphenyl-1,4-phenylenediamine (TPPA), 1,3,5-tris(diphenylamino)benzene (TDAB) and 1,3,5-tris[(3-methylphenyl)-phenylamino]benzene (MTDAB)), respectively, catalyzed by methanesulfonic acid. This reaction has the advantages of low cost, low volatility, no metal catalyst residue and strong positioning ability. The four CMPs were hard bulk solids with good chemical and thermal stability. The BET surface areas of TTPB, TTPPA, TTDAB, and Tm-MTDAB were 222, 512, 1.64 and 2.78 m2 g−1, respectively. The total pore volumes of TTPB and TTPPA were 0.127 and 0.2997 cm3 g−1, respectively. Also, TTDAB, and Tm-MTDAB were almost non-porous. Interestingly, the resulting polymers TTPB, TTPPA, TTDAB and Tm-MTDAB possessed a nitrogen-rich, twisted structure, microporosity and good stability, and displayed very high affinity for guest uptake of 443, 490, 313 and 304 wt% iodine vapor. This is because the triphenylamine derivatives are nonplanar molecules, and the electrons easily transfer to form free radical cations. Besides, the introduction of triazine rings increases the polarity of the porous network, promotes the electron transfer process, and forms a complex with the formation of polyiodide anions by electrostatic attraction, which can significantly improve the adsorption performances for iodine. Similarly, Chen and co-workers developed three hierarchical CMPs (NOP-53, NOP-54 and NOP-55)81 based on a tetraphenyladamantane building block, and their pore features were controlled by adjusting the lengths and rigidities of the linkers. At 348 K and ambient pressure, NOP-53, NOP-54 and NOP-55 could capture iodine of 177, 202, and 139 wt%.


image file: c8mh01656a-f6.tif
Fig. 6 Synthesis of TTPB, TTPPA, TTDAB, and Tm-MTDAB.

Over all, N-containing CMPs82–85 have shown outstanding performances in iodine capture. Their effective uptake of I2 can be ascribed to the combination of high nitrogen content, π-electron conjugated structure, and porosity. The electron-rich N atom is preferred for the sorption of I2 due to the strong interaction between the electron-deficient I2 and N-enriched host absorbents. The π-electron conjugated structure allows for the enrichment of π-electrons clouds, which make them suitable hosts for the adsorption of iodine guests. Also, the existence of a larger number of pores is beneficial for I2 sorption. These studies may provide a useful guidance to manipulate CMPs by varying the constitution of their core building blocks.

Besides the abovementioned functional nitrogen heteroatoms, the introduction of the electron-enriched sulfur and fluorine is also an effective strategy for improving the interaction between host adsorbents and guest adsorbates. For example, four thiophene-based CMPs (SCMP-I, SCMP-II, SCMP-1, and SCMP-2) (Fig. 7)16,86 were prepared via the polymerization of 3,3′,5,5′-tetrabromo-2,2′-bithiophene, 2,3,5-tribromothiophene, and ethynylbenzene monomers through the palladium-catalyzed Sonogashira–Hagihara cross-coupling reaction. SEM (scanning electron microscopy) images clearly showed that the four samples had different morphologies. SCMP-I was composed of agglomerated spheres with different sizes. SCMP-II showed honeycomb-liked porous spheres with penetrated pores on its surface, and SCMP-1 and SCMP-2 were comprised of agglomerated microgel particles. The BET surface areas were 2.72 m2 g−1 for SCMP-I, 119.76 m2 g−1 for SCMP-II, 413 m2 g−1 for SCMP-1, and 855 m2 g−1 for SCMP-2. More interestingly, SCMP-II, which had the lowest surface area among the obtained materials, showed the highest adsorption capacity with an uptake of up to 345 wt%. SCMP-1, SCMP-2, and SCMP-I exhibited iodine uptake of 188, 222, and 102 wt%, respectively. This high amount of iodine adsorption capacity of SCMP-II was attributed to its unique macroscopic honeycomb-liked porous morphology and π–π conjugated structure.


image file: c8mh01656a-f7.tif
Fig. 7 Synthesis of SCMP-I, SCMP-II, SCMP-1, SCMP-2.

Inspired by the excellent properties of CMPs containing thiophene and nitrogen units, Geng and Niu reported another CMP (TTPT)87 (Fig. 8a) with thiophene derivatives and nitrogen groups, which was fabricated via the AlCl3-catalyzed Friedel–Crafts polymerization reaction of 2,4,6-trichloro-1,3,5-triazine and 2,3,4,5-tetraphenylthiophene in dichloromethane. TTPT had a moderate surface area of 315 m2 g−1 and total pore volume of 0.232 cm3 g−1. Its dominant pore width was 1.428 nm. The iodine loading of TTPT was 177 wt% at 350 K and ambient pressure. This value is larger than that of the theoretical maximum adsorption capacity (114 wt%) of TTPT itself, indicating the surface adsorption behavior of the polymer.


image file: c8mh01656a-f8.tif
Fig. 8 Synthesis of TTPT and FCMP-600@1–4.

Besides, our group also reported four fluorine-enriched CMPs (FCMP-600@1–4, Fig. 8b).88 Under the same conditions, FCMP-600@1, 2, 3 and 4 exhibited maximum values for iodine uptake ranging from 90 to 141 wt%. Meanwhile, Song and co-workers designed and synthesized two novel BODIPY-based CMPs (BDP-CPP-1 and BDP-CPP-2)89 (Fig. 9) via Sonogashira cross-coupling of 1,3,5-triethynylbenzene and dibromo-substituted BODIPY derivatives. The BET surface areas and the total pore volumes were 635 m2 g−1 and 0.78 cm3 g−1 for BDP-CPP-1 and 235 m2 g−1 and 0.18 cm3 g−1 for BDP-CPP-2, respectively. SEM images revealed that BDP-CPP-1 was composed of spherical particles, while BDP-CPP-2 showed a block morphology. The significant porous porosity and morphology differences between BDP-CPP-1 and BDP-CPP-2 may be closely related to the ethyl substituents at the 2,6-positions of the BODIPY moieties. BDP-CPP-1 and BDP-CPP-2 could adsorb about 283 and 223 wt% iodine. It is clear that BDP-CPP-1 and BDP-CPP-2 polymers have strong affinity with iodine molecules due to the coexistence of a triple bond, aromatic ring and π-conjugated BODIPY units, which contain heteroatoms (N, B, and F), resulting in a high adsorption capacity for iodine.


image file: c8mh01656a-f9.tif
Fig. 9 Synthesis of BODIPY-based CMPs.

The above results suggest that heteroatoms (such as nitrogen, thiophene, and fluorine) have a positive influence on the iodine capture of CMP materials. The electron-rich CMPs have the potential to increase the capture capacity for iodine molecules because the lone pair electrons of heteroatoms can enhance the interaction between the adsorbents and adsorbates. Despite their low surface areas and pore volumes, the heteroatoms-rich CMPs are highly desirable for the capture of iodine; however, the examples of high uptake of CMPs are quite limited to date.

On the other hand, the integration of free-ions into the CMP skeleton is another strategy to improve its iodine storage capacity. Two non-porous viologen-based cationic CMPs, COP1++ and COP2++, were developed, in which the viologen units crosslink hexatopic cyclotriphosphazene core moieties (Fig. 10).90 COP1++ and COP2++ were obtained by using Menshutkin and Zincke reactions. Then, the radical cationic polymers, COP1˙+ and COP2˙+, and the neutral polymers, COP01 and COP02, were generated via treatment with sodium dithionite (Na2S2O4) or excess cobaltocene under a nitrogen atmosphere. N2 gas adsorption studies exhibited that the two di-cationic polymers, the two radical cationic polymers, and the two neutral polymers were non-porous. At 333 K and ambient pressure, COP1++, COP1˙+, COP01, COP2++, COP2˙+, and COP02 could adsorb about 212, 195, 380, 258, 2.1, and 2.77 wt% iodine, respectively. Interestingly, the sorption rates of iodine vapor for all six non-porous polymers (in the range of minutes) were faster than that of most porous and non-porous materials (in the range of hours) reported to date. In addition, density functional theory (DFT) calculations suggested that polymer reduction to its diradical form leads to the reduction of I2 to I, which eventually is tightly bound to the polymer through electrostatic interactions. Raman spectroscopy and X-ray photoelectron spectroscopy also indicated that iodine exists mainly as polyiodides I3 and I5.


image file: c8mh01656a-f10.tif
Fig. 10 Synthesis of COP1++ and COP2++. Reproduced from ref. 90 with permission from the Royal Society of Chemistry.

Subsequently, the same research group utilized another strategy to explore the iodine capture capacity of CMPs. They synthesized a series of hyper-cross-linked π-bond-rich CMPs (CalP2, CalP3, and CalP4) based on a polycalix[4]arene building block (Fig. 11).91 The CalP2, CalP3, and CalP4 polymers were further lithiated to obtain the corresponding lithiated polymers CalP2-Li, CalP3-Li, and CalP4-Li, respectively. These polymers exhibited excellent thermal stability over 500 °C. The CalPn and CalPn-Li (n = 2–4) polymers displayed type II isothermal curves with H4 type hysteresis loops. The special BET surface areas and total pore volumes were 596 m2 g−1 and 0.73 cm3 g−1 for CalP2, 630 m2 g−1 and 0.639 cm3 g−1 for CalP3, 759 m2 g−1 and 1.08 cm3 g−1 for CalP4, 274 m2 g−1 and 0.31 cm3 g−1 for CalP2-Li, 308 m2 g−1 and 0.558 cm3 g−1 for CalP3-Li, 445 m2 g−1 and 0.588 cm3 g−1 for CalP4-Li, respectively. Compared to the pre-lithiated polymers, these lithium ions tended to agglomerate, which caused their overall surface areas and pore volumes to decrease. However, their pore-size distributions remained unchanged with the average pore diameter in the range of 6.2–7 nm. All the CMPs samples were capable of adsorbing iodine from the solution and vapor phase. At 348 K and ambient pressure, the capture iodine uptake of CalP2, CalP3, CalP4, CalP2-Li, CalP3-Li, and CalP4-Li was 88, 196, 220, 108, 248, and 312 wt%, respectively. Among all the polymers tested, CalP4-Li showed the fastest I2 uptake, which required only 30 min for complete adsorption. This result implies that the lithiated polymers were more efficient absorbers.


image file: c8mh01656a-f11.tif
Fig. 11 Synthesis of CalPn and CalPn-Li (n = 2–4).

Additionally, many studies have shown that the functions of porous polymers are not only affected by their skeleton structures, and the different morphologies of polymers also have an obvious impact on their properties. After constant efforts, to date, the functions of polymers have been well adjusted by controlling their structure. More recently, the morphology of porous polymers was finely tuned by adjusting the reaction conditions, such as solvent polarity, types, monomers ratios, and reaction temperature. Some porous polymers with different morphologies have been reported in succession, for example tubular, spherical, and membranous morphology. These polymers exhibit different functional properties, even though they have the same chemical compositions and similar structure skeletons.

Li and Deng92 synthesized three nanotube-like CMPs (CMPN-1, CMPN-2, and CMPN-3) (Fig. 12) via Sonogashira–Hagihara cross-coupling polymerization using different molecular structures. The pore volumes and the BET surface areas were 0.14 cm3 g−1 and 230 for CMPN-1, 0.39 cm3 g−1 and 339 m2 g−1 for CMPN-2, and 2.36 cm3 g−1 and 1368 m2 g−1 for CMPN-3, respectively. SEM and TEM images clearly displayed that these samples possessed a nanotube-like morphology. The tips of CMPN-1 were open, which exhibited a smooth and neat tubular surface with ca. 150 nm diameter and ca. 20 μm length. CMPN-2 and CMPN-3 displayed a hollow and hierarchical porous surface morphology with honeycomb-like pores in their tube walls. CMPN-1, CMPN-2 and CMPN-3 could capture gaseous iodine of 95, 110 and 208 wt% at 343 K and ambient pressure, which are proportional to their BET surface area. In addition, the π–π conjugated structure also increased the affinity of CMPN-3 for iodine molecules. Due to its special tubular morphology, high mechanical strength, not easily deformed by external forces, combined with better film forming, CMPN-3 is more suitable for industrial applications.


image file: c8mh01656a-f12.tif
Fig. 12 Structure of nanotube-like CMPs (CMPN-1, CMPN-2, CMPN-3, and NCMP-5,6,7,8).

Subsequently, they prepared four NH2-functional porous materials (NCMP-5, NCMP-6, NCMP-7, and NCMP-8)93 (Fig. 12) via the Sonogashira–Hagihara cross-coupling of 2-amino-3,5-dibromopyridine and 1,3,5-triethynylbenzene with different solvent systems. SEM images showed that NCMP-5 was composed of amorphous aggregated nanoparticles. NCMP-6 and NCMP-8 displayed a similar tubular-shape morphology, while NCMP-7 exhibited smooth small spheres. The iodine vapor adsorption capacity followed the order NCMP-6 > NCMP-7 > NCMP-5 > NCMP-8, in which NCMP-6 had the highest iodine adsorption amount of 297 wt% at 350 K and 1.0 bar. The results indicated that all the porous materials were greatly influenced by the component of mixture solvent used in the reaction process, and the introduction of nitrogen atoms enhanced the interaction between guest molecules and CMPs adsorbents, but different reaction conditions could regulate the pore properties and morphology of these porous materials.

The foregoing discussion clearly indicates that, as a new class of porous materials, the structure skeleton, surface area, pore volume, pore size, and morphology of CMPs can be finely controlled through the selection of different building blocks (such as different geometry, size, linkage and component) and different polar solvents (such as single component, binary mixture component, ternary mixture component). Although CMPs are amorphous in nature, their structure and functionalization can be precisely tuned using the abovementioned methods. Owing to their unique π-conjugation structure and permanent porosity, CMPs can serve as outstanding absorbents for iodine in vapor and solution. In addition, iodine-loaded CMPs have been widely studied for their luminescence, magnetism, electrical conductivity, etc. For example, a multifunctional carbazolic porous framework (Cz-TPM) with a tetrahedral core synthesized via FeCl3 oxidative coupling polymerization displayed tandem visual detection of iodide and mercury.94 An iodine-functionalized tetrathiafulvalene (TTF)-containing CMP displayed a strong electron-donor network. Due to the excessive tendency of TTF and its derivatives to form iodide/polyiodide salts, a charged analogue (TTFCMPC+-In) was subsequently formed by the addition of iodine to the porous networks, which is further step towards realizing organic (opto)electronic devices.95

2.2 COFs for iodine capture

Compared to amorphous CMPs, the preparation of crystalline covalent organic framework (COF) materials is more challenging. COFs are an emerging class of porous ordered materials, which are composed of multidentate organic building blocks and linked by strong covalent bonds. Since the first COF was synthesized by Yaghi and co-workers in 2005,38 it has attracted great interest. The unique combination of crystallinity and organic functionality make COFs highly efficient for guest selective storage and separation, sensing, transportation of guest molecules, etc. Especially recently, COFs have been proven to be exceptional iodine capture materials.

Zhao's group reported a hetero-pore COF (SIOC-COF-7)96 (Fig. 13), which consisted two different types of micropores with unprecedented shapes. SIOC-COF-7 had a BET surface area of 618 m2 g−1 and total pore volume of 0.41 cm3 g−1. In addition, SIOC-COF-7 exhibited two main pore size distributions at 0.5 and 1.18 nm, which were almost the same as the theoretical values of 0.5 and 1.19 nm estimated by PM3 calculations, respectively. It existed as hollow microspheres and exhibited a high iodine uptake (up to 481 wt%) by encapsulating iodine in the inner cavities and porous shells of its microspheres. This result indicated that the abundant aromatic rings, high nitrogen content and well-ordered network of heteroporous COFs should be favorable for iodine enrichment, and the inner cavities of the hollow microspheres and the highly ordered channels in the shells also make significant contributions to the extraordinarily high iodine capture capacity.


image file: c8mh01656a-f13.tif
Fig. 13 Synthetic route for SIOC-COF-7.

Jiang's group also reported a stable 3D COF (COF-DL229)97 (Fig. 14), which adopted an extended diamond topology, consisted of eight-fold interwoven skeletons and possessed ordered 1D nanochannels with π-conjugated pore walls. The 3D COF-DL229 was synthesized via the condensation reaction of 1,3,5,7-tetrakis(4-aminophenyl)-adamantane and 1,4-phthalaldehyde under solvothermal conditions. The COF-DL229 was thermally stable up to 500 °C under a nitrogen atmosphere. SEM images showed a uniform morphology of aggregates of rod-shaped crystallites with a minimum dimension of approximately 500 nm. COF-DL229 possessed a BET surface area, total pore volume, and pore size of 1762 m2 g−1, 0.64 cm3 g−1 and 1.4 nm, respectively, and it could adsorb 470 wt% iodine at 348 K.


image file: c8mh01656a-f14.tif
Fig. 14 (a) Synthesis of COF-DL229. (b) Adamantane-knotted cage in the diamond net. (c) 8-fold interpenetration of the diamond net with π-conjugated pore walls. Reprinted with permission from ref. 97 Copyright 2018 Wiley-VCH.

Subsequently, they synthesized five stable 2D COFs (TPB-DMTP, TTA-TTB, TTA-TFB, TFBCz-PDA, and ETTA-TPA)98 (Fig. 15) containing 1D open channels with different shapes, and found their pore volume determined their uptake capacity of iodine. TPB-DMTP and TTA-TTB with 1D hexagonal channels and higher pore volumes (VTPB-DMTP = 1.28 cm3 g−1 and VTTA-TTB = 1.01 cm3 g−1) showed higher iodine capture capacities of up to 620 and 500 wt%, respectively. TTA-TFB also consists of hexagonal channels; however, it showed the lowest iodine uptake capacity of 270 wt% due to its lowest pore volume (VTTA-TFB = 0.55 cm3 g−1). The channel shapes and pore volumes of TFBCz-PDA and ETTA-TPA were tetragonal and 0.74 cm3 g−1 and trigonal and 0.95 cm3 g−1, and they could uptake iodine of 370 and 470 wt%, respectively. In the five 2D COFs, all the 1D channels were fully accessible to iodine, generalizing a new paradigm that the pore volume determines the uptake capacity.


image file: c8mh01656a-f15.tif
Fig. 15 2D hexagonal COFs, 2D tetragonal COFs and 2D trigonal COFs. Reprinted with permission from ref. 98 Copyright 2018 Wiley-VCH.

The ordered structure and high crystallinity of COF materials will greatly increase the properties of materials in iodine adsorption. Therefore, effectively controlling the crystallinity of COFs and regulating their related application performance are practical research subjects, especially for COFs based on flexible building blocks, which usually have large lattice sizes and various monomer sources. Based on the above considerations, Ma and Xia et al. designed and prepared a series of 2D COFs from flexible building blocks with different contents of intralayer hydrogen bonds (Fig. 16).99 The effect of H-bonding on their crystallinity and adsorption properties showed that the partial structure of the COFs was “locked” by the H-bonding interaction, which consequently improved their degree of microscopic order and crystallinity. The obtained COFs exhibited excellent and reversible adsorption properties for volatile iodine with uptake of up to 543 wt%.


image file: c8mh01656a-f16.tif
Fig. 16 Synthesis of 2D COFs. Reprinted with permission from ref. 99 Copyright 2018 American Chemical Society.

It's well-known that the capture and separation of iodine from water at room temperature is very important in practical applications, but remains challenging. Ke and co-workers developed a microporous hydrogen-bonded COF (HCOF-1)100 through photo-irradiated single-crystal-to-single-crystal (SCSC) transformation from its molecular precursor 1crystal (Fig. 17). Firstly, 1crystal was soaked in neat ethanedithiol in the dark overnight, and photo-irradiated under a mercury lamp or ambient light, and then unreacted ethanedithiol was removed, and it was finally activated using supercritical CO2. In this reaction system, ethanedithiol served as cross-linkers to form C–S bonds between two alkynyl groups. The obtained polymer HCOF-1 could adsorb I2 rapidly in an aqueous environment with a high uptake capacity of 290 wt%, which may be attributed to the hydrophobic nature of its pores. Thus, HCOF-1 was prone to forming N–H⋯I hydrogen bonding and N⋯I and S⋯I halogen bonding interactions with I2 molecules. Notably, the loaded iodine molecules could be released from HCOF-1 at different rates. The release speed was very fast in dimethyl sulfoxide, within 30 min; however, it was slower in MeOH because the solvent molecules can compete with the hydrogen-bonding interactions between iodine and HCOF-1. In addition, Thomas and co-workers developed Fe3O4-based COF composites via a facile method (Fig. 18, Fe3O4/COF).101 Fe3O4/COF can be used as a magnetically recoverable adsorbent, which could capture iodine with an uptake of 797 mg g−1 in iodine aqueous solution (0.2 g L−1, 5 mL) with the liquid phase becoming colorless upon sonication for 15 s, indicating the efficient adsorption of iodine. To further illustrate the significance and advantage and the current limitations of COFs for iodine capture, the performances of 187 experimentally reported COFs for gaseous I2 were systematically evaluated by theoretical calculation under real industrial conditions.102 Among them, 3D-Py-COF was identified with the highest I2 uptake of 1670 wt%, outperforming all the adsorbents reported to date.


image file: c8mh01656a-f17.tif
Fig. 17 Synthesis of hydrogen-bonded COF (HCOF-1). Reprinted with permission from ref. 100 Copyright 2017 American Chemical Society.

image file: c8mh01656a-f18.tif
Fig. 18 Synthesis of Fe3O4/COF composites. Reprinted with permission from ref. 101 Copyright 2017 American Chemical Society.

Overall, COFs exhibit excellent performances in capturing iodine. COF materials have many advantages, such as regular pore structure and uniform size, easy to adjust pore size in the range of microporous and mesoporous, and easy to control internal pore environment. Moreover, COFs have high thermal/chemical stability, even in strong acidic or alkaline conditions. COFs are comprised of pure organic building units linked by strong covalent bonds. Thus, their skeleton is chemically robust against iodine to avoid structural collapse. Some researchers suggest that the porous skeletons of COFs are soft by triggering structural fitting to iodine, while keeping the covalent connectivity and degree of interpenetration, which is comparable with that of inorganic porous materials. From the perspective of ordered structure, the ordered arrangement of the skeleton of COFs enables full accessibility of their porous space to iodine, which leads ultrahigh iodine adsorption capacity. The ordered structure of COFs plays a key role in iodine adsorption. Thus, studies on the iodine adsorption of COF materials demonstrate they are new benchmark materials for coping with fission vapor waste issues.

2.3 PAFs for iodine capture

The creation of other members of POPs has broadened the range of porous materials to a large extent. Inspired by the successful construction of artificial porous materials, a number of attempts have been made in exploring new POPs. From the viewpoint of structural characters and performance, porous aromatic framework (PAF) networks generally feature diamond-like structures, ultrahigh surface areas and high physicochemical stability, which are also an important class of adsorbents for iodine capture.

Qiu and co-workers have reported two PAFs with high iodine capture capacities, PAF-1 and JUC-Z2 (Fig. 19).103 PAF-1 showed a diamond topology and robust all-carbon scaffold with a pore size in the range of 0.5–1.0 nm and high physicochemical stability (up to 420 °C) with an ultrahigh surface area (SBET = 5640 m2 g−1). JUC-Z2 with a hcb topology possessed a well-defined uniform micropore distribution (∼1.2 nm), large surface area (SBET = 2081 m2 g−1) and high physicochemical stability (>440 °C). The amount of iodine adsorbed by PAF-1 and JUC-Z2 was 186 and 144 wt%, respectively.


image file: c8mh01656a-f19.tif
Fig. 19 Iodine vapor adsorption of PAF-1 and JUC-Z2. Reproduced from ref. 103 with permission from the Royal Society of Chemistry.

Subsequently, Zhu and co-workers104 synthesized a series of charged PAFs based on a tetrahedral building unit, lithium tetrakis(4-iodophenyl)borate, and different alkyne monomers as linkers via Sonogashira–Hagihara coupling reaction (PAF-23, PAF-24, and PAF-25, Fig. 20). PAF-23, PAF-24 and PAF-25 showed permanent porosities and excellent thermal/chemical stabilities. In addition, PAF-23, PAF-24 and PAF-25 exhibited high affinities and capacities for iodine molecules due to their effective sorption sites, namely ion bond, phenyl ring, and triple bond together. These materials showed strong polarization–polarization interaction with iodine and displayed very high affinity for iodine in vapor and solution with an adsorption capacity in the range of 260–276 wt%.


image file: c8mh01656a-f20.tif
Fig. 20 Synthesis of PAF-23, PAF-24, and PAF-25. Reprinted with permission from ref. 104 Copyright 2015 Wiley-VCH.

As stated above, POPs have been found to be a class of potential and effective iodine sorbents due to their high surface area, inherent porosity (pore size, pore volume, and pore shape), and light weight. The iodine adsorption capacity of POPs can be enhanced via several techniques such as selecting core building blocks with different functional groups (porphyrin, phthalocyanine, thiophene, –NH2, –NH–, OH, –CO–NH–, fluorine, etc.) for their synthesis, different preparation methods (Yamamoto reaction, Sonogashira coupling polymerization, Suzuki coupling polymerization, azo-linked reaction, Friedel–Crafts reaction, and Schiff-base reaction) and activation routes (sodium dithionite, cobaltocene, and nBu-Li), and changing the functional groups and electronegativity of the polymers. POPs have prominent iodine sorption capacity. However, in laboratory studies, the test conditions and corresponding performance evaluations do not always represent the expected real conditions during a serious nuclear accident (temperature > 80 °C and relative humidity > 40%). Therefore, the main factors such as temperature and relative humidity should be considered in data analysis, which will provide an important basis for the practical application of porous adsorbents in the iodine industry. Moreover, it can be noted that the capture of iodine by POPs sorbents has been depicted as physisorption and chemisorption mechanisms. Physisorption has been found to be reversible, in contrast, chemisorption is irreversible due to the strong chemical bond formation between the adsorbate and adsorbent. In addition, although COFs show excellent performances as iodine adsorbents, there are still several problematic issues, including the synthesis of COFs is cumbersome, monomers are not easily obtain, and COFs cannot be produced on an industrial scale. Thus, these factors will limit the practical application of POP iodine adsorbents.

3. Metal–organic frameworks (MOFs) for iodine capture

As is known, MOFs are effective porous adsorbents and are widely applied in the adsorption/separation of guest molecules such as gases,105 water,106 organic vapors,107 and dyes,108 and others.109 The use of porous MOFs for iodine sorption is motivated by the capturing and sequestering of the volatile radioactive iodine produced in nuclear energy enterprises. Porous MOFs are good candidates for iodine capture, and many excellent explorations of iodine adsorption have been carried out, in which iodine was introduced into MOFs by guest exchange in iodine solution or by diffusion in high concentration iodine vapor. Moreover, some researchers indicate most MOF materials show good recycling because the encapsulated iodine can be released, and the frameworks can be retained. In addition, iodine assembly in their restricted spaces has been widely studied for luminescence,110 magnetism,111 and electrical conductivity.112

Herein, we simply divide MOFs for iodine adsorption into two categories of non-iodine MOFs and iodine-templated MOF materials based on the presence or absence of iodine. In some systems, we found that iodine species are introduced as raw materials to obtain iodine-templated MOF materials. We will present the iodine adsorption performances of these two types of MOFs. The relevant structural characteristics and iodine adsorption capacities of the selected MOFs are summarized and tabulated in Table 1.

3.1 Iodine capture in non-iodine MOFs

In the field of capture iodine, a large number of MOFs have been developed and utilized, of which non-iodine MOFs account for the majority. A wide variety of metals and organic ligand units offer the possibility of constructing diverse and versatile MOFs materials. Their large surface areas, adjustable skeletal structures, facile post-processing and functionalization provide an excellent platform for adsorbing iodine. In addition, the morphology, size, and further processing of MOFs can also affect their uptake capacities of iodine.

As early as 2003, Robson and co-workers113 reported a chiral zinc d-saccharate network [Zn(C6H8O8)]˙ ≈ 2H2O containing two different characteristics of hydrophobic and hydrophilic parallel channels, which were isolated from each other by seemingly impenetrable walls. The color of the crystals became deep red after exposure to I2 vapor for a few days. Single crystal X-ray diffraction indicated that the hydrophilic channels were still assigned to the H2O molecules; however, disordered iodine centres were defined in the hydrophobic channels. The absorption amount was approximately 0.30 I2 molecules per zinc atom or 2.6 I2 molecules per unit cell. The results showed that the hydrophobic/hydrophilic characteristics of its pores and channels affected the adsorption of iodine. Similarly, Ghosh et al.112 synthesized a 3D biporous anionic MOF {[(Me2NH2)2]·[Cd3(5-tbip)4]·2DMF}n (5-tbipH2 = 5-tert-butylisophthalic acid), which possessed hydrophobic and hydrophilic channels due to the 5-tbip linker containing a hydrophobic t-butyl group and hydrophilic carboxylate groups. The activated sample exhibited reversible iodine uptake in vapor and solution, and 58 wt% iodine was included. Furthermore, the iodine-loaded MOF showed ∼76 times higher electrical conductivity compared with the guest-free MOF.

Inspired by the potential iodine adsorption of MOFs, some well-known classic MOFs have been extended as iodine sorbents. For example, zeolitic imidazolate framework ZIF-8,114 which was purposely selected due to its relatively large cavities (11.6 Å), large specific surface area, high chemical and thermal stability and suitable apertures (3.4 Å) close to the I2 kinetic diameter (∼3.35 Å). Nenoff et al.115 presented detailed structural evidence of I2 capture within ZIF-8 by employing a combination of experiments and molecular simulations. ZIF-8 had a high sorption capacity for I2, up to 125 wt%, where only 25 wt% I2 binds to the surface, while ∼100 wt% I2 is efficiently contained within the sodalite cages of ZIF-8. The I2 adsorption is mainly due to the favorable interactions with the imidazolate-based linker. Moreover, cage-trapped I2 is secured in the pores of ZIF-8 until the framework decomposes at ∼575 K. Thus, the ZIF-8 material can be used to sequestrate iodine. The cages of ZIF-8 strongly bind iodine within the framework, while I2 bound on the surface mirrors binding energies of traditional iodine sorbents. ZIF-8 binds iodine 4 times more strongly than activated carbon (Fig. 21a).116 In a broader sense, this thermodynamic analysis of iodine-ZIF-8 binding offers clues for designing effective functional MOFs for specific applications, which are centered on guest–host interactions.


image file: c8mh01656a-f21.tif
Fig. 21 (a) I2 capture within ZIF-8 and (b) competitive I2 sorption by Cu-BTC from Humid Gas Streams. Reproduced with permission from ref. 116 and 118 Copyright and 2013 American Chemical Society.

The other classic MOF Cu-BTC, also known as HKUST-1,117 is a hydrophilic open-pored MOF containing coordinatively unsaturated metal centers, which offer the potential for open metal sites to directly interact with the targeted capture molecule I2. Cu-BTC has been used as I2 “getters” for competitive sorption of molecular I2 from a mixed stream of iodine and water vapor (Fig. 21b).118 The combination of simulation with crystallography showed that HKUST-1 substantially adsorbs I2 in preference to water vapor because I2 forms an effective hydrophobic barrier to minimize water sorption. I2 molecules are adsorbed in two principal areas, first, in the smallest cage close to the copper paddle wheel, then second, it adsorbs within the main pore with close interactions with the benzene tricarboxylate organic linker. The maximum I2 uptake was ∼175 wt%. Furthermore, the wider 6.5 Å window openings, as well as interactions of I2 with the framework are sufficiently weak to allow the “on-demand” release of I2 upon moderate heat treatment. Also, Lobanov and co-workers119 reported the iodine adsorption by two microporous MOFs, SBMOF-1 (Ca(sdb), sdb = 4,4′-sulfonyldibenzoate) and SBMOF-2 (Ca(tcpb), tcpb = 1,2,4,5-tetrakis(4-carboxyphenyl)-benzene),120,121 which preferentially adsorbed I2 over water in a humid environment due to the specific phenyl ring-halogen interactions in their channels. The competitive I2 sorption from Humid Gas Streams is relevant to mixed gas streams in nuclear energy industrial processes and accident remediation. Thus, moisture-stable porous MOFs that can selectively adsorb I2 in the presence of water vapor are of great interest. Besides ZIF-8 and HKUST-1, MIL-101 has been used as a sorbent due to its large surface area and high stability, and it shows efficacy for the capture of radioiodine.122

In addition to the classic MOF materials, many outstanding MOFs with distinct structures have been designed and synthesized for iodine adsorption. In 2010, a highly stable microporous MOF, {[Zn3(DL-lac)2(pybz)2]·2.5DMF}n (DL-lactic acid (DL-lac), 4-(pyridin-4-yl)benzoic acid (pybz)), was constructed by rigid metal–organic pillars (Zn-lactate) connected through organic groups (pybz) forming double walls (Fig. 22).123 The desolvated crystals could uptake iodine (101 wt%) in a cyclohexane solution of I2, and the excellent and promising I2 affinity may be attributed to the structural character of the regular π-electron walls made of pybz. Its uptake capacity (101 wt%) was remarkable and clearly exceeded that of zeolite 13× (0.32–0.38 g g−1, with 10 Å pore)124 and activated carbon (∼0.84 g g−1). Conversely, the I2 could be released, which was governed by the host–guest interaction. Furthermore, donor–acceptor (I2-π-electron of the surface of the walls) interactions resulted in cooperative anisotropic electrical conductivity. Recently, a 4-fold interpenetrated double-walled MOF {[Ni4(44pba)8]·sol}n, (44pba = 4-(4-pyridyl)benzoate)125 was reported to absorb 110 wt% iodine. Its high uptake for I2 was attributed to the high porosity of its network, in which the aromatic π-electron-rich walls formed by 44pba ligands allowed the interaction of iodine molecules with the framework. Another porous MOF, [Zn2(tptc)(apy)2−x (H2O)x]·H2O (where x ≈ 1, apy = aminopyridine, H4tptc = terphenyl-3,3′′,5,5′′-tetracarboxylic acid), displayed efficient, reversible adsorption of I2 in vapor and solution.126 The I2 loading was 216 wt%. Its strong affinity for I2 was ascribed to its large porosity, conjugated π-electron aromatic system, halogen bonds, electron-donating aminos, and unsaturated Zn(II) sites. These results indicated that π-electron-rich MOFs facilitate the binding of their skeleton and guest iodine, and further support that the judiciously assembled MOFs with predesigned regular and suitable structures can find many applications in the encapsulation of iodine substances.


image file: c8mh01656a-f22.tif
Fig. 22 (a)–(c) Structure of {[Zn3(DL-lac)2(pybz)2]·2.5DMF}n. (d) Schematic of I2 molecules diffusing in the channels. Reproduced with permission from ref. 123 Copyright 2010 American Chemical Society.

In the field of iodine adsorption, an increasing number of MOFs have been studied, showing varying capacity, reversibility and stability. The thermally stable Hofmann-type clathrate NiII(pz)[NiII(CN)4] (pz = pyrazine)127 could efficiently and reversibly capture I2 both in solution and the gaseous phase of 83 wt% and 59 wt%, respectively. The seemingly nonporous highly permeable discrete metallocyclic complex ([Cu2(bitmb)2Cl4]·CH3OH·H2O, bitmb = 1,3-bis(imidazol-1-ylmethyl)-2,4,6-trimethylbenzene)128 could absorb iodine from iodine vapor, which is because the host molecules may cooperate with each other in a dynamic and concerted fashion to create windows of opportunity, thus allowing mobile guest iodine molecules to freely traverse the crystals. Moreover, it is worth noting that the locations of I2 in the pores of Cu-MOF ({[CuII(btz)]·0.5H2O}n, H2btz = 1,5-bis(5-tetrazolo)-3-oxapentane)129 and porphyrin-based Co-MOF, [Co(DpyDtolP)]6·12H2O (DpyDtolP, 5,15-di(4-pyridyl)-10,20-di(4-methylphenyl)porphyrin)130 were unambiguously determined by single crystal diffraction. More importantly, the successful location determination of the I2 molecules may provide new insight for the design and application of MOF materials. Besides, many other MOFs have been used for the capture of iodine molecules, including calixarene-based cages CIAC-103 and CIAC-104,131 MOFs, IFMC-10,132 IFMC-15,133 IFMC-69,134 members of the [M3(HCOO)6] family,111,135,136 JLU-Liu14,137 CityU-7,138 and others.110,139–152 They consist of interpenetrating, non-interpenetrating, cationic, anionic, chiral, achiral, and other networks.

The above-mentioned MOFs are all composed of common transitions metals (such as Zn, Cd, Cu, Co, and Ni). Besides, some MOFs constructed from the other relatively rare metals (for instance Pb, In, and Zr) have shown excellent performances in the field of iodine adsorption. Three 3D porous lead(II) coordination polymers, HMTI-1153 ([Pb(4-bpdh)(NO3)(H2O)]n, 4-bpdh = 2,5-bis(4-pyridyl)-3,4-diaza-2,4-hexadiene), HMTI-2154 ([Pb(4-bpdh)(NO3)2]n), and TMU-15155 ([Pb(4-bpdb)(μ-NO3)(μ-SCN)]n·1.5CH3OH, 4-bpdb = 1,4-bis(4-pyridyl)-2,3-diaza-1,3-butadiene) could reversibly absorb approximately 3I2 (150 wt%), 2I2 (100 wt%), and 3I2 for one lead(II) atom, respectively. Their exceptional affinity for I2 may be attributed to the structural character of their regular π-electron walls made of organic ligands. Moreover, HMTI-2 could complete a reversible crystal to crystal transformation to HMTI-1 by hydration and dehydration. The presence of H2O molecules in HMTI-1 plays an effective role in increasing its iodine adsorption and speeding up the adsorb rate, while the delivery of iodine from HMTI-2 was faster than HMTI-1 because the coordinated H2O can increase the affinity for I2.

Zr(IV) MOFs, as in the UiO-66 series,156 have been receiving widespread research interest due to their enhanced chemical and mechanical stabilities.157,158 Forgan et al.159 investigated the iodine storage capacities from iodine vapors of the stable Zr-stilbene MOFs [Zr6O4(OH)4(sdc)6]n, [Zr6O4(OH)4(edb)6]n,160 [Zr6O4(OH)4(bdb)6]n, and [Zr6O4(OH)4(peb)6]n (sdc = 4,4′-stilbenedicarboxylate, edb = 4,4′-ethynylenedibenzoate, bdb = 4,4′-(buta-1,3-diyne-1,4-diyl)-dibenzoate, and peb = 4,4′-[1,4-phenylenebis(ethyne-2,1-diyl)]-dibenzoate) (Fig. 23). Unsaturated alkene, alkyne and butadiyne units can serve as reactive sites for PSM by halogenation. Although both bromination and bromohydrination can occur in a facile manner, [Zr6O4(OH)4(sdc)6]n is unable to chemisorb I2 across its integral alkene units, and its maximum I2 physisorption storage capacity was 107 wt%. In contrast, [Zr6O4(OH)4(edb)6]n and [Zr6O4(OH)4(bdb)6]n could both chemi- and physisorb I2, demonstrating a maximum of 57 wt% and 88.5 wt% trapping of I2. The higher adsorption of [Zr6O4(OH)4(bdb)6]n is due to the presence of twice as many alkyne units. Among them, the interpenetrated [Zr6O4(OH)4(peb)6]n showed the highest storage value, with a maximum uptake of 279 wt%, which is comparable to the benchmark capacity of 276 wt% set by the charged PAF-24.104 The superior uptake of [Zr6O4(OH)4(peb)6]n is the result of the high tendency of I2 to physisorb within pores, possibly due to the high density of Zr6 clusters as a result of interpenetration. By comparison, the I2 chemisorption capacity of [Zr6O4(OH)4(peb)6]n was lower than that of [Zr6O4(OH)4(bdb)6]n, even though both materials contained a similar alkyne content. Chemisorption by [Zr6O4(OH)4(bdb)6]n may also lead to partial pore collapse, thus hindering later I2 uptake. Generally, the chemical stability of Zr(IV) MOFs, combined with their high porosity in concert with reactive chemisorptive sites on the linkers, generates potential candidates for I2 storage/capture applications.


image file: c8mh01656a-f23.tif
Fig. 23 (a) Scheme of the ligands used and (b) representation of the crystal structure of [Zr6O4(OH)4(sdc)6]n. (c) Summary of the physisorption and chemisorption I2 storage capacities of the four [Zr6O4(OH)4L6]n MOFs. Reproduced with permission from ref. 159 Copyright 2016 Wiley-VCH.

More recently, MFM-300(M) (M = Al, Sc, Fe, and In)161–164 with a high storage density have been developed for I2 capture. The maximum uptake of I2 by MFM-300(M) was 94, 154, 129, and 116 wt% for Al, Sc, Fe, and In, respectively. Among them, MFM-300(Sc) exhibited fully reversible I2 uptake of 154 wt%, and its structure remained completely unperturbed upon the inclusion/removal of I2. Both experimental and theoretical investigations rationalized that at a low loading, the adsorbed I2 molecules form specific interactions with the metal-bound hydroxyl groups within the pores, supplemented by intermolecular interactions between adsorbed I2 molecules. In contrast, at a high loading, the intermolecular guest–guest interaction becomes the dominant feature and leads to the unusual self-aggregation of adsorbed I2 molecules into ordered triple-helical chains within the confined channels, generating an exceptional density of 3.08 g cm−3 in MFM-300(Sc), 63% of that for solid I2 (4.93 g cm−3 at 298 K). The widespread implementation of sustainable nuclear energy requires the development of efficient iodine stores that have simultaneously high capacity, stability and more importantly, storage density (and hence minimized system volume).

The choice and design of organic ligand units are crucial for the synthesis of MOFs, which directly affect the structures of the obtained MOFs, and then influence their ability to adsorb iodine. Using organic linkers with functional groups to construct MOFs or functionalizing post-treatments of MOFs can both change the porous environments, and thus improve the interaction between the host and the guests. The use of amine-tagged ligands can enhance the adsorbent-adsorbate interactions through host–guest hydrogen-bonding to enhance iodine adsorption. Morsali et al.165 constructed two isoreticular two-fold interpenetrated MOFs, [Zn2(BDC)2(4-bpdh)]·3DMF (TMU-16) and [Zn2(NH2-BDC)2(4-bpdh)]·3DMF (TMU-16-NH2, NH2-BDC = amino-1,4-benzenedicarboxylate) (Fig. 24a). Both unmodified and amino functionalized MOFs could serve as hosts for encapsulating I2. Their adsorption amounts of I2 were similar (∼45 wt%), while, the amino functionalized TMU-16-NH2 could absorb iodine 1.4 times faster than the non-functionalized analogue. The differences in adsorption rate between the non- and amine-functionalized frameworks probably resulted from the different host–guest interactions. Thus, assembling MOFs with –NH2 groups tagged on the linkers as a hydrogen bond donor can find more applications for the encapsulation of iodine. Subsequently, using a similar strategy, two isoreticular 3D pillared-bilayer interpenetrated MOFs, {[Cd(bdc)(4-bpmh)]}n·2n(H2O) and {[Cd(2-NH2bdc)(4-bpmh)]}n·2n(H2O) (4-bpmh = N,N-bis-pyridin-4-yl-methylene-hydrazine),166 were developed (Fig. 24b), in which the iodine capture capacity of the amino-functionalized MOF was two times that of the non-functionalized MOF. A similar phenomenon was observed in the amino-functionalized Al-based MOF.167 These works demonstrate that amino-functionalized MOFs exhibit enhanced iodine adsorption, faster I2 adsorption rate and controlled delivery of I2 over the non-functionalized homologs, which indicate that the electro-donor –NH2 group is favorable for the adsorption and storage of iodine.


image file: c8mh01656a-f24.tif
Fig. 24 (a) and (b) Schematic views of the comparative synthesis of amino-functionalized MOF and non-functionalized MOF, respectively. Reproduced from ref. 165 with permission from the Royal Society of Chemistry, and ref. 166 with permission from 2015 Wiley-VCH.

The introduction of thiol (–SH) groups influence iodine capture. The thiol-modified MIL-53(Al)168 was capable of iodine adsorption from the vapor phase and from solution with an equilibrium uptake of ∼32.5 wt%, which is higher than that of other reported modified forms of MIL-53. The formation of covalent S–I bonds allows the sequestration of iodine by the porous solid. Analogously, ZrDMBD (Zr6O4(OH)4·(C8H6O4S2)6·(DMF)7·(H2O)18) appended with free-standing thiol (–SH) groups could uptake iodine from solution to form Zr6O4(OH)4·(C8H6O4S2)6·(DMF)3·(I)15·(H2O)25.169 ZrDMBD could react readily with I2 to form sulfenyl iodide (S–I) units (Fig. 25a), which anchored onto the rigid MOF skeleton and thus prevented them from approaching each another to undergo the dismutation reaction, exhibiting distinct stability even at elevated temperatures (e.g. 90 °C). Besides, Gu and co-workers170 developed UiO-66-PYDC constructed by utilizing the pyridine-containing linker PYDC (Fig. 25b). The abundant and inherent pyridine moieties in the developed adsorbent worked as active adsorption sites to capture I2. Due to the strong affinity of pyridine-containing ligands to I2 and high porosity, the adsorption capacities of UiO-66-PYDC for I2 could reach as high as 125 wt%. Additionally, UiO-66-PYDC exhibited excellent renewable adsorption properties, prefiguring their great promise as green adsorbents for I2 removal in nuclear waste management. These results further prove that the introduction of functional groups (–NH2, –SH, and pyridine) can enhance iodine adsorption.


image file: c8mh01656a-f25.tif
Fig. 25 (a) Schematic of I2 oxidation of the –SH groups in the ZrDMBD net (left) into the ZrDMBD-I2 net (right). The host net is simplified as a tetrahedral cage, with each Zr6O4(OH)4 cluster shown as a sphere. (b) Schematic illustration of the I2 removal mechanism with the developed adsorbent of UiO-66-PYDC using 2,5-pyridine-dicarboxylic acid (PYDC) as the inherent bridging struts and active adsorption sites. Reproduced from ref. 169 and 170 with permission from the Royal Society of Chemistry.

Besides, due to their adjustable structures, some MOFs can be easily subjected to post-synthetic modification, which plays an important role in adjusting the interaction between host and guests. Zeng's grpoup171 described an MOF containing uncoordinated hydroxyethyl groups, {Zn3[(L)22-OH)2]·6H2O}n (1) (H2L = 2-(1-hydroxyethyl)-1H-benzo[d]imidazole-5-carboxylic acid), which could undergo unusual tandem PSM reactions (Fig. 26). Heating crystals of 1 at 120 °C gave rise to dehydrated 1′, and then heating crystals of 1′ at 250 °C resulted in a new MOF 1′′ with vinyl groups in its channels, which was associated with the elimination of H2O from the hydroxyethyl groups to give vinyl groups. After PSMs, 1′ and 1′′ could absorb comparable iodine of 28 wt% and 26 wt%, respectively, from solution, while the iodine sorption rate and release rate of 1′ were quicker than that of 1′′.


image file: c8mh01656a-f26.tif
Fig. 26 Scheme of the transformations of 1: post-synthetic elimination of H2O from the hydroxyethyl groups in the channels to form vinyl groups, and photos showing the iodine sorption of 1′ from a saturated solution of iodine in cyclohexane. Reproduced with permission from ref. 171 Copyright 2013 Wiley-VCH.

There are also reports that the frameworks of MOF materials can react with guest iodine (e.g. redox reaction) to realize iodine capture. This type of MOF is usually constructed of metals with redox activity (such as Fe and Ni). The pillared bilayer open framework BOF-1 ([Ni2(C26H52N10)]3 [BTC]4·6C5H5N·36H2O) had unprecedented sponge-like dynamic behavior in response to guest molecules.172 Upon reaction with I2, two-thirds of the Ni(II) ions in the BOF-1 framework were oxidized to low spin Ni(III) and the I2 molecules were reduced to I3 anions, which were included in the channels. Previously, the Hofmann clathrate FeII(pz) [PtII(CN)4]173 was used as a host for the chemisorptive uptake of I2 molecules involving associative oxidation of PtII to PtIV and reduction of the dihalogen to the corresponding halide to give to FeII(pz)[PtII/IV(CN)4](I) with the formation of a PtIV–I bond. Also, Kitagawa and co-workers174 prepared a 2D layer porous framework [Fe(isophthalate)(bpy)] (bpy = 4,4′-bipyridyl) containing redox active Fe2+ centers. The activated samples could uptake I2 after exposure to I2 vapor, and the adsorbed amount corresponded to five iodine atoms per Fe2+ dimer. Of importance is that there was no adsorption of I2 for isostructural frameworks that included Zn2+, Mn2+ or Ni2+ with the same ligand systems. The iodine accommodation by only the Fe2+ compound is presumably due to the electron-donating property of the Fe2+ frameworks. The charge transferred from the Fe2+ framework to iodine, and half of the Fe ions in the framework were oxidized to Fe3+. The iodine inside was predominantly of the pentaiodide (I5) form.175 Furthermore, almost all the Fe3+ was reduced to Fe2+ as iodine was released, which proved its redox reversibility. These results demonstrate that some frameworks with redox active metal centers can react with iodine to achieve the trapping of iodine. These designs may be applied to prepare a new class of versatile multifunctional crystalline materials, including host lattices for guest molecular and chemical reactions.

Besides the storage of molecular iodine, porous MOFs can also be applied to uptake other iodine species such as I, I3 IO3, and organic iodides. In 2011, Dong et al.176 developed a porous 3D MOF, [Cd(L)2(ClO4)2]·H2O (L = 4-amino-3,5-bis(4-pyridyl-3-phenyl)-1,2,4-triazole), for iodine enrichment in vapor- and liquid-phases via both molecular sorption and anion-exchange approaches (Fig. 27). The Cd(II)-triazole MOF could uptake around 46 wt% of I2 to generate 2I2⊂CdL2 in vapor, and is also capable of trapping I2 in solution (0.8I2⊂CdL2). More interestingly, the MOF was capable of enriching other iodine anions. I3 and IO3 were included via ion-exchange by immersion in an aqueous solution of I2/KI and an aqueous solution of NaIO3. Also the other porous hydrogen-bonding Cu(II) coordination framework [Cu6(AcNTB)6]·6ClO4·nH2O177 was used for accumulating iodine species I, I3 or I2 in an aqueous environment and another robust and water-stable MOF TMBP·CuI (TCuI, TMBP = 3,3′,5,5′-tetramethyl-4,4′-bipyrazol)178 was capable of reversibly binding the volatile molecules of ICl and I2.


image file: c8mh01656a-f27.tif
Fig. 27 Schematic representation of iodine species enrichment based on [Cd(L)2(ClO4)2]·H2O via sorption and anion-exchange approaches. Reproduced from ref. 176 with permission from the Royal Society of Chemistry.

Organic iodides, as a class of radioactive species, should also be selectively captured and sequestered to ensure safe nuclear energy usage. The different MOFs, MIL-53(Al), MIL-100(Al), MIL-120(Al), CAU-1(Al), UiO-66(Zr), HKUST-1(Cu) and ZIF-8(Zn) have been studied for CH3I adsorption.179 Li and co-workers180 recently created organic iodide molecular traps through post-functionalization of the MIL-101-Cr MOF with tertiary amine-binding sites. The obtained molecular trap MIL-101-Cr-TED (TED = triethylenediamine) exhibited a high CH3I uptake capacity of 71 wt% at 150 °C, which is more than 340% higher than that of the industrial adsorbent Ag0@MOR (15 wt%, MOR = mordenite) under identical conditions. Furthermore, the resulting adsorbent could be recycled multiple times without loss in capacity. In combination with its chemical and thermal stability, high capture efficiency, and relatively low cost, the adsorbent demonstrated promise for industrial radioactive organic iodide capture from nuclear waste. Then they investigated the effect of amine molecules (N,N-dimethylethylenediamine (DMEDA), N,N-dimethyl-1,3-propanediamine (DMPDA), and N,N-dimethyl-1,4-butanediamine (DMBDA) selected for MOF functionalization) with varying lengths on organic iodide capture.181 The results revealed that DMEDA functionalized MIL-101-Cr gave rise to a record-high CH3I saturation uptake capacity of 80% at 150 °C, which is 5.3 times that of Ag0@MOR, a leading adsorbent material for capturing radioactive organic iodides during nuclear fuel reprocessing.

Furthermore, the size,130,153 crystallinity, morphology182–184 and further processing of MOF materials185–188 may affect their uptake of guests. For example, the adsorption and desorption rates of Nano-HMTI-1 are faster than that of HMTI-1 with a larger size. It seems that the pores in nano-HMTI-1 are more accessible than HMTI-1. Usually, amorphous crystals do not exhibit good adsorption capacity. However, ZIF-8 maintains its high adsorption capacity in extruded pellet form. Besides, there are some works of further processing of MOFs to prepare MOF-based composites for iodine uptake.185–188 For instance, the growth of HMTI-1 on silk fibers resulted in superior adsorption capacity and recovery of I2 under ambient conditions.185 Also, Ag@MIL-101 is a potentially attractive material for the highly efficient adsorption of radioactive iodide anions from water.188

3.2 Iodine capture in iodine-templated MOFs

One effective strategy to obtain MOFs is the introduction of templates (for example polyoxometalate anion and iodine molecule) introduced into the assembly process. Compared with polyoxometalate anions, the iodine molecule possesses particular chemistry such as oxidizability and conductivity. Besides, iodine has a lower melting point and good solubility, and can be removed easily through heating or guest molecule exchange. From another perspective, polyiodide-MOFs have earned much deserved attention not only due to their tremendous structural variety but also their applications in a wide variety of areas such as electrical devices, fuel cells, batteries, solar cells, and optical devices. To date, numerous MOFs containing small polyiodides such as I3, I42−, and I5 have been reported, and even a few large polyiodide species including I7, I9, I10, and I122− are also known.189

More than ten years ago, Horn et al. described a system combining the supramolecularity of M-phen (phen = 1,10-phenathroline) complexes and polyiodides, [CuI(phen)2]I3,190 [Fe(phen)3]I12,191 [M(phen)3]I7 (M = Mn, Fe),192 [M(phen)3]I8 (M = Mn, Fe),189 [Fe(phen)3](I3)2193 [Fe(phen)3]I14 and [M(phen)3]I18(M = Mn and Fe).194 The [Fe(phen)3]2+-polyiodide crystal system is particularly superior. Successively, three iodine-containing coordination polymers, [CuI(C5H3NI2)·1/2I2],195 {[Cu2(IN)3]·I5·2/3I2·H2O} (IN: isonicotinato)196 and [Cu(IN)2]·I2197 were obtained in the presence of iodine under hydrothermal conditions. [CuI(C5H3NI2)·1/2I2] is the first novel copper(I) diiodopyridine coordination polymer produced via the simultaneous reduction of copper(II) to copper(I) and substitution of carboxylato groups by iodo nucleophiles in a self-assembly chemical process. {[Cu2(IN)3]·I5·2/3I2·H2O} is a unique 3D nano-size open-channel framework polymer with polyiodide guests synthesized via an oxidation reaction route. Interestingly, [Cu(IN)2]·I2 is the removable iodine included open-framework polymer with an eclipsed 2D structure, in which each iodine molecule has weak donor–acceptor electrostatic attractions to oxygen atoms in the neighbouring layers. By evacuating the crystals at 195 °C, the iodine molecules could be removed to form [Cu(IN)2], which was still stable after the removal of the iodine molecules from the open-channel of its eclipsed layers. Besides, a copper(I) halide-based cationic framework, [Cu4I3(DABCO)2]I3 (DABCO = N,N′-dimethyl-1,4-diazabicyclo[2.2.2]octane),198 was constructed, and its channels were occupied by I2 and I. The guest I2 molecules could move freely in and out of the host-framework and the guest I could be exchanged by SCN. Thus, this compound exhibited iodine release/adsorption and ion-exchange properties.

By attempting to use iodine molecules as precursor templates, a Cu(II)-MOF with polyiodides {[Cu6(pybz)8(OH)2]·I5·I7}n (pybz, 4-pyridyl benzoate)199 was obtained successfully. It is an unexpectedly interdigitated and 2-fold interpenetrated bipillared-bilayer MOF (Fig. 28). Its cationic framework contains two types of zigzag 1D channels with the largest dimensions of ca. 4.8 and 5.1 Å, and segregate I5 and I7 polyiodides in its channels, respectively. Its iodine content was about 43.2 wt%, which is competitive with the traditional methods of introducing iodine. Moreover, it could slowly release iodine in dry methanol to give [Cu6(pybz)8(OH)2](I)2·3.5CH3OH and partially recover iodine from cyclohexane to form [Cu6(pybz)8(OH)2](I)2·xI2. The three MOFs could mutually transform into one another by the release and recovery of I2. In addition, the polyiodide anion-containing {[Cu6(pybz)8(OH)2]·I5·I7}n had good electrical conductivity and nonlinear optical activity. Therefore, rational synthesis using target guest (iodine) opens a promising approach for the construction of functional guest-encapsulated MOFs with new structural topologies and interesting properties.


image file: c8mh01656a-f28.tif
Fig. 28 Schematic presentation of the Cu(II)-MOF {[Cu6(pybz)8(OH)2]·I5·I7}n using iodine as the precursor template. Reproduced with permission from ref. 199 Copyright 2012 American Chemical Society.

Additionally, Cu(II)-MOF {[Cu4(TTTMB)4(NO3)4−(H2O)5]·(NO3)2·(I3)2·H2O}n (TTTMB, 1,3,5-tris(triazol-1-ylmethyl)-2,4,6-trimethylbenzene)200 containing polyiodide I3 was obtained through SCSC by immersing a CdII complex into an aqueous solution of Cu(NO3)2. Also, Ma and co-workers201 reported a 2D I2-containing [Cu(IN)2]·I2, which strikingly could transform into a known 3D compound [Cu(IN)2] through lattice iodine release. This SCSC transformation process represents an example involving the breakage/formation of chemical bonds via lattice iodine release. The iodine content was high at about 45.2 wt%. Subsequently, they have presented an I2-containing In(III)-MOF, [In2(pydc)3(H2O)]·0.5I2·0.5H2O (H2pydc = pyridine-2,5-dicarboxylic acid),202 which could release its lattice iodine molecules.

Besides, the presence of iodide groups in the frameworks also can affect iodine adsorption. Kawano et al.203 obtained two networks using a labile Cu4I4 cubane cluster [Cu4I4(PPh3)4] (1) and Td-symmetry ligand tetra-4-(4-pyridyl)phenylmethane (2). By controlling the cooling rate, a yellow needle crystal isomer 3a {[(CuI)2(2)]·solvent}n was produced exclusively under rapid cooling, while slow cooling produced orange prism crystals {[(Cu2I2)(2)]·solvent}n (isomer 3b) (Fig. 29). The isomer 3a is a non-interpenetrated porous network and has 1D channels with a pore window of 5.8 × 5.5 Å. The salient feature of this network is that the bridging iodide in the framework faces into the 1D channel. Isomer 3b is a quadruply interpenetrated network and possesses 1D channels with a pore window of 4.0 × 3.9 Å. Upon exposure to iodine vapor, the desolvated isomer 3a′ and 3b′ adsorbed iodine to give iodine-loaded samples. The crystal-structure analysis revealed that the pores of the isomer 3a encapsulated I2 by chemisorption through the formation of I3 groups from bridging iodide units, giving a reasonable geometry for the I3 ion. However, isomer 3b showed physisorption and the I2 molecules were arranged linearly along the 1D channels and highly disordered. Their work indicates that frameworks with iodine groups can encapsulate I2 by chemisorption.


image file: c8mh01656a-f29.tif
Fig. 29 Selective preparation of network isomers 3a and 3b. Structure of iodine-encapsulating network isomer 3a′ and 3b′. Reproduced with permission from ref. 203 Copyright 2013 Wiley-VCH.

The well-known crystalline sponge porous framework [(ZnI2)3(tpt)2]204 was assembled from the reaction of ZnI2 nodes with 2,4,6-tris(4-pyridyl)-1,3,5-triazine (tpt) ligand, in which iodide functional groups lining the one-dimensional channels could act as anchor sites for I2 adsorption. Woo, Murugesu and co-workers205 presented the stepwise crystallographic visualization of the incorporation of I2 and identified the preferred binding motifs throughout the uptake process. The guest I2 molecules initially bind with terminal iodide atoms of the framework to form [I4]2− units, and similar I42− ions were observed in MOF [Tb(Cu4I4)(ina)3(DMF)] (ina = isonicotinate) containing Cu4I4 nodes.206 However, as the adsorption progresses, the I2 molecules form less energetically favorable I3 groups with the same framework iodide, thereby allowing more guest molecules to be chemisorbed. At near saturation, even more binding motifs are observed in the same pores, including both physisorbed and chemisorbed guest molecules. The participation of both physisorption and chemisorption in the uptake of gaseous guests is another fascinating feature of MOFs since chemisorption is generally associated with frameworks exhibiting open metal sites that can bind guest molecules.207 The MOF [(ZnI2)3(tpt)2] could load an excess of 173 wt% of I2 at room temperature, which is comparable with the high I2 uptake capacity reported for Cu-BTC (175 wt%).118 This high uptake of I2 can be attributed to a number of factors, including favorable guest–host interactions, framework flexibility, and high porosity (∼60%). More specifically, this MOF offers strong adsorption site for I2 through the terminal iodide ion of the framework as well as favorable π-halogen interaction with the tpt ligand. Their study presented the successful identification of a unique set of host–guest interactions, which will drive the improvement of high capacity iodine capture materials. In contrast, when the reaction between ZnI2 and tpt took place in the presence of triphenylene, a different triphenylene-loaded network [(ZnI2)3(tpt)2]·triphenylene208 with two types of channels was formed. Recently, Burrows, Raithby et al.209 evaluated the iodine uptake in a triphenylene-loaded framework using dynamic X-ray crystallography. The included iodine molecules interacted with the iodine atoms of the ZnI2 nodes, forming coordinated I3 ions. Also, the porous polymer [(Na2I2CB[6])·8H2O]n (CB[6] = Cucurbit[6]uril)210 has discrete iodide anions located in its channel regions, which are surrounded by {Na2CB[6]} chains, playing both space-filling and charge-compensating roles. After I2 vapor exposure, it could uptake ca. 25 wt% I2. The formation of halogen bonds between discrete iodide ions and iodine molecules is predominant driving forces for iodine adsorption. These works further indicate that iodine groups in the crystal lattice are ideal and preferential molecular docks for iodine accommodation.

Briefly, MOF materials possess large surface areas, high porosities, tunable chemical compositions, potential for post-synthetic modification and other structural and chemical features, which make MOFs promising candidates for iodine capture. With the research progress, more and more MOFs with different iodine adsorption capacities have been reported. Some well-known classic MOFs (ZIF-8 and HKUST-1) have been explored as iodine sorbents. The introduction of functional groups (–NH2 and –SH, pyridine) can enhance iodine adsorption. Besides the storage for molecular iodine, porous MOFs can also be applied to uptake other iodine species such as I and organic iodides. In addition, the size, morphology, and further processing of MOFs can also affect their uptake capacities for iodine. Also, rational synthesis using iodine species as templates is a promising approach for constructing functional guest-encapsulated MOFs. However, only a few materials can remain stable in wet conditions, and these materials have limited thermal stability and are not produced on a large scale, which limit their application.

4. Analysis and improvement strategies for iodine capture of POP and MOF materials

POP and MOF materials are becoming promising candidates for iodine capture due to their high surface areas, large pore volumes, precisely adjustable pore structures, good physicochemical stabilities, facile pre- or post-synthetic modification, etc. More relevant to the present study is the increasing number of POPs and MOFs reported to possess various iodine uptake properties, including irreversible and reversible trapping and even chemisorption. MOFs are organic–inorganic hybrid materials that consist of inorganic metal ions and organic ligands linked by coordination bonds, while POPs are constructed from various pure organic building blocks by strong covalent bonds. Most MOFs are unstable under humid conditions, which limits their application. POPs are usually amorphous solids, unlike crystalline MOFs; however, their excellent physicochemical stability makes POPs more suited for real applications. In addition, compared to MOFs, POPs usually have higher iodine adsorption capacity. However, in the highly ordered crystalline MOF system, it is helpful to elucidate the relationship between the structure and the adsorption behavior of iodine. Furthermore, the synthetic costs of POP and MOF materials are high and their production is in the laboratory stage; thus, their actual application is far away.

It has been demonstrated that both molecular porosity and preferential binding sites in porous materials play crucial roles in determining their efficiency and capacity for iodine adsorption. There are a number of potential routes for increasing the iodine uptake capacity in POPs and MOFs, which include: (i) increasing their surface areas and pore volumes. Generally, the high surface areas and large pore volumes of POPs and MOFs can be beneficial in improving iodine capture by providing more sites with interactions between sorbent and iodine. (ii) Tuning pore size and pore structure. A combination of suitable pore size (6–8 Å) and shape/geometry (squared-shaped pore) of channels and suitable pore decorations (e.g., hydroxyl groups) can provide a unique platform to stabilize the formation of a complex assembly of I2, resulting in highly efficient packing and hence exceptional storage of I2. (iii) Increasing the affinity of host materials to iodine. Some networks with lower surface areas and pore volumes but strong affinity have led to increased adsorption uptake. Thus, the introduction of desirable functional groups and open metal sites into the pore structures to modify the environment of the pores and surfaces will allow tuning of the interactions with guests, which is regarded as an effective design to enhance affinity. Open-pored POPs and MOFs containing unsaturated active centers offer potential for active sites to directly interact with the targeted capture iodine. Porous materials with pore functionality of electron-donating groups, for example –OH, –NH2, –SH, and pyridine, typically achieve high adsorption efficiency for iodine molecules. Accordingly, materials with discrete iodide moieties can be an excellent choice to realize high iodine adsorption efficiency since it is well-known that electron-rich iodide moieties can serve as electron-donating groups to bind iodine molecules via strong halogen bonds to form polyiodide, in which, iodide serves as a halogen bond acceptor and iodine is the halogen bond donor. (iv) Increasing the physicochemical stability. The excellent stability of materials is a prerequisite for ensuring their practical application. Especially, the strong covalent bonds of POP materials afford high thermal and chemical stability, and their constituent light elements add gravimetric advantage. POPs are stable in different polar solvents, and moisture and water conditions, and they can be generally reused for manifold cycles via the use of a simple filter. The use of specific porous materials has often been dependent on sorption material characteristics such as size selectivity into a pore opening, chemical bonding with framework sites, such as metal centers or functional sites, and geometric “pockets” that hold the guest species preferentially.

To date, the research on POPs and MOFs for high efficiency capture of iodine still faces the following problems and challenges: (i) presently, the preparation of porous materials requires complex monomers, expensive transition metal catalysts and harsh reaction conditions. Besides, the synthesis and screening have a certain degree of randomness, and the understanding of the various factors that affect the final structure is still in its infancy. Thus, the controllable synthesis and assembly of POPs and MOFs are still key scientific issues to be solved. (ii) Compared with activated carbon with organic amine impregnated, there is still a large gap between the binding capacity and adsorption capacity for iodine. The interaction between POP and MOF materials and iodine molecules is weak, and iodine adsorption in the low-pressure zone is low, and thus it is difficult to meet practical application requirements. (iii) Most of the current research focuses on the synthetic strategies and functions of materials. The structure–activity relationship between the structure and the adsorption properties of iodine, and the adsorption behavior of iodine have not been elucidated. The mechanism of action on iodine, the interaction between absorbent and adsorbate, and the relationship between properties and macroscopic solid materials also need to be further studied. There has been a small number of theoretical research articles focused on iodine adsorption by MOFs using periodic dispersion density functional theory and Grand Canonical Monte Carlo simulations,211–213 which help deepen our understanding. For POP materials, most are amorphous or microcrystalline and their structures cannot be accurately characterized by experimental means. For MOFs, although their crystal structures can be unambiguously determined, host–iodine interactions are still difficult to characterize due to the loss of crystallinity and/or the poor crystal quality after the loading of guest molecules, the serious disorder of guest molecules, etc. Presently, it is rather rare to determine the location of guest iodine within pores. Nevertheless, to better understand and improve the adsorption function of POP and MOF materials, it is necessary to investigate the locations of the guest molecules within their pores and the interactions between the guest molecules and pore structures.

5. Summary and concluding remarks

Presently, the storage of iodine and the treatment of radioactive iodine have caused great concerns to the population. POP and MOF materials show outstanding application prospects in the field of adsorption of iodine. Compared with traditional porous materials, POP and MOF materials have some obvious advantages. The types of building blocks are diversified, which can form various structures. Their surface areas, pore sizes and pore volumes are adjustable, and it is easy to modify their inner pores and surfaces. In recent years, a large number of POPs and MOFs have been reported, and these materials have demonstrated potential applications in the capture and storage of iodine. Herein, we focused on two relatively new classes of materials, POPs and MOFs, as platforms for iodine capture.

Although POP and MOF sorbents are highly praised for their good performance in iodine capture, they are difficult to apply industrially for the trapping and treatment of iodine. Furthermore, they are still new laboratory materials and are therefore produced in very small quantities at high prices. Thus, the further development of efficient sorbents that combine a high iodine adsorption capacity over a wide temperature range, high stability and selectivity in the capture of iodine at high temperatures and humidity and in the presence of other molecules, and low cost still remains a challenge. We recommend that new sorbent materials should have controlled morphology, well-defined particle sizes, the absence of any blocked bulk volume, high sensitivity with fairly reduced purge volumes, low energy and overall costs, and high chemical durability. Thus, the synthesis of composite materials combining the advantageous properties of several types of iodine sorbents can be a promising way to fulfil the requirements of real applications. We hope this review can provide some useful references for researchers, and further broaden the research scope of adsorbent materials.

We tried to present an up-to-date overview of this rapidly growing field; however, this subject is very active, and many papers are contributed each year (even during the preparation of this article) from chemists, energy, environmental and materials scientists. Therefore, it is hard to consider all publications in this field in limited pages. We apologize here if some significant contributions were omitted.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

The financial support of the Science and Technology Program of Jilin Province (Grants No. 20180520159JH, 20180520152JH), the Research Program on Science and Technology from the Education Department of Jilin Province (JJKH20190995KJ, JJKH20191011KJ), the National Natural Science Foundation of China (Grants No. 21501065), and the Changbai Mountain Scholars Program (Grant No. 2013073) are acknowledged.

Notes and references

  1. E. Kintisch, Science, 2005, 310, 1406 CrossRef CAS PubMed.
  2. P. A. Kharecha and J. E. Hansen, Environ. Sci. Technol., 2013, 47, 4889 CrossRef CAS PubMed.
  3. J. Y. Lu, Coord. Chem. Rev., 2003, 246, 327 CrossRef CAS.
  4. J. E. T. Hoevea and M. Z. Jacobson, Energy Environ. Sci., 2012, 5, 8743 RSC.
  5. K. S. Subrahmanyam, D. Sarma, C. D. Malliakas, K. Polychronopoulou, B. J. Riley, D. A. Pierce, J. Chun and M. G. Kanatzidis, Chem. Mater., 2015, 27, 2619 CrossRef CAS.
  6. Y. S. Shimamoto, Y. Takahashi and Y. Terada, Environ. Sci. Technol., 2011, 45, 2086 CrossRef CAS PubMed.
  7. F. C. Kupper, M. C. Feiters, B. Olofsson, T. Kaiho, S. Yanagida, M. B. Zimmermann, L. J. Carpenter, G. W. Luther, Z. L. Lu, M. Jonsson and L. Kloo, Angew. Chem., Int. Ed., 2011, 50, 11598 CrossRef PubMed.
  8. A. Saiz-Lopez, J. M. C. Plane, A. R. Baker, L. J. Carpenter, R. von Glasow, J. C. G. Martin, G. McFiggans and R. W. Saunders, Chem. Rev., 2012, 112, 1773 CrossRef CAS PubMed.
  9. J. L. Schnoor, Lessons from Fukushima, Environ. Sci. Technol., 2011, 45, 3820 CrossRef CAS PubMed.
  10. S. Xu, S. P. H. T. Freeman, X. Hou, A. Watanabe, K. Yamaguchi and L. Zhang, Environ. Sci. Technol., 2013, 47, 10851 CrossRef CAS PubMed.
  11. S. Liu, N. Wang, Y. Zhang, Y. Li, Z. Han and P. Na, J. Hazard. Mater., 2015, 284, 171 CrossRef CAS PubMed.
  12. K. S. Subrahmanyam, C. D. Malliakas, D. Sarma, G. S. Armatas, J. Wu and M. G. Kanatzidis, J. Am. Chem. Soc., 2015, 137, 13943 CrossRef CAS PubMed.
  13. T. C. T. Pham, S. Docao, I. C. Hwang, M. K. Song, D. Y. Choi, D. Moon, P. Oleynikov and K. B. Yoon, Energy Environ. Sci., 2016, 9, 1050 RSC.
  14. K. W. Chapman, P. J. Chupas and T. M. Nenoff, J. Am. Chem. Soc., 2010, 132, 8897 CrossRef CAS.
  15. T. M. Geng, Z. M. Zhu, W. Y. Zhang and Y. Wang, J. Mater. Chem. A, 2017, 5, 7612 RSC.
  16. X. Qian, Z. Q. Zhu, H. X. Sun, F. Ren, P. Mu, W. Liang, L. Chen and A. Li, ACS Appl. Mater. Interfaces, 2016, 8, 21063 CrossRef CAS PubMed.
  17. N. V. Nguyen, J. Jeong, D. Shin, B. S. Kim, J. Lee and B. D. Pandey, Mater. Trans., 2012, 53, 760 CrossRef CAS.
  18. G. M. Adams and A. S. Weller, Coord. Chem. Rev., 2018, 355, 150 CrossRef CAS.
  19. C. S. Diercks and O. M. Yaghi, Science, 2017, 355, 1585 CrossRef PubMed.
  20. N. Huang, L. Zhai, H. Xu and D. Jiang, J. Am. Chem. Soc., 2017, 139, 2428 CrossRef CAS.
  21. G. M. Espallargas and E. Coronado, Chem. Soc. Rev., 2018, 47, 533 RSC.
  22. P. A. Kobielska, A. J. Howarth, O. K. Farha and S. Nayak, Coord. Chem. Rev., 2018, 358, 92 CrossRef CAS.
  23. K. Shen, L. Zhang, X. Chen, L. Liu, D. Zhang, Y. Han, J. Chen, J. Long, R. Luque, Y. Li and B. Chen, Science, 2018, 359, 206 CrossRef CAS PubMed.
  24. S. U. Nandanwar, K. Coldsnow, V. Utgikar, P. Sabharwall and D. E. Aston, Chem. Eng. J., 2016, 306, 369 CrossRef CAS.
  25. K. Vellingiri, K.-H. Kim, A. Pournara and A. Deep, Prog. Mater. Sci., 2018, 94, 1 CrossRef CAS.
  26. J. Huve, A. Ryzhikov, H. Nouali, V. Lalia, G. Augé and T. J. Daou, RSC Adv., 2018, 8, 29248 RSC.
  27. Y. Lin, Y. Zhu, G. Kuang, G. Yu and R. Jin, Prog. Chem., 2017, 29, 766 Search PubMed.
  28. Y. Xu, S. Jin, H. Xu, A. Nagai and D. Jiang, Chem. Soc. Rev., 2013, 42, 8012 RSC.
  29. X. Zhuang, D. Gehrig, N. Forler, H. Liang, M. Wagner, M. R. Hansen, F. Laquai, F. Zhang and X. Feng, Adv. Mater., 2015, 27, 3789 CrossRef CAS.
  30. C. Gu, N. Huang, J. Gao, F. Xu, Y. Xu and D. Jiang, Angew. Chem., Int. Ed., 2014, 53, 4850 CrossRef CAS PubMed.
  31. M. P. Tsyurupa and V. A. Davankov, React. Funct. Polym., 2006, 66, 768 CrossRef CAS.
  32. T. A. Makal, J. R. Li, W. G. Lu and H. C. Zhou, Chem. Soc. Rev., 2012, 41, 1400 RSC.
  33. Y. Zhang and S. N. Riduan, Chem. Soc. Rev., 2012, 41, 2083 RSC.
  34. N. B. McKeown and P. M. Budd, Chem. Soc. Rev., 2006, 35, 675 RSC.
  35. P. M. Budd, B. S. Ghanem, S. Makhseed, N. B. McKeown, K. J. Msayib and C. E. Tattershall, Chem. Commun., 2004, 230 RSC.
  36. N. B. McKeown and P. M. Budd, Macromolecules, 2010, 43, 5163 CrossRef CAS.
  37. X. Feng, X. Ding and D. Jiang, Chem. Soc. Rev., 2012, 41, 6010 RSC.
  38. A. P. Côté, A. I. Benin, N. W. Ockwig, M. O’Keeffe, A. J. Matzger and O. M. Yaghi, Science, 2005, 310, 1166 CrossRef PubMed.
  39. J. J. Perry IV, J. A. Perman and M. J. Zaworotko, Chem. Soc. Rev., 2009, 38, 7761 RSC.
  40. J. R. Holst, A. Trewin and A. I. Cooper, Nat. Chem., 2010, 2, 915 CrossRef CAS PubMed.
  41. J. T. A. Jones, T. Hasell, X. Wu, J. Bacsa, K. E. Jelfs, M. Schmidtmann, S. Y. Chong, D. J. Adams, A. Trewin, F. Schiffman, F. Cora, B. Slater, A. Steiner, G. M. Day and A. I. Cooper, Nature, 2011, 474, 367 CrossRef CAS PubMed.
  42. D. C. Wu, F. Xu, B. Sun, R. W. Fu, H. K. He and K. Matyjaszewski, Chem. Rev., 2012, 112, 3959 CrossRef CAS PubMed.
  43. B. Li, Y. Zhang, R. Krishna, K. Yao, Y. Hang, Z. Wu, D. Ma, Z. Shi, T. Pham, B. Space, J. Liu, P. K. Thallapally, J. Liu, M. Chrzanowski and S. Ma, J. Am. Chem. Soc., 2014, 136, 8654 CrossRef CAS.
  44. X. Wu, M. Shaibani, S. J. D. Smith, K. Konstas, M. R. Hill, H. Wang, K. Zhang and Z. Xie, J. Mater. Chem. A, 2018, 6, 11327 RSC.
  45. S. Kotha, E. Brahmachary and K. Lahiri, Eur. J. Org. Chem., 2005, 4741 CrossRef CAS.
  46. V. Gevorgyan and Y. Yamamoto, J. Organomet. Chem., 1999, 576, 232 CrossRef CAS.
  47. B. Karami and S. Khodabakhshi, J. Serb. Chem. Soc., 2011, 76, 1191 CrossRef CAS.
  48. M. Yamada, Y. Tanaka, Y. Yoshimoto, S. Kuroda and I. Shimao, Bull. Chem. Soc. Jpn., 1992, 65, 1006 CrossRef CAS.
  49. J. X. Jiang, F. B. Su, A. Trewin, C. D. Wood, N. L. Campbell, H. J. Niu, C. Dickinson, A. Y. Ganin, M. J. Rosseinsky, Y. Z. Khimyak and A. I. Cooper, Angew. Chem., Int. Ed., 2007, 46, 8574 CrossRef CAS.
  50. J. X. Jiang, A. Trewin, D. J. Adams and A. I. Cooper, Chem. Sci., 2011, 2, 1777 RSC.
  51. J. Schmidt, M. Werner and A. Thomas, Macromolecules, 2009, 42, 4426 CrossRef CAS.
  52. R. Dawson, A. Laybourn, Y. Z. Khimyak, D. J. Adams and A. I. Cooper, Macromolecules, 2010, 43, 8524 CrossRef CAS.
  53. Y. H. Xu, L. Chen, Z. Q. Guo, A. Nagai and D. Jiang, J. Am. Chem. Soc., 2011, 133, 17622 CrossRef CAS.
  54. X. Liu, H. Li, Y. Zhang, B. Xu, A. Sigen, H. Xia and Y. Mu, Polym. Chem., 2013, 4, 2445 RSC.
  55. A. Bhunia, V. Vasylyeva and C. Janiak, Chem. Commun., 2013, 49, 3961 RSC.
  56. P. Kuhn, M. Antonietti and A. Thomas, Angew. Chem., Int. Ed., 2008, 47, 3450 CrossRef CAS.
  57. R. Palkovits, M. Antonietti, P. Kuhn, A. Thomas and F. Schüth, Angew. Chem., Int. Ed., 2009, 48, 6909 CrossRef CAS PubMed.
  58. L. Qin, G. Xu, C. Yao and Y. Xu, Chem. Commun., 2016, 52, 12602 RSC.
  59. P. Kuhn, A. Thomas and M. Antonietti, Macromolecules, 2009, 42, 319 CrossRef CAS.
  60. W. Zhang, F. Liang, C. Li and L. G. Qiu, J. Hazard. Mater., 2011, 186, 984 CrossRef CAS PubMed.
  61. Z. H. Xiang, D. P. Cao, W. C. Wang, W. T. Yang, B. Y. Han and J. M. Lu, J. Phys. Chem. C, 2012, 116, 5974 CrossRef CAS.
  62. G. Cheng, T. Hasell, A. Trewin, D. J. Adams and A. I. Cooper, Angew. Chem., Int. Ed., 2012, 51, 12727 CrossRef CAS.
  63. J. X. Jiang, F. B. Su, A. Trewin, C. D. Wood, H. J. Niu, J. T. A. Jones, Y. Z. Khimyak and A. I. Cooper, J. Am. Chem. Soc., 2008, 130, 7710 CrossRef CAS PubMed.
  64. J. Weber, J. Schmidt, A. Thomas and W. Böhlmann, Langmuir, 2010, 26, 15650 CrossRef CAS PubMed.
  65. D. Z. Tan, W. J. Fan, W. N. Xiong, H. X. Sun, Y. Q. Cheng, X. Y. Liu, C. G. Meng, A. Li and W. Q. Deng, Macromol. Chem. Phys., 2012, 213, 1435 CrossRef CAS.
  66. B. Kiskan and J. Weber, ACS Macro Lett., 2012, 1, 37 CrossRef CAS.
  67. A. Sigen, Y. Zhang, Z. Li, H. Xia, M. Xue, X. Liu and Y. Mu, Chem. Commun., 2014, 50, 8495 RSC.
  68. K. C. Park, J. Cho and C. Y. Lee, RSC Adv., 2016, 6, 75478 RSC.
  69. H. Li, X. S. Ding and B. H. Han, Chem. – Eur. J., 2016, 22, 11863 CrossRef CAS PubMed.
  70. K. Su, W. Wang, B.-B. Li and D. Yuan, ACS Sustainable Chem. Eng., 2018, 6, 17402 CrossRef CAS.
  71. Q. Q. Dang, X. M. Wang, Y. F. Zhan and X. M. Zhang, Polym. Chem., 2016, 7, 643 RSC.
  72. Y. Z. Liao, J. Weber, B. M. Mills, Z. H. Ren and C. F. J. Faul, Macromolecules, 2016, 49, 6322 CrossRef CAS.
  73. X. Qian, B. Wang, Z. Q. Zhu, H. X. Sun, F. Ren, P. Mu, C. Ma, W. D. Liang and A. Li, J. Hazard. Mater., 2017, 338, 224 CrossRef CAS PubMed.
  74. Y. Liao, Z. Cheng, W. Zuo, A. Thomas and C. F. J. Faul, ACS Appl. Mater. Interfaces, 2018, 10, 16049 CrossRef PubMed.
  75. C. Yao, G. Li, J. Wang, Y. Xu and L. Chang, Sci. Rep., 2018, 8, 1867 CrossRef PubMed.
  76. J. Y. Weng, Y. L. Xu, W. C. Song and Y. H. Zhang, J. Polym. Sci., Polym. Chem., 2016, 54, 1724 CrossRef CAS.
  77. Y. H. Abdelmoaty, T. D. Tessema, F. A. Choudhury, O. M. El-Kadri and H. M. El-Kaderi, ACS Appl. Mater. Interfaces, 2018, 10, 16049 CrossRef CAS PubMed.
  78. L. Lin, H. D. Guan, D. L. Zou, Z. J. Dong, Z. Liu, F. F. Xu, Z. G. Xie and Y. X. Li, RSC Adv., 2017, 7, 54407 RSC.
  79. Z. Guo, P. Sun, X. Zhang, J. Lin, T. Shi, S. Liu, A. Sun and Z. Li, Chem. – Asian J., 2018, 13, 2046 CrossRef CAS.
  80. T. Geng, S. Ye, Z. Z. Zhu and W. Hang, J. Mater. Chem. A, 2018, 6, 2808 RSC.
  81. D. Chen, Y. Fu, W. Yu, G. Yu and C. Pan, Chem. Eng. J., 2018, 334, 900 CrossRef CAS.
  82. H. Ma, J. J. Chen, L. X. Tan, J. H. Bu, Y. H. Zhu, B. Tan and C. Zhang, ACS Macro Lett., 2016, 5, 1039 CrossRef CAS.
  83. J. K. Zang, Z. Q. Zhu, P. Mu, H. X. Sun, W. D. Liang, F. C. Yu, L. H. Chen and A. Li, Macromol. Mater. Eng., 2016, 301, 1104 CrossRef CAS.
  84. P. Mu, H. Sun, T. Chen, W. Zhang, Z. Zhu, W. Liang and A. Li, Macromol. Mater. Eng., 2017, 302, 1700156 CrossRef.
  85. D. K. L. Harijan, V. Chandra, T. Yoon and K. S. Kim, J. Hazard. Mater., 2018, 344, 576 CrossRef CAS PubMed.
  86. F. Ren, Z. Q. Zhu, X. Qian, W. D. Liang, P. Mu, H. X. Sun, J. H. Liu and A. Li, Chem. Commun., 2016, 52, 9797 RSC.
  87. T. Geng, W. Zhang, Z. Zhu, G. Chen, L. Ma, S. Ye and Q. Niu, Polym. Chem., 2018, 9, 777 RSC.
  88. G. Y. Li, C. Yao, J. K. Wang and Y. H. Xu, Sci. Rep., 2017, 7, 13972 CrossRef PubMed.
  89. Y. L. Zhu, Y. J. Ji, D. G. Wang, Y. Zhang, H. Tang, X. R. Jia, M. Song, G. P. Yu and G. C. Kuang, J. Mater. Chem. A, 2017, 5, 6622 RSC.
  90. G. Das, T. Prakasam, S. Nuryyeva, D. S. Han, A. Abdel-Wahab, J. C. Olsen, K. Polychronopoulou, C. Platas-Iglesias, F. Ravaux, M. Jouiadf and A. Trabolsi, J. Mater. Chem. A, 2016, 4, 15361 RSC.
  91. D. Shetty, J. Raya, D. S. Han, Z. Asfari, J. C. Olsen and A. Trabolsi, Chem. Mater., 2017, 29, 8968 CrossRef CAS.
  92. Y. F. Chen, H. X. Sun, R. X. Yang, T. T. Wang, C. J. Pei, Z. T. Xiang, Z. Q. Zhu, W. D. Liang, A. Li and W. Q. Deng, J. Mater. Chem. A, 2015, 3, 87 RSC.
  93. Y. J. Lin, Y. L. Zhu, G. C. Kuang, G. M. Yu and R. G. Jin, Prog. Chem., 2017, 29, 766 Search PubMed.
  94. H. Bildirir, J. P. Paraknowitsch and A. A. Thomas, Chem. – Eur. J., 2014, 20, 9543 CrossRef CAS.
  95. Q. Q. Dang, H. J. Wan and X. M. Zhang, ACS Appl. Mater. Interfaces, 2017, 9, 21438 CrossRef CAS PubMed.
  96. Z. J. Yin, S. Q. Xu, T. G. Zhan, Q. Y. Qi, Z. Q. Wu and X. Zhao, Chem. Commun., 2017, 53, 7266 RSC.
  97. C. Wang, Y. Wang, R. Ge, X. D. Song, X. Q. Xing, Q. K. Jiang, H. Lu, C. Hao, X. W. Guo, Y. A. Gao and D. L. Jiang, Chem. – Eur. J., 2018, 24, 585 CrossRef CAS PubMed.
  98. P. Wang, Q. Xu, Z. Li, W. Jiang, Q. Jiang and D. Jiang, Adv. Mater., 2018, 30, 1801991 CrossRef PubMed.
  99. X. Guo, Y. Tian, M. Zhang, Y. Li, R. Wen, X. Li, X. Li, Y. Xue, L. Ma, C. Xia and S. Li, Chem. Mater., 2018, 30, 2299 CrossRef CAS.
  100. Y. X. Lin, X. F. Jiang, S. T. Kim, S. B. Alahakoon, X. S. Hou, Z. Y. Zhang, C. M. Thompson, R. A. Smaldone and C. F. Ke, J. Am. Chem. Soc., 2017, 139, 7172 CrossRef CAS PubMed.
  101. Y. Liao, J. Li and A. Thomas, ACS Macro Lett., 2017, 6, 1444 CrossRef CAS.
  102. Y. S. Lan, M. M. Tong, Q. Y. Yang and C. L. Zhong, CrystEngComm, 2017, 19, 4920 RSC.
  103. C. Y. Pei, T. Ben, S. X. Xu and S. L. Qiu, J. Mater. Chem. A, 2014, 2, 7179 RSC.
  104. Z. J. Yan, Y. Yuan, Y. Y. Tian, D. M. Zhang and G. S. Zhu, Angew. Chem., Int. Ed., 2015, 19, 12733 CrossRef.
  105. J. Yu, L. H. Xie, J. R. Li, Y. Ma, J. M. Seminario and P. B. Balbuena, Chem. Rev., 2017, 117, 9674 CrossRef CAS PubMed.
  106. A. Cadiau, Y. Belmabkhout, K. Adil, P. M. Bhatt, R. S. Pillai, A. Shkurenko, C. Martineau, G. Maurin and M. Eddaoudi, Science, 2017, 735, 731 CrossRef PubMed.
  107. J. B. Decoste and G. W. Peterson, Chem. Rev., 2014, 114, 5695 CrossRef CAS PubMed.
  108. P. Li, N. A. Vermeulen, X. Gong, C. D. Malliakas, J. F. Stoddart, J. T. Hupp and O. K. Farha, Angew. Chem., Int. Ed., 2016, 55, 1 CrossRef.
  109. J. R. Li, Y. G. Ma, M. C. McCarthy, J. Sculley, J. M. Yu, H. K. Jeong, P. B. Balbuena and H. C. Zhou, Coord. Chem. Rev., 2011, 255, 1791 CrossRef CAS.
  110. L. L. Lv, J. Yang, H. M. Zhang, Y. Y. Liu and J. F. Ma, Inorg. Chem., 2015, 54, 1744 CrossRef CAS PubMed.
  111. Z. M. Wang, Y. J. Zhang, T. Liu, M. Kurmoo and S. Gao, Adv. Funct. Mater., 2007, 17, 1523 CrossRef CAS.
  112. A. K. Chaudhari, S. Mukherjee, S. S. Nagarkar, B. Joarder and S. K. Ghosh, CrystEngComm, 2013, 15, 9465 RSC.
  113. B. F. Abrahams, M. Moylan, S. D. Orchard and R. Robson, Angew. Chem., Int. Ed., 2003, 42, 1848 CrossRef CAS.
  114. K. S. Park, Z. Ni, A. P. Côté, J. Y. Choi, R. Huang, F. J. Uribe-Romo, H. K. Chae, M. O’Keeffe and O. M. Yaghi, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 10186 CrossRef CAS PubMed.
  115. D. F. Sava, M. A. Rodriguez, K. W. Chapman, P. J. Chupas, J. A. Greathouse, P. S. Crozier and T. M. Nenoff, J. Am. Chem. Soc., 2011, 133, 12398 CrossRef CAS PubMed.
  116. J. T. Hughes, D. F. Sava, T. M. Nenoff and A. Navrotsky, J. Am. Chem. Soc., 2013, 135, 16256 CrossRef CAS PubMed.
  117. S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen and I. D. Williams, Science, 1999, 283, 1148 CrossRef CAS PubMed.
  118. D. F. Sava, K. W. Chapman, M. A. Rodriguez, J. A. Greathouse, P. S. Crozier, H. Zhao, P. J. Chupas and T. M. Nenoff, Chem. Mater., 2013, 25, 2591 CrossRef CAS.
  119. D. Banerjee, X. Chen, S. S. Lobanov, A. M. Plonka, X. Chan, J. A. Daly, T. Kim, P. K. Thallapally and J. B. Parise, ACS Appl. Mater. Interfaces, 2018, 10, 10622 CrossRef CAS PubMed.
  120. D. Banerjee, Z. Zhang, A. M. Plonka, J. Li and J. B. A. Parise, Calciumin, Cryst. Growth Des., 2012, 12, 2162 CrossRef CAS.
  121. X. Y. Chen, A. M. Plonka, D. Banerjee, R. Krishna, H. T. Schaef, S. Ghose, P. K. Thallapally and J. B. Parise, J. Am. Chem. Soc., 2015, 137, 7007 CrossRef CAS PubMed.
  122. T. Assaad and B. Assfour, J. Nucl. Mater., 2017, 493, 6 CrossRef CAS.
  123. M. H. Zeng, Q. X. Wang, Y. X. Tan, S. Hu, H. X. Zhao, L. S. Long and M. Kurmoo, J. Am. Chem. Soc., 2010, 132, 2561 CrossRef CAS PubMed.
  124. M. Murthi and R. Q. Snurr, Langmuir, 2004, 20, 2489 CrossRef CAS.
  125. G. Mehlana, G. Ramon and S. A. Bourne, Microporous Mesoporous Mater., 2016, 231, 21 CrossRef CAS.
  126. R. X. Yao, X. Cui, X. X. Jia, F. Q. Zhang and X. M. Zhang, Inorg. Chem., 2016, 55, 9270 CrossRef CAS PubMed.
  127. G. Massasso, J. Long, J. Haines, S. Devautour-Vinot, G. Maurin, A. Grandjean, B. Onida, B. Donnadieu, J. Larionova, C. Guérin and Y. Guari, Inorg. Chem., 2014, 53, 4269 CrossRef CAS PubMed.
  128. L. Dobrzańska, G. O. Lloyd, H. G. Raubenheimer and L. J. Barbour, J. Am. Chem. Soc., 2006, 128, 698 CrossRef PubMed.
  129. P. Cui, L. Ren, Z. Chen, H. Hu, B. Zhao, W. Shi and P. Cheng, Inorg. Chem., 2012, 51, 2303 CrossRef CAS PubMed.
  130. S. H. Chae, H. C. Kim, Y. S. Lee, S. Huh, S. J. Kim, Y. Kim and S. J. Lee, Cryst. Growth Des., 2015, 15, 268 CrossRef CAS.
  131. M. Liu, W. Liao, C. Hu, S. Du and H. Zhang, Angew. Chem., Int. Ed., 2012, 51, 1585 CrossRef CAS PubMed.
  132. L. Chen, K. Tan, Y. Q. Lan, S. Li, K. Z. Shao and Z. M. Su, Chem. Commun., 2012, 48, 5919 RSC.
  133. W. W. He, S. Li, G. S. Yang, Y. Q. Lan, Z. M. Su and Q. Fu, Chem. Commun., 2012, 48, 10001 RSC.
  134. P. Shen, W. W. He, D. Y. Du, H. L. Jiang, S. L. Li, Z. L. Lang, Z. M. Su, Q. Fu and Y. Q. Lan, Chem. Sci., 2014, 5, 1368 RSC.
  135. Z. Wang, B. Zhang, H. Fujiwara, H. Kobayashi and M. Kurmoo, Chem. Commun., 2004, 416 RSC.
  136. Z. Wang, Y. Zhang, M. Kurmoo, T. Liu, S. Vilminot, B. Zhao and S. Gao, Aust. J. Chem., 2006, 59, 617 CrossRef CAS.
  137. J. Wang, J. Luo, X. Luo, J. Zhao, D. S. Li, G. Li, Q. Huo and Y. Liu, Cryst. Growth Des., 2015, 15, 915 CrossRef CAS.
  138. J. Liu, R. Xiao, Y. L. Wong, X. P. Zhou, M. Zeller, A. D. Hunter, Q. Fang, L. Liao and Z. Xu, Inorg. Chem., 2018, 57, 4807 CrossRef CAS PubMed.
  139. X. M. Zhang, C. W. Zhao, J. P. Ma, Y. Yu, Q. K. Liu and Y. B. Dong, Chem. Commun., 2015, 51, 839 RSC.
  140. N. X. Zhu, C. W. Zhao, J. C. Wang, Y. A. Li and Y. B. Dong, Chem. Commun., 2016, 52, 12702 RSC.
  141. Y. Rachuri, K. K. Bisht and E. Suresh, Cryst. Growth Des., 2014, 14, 3300 CrossRef CAS.
  142. Y. Rachuri, K. K. Bisht, B. Parmar and E. Suresh, J. Solid State Chem., 2015, 223, 23 CrossRef CAS.
  143. J. Li, P. Huang, X. R. Wu, J. Tao, R. B. Huang and L. S. Zheng, Chem. Sci., 2013, 4, 3232 RSC.
  144. Y. H. Luo, X. Y. Yu, J. J. Yang and H. Zhang, CrystEngComm, 2014, 16, 47 RSC.
  145. H. M. Zhang, J. Yang, Y. C. He and J. F. Ma, Chem. – Asian J., 2013, 8, 2787 CrossRef CAS PubMed.
  146. Y. Chen, L. Li, J. Yang, S. Wang and J. Li, J. Solid State Chem., 2015, 226, 114 CrossRef CAS.
  147. G. P. Li, K. Zhang, H. Y. Zhao, L. Hou and Y. Y. Wang, ChemPlusChem, 2017, 82, 716 CrossRef CAS.
  148. Y. J. Wu, D. C. Hu, X. Q. Yao, Y. X. Yang and J. C. Liu, Inorg. Chim. Acta, 2016, 453, 488 CrossRef CAS.
  149. J. P. Lang, Q. F. Xu, R. X. Yuan and B. F. Abrahams, Angew. Chem., Int. Ed., 2004, 43, 4741 CrossRef CAS PubMed.
  150. Z. J. Zhang, W. Shi, Z. Niu, H. H. Li, B. Zhao, P. Cheng, D. Z. Liao and S. P. Yan, Chem. Commun., 2011, 47, 6425 RSC.
  151. S. Yao, X. Sun, B. Liu, R. Krishna, G. Li, Q. Huo and Y. Liu, J. Mater. Chem. A, 2016, 4, 15081 RSC.
  152. J. Yuan, J. Li, L. Kan, L. Zou, J. Zhao, D.-S. Li, G. Li, L. Zhang and Y. Liu, Cryst. Growth Des., 2018, 18, 5449 CrossRef CAS.
  153. L. Hashemi and A. Morsali, CrystEngComm, 2012, 14, 779 RSC.
  154. L. Hashemi and A. Morsali, CrystEngComm, 2012, 14, 8349 RSC.
  155. L. Hashemi and A. Morsali, CrystEngComm, 2014, 16, 4955 RSC.
  156. J. H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga and K. P. Lillerud, J. Am. Chem. Soc., 2008, 130, 13850 CrossRef PubMed.
  157. J. E. Mondloch, M. J. Katz, N. Planas, D. Semrouni, L. Gagliardi, J. T. Hupp and O. K. Farha, Chem. Commun., 2014, 50, 8944 RSC.
  158. B. Van de Voorde, I. Stassen, B. Bueken, F. Vermoortele, D. De Vos, R. Ameloot, J. C. Tan and T. D. Bennett, J. Mater. Chem. A, 2015, 3, 1737 RSC.
  159. R. J. Marshall, S. L. Griffin, C. Wilson and R. S. Forgan, Chem. – Eur. J., 2016, 22, 4870 CrossRef CAS PubMed.
  160. R. J. Marshall, S. L. Griffin, C. Wilson and R. S. Forgan, J. Am. Chem. Soc., 2015, 137, 9527 CrossRef CAS PubMed.
  161. 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 and M. Schröder, Nat. Chem., 2012, 4, 887 CrossRef CAS PubMed.
  162. J. Qian, F. Jiang, D. Yuan, M. Wu, S. Zhang, L. Zhang and M. Hong, Chem. Commun., 2012, 48, 9696 RSC.
  163. I. A. Ibarra, S. Yang, X. Lin, A. J. Blake, P. J. Rizkallah, H. Nowell, D. R. Allan, N. R. Champness, P. Hubberstey and M. Schröder, Chem. Commun., 2011, 47, 8304 RSC.
  164. X. Zhang, I. Silva, H. G. W. Godfrey, S. K. Callear, S. A. Sapchenko, Y. Cheng, I. Vitórica-Yrezábal, M. D. Frogley, G. Cinque, C. C. Tang, C. Giacobbe, C. Dejoie, S. Rudić, A. J. Ramirez-Cuesta, M. A. Denecke, S. Yang and M. Schröder, J. Am. Chem. Soc., 2017, 139, 16289 CrossRef CAS PubMed.
  165. V. Safarifard and A. Morsali, CrystEngComm, 2014, 16, 8660 RSC.
  166. S. Parshamoni, S. Sanda, H. S. Jena and S. Konar, Chem. – Asian J., 2015, 10, 653 CrossRef CAS.
  167. C. Falaise, C. Volkringer, J. Facqueur, T. Bousquet, L. Gasnot and T. Loiseau, Chem. Commun., 2013, 49, 10320 RSC.
  168. A. S. Munn, F. Millange, M. Frigoli, N. Guillou, C. Falaise, V. Stevenson, C. Volkringer, T. Loiseau, G. Cibin and R. I. Walton, CrystEngComm, 2016, 18, 8108 RSC.
  169. K. K. Yee, Y. L. Wong and Z. Xu, Dalton Trans., 2016, 45, 5334 RSC.
  170. Z. Wang, Y. Huang, J. Yang, Y. Li, Q. Zhuang and J. Gu, Dalton Trans., 2017, 46, 7412 RSC.
  171. F. Sun, Z. Yin, Q. Q. Wang, D. Sun, M. H. Zeng and M. Kurmoo, Angew. Chem., Int. Ed., 2013, 52, 4538 CrossRef CAS PubMed.
  172. H. J. Choi and M. P. Suh, J. Am. Chem. Soc., 2004, 126, 15844 CrossRef CAS PubMed.
  173. G. Agustí, R. Ohtani, K. Yoneda, A. B. Gaspar, M. Ohba, J. F. Sánchez-Royo, M. C. Muñoz, S. Kitagawa and J. A. Real, Angew. Chem., Int. Ed., 2009, 48, 8944 CrossRef PubMed.
  174. S. Horike, M. Sugimoto, K. Kongpatpanich, Y. Hijikata, M. Inukai, D. Umeyama, S. Kitao, M. Seto and S. Kitagawa, J. Mater. Chem. A, 2013, 1, 3675 RSC.
  175. R. Ohtani, K. Yoneda, S. Furukawa, N. Horike, S. Kitagawa, A. B. Gaspar, M. C. Munoz, J. A. Real and M. Ohba, J. Am. Chem. Soc., 2011, 133, 8600 CrossRef CAS PubMed.
  176. Q. K. Liu, J. P. Ma and Y. B. Dong, Chem. Commun., 2011, 47, 7185 RSC.
  177. L. Fu, Y. Liu, M. Pan, X. J. Kuang, C. Yan, K. Li, S. C. Wei and C. Y. Su, J. Mater. Chem. A, 2013, 1, 8575 RSC.
  178. J. He, J. Duan, H. Shi, J. Huang, J. Huang, L. Yu, M. Zeller, A. D. Hunter and Z. Xu, Inorg. Chem., 2014, 53, 6837 CrossRef CAS PubMed.
  179. M. Chebbi, B. Azambre, C. Volkringer and T. Loiseau, Microporous Mesoporous Mater., 2018, 259, 244 CrossRef CAS.
  180. B. Li, X. Dong, H. Wang, D. Ma, K. Tan, S. Jensen, B. J. Deibert, J. Butler, J. Cure, Z. Shi, T. Thonhauser, Y. J. Chabal, Y. Han and J. Li, Nat. Commun., 2017, 8, 485 CrossRef PubMed.
  181. B. Li, X. Dong, H. Wang, D. Ma, K. Tan, Z. Shi, Y. J. Chabal, Y. Han and J. Li, Faraday Discuss., 2017, 201, 47 RSC.
  182. K. W. Chapman, D. F. Sava, G. J. Halder, P. J. Chupas and T. M. Nenoff, J. Am. Chem. Soc., 2011, 133, 18583 CrossRef CAS PubMed.
  183. T. D. Bennett, P. J. Saines, D. A. Keen, J. C. Tan and A. K. Cheetham, Chem. – Eur. J., 2013, 19, 7049 CrossRef CAS PubMed.
  184. T. He, X. Xu, B. Ni, H. Lin, C. Li, W. Hu and X. Wang, Angew. Chem., Int. Ed., 2018, 57, 10148 CrossRef CAS PubMed.
  185. S. Khanjani and A. Morsali, CrystEngComm, 2012, 14, 8137 RSC.
  186. D. F. Sava, T. J. Garino and T. M. Nenoff, Ind. Eng. Chem. Res., 2012, 51, 614 CrossRef.
  187. B. Valizadeh, T. N. Nguyen, B. Smit and K. C. Stylianou, Adv. Funct. Mater., 2018, 28, 1801596 CrossRef.
  188. P. Mao, B. Qi, Y. Liu, L. Zhao, Y. Jiao, Y. Zhang, Z. Jiang, Q. Li, J. Wang, S. Chen and Y. Yang, J. Solid State Chem., 2016, 237, 274 CrossRef CAS.
  189. C. Horn, M. Scudder and I. Dance, CrystEngComm, 2001, 3, 1 RSC.
  190. C. Horn, B. Ali, I. Dance, M. Scudder and D. Craig, CrystEngComm, 2000, 2, 6 RSC.
  191. C. Horn, M. Scudder and I. Dance, CrystEngComm, 2000, 2, 53 RSC.
  192. C. Horn, M. Scudder and I. Dance, CrystEngComm, 2000, 2, 196 RSC.
  193. C. Horn, L. Berben, H. Chow, M. Scudder and I. Dance, CrystEngComm, 2002, 4, 7 RSC.
  194. C. Horn, M. Scudder and I. Dance, CrystEngComm, 2001, 3, 9 RSC.
  195. J. Y. Lu and A. M. Babb, Inorg. Chem., 2002, 41, 1339 CrossRef CAS PubMed.
  196. J. Y. Lu and V. Schauss, Eur. J. Inorg. Chem., 2002, 1945 CrossRef CAS.
  197. J. Y. Lu and A. M. Babb, Chem. Commun., 2003, 1346 RSC.
  198. B. Xin, G. Zeng, L. Gao, Y. Li, S. Xing, J. Hua, G. Li, Z. Shi and S. Feng, Dalton Trans., 2013, 42, 7562 RSC.
  199. Z. Yin, Q. X. Wang and M. H. Zeng, J. Am. Chem. Soc., 2012, 134, 4857 CrossRef CAS PubMed.
  200. H. Li, X. Wang, Y. Jia, B. Zhao, R. Ding, H. Hou and Y. Fan, CrystEngComm, 2012, 14, 5155 RSC.
  201. Y. C. He, J. Yang, G. C. Yang, W. Q. Kan and J. F. Ma, Chem. Commun., 2012, 48, 7859 RSC.
  202. Y. C. He, W. Q. Kan, J. Guo, Y. Yang, P. Du, Y. Y. Liu and J. F. Ma, CrystEngComm, 2013, 15, 7406 RSC.
  203. H. Kitagawa, H. Ohtsu and M. Kawano, Angew. Chem., Int. Ed., 2013, 52, 12395 CrossRef CAS PubMed.
  204. K. Biradha and M. Fujita, Angew. Chem., Int. Ed., 2002, 41, 3392 CrossRef CAS.
  205. G. Brunet, D. A. Safin, M. Z. Aghaji, K. Robeyns, I. Korobkov, T. K. Woo and M. Murugesu, Chem. Sci., 2017, 8, 3171 RSC.
  206. Y. Q. Hu, M. Q. Li, Y. Wang, T. Zhang, P. Q. Liao, Z. Zheng, X. M. Chen and Y. Z. Zheng, Chem. – Eur. J., 2017, 23, 8409 CrossRef CAS PubMed.
  207. J. S. Anderson, A. T. Gallagher, J. A. Mason and T. D. Harris, J. Am. Chem. Soc., 2014, 136, 16489 CrossRef CAS PubMed.
  208. O. Ohmori, M. Kawano and M. Fujita, Angew. Chem., Int. Ed., 2005, 44, 1962 CrossRef CAS PubMed.
  209. W. J. Gee, L. E. Hatcher, C. A. Cameron, C. Stubbs, M. R. Warren, A. D. Burrows and P. R. Raithby, Inorg. Chem., 2018, 57, 4959 CrossRef CAS PubMed.
  210. J. X. Lin, J. Liang, J. F. Feng, B. Karadeniz, J. Lü and R. Cao, Inorg. Chem. Front., 2016, 3, 1393 RSC.
  211. B. Assfour, T. Assaad and A. Odeh, Chem. Phys. Lett., 2014, 610–611, 45 CrossRef CAS.
  212. S. Chibani, F. Chiter, L. Cantrel and J.-F. Paul, J. Phys. Chem. C, 2017, 121, 25283 CrossRef CAS.
  213. D. F. S. Gallis, I. Ermanoski, J. A. Greathouse, K. W. Chapman and T. M. Nenoff, Ind. Eng. Chem. Res., 2017, 56, 2331 CrossRef.

This journal is © The Royal Society of Chemistry 2019