Chem Soc Rev View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. REVIEW ARTICLE Cite this: DOI: 10.1039/c5cs00837a View Journal Zr-based metal–organic frameworks: design, synthesis, structure, and applications Yan Bai,a Yibo Dou,a Lin-Hua Xie,a William Rutledge,b Jian-Rong Li*a and Hong-Cai Zhoub Among the large family of metal–organic frameworks (MOFs), Zr-based MOFs, which exhibit rich structure types, outstanding stability, intriguing properties and functions, are foreseen as one of the most promising MOF materials for practical applications. Although this specific type of MOF is still in its early stage of development, significant progress has been made in recent years. Herein, advances in Zr-MOFs since 2008 are summarized and reviewed from three aspects: design and synthesis, structure, and applications. Four synthesis strategies implemented in building and/or modifying Zr-MOFs as well as their scale-up preparation under green and industrially feasible conditions are illustrated first. Zr-MOFs with various structural types are Received 6th November 2015 then classified and discussed in terms of different Zr-based secondary building units and organic ligands. DOI: 10.1039/c5cs00837a Finally, applications of Zr-MOFs in catalysis, molecule adsorption and separation, drug delivery, and fluorescence sensing, and as porous carriers are highlighted. Such a review based on a specific type of MOF www.rsc.org/chemsocrev is expected to provide guidance for the in-depth investigation of MOFs towards practical applications. 1. Introduction As an emerging class of highly ordered crystalline porous materials, metal–organic frameworks (MOFs) with various potential a Beijing Key Laboratory for Green Catalysis and Separation and Department of Chemistry and Chemical Engineering, College of Environmental and Energy Engineering, Beijing University of Technology, Beijing 100124, P. R. China. E-mail: jrli@bjut.edu.cn b Department of Chemistry, Texas A&M University, College Station, Texas 77842-3012, USA applications have become a new research hotspot in chemistry and materials science during the last few decades, which is mainly attributed to their exceptionally high surface area, tunable pores, as well as intriguing functionalities.1–5 Many studies have demonstrated that MOFs perform excellently in various applications when compared to traditional porous solid materials, such as zeolites and carbon-based porous materials. However, most of these results are believed to be conceptual, because performances under ideal conditions and lack of economical evaluation do not guarantee the applicability of the materials. Unlike many industrially applied materials with Yan Bai obtained her BS degree in 2013 from the Shandong University of Technology. She is currently a graduate student at the Beijing University of Technology and pursues her study under the supervision of Prof. J.-R. Li. Her research focuses on the design, synthesis, and application of stable MOFs. Yan Bai This journal is © The Royal Society of Chemistry 2016 Yibo Dou Yibo Dou obtained his BS degree from the Beijing University of Chemical Technology (BUCT) in 2009, and then continued his postgraduate study under the supervision of Prof. X. Duan at BUCT. In 2013, as a visiting student, he studied at the University of Oxford under the supervision of Prof. D. O’Hare. After obtaining his PhD degree in 2015, he worked at the Beijing University of Technology. Currently, his main research topic is the design and fabrication of MOF-based composite functional materials. Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Chem Soc Rev uniform covalent bonds, MOFs are constructed by the coordination bonding of metal ions and organic ligands. The poor stability of MOFs derived from the reversible nature of coordination bonds is commonly regarded as the major drawback for their practical applications in most research fields.6–10 Much effort has been devoted to constructing MOFs with improved stability.11–22 Among them, post-treatments or postsynthetic modifications have proven effective.23–26 For example, IRMOF-1 (MOF-5) showed enhanced stability against moisture after thermal modification and formation of an amorphous carbon coating on the MOF particle surfaces, or after coating the MOF surface with a thin hydrophobic polydimethysiloxane (PDMS) layer via the thermal vapor deposition technique.25,26 However, the improved stability of the MOFs after posttreatment is always at the price of reduction of their micropore surface areas or functionality. In addition, most experimental steps or physical techniques in post-treatments are labor and resource expensive. On the other hand, much research has been invested to directly construct MOFs with inherent stability in structure and composition. A remarkable step forward to this end was the discovery of Zr6(m3-O)4(m3-OH)4(BDC)6 (UiO-66, UiO stands for the University of Oslo) with 12-coordinated Zr6(m3-O)4(m3-OH)4(CO2)12 clusters by Cavka et al.27 The structure of UiO-66 exhibits unprecedented stability, especially hydrothermal stability beyond most reported MOFs.28–31 From then on, the discovery of MOFs based on Zr(IV) ions (Zr-MOFs), mostly Zr(IV) carboxylates, ushered in new structural types being reported, diverse strategies being adopted to modify their structures and properties, and various functions and applications being explored (Fig. 1). It has been proposed that the stability of MOFs is governed by multiple factors, including the pKa of ligands, the oxidation state, reduction potential and ionic radius of metal ions, metalligand coordination geometry, hydrophobicity of the pore Lin-Hua Xie obtained his PhD in 2010 from Sun Yat-Sen University under the supervision of Prof. X.-M. Chen. After postdoctoral research at the Seoul National University in Prof. M. P. Suh’s group from 2010 to 2012, and at the King Abdullah University of Science and Technology in Prof. Z. Lai’s group from 2012 to 2015, he joined the Beijing University of Technology as an associate professor. His research interests include the functionalization and separation application of new porous materials. William Rutledge obtained his BS degree in 2014 from Millersville University, where he worked on iridium N-heterocyclic carbenebased transfer hydrogenation catalysts. In 2015, he joined Dr Hong-Cai Zhou’s group as a PhD student at Texas A&M University. His research interest is focused on the development of MOF materials for use in clean energy technologies. Lin-Hua Xie Jian-Rong ‘‘Jeff’’ Li obtained his PhD in 2005 from Nankai University under the supervision of X.-H. Bu. Until 2007, he was an assistant professor at the same University. From 2008 to 2009, he was a postdoctoral research associate, first at Miami University and later at Texas A&M University in Prof. H.-C. Zhou’s group; since 2010 he has been an assistant research scientist at the same university. Jian-Rong Li Since 2011, he has been a full professor at the Beijing University of Technology. His research interest focuses on porous materials for chemical engineering, energy and environmental science. Chem. Soc. Rev. William Rutledge Hong-Cai ‘‘Joe’’ Zhou obtained his PhD in 2000 from Texas A&M University under the supervision of F. A. Cotton. After a postdoctoral stint at Harvard University with R. H. Holm, he joined the faculty of Miami University, Oxford, in 2002. He moved back to Texas A&M University and became a full professor in 2008. He was promoted to a Davidson Professor of Science in 2014 and a Robert A. Welch Chair in Hong-Cai Zhou Chemistry in 2015. He was recognized as a Thomson Reuters ‘‘Highly Cited Researcher’’ in both 2014 and 2015. His research focuses on the discovery of synthetic methods to obtain robust framework materials with unique catalytic activities or desirable properties for clean-energy-related applications. This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Fig. 1 The year-by-year increase of reported Zr-MOFs in the last eight years (by SciFinder). Review Article determination, as well as scale-up preparation under green and industrially feasible conditions are illustrated. Then, Zr-MOFs with various compositions and structural types are classified and discussed in terms of different Zr-based clusters and organic ligands. Finally, the applications of Zr-MOFs in catalysis, molecule adsorption and separation, drug delivery, and fluorescence sensing, and as porous carriers are highlighted. Most of the reported Zr-MOFs and their general information are summarized in Table 1. We have tried our best to present a comprehensive and up-to-date generalization; however, this topic is highly active and extensively investigated. Therefore, some studies may be overlooked and some papers will emerge in the final stages of manuscript preparation/submission. Here we apologize in advance to the authors whose studies are related but missed in the review process. 2. Design and synthesis surface, etc.11,14,32,33 In particular, metal–ligand bond strength is believed to be crucial for determining the hydrothermal stability of MOFs, and it is determined by the nature of both the metal ion and the ligand. Thus, comparing the bonding strength of MOFs with different metals and ligands is not a straightforward affair. Nevertheless, it is well established that metal–ligand bond interactions in MOFs with a given ligand are greatly affected by the oxidation state and the ionic radius, or the charge density of the metal ions. One key feature of Zr-MOFs is the high oxidation state of Zr(IV) compared with M(I), M(II), and M(III)-based MOFs (M stands for metal elements). Due to high charge density and bond polarization, there is a strong affinity between Zr(IV) and carboxylate O atoms in most carboxylate-based Zr-MOFs.34 This is in line with Pearson’s hard/soft acid/base concept.35 Zr(IV) ions and carboxylate ligands are considered hard acid and hard base, respectively, and their coordination bonds are strong. As a result, most Zr-MOFs are stable in organic solvents and water, and even tolerable to acidic aqueous solution. However, it should be mentioned that reported Zr-MOFs are still less stable in basic aqueous solution, which can be well explained by natural bond orbital (NBO) theory.33,36 The NBO charge of O atoms in OH (1.403) is significantly larger than that of O atoms in the carboxylate group (0.74), which implies that OH can form a stronger bond with Zr(IV) than a carboxylate O atom, leading to the decomposition of Zr-MOFs under basic conditions. The improvement in the stability of MOFs can expand their applicable fields, and make the research in some topics such as catalysis, adsorption and separation, and biomedicine more reliable and meaningful.28,37–43 Zirconium is widely distributed in nature and is found in all biological systems. The rich content and low toxicity of Zr further favor the development and application of Zr-MOFs.44 As related studies have been increasing rapidly, a need to review recent research advances of Zr-MOFs has arisen. In this contribution, advances in Zr-MOFs since 2008 are summarized based on three aspects: design and synthesis, structure, and application. First, four synthesis strategies used in building and/or modifying Zr-MOFs, particularly for growing single crystals for X-ray diffraction structure This journal is © The Royal Society of Chemistry 2016 Due to the in situ formation of Zr-clusters and Zr–O coordination bonds, the structural predication of Zr-MOFs faces a tremendous challenge. Even so, some design strategies and synthetic methods have been developed, which speed up the discovery of new Zr-MOFs. In the following section, we have highlighted four effective synthetic strategies, which are frequently implemented to construct Zr-MOFs, including modulated synthesis, isoreticular expansion, topology-guided design, and postsynthetic modification. 2.1 Modulated synthesis Despite being a focus of attention, a significant synthetic issue existed in the early stages of Zr-MOF development. The crystallinity of synthesized Zr-MOFs was poor, and microcrystalline powder samples were commonly obtained before introducing the modulated synthesis strategy. This can be explained by the high charge density of Zr(IV) which polarizes the Zr–O bond to present covalent character, slowing down the ligand exchange reaction between Zr clusters and carboxylate ligands.34 In this case, it is unfavorable for defect repair during the crystallization process to form high quality crystals. In order to obtain big phase-pure single crystals, the modulated synthesis strategy was employed in Zr-MOF synthesis. Modulated synthesis refers to regulating the coordination equilibrium by introducing modulators with similar chemical functionality as the organic ligands used to hinder the coordination interaction between metal ions and the organic ligands.45 As a result, the competitive reaction can modulate the rate of nucleation and crystal growth. Simultaneously the selective regulation of coordination interactions in one of the coordination modes makes it possible to control the morphology of resulting crystals, leading to the anisotropic growth of MOF crystals.46 Moreover, it has been proven that the coordination modulation method can greatly improve the reproducibility of synthesis procedures and tune crystal features such as size, morphology, and crystallinity. In particular, it plays an important role in dominating the degree of agglomeration/aggregation of crystals. Chem. Soc. Rev. View Article Online Review Article Table 1 Summary of the most reported Zr-MOFs Zr-MOF Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev UiO-66 UiO-67 UiO-68 NU-1000 NU-1100 NU-1101 NU-1102 NU-1103 NU-1104 MOF-801 MOF-802 MOF-804 MOF-805 MOF-806 MOF-808 MOF-812 MOF-841 MOF-525 MOF-535 MOF-545 MOF-867 PCN-56 PCN-57 PCN-58 PCN-59 PCN-94 PCN-221 PCN-222 PCN-223 PCN-224 PCN-225 PCN-228 PCN-229 PCN-230 PCN-521 PCN-700 PCN-777 DUT-51 DUT-52 DUT-84 DUT-67 DUT-68 DUT-69 BUT-10 BUT-11 BUT-30 MIL-140A MIL-140B MIL-140C MIL-140D MIL-153 MIL-154 MIL-163 CPM-99 UMCM-309a Zr-TTMC Zr-BTB Zr-ABDC PIZOF BPV-MOF mBPV-MOF mPT-MOF ZrBTBP UPG-1 UiO-66(AN) UiO-66-1,4-Naph Zr-ADC Zr-DTDC Zr-TCPS Zr-BTDC Chem. Soc. Rev. Liganda 2 BDC BPDC2 TPDC2 TBAPy4 PTBA4 Py-XP4 Por-PP4 Py-PTP4 Por-PTP4 FUM2 PZDC2 BDC-2OH2 NDC-2OH2 BPDC-2OH2 BTC3 MTB4 MTB4 TCPP4 XF4 TCPP4 BPYDC2 TPDC-2CH32 TPDC-4CH32 TPDC-2CH2N32 TPDC-4CH2N32 ETTC4 TCPP4 TCPP4 TCPP4 TCPP4 TCPP4 TCP-14 TCP-24 TCP-34 MTBC4 Me2BPDC2 TATB3 DTTDC2 2,6-NDC2 2,6-NDC2 TDC2 TDC2 TDC2 FDCA2 DTDAO2 EDDB2 BDC2 2,6-NDC2 BPDC2 Cl2ABDC2 pgal3 Hgal3, Hsal TzGal6 TCBPP4 BTB3 TTMC2 BTB3 ABDC2 PEDC2 BPV2 BPHV2, BPV2 PT2, TPHN2 H3BTBP3 H4TTBMP2 AN2 1,4-NDC2 ADC2 DTDC2 TCPS4 BTDC2- Zr cluster/core Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-OH)8 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)8 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr8(m4-O)6 Zr6(m3-OH)8 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-OH)8 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)6(m3-OH)2 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)8 Zr6(m3-O)6(m3-OH)2 Zr6(m3-O)6(m3-OH)2 Zr6(m3-O)4(m3-OH)42 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr(m3-O)3O4 Zr(m3-O)3O4 Zr(m3-O)3O4 Zr(m3-O)3O4 ZrO8 ZrO8 ZrO8 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 ZrO6 ZrO6 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Topology fcu, 12-connected fcu, 12-connected fcu, 12-connected csq, (4,8)-connected ftw, (4,12)-connected ftw, (4,12)-connected ftw, (4,12)-connected ftw, (4,12)-connected ftw, (4,12)-connected fcu, 12-connected bct, 10-connected, fcu, 12-connected fcu, 12-connected fcu, 12-connected spn, (3,6)-connected ith, (4,12)-connected flu, (4,8)-connected ftw, (4,12)-connected ftw, (4,12)-connected csq, (4,8)-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected ftw, (4,12)-connected ftw, (4,12)-connected csq, (4,8)-connected shp, (4,12)-connected she, (4,6)-connected sqc, (4,8)-connected ftw, (4,12)-connected ftw, (4,12)-connected ftw, (4,12)-connected flu, (4,8)-connected bcu, 8-connected spn, (3,6)-connected reo, 8-connected fcu, 12-connected (4,4)IIb, 6-connected reo, 8-connected 8-connected bct, 10-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected — — — — — — — ftw, (4,12)-connected kgd, (3,6)-connected fcu, 12-connected kgd, (3,6)-connectedc fcu, 12-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected — — fcu, 12-connected fcu, 12-connected fcu, 12-connected fcu, 12-connected flu, (4,8)-connected fcu, 12-connected BET surface area (m2 g1) b 1187 3000b 4170b 2320 4020 4422 4712 5646 5290 990 o20 1145 1230 2220 2060 2335 1390 2620 1120 2260 2646 3741 2572 2185 1279 3377 1936 2223 1600 2600 1902 4510 4619 4455 3411 1807 2008 2335 1399 637 1064 891 560 1848 1310 3940.6 415 460 670 701 — — 90–170 1030 810 705 613 3000 2080 373 1207 3834 o10 410 627 — 2776 1200 1039 2207 Ref. 27 27 27 121 83 63 63 63 63 155 155 155 155 155 155 155 155 159 159 159 282 and 154 97 97 97 97 163 135 144 143 146 145 65 65 65 34 89 147 49 149 149 50 50 50 205 205 214 139 139 139 139 141 141 142 161 51 54 157 72 73 100 100 100 136 137 150 148 151 152 36 153 This journal is © The Royal Society of Chemistry 2016 View Article Online Chem Soc Rev Table 1 Zr-MOF Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Zr-BTBA Zr-PTBA MMPF-6 Zr-AP-1 Zr-AP-2 Zr-AP-3 Review Article (continued) Liganda 4 BTBA PTBA4 TCPP4 AP-CH32 AP2 AP2 Zr cluster/core Topology BET surface area (m2 g1) Ref. Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)4(m3-OH)4 Zr6(m3-O)8 Zr6(m3-OH)8 Zr6(m3-OH)8 Zr6(m3-OH)8 ftw, (4,12)-connected ftw, (4,12)-connected csq, (4,8)-connected bcu, 8-connected dia, 4-connected bcu, 8-connected 4342 4116 2100 — — — 162 162 160 156 156 156 a Ligands (see their structures in Scheme 1) are abbreviated as: BDC2 = terephthalate; BPDC2 = biphenyl-4,4 0 -dicarboxylate; TPDC2 = [1,1 0 :4 0 ,100 terphenyl]-4,400 -dicarboxylate; TBAPy4 = 1,3,6,8-tetrakis(p-benzoate)pyrene; Py-XP4 = 4 0 ,4 0 0 0 ,4 0 0 0 0 0 ,4 0 0 0 0 0 0 0 -(pyrene-1,3,6,8-tetrayl) tetrakis(2 0 ,5 0 Por-PP4 = meso-tetrakis-(4-carboxylatebiphenyl)-porphyrin; PTBA4 = 4-[2-[3,6,8-tris[2-(4dimethyl-[1,1 0 -biphenyl]-4-carboxylate); carboxylatephenyl)-ethynyl]-pyren-1-yl]ethynyl]-benzoate; Py-PTP4 = 4,4 0 ,400 ,4 0 0 0 -((pyrene-1,3,6,8-tetrayltetrakis(benzene-4,1-diyl))tetrakis(ethyne4 2 2 = meso-tetrakis-(4-((phenyl)ethynyl)benzoate)porphyrin; FUM = fumarate; PZDC = 1H-pyrazole-3,52,1-diyl))tetrabenzoate; Por-PTP dicarboxylate; BTC3 = benzene-1,3,5-tricarboxylate; MTB4 = 4,4 0 ,400 ,4 0 0 0 -methanetetrayltetrabenzoate; TCPP4 = meso-tetrakis(4-carboxylatephenyl)porphyrin; XF4 = 4,4 0 -((1E,10 E)-(2,5-bis((4-carboxylatephenyl)ethynyl)-1,4-phenylene)bis(ethene-2,1-diyl))dibenzoate; BPYDC2 = 2,20 -bipyridine5,5 0 -dicarboxylate; ABDC2 = 4,4-azobenzenedicarboxylate; TCBPP4 = tetrakis(4-carboxylatebiphenyl)porphyrin; ETTC4 = 40 ,400 ,40 0 0 ,40 0 0 0 -(ethene-1,1,2,2tetrayl)tetrabiphenyl-4-carboxylate; TDC2 = 2,5-thiophenedicarboxylate; MTBC4 = 4 0 ,400 ,40 0 0 ,4 0 0 0 0 -methanetetrayltetrabiphenyl-4-carboxylate; TATB3 = 4,4 0 ,400 -s-triazine-2,4,6-triyl-tribenzoate; DTTDC2 = dithieno[3,2-b;2 0 ,30 -d]-thiophene-2,6-dicarboxylate; 2,6-NDC2 = naphthalene-2,6-dicarboxylate; FDCA2 = 9-fluorenone-2,7-dicarboxylate; DTDAO2 = dibenzo[b,d]thiophene-3,7-dicarboxylate 5,5-dioxide; EDDB2 = 4,4 0 -(ethyne-1,2-diyl)dibenzoate; H3pgal = pyrogallol; H4gal = gallic acid; H2sal = salicylic acid; H6TzGal = 5,5 0 -(1,2,4,5-tetrazine-3,6-diyl)bis(benzene-1,2,3-triol); BTB3 = 5 0 -(4carboxyphenyl)[1,1 0 :300 ,100 -terphenyl]-4,400 -dicarboxylate; TTMC2 = (2E,4E)-hexa-2,4-dienedioate; ABDC2 = azobenzenedicarboxylate; H6BTBP = 1,3,5-tris(4-phosphonophenyl)benzene; H6TTBMP = 2,4,6-tris(4-(phosphonomethyl)phenyl)-1,3,5-triazine; PEDC2 = 4,4 0 -(1,4-phenylenebis(ethyne-2,1-diyl))dibenzoate; BPV2 = 5,5 0 -bis(carboxylateethenyl)-2,2 0 -bipyridine; BPHV2 = 4,4 0 -bis(carboxylateethenyl)-1,1 0 -biphenyl; TPHN2 = 4,4 0 -bis(carboxylatephenyl)-2-nitro-1,1 0 -biphenyl; ADC2 = 9,10-anthacenyl bis(benzoate); DTDC2 = 3,4-dimethylthieno[2,3-b]thiophene-2,5dicarboxylate; TCPS4 = tetrakis(4-carboxyphenyl) silane; BTBA4 = 4,4 0 ,400 ,4 0 0 0 -(biphenyl-3,3 0 ,5,5 0 -tetrayltetrakis(ethyne-2,1-diyl))tetrabenzoate; AP2 = 1,6-adipate. b Langmuir surface area. c The structure of Zr-BTB is constructed from two-dimensional (2D) - three-dimensional (3D) interpenetration based on a 3,6-connected kgd net. The first example of applying the modulated synthesis strategy to prepare Zr-MOFs was reported by Schaate et al. in 2011.47 In this work they studied the effects of modulator benzoic acid (HBC), acetic acid (AcOH) and water on the formation of Zr-BDC (UiO-66), Zr-BDC-NH2, Zr-BPDC (UiO-67), and Zr-TPDC-NH2 (UiO-68-NH2) crystals. It was found that the synthesis of UiO-66 could be modulated by varying the amount of HBC. As the concentration of modulator increased, intergrown products of UiO-66 turned into individual and bigger crystals. Similarly, the addition of HBC could also tune the crystal size and morphology of Zr-BPDC and make the products highly reproducible. In addition, Schaate et al. also proved that water was crucial for the formation of Zr-BDC-NH2 crystals. By this synthetic control, they obtained large single crystals of Zr-TPDC-NH2 and determined its structure by the single-crystal X-ray diffraction technique, which represented the first single-crystal structure of a Zr-MOF. Following this, Zr6(m3-O)4(m3-OH)4(FUM)6 (Zr-FUM) was prepared by Wißmann et al. with formic acid as a modulator and fumarate as a linker.48 Different from the modulators in the previous work which slowed the crystallization rate of the Zr-MOFs, it was found that formic acid accelerated the formation of Zr-FUM crystals. A possible reason suggested by the authors is that formic acid is a direct product of the decomposition of the solvent N,N 0 -dimethylformamide (DMF) with water, and the presence of formic acid affects the reaction equilibrium and water content in the reaction system. In situ formation of coordination complexes between the Zr(IV) cation and monotopic carboxylic acid modulators was proposed as the reason why bigger crystals can be obtained with an increasing amount of modulator. The structures of these complexes should be similar to secondary building units (SBUs) in Zr-MOFs.47 These complexes could thus serve as intermediates in the construction of This journal is © The Royal Society of Chemistry 2016 final Zr-MOFs, which are formed through an exchange reaction between the modulators and used bridging ligands. So an increasing number of modulators would decrease the chances of dicarboxylic acid bridging with the intermediate and consequently inhibit the formation of nuclei and nuclei growth, leading to the formation of bigger MOF crystals. Encouraged by previous studies, Bon et al. used bent-type dicarboxylate organic linkers to synthesize a Zr-MOF in the presence of HBC as a modulator.49 A new structure Zr6(m3-O)6(m3-OH)2(DTTDC)4(BC)2(DMF)6 (DUT-51) with large pores was obtained. In this work the modulator also behaved as a structure-directing agent and decreased the connectivity of SBUs from twelve to eight in the final MOF. Later, with AcOH as a modulator, they synthesized three new Zr-MOFs, Zr6(m3-O)6(m3-OH)2(TDC)4(AcO)2 (DUT-67), Zr6(m3-O)6(m3-OH)2(TDC)4.5(AcO) (DUT-68), and Zr6(m3-O)4(m3-OH)4(TDC)5(AcO)2 (DUT-69).50 Recently, Ma et al. reported a 6-connected Zr-MOF, Zr6(m3-O)4(m3-OH)4(BTB)2(OH)6(H2O)6 (UMCM-309a), with H3BTB as the linker and HCl as a modulator.51 UMCM-309a features a stable two-dimensional (2D) layered network and the structure remained intact for over 4 months in aqueous HCl (1 M). Similarly, with HBC and biphenyl-4-carboxylic acid as modulators, UMCM-309b and UMCM-309c were synthesized with layered structures as in UMCM-309a but had different inter-layer distances, because these monocarboxylic acids substituted free hydroxyl sites in the Zr6 clusters and pointed to the space between layers. Consequently, when the longer modulator biphenyl-4-carboxylic acid was used, the layer space (14.8 Å) in UMCM-309c was greatly stretched compared with that (7.04 Å) in UMCM-309a. Moreover, it was found that after the removal of the modulator, the parent UMCM-309a was obtained. The modulators thus played another role of tuning the whole crystal structure of resulting Zr-MOFs. Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Early MOFs synthesized from the direct self-assembly of discrete metal ions and organic ligands were often unstable after the removal of guest molecules. Metal clusters were then introduced into MOF synthesis to improve the stability of resulting frameworks.52 Some pre-built polynuclear coordination complexes, having identical structures and functions with SBUs of target MOFs, were also utilized as so-called ‘‘precursors’’ to construct MOFs.53 Guillerm et al. used a known complex Zr6(m3-O)4(m3-OH)4(OMc)12 (OMc = methacrylate) as a precursor to react with dicarboxylic acid H2TTMC and synthesized Zr6(m3-O)4(m3-OH)4(TTMC)6 with the same topology of UiO-66 (Fig. 2).54 During the reaction process, the OMc ligands were substituted by the dicarboxylate ligands, which may be also regarded as a modulated synthesis. Rationally selecting modulators could also control the number and nature of defect sites of Zr-MOFs, which have been applied to tune the pore properties of Zr-MOFs for adsorption applications, as well as achieving open metal sites for catalysis. By varying the concentration of modulator AcOH, Wu et al. demonstrated that linker vacancies could be tuned systematically to achieve pore volume modulation ranging from 0.44 to 1.0 cm3 g1, and Brunauer–Emmett–Teller (BET) surface areas changing from 1000 to 1600 cm3 g1 of the obtained UiO-66 with different concentrations of linker vacancies.55 On the other hand, Vermoortele et al. used trifluoroacetic acid (TFA) as a modulator and HCl as a crystallizing agent for the synthesis of UiO-66 with ligand defects and open Zr sites.56 The resulting UiO-66 showed high catalytic activity for Lewis acid catalytic reactions. It was also suggested that HCl in this system displayed a double duty of slowing down the hydrolysis of ZrCl4 and counteracting the deprotonation of involved carboxylic acids. Consequently, the presence of HCl favored the incorporation of TFA in the final sample of UiO-66. Through thermal activation, a defected UiO-66 with a large number of open Zr sites was thus generated after the removal of TFA. 2.2 Isoreticular expansion Two of the most desirable characteristics of MOFs are their permanent porosity and ultrahigh surface areas. The realization of MOFs with such characteristics, in turn, can significantly advance the development of their applications such as in gas sorption and storage.57–59 Isoreticular expansion is a strategy for enlarging pore size, increasing surface areas, and/or tuning pore surface functionality for MOFs by propagating the structure of a prototypical MOF structure. This particular method takes advantage of experimental conditions of the original MOF to synthesize new isoreticular frameworks, however with desired pore surface, larger porosity and/or higher surface areas than the prototypical structure through modifying the linkers. In this manner, predesigned building blocks are purposefully incorporated into extended networks with targeted specific features.60–62 One of the most successful examples using the isoreticular expansion strategy is the synthesis of a series of Zr-MOFs, Zr6(m3-O)4(m3-OH)4(L1–4)3(NU-1101–1104, L1 = Py-XP4, L2 = Por-PP4, L3 = Py-PTP4 and L4 = PorPTP4) by Wang et al. (Fig. 3).63 Among these, Py-PTP4 and Por-PTP4 were obtained Chem. Soc. Rev. Chem Soc Rev Fig. 2 Schematic representation of the modulated synthesis of Zr6(m3O)4(m3-OH)4(TTMC)6 (right) starting from the precursor Zr6(m3-O)4(m3OH)4(OMc)12, in which the OMc ligands were replaced by TTMC2 (left). Reprinted with permission from ref. 54. Copyright 2010 The Royal Society of Chemistry. through inserting a triple bond in the arms of original tetratopic ligands Py-XP4 and Por-PP4, respectively. The geometric surface areas of NU-1101 to NU-1104 calculated by using a rolling probe method independent of BET theory64 are 4422, 4712, 5646, and 5290 m2 g1, respectively. It is worth mentioning that NU-1103 with Py-PTP4 as the linker has a pore volume of 2.91 cm3 g1 and a BET surface area of 6650 m2 g1, being the highest value reported to date for Zr-MOFs. Such high surface areas render them attractive for energy gas storage, such as H2 and CH4. Similarly, Liu et al. inserted a benzene ring, and one and two triple-bond spacers in the arms of the ligand to synthesize a series of extended porphyrinic linkers, H4TCP-1, H4TCP-2, and H4TCP-3.65 Using them, a series of Zr6(m3-O)4(m3-OH)4(TCP)4 from PCN-228 to -230 were obtained, which have BET surface areas of 4510, 4619, and 4455 m2 g1, and pore sizes of 25, 28, and 38 Å, respectively. Note that although the surface area of MOFs can be increased with the extension of the ligand length, the framework stability is usually inversely correlated. Some MOFs synthesized by isoreticular expansion with increased surface areas became unstable.66–70 For example, the Langmuir surface areas of UiO-66 and -67 are 1187 and 3000 m2 g1, respectively. However, the stability became worse for UiO-67.29 Another phenomenon often encountered in isoreticular expansion is network interpenetration, especially when using a long ligand with small steric hindrance.71 Interpenetration of networks always leads to a large reduction of surface area and pore volume of the MOFs, which is not favorable for some applications, such as gas storage, separation and catalysis where large molecules are involved. For example, Peter Behrens’s group synthesized three Zr-MOFs with three different ligands in length. Among them, Zr-FUM was isostructural to UiO-66 with a BET surface area of 856 m2 g1.48 Due to the short length of the FUM2 ligand, it is not surprising that the surface area is smaller than that of UiO-66 (1069 m2 g1). Later, they reported another isostructural Zr-MOF, Zr6(m3-O)4(m3-OH)4(ABDC)6 (Zr-ABDC), with a high surface area of 3000 m2 g1 and a large pore volume of 1.41 cm3 g1.72 With an even more elongated triphenylethynyl dicarboxylate ligand PEDC-2OMe2, Zr6(m3-O)4(m3-OH)4(PEDC2OMe)4 (PIZOF-2) with a 2-fold interpenetrated structure of the same topology was obtained.73 As a result, PIZOF-2 showed This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article Fig. 3 Tetratopic organic linkers of Py-XP4, Por-PP4, Py-PTP4, and Por-PTP4 used for constructing NU-1101–1104 (top). The increasing dimensions of the organic linkers result in the isoreticular expansion of the MOF structures, with surface areas of 4422, 4712, 5646, and 5290 m2 g1, respectively (bottom). Reprinted with permission from ref. 63. Copyright 2015 American Chemical Society. a BET surface area of about 1250 m2 g1, significantly lower than Zr-ABDC (3000 m2 g1) regardless of its much longer ligand. 2.3 Topology-guided design In order to avoid interpenetration in MOFs, some methods such as changing synthesis conditions and increasing steric hindrance of ligands have been adopted.71,74–77 However, the precise control of these synthetic factors is difficult. Take into consideration that interpenetration probability is closely related to the framework topology. As a result, much attention has been directed on MOF topological studies.78–81 Topologyguided design strategies were thus applied in the design and synthesis of MOFs, usually based on the ‘‘non-interpenetrated topological structure’’ as the ‘‘template’’.61 By this method, Zr-MOFs with preconceived structures and desirable properties have been designed with high synthetic accessibility. Here, for clarity, some representative topological networks appearing in Zr-MOFs are shown in Fig. 4. Apparently, the topology-guided design of MOFs depends on the judicious selection of metal-containing building units and organic ligands. At the same time, the symmetrical complementary is applied as the selection criteria for the synthesis of the desired network as well. In this regard, Gomez-Gualdron et al. demonstrated that Zr-MOFs with csq, scu, or ftw topologies can be constructed with a square-shaped tetratopic ligand and a Zr6 cluster, and the ftw topology Zr-MOF exhibits the highest surface area and has the lowest propensity for structure catenation as well.82 An ideal ftw Zr-MOF crystallizes in the Pm3% m space group of the cubic crystal system, where the cube vertices are occupied by 12-connected [Zr6(m3-O)4(m3-OH)4]12+ clusters, which are connected by planar tetratopic linkers across faces of the cubic unit cell. It has been found that in order to achieve such a MOF, two design points should be considered in the selection of This journal is © The Royal Society of Chemistry 2016 organic linkers: (1) the ligand should display planar geometry and (2) the relative dimensions of the linker should meet the requirement of four carboxylate groups being in a rectangular shape (close to a square).83 Hereafter, based on these selection criteria Gutov et al. designed and incorporated a pyrene-based elongated tetratopic ligand PTBA4 into a desired ftw net to obtain Zr6(m3-O)4(m3-OH)4(PTBA)3 (NU-1100).83 This MOF features two types of pores: (1) large pores located in the center of the cubic unit cell and (2) small ones located at the edges of the cubic unit cell. It is clear that a fully occupied 12-connected Zr6(m3-O)4(m3-OH)4(CO2)12 cluster has the Oh symmetry and TCPP4 is a 4-connected linker with the D4h symmetry. These two nodes are connected to each other and thus theoretically should lead to an ftw network. However, the topology-guided design strategy failed to realize the target for this system. The reason is that the spatial orientation of the ligand is as important as its connectivity and symmetry for construction of MOFs with specific topologies. In fact, when the 12-connected Zr6 clusters and TCPP4 ligands substitute corresponding nodes in the ftw network, the peripheral phenyl rings have no choice but to revolve to the same plane with the porphyrin center. However, this conformation is energically unfavorable due to steric hindrance and thus was difficult to realize. To solve this problem, Liu et al. introduced triple-bond spacers between adjacent phenyl rings and carboxylate groups to realize the controllable conformation.65 As a result, a series of Zr-MOFs, from PCN-228 to -230, were successfully prepared by using these ligands. The triple-bond spacers in this system not only elongate the ligands but also alleviate the steric constrain, which permits the free rotation of peripheral benzoates to adopt a compatible direction for the Zr6 clusters in the ftw net. Besides the ftw topology, the default and high symmetry edge transitive network of the fluorite (flu) topology84–86 is also Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Chem Soc Rev Fig. 5 (a) Schematic representation of the fluorite structure where the F anions fill all the tetrahedral interstitial cavities; (b) the unoccupied octahedral interstitial cavities (turquoise) in the fluorite structure; (c) the structure of PCN-521 with the flu topology in which all the tetrahedral interstitial cavities are occupied by the MTBC4 ligands; and (d) the representation of an unoccupied octahedral cavity in PCN-521 with the size of 20.5 20.5 37.4 Å. Reprinted with permission from ref. 34. Copyright 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. In addition, the combination of a 12-connected Zr6 cluster and a linear linker usually gives rise to the fcu topology. However, if the two carboxylate groups of the linear ditopic linker are noncoplanar, it is possible that other topological nets form, such as a bcu net. Following this topology-guided design strategy, Yuan et al. introduced two methyl groups into the 2- and 2 0 -positions of the BPDC2 ligand, leading to two carboxylate groups of the resulting Me2BPDC2 ligand with an off-plane state.89 With this ligand, a new Zr-MOF, Zr6(m3-O)4(m3-OH)4(OH)4(H2O)4(Me2-BPDC)4 (PCN-700), with a bcu network structure was obtained. 2.4 Fig. 4 Representative network topologies in reported Zr-MOFs. intriguing for constructing Zr-MOFs with high permanent porosity. The flu structure is conceived as cubic close packing of Ca(II), in which the tetrahedral cavities are filled with F anions, leaving the octahedral cavities unoccupied. For the flu topology, it is impossible to translate in all directions without overlapping with itself, which is favorable for avoiding interpenetration.87,88 Based on these considerations, Zhang et al. designed and synthesized Zr6(m3-OH)8(OH)8(MTBC)2 (PCN-521) by using a tetrahedral ligand MTBC4.34 The structure of PCN-521 consists of cubic close packing of 8-connected Zr6(m3-OH)8 clusters, in which all the tetrahedral interstitial cavities are occupied by the MTBC4 ligands (Fig. 5). PCN-521 displayed a BET surface area of 3411 m2 g1, a pore size of 20.5 20.5 37.4 Å, and a void volume of 78.5%. Chem. Soc. Rev. Postsynthetic functionalization Similar to conventional linker pre-functionalization, the functional groups of organic linkers and metal clusters in Zr-MOFs are also tailorable based on postsynthetic functionalization. It has been proved that the postsynthetic method is a versatile tool to prepare topologically identical Zr-MOFs with diverse functionalities and uncompromised structural stability.90,91 Covalent modification is the most common route of postsynthetic functionalization that endows MOFs with tailor-made internal surfaces to meet the requirements for specific applications.92–96 Jiang et al. reported the preparation and covalent modifications of a series of highly stable isoreticular Zr-MOFs, Zr3O2(OH)2(TPDC-R)3 (PCN-56, R = 2CH3, PCN-57, R = 4CH3; PCN-58, R = 2CH2N3, and PCN-59, R = 4CH2N3).97 In PCN-58 and -59 the accessible and reactive azide groups enable the obtained MOFs to undergo a quantitative click reaction98 with alkynes to form various triazole groups containing substituents This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Fig. 6 Schematic representation of the click reaction for the pore surface engineering of PCN-58–59 with precise control over the composition, density, and functionality by two steps. Reprinted with permission from ref. 97. Copyright 2012 American Chemical Society. on their pore surfaces (Fig. 6). It was revealed that the loading of azide groups can be accurately tuned by varying the ratio of ligands with and without azide groups during synthesis. Thus, a variety of functional groups were anchored onto the pore walls of obtained MOFs with precise control over loading, density, and functionality. Therefore, the resulting materials showed tunable CO2 selective adsorption ability over N2, being potentially useful in CO2 capture. Besides covalent modification, coordination modification of ligands in postsynthetic metalation (PSM) was also adopted. Toyao et al. reported the coordination modification in free chelating sites of 2,2 0 -bipyridine in Zr6(m3-O)4(m3-OH)4(BPYDC)6 (Zr-MOF-BPYDC) with CuBr2.99 It was found that after loading B3.3 wt% of CuBr2 to the MOF, 20.6% of BPYDC units could be metalated with Cu(II). Another typical example was reported by Manna et al. In their work, a series of bipyridyl- and phenanthrylbased Zr-MOFs, Zr6(m3-O)4(m3-OH)4(BPHV)6 (BPHV-MOF), Zr6(m3-O)4(m3-OH)4(BPHV)4(BPV)2 (mBPV-MOF), and Zr6(m3-O)4(m3-OH)4(TPHN)4(PT)2 (mPT-MOF), were prepared.100 Treatment of these Zr-MOFs with different equivalences of [Ir(COD)(OMe)]2 in tetrahydrofuran (THF) afforded Ir-functionalized Zr-MOFs, BPV-MOF-Ir, mBPV-MOF-Ir, and mPT-MOF-Ir. Inductively coupled plasma-mass spectroscopy (ICP-MS) analyses of the Ir/Zr ratio revealed that the Ir loadings could reach 65%, 16%, and 20% for the three MOFs, respectively. Significantly, these Ir-functionalized Zr-MOFs showed high catalytic activity for tandem hydrosilylation of aryl ketones and aldehydes. Additionally, based on the postsynthetic metalation, Manna et al. also treated Zr6(m3-O)4(m3-OH)4(BPYDC)6 (BPY-UiO) with [Ir(COD)(OMe)]2 in THF to obtain BPY-UiO-Ir, and with [Pd(CH3CN)4][BF4]2 in dimethyl sulfoxide (DMSO) to obtain BPYUiO-Pd, in which Zr/Ir in BPY-UiO-Ir and Zr/Pd ratios in BPYUiO-Pd reached 30 and 24%, respectively.101 This journal is © The Royal Society of Chemistry 2016 Review Article Amine-tagged MOFs can also be used to graft functional groups.102–104 After the covalent modifications of amine groups, the resulting MOFs are also well suited for the coordination modification. Rasero-Almansa et al. treated UiO-66-NH2 with aldehyde in dichloromethane (CH2Cl2), followed by another treatment with [IrCl(COD)]2, leading to quantitative conversion of chelating Ir-functionalized Zr-MOF materials.105 Obviously, the presence of functional group ‘‘tags’’ in parent frameworks is essential for the subsequent covalent and coordination modifications. It should be noted that not all Zr-MOFs contain functional group ‘‘tags’’, or are endurable for covalent and coordination modifications. In such cases, postsynthetic ligand exchange (PSE) of parent MOFs is exploited as an alternative means to introduce functionalized ligands into MOFs under mild conditions. Han et al. have reviewed such substitution/exchange reactions occurring in MOFs and metal–organic polyhedra (MOPs) in detail.106 In this review, a strategy combining PSE and PSM is highlighted. Based on this strategy, the linkers without functional groups can be exchanged by those with functional groups in Zr-MOFs, and then the resulting MOFs are metallated. As a result, various combinations of chelators and metal ions can be introduced into the MOFs. Catechol is one of the most common metal chelators applied in synthesizing coordination complexes. Fei et al. obtained Zr6(m3-O)4(m3-OH)4(CAT-BDC)6 (UiO-66-CAT, CAT-H2BDC = 2,3-dihydroxyterephthalic acid) via the PSE by exposing UiO66 into DMF/water solution of CAT-H2BDC.107,108 The highest loading value of CAT-BDC could be up to 90% after two rounds of the PSE process. UiO-66-FeCAT and UiO-66-CrCAT were then obtained by metalizing UiO-66-CAT with Fe(ClO4)3 and K2CrO4, respectively. The atomic ratios confirmed by energy-dispersed X-ray spectroscopy (EDX) were 1.00 : 0.21 (Zr : Fe) and 1.00 : 0.04 (Zr : Cr) for UiO-66-FeCAT and UiO-66-CrCAT, respectively. In addition, they also replaced the ligand in UiO-66 with thiocatecholate, which thus provided the metal-chelating motif for PSM. As a result, UiO-66-PdTCAT (TCAT-H2BDC = 2,3-dimercaptoterephthalic acid) was obtained after the PSE and PSM processes.108 This functionalized Zr-MOF was demonstrated to be a recyclable catalyst for the regioselective functionalization of sp2 C–H bonds. Incorporating multiple functionalities with synergistic effects within one MOF is particularly interesting for its unique properties and specific applications.97,109–112 Different from the conventional one-pot synthesis approach with mixed linkers, Yuan et al. developed sequential ligand installation (SLI) as a controllable strategy to construct multivariate MOFs (MTV-MOFs), which could precisely arrange the positions of functional groups (Fig. 7).89 In this study, a prototype Zr-MOF, PCN-700, built from 8-connected Zr6(m3-O)4(m3-OH)4(OH)4(H2O)4 clusters served as the platform. After removing terminal OH/H2O ligands on the adjacent Zr6 clusters, two types of natural ‘‘pockets’’ with different sizes were formed, 16.4 Å for pocket A and 7.0 Å for pocket B, which allowed for sequential installation of two types of linkers with different lengths. Then, BDC2 and TPDC-Me22 ligands were selected based on the pocket size. Next, the SLI process was performed: PCN-700 crystals were exposed to solutions of H2BDC and Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Chem Soc Rev Fig. 8 Schematic representation of Zr(IV) substitution in UiO-66 by Ti(IV) and Hf(IV) for the synthesis of UiO-66(Zr/Ti) and UiO-66(Zr/Hf).117 Fig. 7 Schematic representation for synthesis of PCN-701–703 based on the sequential ligand installation in PCN-700.89 H2TPDC-Me2 in DMF at enhanced temperature, respectively, resulting in mixed-ligand Zr-MOFs, Zr6(m3-O)4(m3-OH)2(OH)4(BPDC-Me2)4(BDC) (PCN-701) and Zr6(m3-O)4(m3-OH)2(OH)4(BPDC-Me2)4(TPDC-Me2) (PCN-702). Then the resultant PCN-701 was soaked in a DMF solution of H2TPDC-Me2 sequentially to obtain Zr6(m3-O)5(m3-OH)3(OH)2(BPDC-Me2)4(BDC)(TPDC-Me2)0.5 (PCN-703). Note that after the incorporation of TPDC-Me22, the pocket is elongated to 8.2 Å which is slightly longer than the size of H2BDC (6.9 Å). Thus PCN-703 could not be obtained from PCN-702. Similarly, based on this synthetic strategy, functionalized linkers of 2-amino-1,4-benzenedicarboxylate (BDC-NH22) and 2 0 ,5 0 -dimethoxyterphenyl-4,400 -dicarboxylate (TPDC-(CH3O)22) were also introduced successfully into MTV-MOFs to obtain Zr6(m3-O)5(m3-OH)3(OH)2 (BPDC-Me2)4(BDC-NH2)(TPDC-(CH3O)2)0.5 (PCN-704). Apart from the ligand substitution/exchange, the metal nodes in Zr-MOFs can be exchanged through post-synthetic modifications. Ti(IV) is considered as an excellent candidate for MOF construction because of its low toxicity, high redox activity, and extraordinary photocatalytic properties. However, due to a strong tendency toward the hydrolysis of Ti(IV) ions, directly synthesizing Ti-based MOFs is quite challenging, and only a few examples have been reported so far.113–116 Alternately, Kim et al. obtained Ti-containing UiO-66 MOFs through the central metal ion substitution approach (Fig. 8).117 Experimentally, UiO-66 samples were soaked in DMF solutions of three different Ti(IV) salts, TiCp2Cl2, TiCl4(THF)2, and TiBr4 for 5 days at 85 1C, respectively. The Ti-substituted MOF UiO-66(Zr/Ti) was obtained with different exchange degrees, which relied on the used Ti(IV) salts. Among them, TiCl4(THF)2 showed the highest exchange level, with about 38 wt% of Zr(IV) ions being substituted by Ti(IV) ions. Using a similar experimental procedure, UiO-66(Zr/Hf) was obtained Chem. Soc. Rev. after UiO-66 was treated with HfCl4. But, only about 20% of Hf(IV) ions were introduced into UiO-66 even at an enhanced temperature. On the other hand, substitution of coordinated guest molecules represents another method for the pore modification of MOFs. Deria et al. used mesoporous Zr6(m3-OH)8(OH)8(TBAPy)2 (NU-1000) as the platform to prepare a series of functionalized Zr-MOFs (Fig. 9). NU-1000 contains octahedral [Zr6(m3-OH)8(OH)8]8+ nodes, where eight of twelve octahedral edges are connected with carboxylate groups of TBAPy4 ligands, while the remaining coordination sites of Zr(IV) ions are occupied by eight terminal –OH groups. First, based on a solvent-assisted ligand incorporation (SALI) method, these –OH groups were substituted by a series of perfluoroalkyl carboxylate entities with varying chain length by exposing a NU-1000 sample into a DMF solution of respective fluoroalkyl carboxylic acids.118 It was found that about 3.4–4.0 perfluoroalkyl carboxylates per Zr6 node could be incorporated within NU-1000. As a result, these modified MOFs showed enhanced CO2 capture capacities due to the presence of C–F dipoles. Then, they also incorporated other carboxylate-based alkyl and aromatic functional groups into NU-1000 by a similar procedure.119 Chemical reactions such as click chemistry, imine condensation, and pyridine quaternization were then performed for the further modification of these materials. Similarly, NU-1000 was modified by phenylphosphonate (PPA) through the SALI method.120 It was found that the phosphonate displaced OH/H2O ligands and meanwhile chelated with Zr(IV) atoms of the MOF through phosphonate O atoms. Moreover, they also demonstrated that the incorporation of PPA rendered parent NU-1000 more renitent to hydroxide attack and framework dissolution. Except for the above SALI method, reactive incorporation of metal ions via atomic layer deposition in MOFs (AIM) was proposed by the same group as well.121 They exposed the NU-1000 microcrystalline sample into diethylzinc (ZnEt2) or trimethylaluminum (AlMe3) vapors, and This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Fig. 9 Molecular representations of NU-1000 with TBAPy4 as a linker (top) and the schematic representation of the SALI (bottom). R refers to the phenylphosphonate, carboxylic functional groups, perfluoroalkyl carboxylate, respectively. Reprinted with permission from ref. 119. Copyright 2014 The Royal Society of Chemistry. new materials, Zn-AIM and Al-AIM, were prepared, respectively. On average, 0.5 Zn or 1.4 Al atoms per Zr atom were observed in the resulting materials. In-depth studies demonstrated that multiple metals, binary metal merge, and metal oxide clusters could also be incorporated into NU-1000 based on the AIM process. In addition, some acidic or basic active sites can also be introduced into Zr-MOFs by PSM. For example, Ameloot et al. functionalized m3-OH groups of Zr6 clusters in UiO-66 with basic lithium t-butoxide (LiOtBu).122 After the dehydration of UiO-66 at high temperature, the exposed coordinatively unsaturated Zr sites in the Zr6O6 units were brought subsequently in contact with LiOtBu in a m3-capping fashion. Assuming that per cluster possessed two possible grafting sites, approximately 25% of available sites were occupied by tBuO anions. The obtained LiOtBu grafted UiO-66 was also demonstrated to be a superior solid ionic conductor. After that, López-Maya et al. reported the synthesis of UiO-66@LiOtBu using the above method.123 Further, they prepared missing-linker-defect Zr6(m3-O)4(AcO)6(BDC)5 (UiO-66@AcO) through the incorporation of AcO. By PSM, an acidic HSO4 group was further introduced into this MOF by soaking activated UiO-66@AcO in a KHSO4 aqueous solution to obtain functionalized Zr6(m3-O)4(AcO)4(BDC)5(HSO4)2 (UiO-66@ HSO4) with Brønsted acidic sites. To improve the CO2 trapping capacity, Li et al. also successfully grafted ethanolamine (Hea) onto the pore surface of UiO-66 by a two-step PSM.124 First, UiO-66 was dehydrated to form Zr6(m3-O)6(BDC)6 with a distorted Zr6(m3-O)6(COO)12 cluster, which then reacted with anhydrous Hea to form Zr6(m3-O)4(m3-OH)2(EA)2(BDC)6 (UiO-66-EA). Due to the charge-assisted coordination bond between m3-OCH2CH2NH2 and Zr(IV) as well as the steric hindrance protection effect, the –NH2 groups could be firmly immobilized even under thermal and humid conditions. As predicted, UiO-66-EA exhibited higher CO2 adsorption enthalpy and selectivity of CO2/N2 than UiO-66. It should be pointed out that although great efforts have been devoted to synthesizing Zr-MOFs, most reported preparations This journal is © The Royal Society of Chemistry 2016 Review Article were carried out at high temperature and high pressure with flammable toxic and expensive organic solvents. As a result, feasible approaches for scale-up production of Zr-MOFs are still lacking. It is highly desirable to develop facile and industrially applicable synthetic methods to prepare Zr-MOFs. In 2013, Yang et al. put forward a feasible water-reflux method to synthesize Zr6(m3-O)4(m3-OH)4(BDC-(COOH)2)6 (UiO-66-(COOH)2) with a relatively high-scale production.125 In this synthesis, water instead of DMF was used, being important in consideration of cost and regeneration issues. Interestingly, due to the introduction of polar free carboxylic groups into the pores, the obtained UiO-66-(COOH)2 showed strong interaction with CO2 and high selectivity of CO2/N2. Following this work, Reinsch et al. successfully prepared Zr6(m3-O)2(m3-OH)6(BDC-F4)6(SO4) (H2BDC-F4 = tetrafluoroterephthalic acid) and Zr6(m3-OH)8(OH)2.8(SO4)3.6(BDC-NH2)3(H2O)7.4 (H2BDC-NH2 = 2-aminoterephthalic acid) using Zr(SO4)2 as the metal salt and water as the solvent under mild conditions.126 Hu et al. also developed a modulated hydrothermal (MHT) approach for the facile, green (aqueous solutions), and scale-up synthesis of UiO-66-F4, UiO-66-(OCH2CH3)2, and UiO-66-(COOH)4 with Zr(NO3)4 as the metal salt, which could not be isolated by common solvothermal methods.127 In addition, Taddei et al. reported the microwave assisted synthetic method for the scale-up preparation of UiO-66, with high quality.128 Although it is still a challenge to develop industrially applicable synthetic approaches, the significant achievements mentioned above delineated an explicit blueprint for the scale-up production and commercialization of Zr-MOFs in the future. 3. Structures Zr-MOFs are constructed by interconnection of polyatomic inorganic Zr-containing clusters and polytopic organic ligands. The structural diversity of the two components and various ways the two components connect contribute to the variety of structures in Zr-MOFs. The structures of reported Zr-MOFs are discussed in this section, in terms of assorted Zr-based clusters generated and organic ligands utilized. 3.1 Zr-based cluster/core The progression from simple metal ions to complicated SBUs proved to be a milestone in the development of MOFs.12,61,129,130 As of this writing, Zr3, Zr4, Zr5, Zr6, Zr8, Zr10, and Zr18 discrete clusters have been synthesized and characterized in cluster chemistry.131–134 However, only two types of discrete Zr-based clusters have been observed in reported Zr-MOFs: Zr6O8 clusters with a variety of coordination environments and Zr8O6 clusters as a 12-connected node only observed in PCN-221.135 In addition, there are some other forms of Zr cores in Zr-MOFs, such as single Zr(IV) ions and chain structures formed from Zr(IV) ions and ligands. Here, we classify these structures according to coordination geometries of the Zr(IV) ions, including ZrO6 in phosphate-based Zr-MOFs,136–138 ZrO7 in MIL-140 series,139,140 and ZrO8 in phenolic Zr-MOFs141,142 (Fig. 10). Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Fig. 10 Observed Zr-based clusters/cores in Zr-MOFs: (a) Zr6O8 cluster, (b) Zr8O6 cluster, (c) ZrO6, (d) ZrO7, and (e) ZrO8. 3.1.1 Zr6O8 cluster. The Zr6(m3-O)4(m3-OH)4 octahedral cluster is the most commonly observed in Zr-SBUs, in which the six vertices of the octahedron are occupied by Zr(IV) centers and eight triangular faces are alternatively capped by four m3-OH and four m3-O groups. Each Zr(IV) is eight-coordinated by O atoms in a square-antiprismatic coordination geometry. One square face of the square-antiprism is formed by four O atoms supplied by carboxylate groups while the other one consists of four O atoms from two m3-O and two m3-OH groups. It should be pointed out that upon activation at high temperature, the Zr6(m3-O)4(m3-OH)4 cluster was able to lose two water molecules, leading to a distorted Zr6(m3-O)6 cluster with the Zr–O coordination numbers reduced to seven (Fig. 11). The symmetry of the cluster thus changes from Td to D3d. These two forms are commonly referred to as hydroxylated and dehydroxylated phases, respectively. When exposed to water vapor, the dehydroxylated form readily converts back to the hydroxylated form.27,31 It is well known that the connectivity and symmetry of a SBU are the most important factors affecting the net topology of associated MOFs. When the Zr6(m3-O)4(m3-OH)4 cluster is fully Fig. 11 (a) The dehydroxylation of the Zr6(m3-O)4(m3-OH)4 cluster upon thermal treatment at 300 1C in a vacuum, leading to a distorted Zr6(m3-O)6 cluster, (b) stick and ball representation of the perfect Zr6 octahedron, and (c) stick and ball representation of a squeezed Zr6 octahedron. Reprinted with permission from ref. 31. Copyright 2011 American Chemical Society. Chem. Soc. Rev. Chem Soc Rev coordinated by twelve carboxylate groups, a Zr6(m3-O)4(m3-OH)4(CO2)12 SBU is generated with the Oh-symmetry, which is observed in most of the reported Zr-MOFs (Fig. 12a). However, in Zr6(m3-O)4(m3-OH)4(TCPP)3 (PCN-223) reported by Feng et al., an unprecedented 12-connected D6h symmetric Zr6 cluster was found.143 For the D6h Zr6 cluster, each Zr(IV) located in the equatorial plane of the Zr6 octahedron is coordinated by eight O atoms, four O atoms from two m3-O and two m3-OH groups, and four O atoms only supplied by three carboxylates. So, only eight carboxylates are bridging two adjacent Zr atoms, and the other four in the equatorial plane are chelating single Zr atoms, respectively (Fig. 12b). The cluster was initially regarded as a 12-connected Zr18 cluster in this MOF structure, but additional experiments confirmed that it was a 12-connected Zr6 cluster instead of Zr18. The Zr18 cluster was actually a crystallographic disordered model where three Zr6 clusters oriented in different directions. The Oh symmetry includes many symmetry subgroups. From a topological point of view, there is a large possibility for construction of three-dimensional (3D) frameworks from the highly symmetric Zr6(m3-O)4(m3-OH)4(CO2)12 SBU and organic ligands by reducing the symmetry and/or connectivity. For example, in DUT-69, only ten edges of the Zr6 octahedral cluster are occupied by carboxylate groups while the remaining four coordination sites in the equatorial plane are occupied by two AcO and two water molecules, leading to the formation of a 10-connected [Zr6(m3-O)4(m3-OH)4(CO2)10]2+ (Fig. 12c).50 In addition, Feng et al. also reported Zr6(m3-OH)8(OH)8(TCPP)2 (PCN-222), in which only eight edges of the Zr6 octahedral cluster are occupied by carboxylate groups while the remaining eight coordination sites of Zr(IV) centers in the equatorial plane are occupied by terminal OH groups.144 Consequently, the SBU becomes 8-connected Zr6(m3-OH)8(OH)8(CO2)8 with the symmetry being reduced from Oh to its subgroup D4h (Fig. 12d). In addition to PCN-222, another 8-connected SBU was observed in Zr6(m3-O)4(m3-OH)4(OH)4(H2O)4(TCPP)2 (PCN-225) reported by Jiang et al.145 Interestingly, in this MOF six Zr atoms formed an exact octahedron, while eight m3-O capped on the triangular faces formed a highly distorted polyhedron instead of an ideal one as observed in PCN-222. As a result, the cluster symmetry of D4h in PCN-222 reduces to D2d in PCN-225, leading to lower symmetry of the latter (Fig. 12e). Furthermore, a 6-connected Zr6 octahedron SBU was observed in Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(TCPP)1.5 (PCN-224).146 Only six edges of the Zr6 octahedron are occupied by carboxylate groups in the structure, leading to the formation of a hexagonal planar SBU with the D3d symmetry (Fig. 12f). In addition, Feng et al. also synthesized another Zr-MOF, Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(TATB)2 (PCN-777), in which another 6-connected trigonal prismatic Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(CO2)6 SBU was observed (Fig. 12g).147 It was suggested that switching positions of carboxylate groups and terminal OH/H2O groups in the Zr6 cluster could realize the cluster interconversion between PCN-777 and PCN-224. 3.1.2 Zr8O6 cluster. The appearance of the Zr8(m4-O)6 cubic cluster (Fig. 10b) in Zr8(m4-O)6(OH)8(TCPP)3 (PCN-221) reported by Feng et al. further enriched the structural diversity of Zr-MOFs.135 This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article Fig. 12 The observed SBU based Zr6(m3-O)4(m3-OH)4 cluster in reported Zr-MOFs: (a) 12-connected Zr6(m3-O)4(m3-OH)4(CO2)12 in UiO-66; (b) 12-connected Zr6(m3-O)4(m3-OH)4(CO2)12 in PCN-223; (c) 10-connected Zr6(m3-O)4(m3-OH)4(CO2)10(AcO)2 in DUT-69; (d) 8-connected Zr6(m3-OH)8(OH)4(CO2)8 in PCN-222; (e) distorted 8-connected Zr6(m3-O)4(m3-OH)4(OH)4(H2O)4(CO2)8 in PCN-225; (f) 6-connected Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(CO2)6 in PCN-224; and (g) 6-connected Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(CO2)6 in PCN-777. Atom color scheme: C, black; O, red; and Zr, blue. In the Zr8(m4-O)6 cluster, each Zr atom coordinates with three O atoms from carboxylates and three m4-O atoms, forming a distorted octahedral coordination environment. Eight Zr atoms connecting six m4-O atoms lead to the formation of a cubic cluster based on the Zr arrangement, in which eight vertices are occupied by Zr atoms and six faces are capped by m4-O atoms. In the structure of PCN-221, each edge of the Zr8O6 cube cluster is bridged by a carboxylate group from a TCPP4 ligand to yield a Zr8(m4-O)6(OH)8(CO2)12 SBU with Oh symmetry. Comparing the structures of Zr8 and Zr6 clusters, it is interesting to find that the exchange of the positions of Zr and m4-O atoms in a Zr8(m4-O)6 cluster results in a cluster similar to Zr6(m4-O)4(m4-OH)4 discussed above. 3.1.3 ZrO6 core. A ZrO6 octahedron coordination core was observed in Zr-phosphonate MOFs, in which each Zr atom is coordinated by six O atoms from six different phosphonate groups.136–138 With various phosphonate ligands, the ZrO6 core presented different connection modes, leading to the formation of various SBUs. For instance, in dehydrated Zr3(H3BTBP)4 (ZrBTBP) the SBU is composed of only three ZrO6 cores (see Fig. 22 of Section 3.2.4).136 In this SBU, the center ZrO6 core connects with two other crystallographically equivalent terminal ZrO6 via six bidentate PO3C and six monodentate PO3C groups. In addition, in another Zr-phosphonate MOF, Zr(H4TTBMP)2 (UPG-1), the SBU exhibits a one-dimensional (1D) chain structure extending along the c-axis.137 In such a catenulate SBU, every ZrO6 octahedron is coordinated by six PO3C groups. However, different from that in Zr3(H3BTBP)4, only two of the six PO3C connecting adjacent ones are bidentate, while the other four are monodentate. 3.1.4 ZrO7 core. A ZrO7 core was observed in an infinite chain type SBU of the MOF ZrO(O2C-R-CO2) (R = C6H4 (MIL-140A), C10H6 (MIL-140B), C12H8 (MIL-140C), and C12N2H6Cl2 (MIL-140D)) reported by Guillerm et al. (Fig. 13).139,140 Every Zr atom coordinates with three m3-oxo groups and four carboxylate O atoms from the ligands in the ZrO7 core. The chain SBU can be considered as the linkage of two parallel corner-sharing or edge-sharing This journal is © The Royal Society of Chemistry 2016 dimers of the Zr(m3-O)3O4 polyhedra. In MIL-140, each chain SBU connects with six other chains via the dicarboxylate linkers, delimiting triangular channels in its structure.139 3.1.5 ZrO8 core. Zr(IV) with a ZrO8 coordination geometry was found in the chain SBU series of phenolic Zr-MOFs recently reported by Devic et al. (see Fig. 23 of Section 3.2.4).141,142 The Zr(IV) ions in these Zr-MOFs are eight coordinated with eight O atoms, but there are some slight differences in the coordination environment for different compounds. For example, in Zr(Hgal)(gal)0.5(DMF)0.5(DMAH2) (MIL-153, DMAH2 = dimethylammonium), Zr(IV) is coordinated by eight O atoms of four pyrogallates in a chelated way and each ligand is chelated to two Zr(IV) atoms.141 While, in Zr(Hgal)(Hsal)(DMF) (MIL-154), the Zr(IV) atom is coordinated by five phenolic O atoms of three gallate ligands, two carboxylate O atoms of salicylate ligands, and one DMF O atom.142 Similarly, both SBUs in the two MOFs are composed of edge-sharing ZrO8 polyhedra. As discussed above, rich structures of the Zr-based clusters and the single Zr coordination core lead to diverse Zr-MOFs. In particular, different ways that certain clusters and ligands combine further increase the unpredictability of resulting MOFs, such as those observed in the Zr6(m3-O)4(m3-OH)4-TCPP system. Meanwhile, there is infinite opportunity to develop new Zr-MOFs with various properties and functions. 3.2 Diverse structures with different ligands The combination of various ligands with above five types of Zr-clusters/cores has led to dozens of reported Zr-MOFs. Among them, aromatic polycarboxylates are the most commonly used ligands due to their rich coordination modes and feasible tailorability. Simultaneously, different coordination ways of the carboxylate group have led to the generation of different SBUs. Besides carboxylate ligands, several other ligands have also been used to construct Zr-MOFs. In this section, Zr-MOFs constructed from the following five types of ligands are discussed: (1) ditopic carboxylate ligands, (2) tritopic carboxylate ligands, (3) tetratopic Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Fig. 13 (a) Structures of a series of MIL-140 viewed along the c axis and (b) the chain SBU based on Zr(m3-O)3O4 cores in MIL-140. Reprinted with permission from ref. 139. Copyright 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. carboxylate ligands, (4) phosphate ligands, and (5) phenolate ligands. Most of the used ligands are summarized briefly in Scheme 1. 3.2.1 Ditopic carboxylate ligands. Linear dicarboxylate ligands have been widely used to build Zr-MOFs. A representative one is terephthalate (BDC2) which is used for the construction of UiO-66, the first reported Zr-MOF.27 In UiO-66, all twelve edges of the octahedral [Zr6(m3-O)4(m3-OH)4]12+ cluster are occupied by twelve carboxylates from twelve BDC2 ligands to yield the Zr6(m3-O)4(m3-OH)4(CO2)12 SBU. Each BDC2 links two SBUs to form a 3D open framework with tetrahedral and octahedral cages, where each octahedral cage is connected to eight tetrahedral cages through sharing triangular windows (Fig. 14). With increasing length of used ligands, several isoreticular Zr-MOFs including UiO-67 and UiO-68 were also reported early. With H2PEDC-2OMe, Schaate et al. later reported the synthesis of PIZOF-2 (Fig. 15).73 The structure and connectivity of the Zr6(m3-O)4(m3-OH)4(CO2)12 SBU and the ligand in PIZOF-2 are the same as those in UiO series. However, PIZOF-2 has a twofold interpenetrating network structure. In each single net, each SBU is connected to twelve other SBUs by the ligands to form a 3D cubic framework with octahedral and tetrahedral cages. It is interesting that there exist both convex and concave tetrahedral cavities of 19 and 14 Å in diameter since the ligand is slightly bent. The concave tetrahedral voids accommodate Zr6(m3-O)4(m3-OH)4(CO2)12 SBUs of the other identical net, while the convex ones are empty. It is clear that the triangular windows formed by the extended linkers are big enough to let other linkers pass, thus leading to the framework interpenetration. Furthermore, using ditopic H2FUM and H2ABDC ligands Chem. Soc. Rev. Chem Soc Rev the same group also synthesized Zr-FUM and Zr-ABDC MOFs, which have 12-connected fcu network structures as in UiO-66.48,72 All of the above Zr-MOFs constructed with linear ditopic carboxylate ligands have a fcu topological network, where the Zr-based SBUs are all 12-connected. In contrast, PCN-700 with bcu topology was constructed from 8-connected Zr clusters and linear ditopic ligand Me2-BPDC2.89 In PCN-700, each Zr6 cluster is linked by eight Me2-BPDC2 ligands: four above and four below the equatorial plane of the Zr6 octahedron. Eight terminal OH/H2O groups occupy the equatorial plane. Interestingly, after the SLI process, 10-connected frameworks of PCN-701 and PCN-702, and 11-connected frameworks of PCN-703 and PCN-704 were obtained. Apart from benzene-based ligands, naphthalene and anthracenebased linear ditopic carboxylate ligands have also been used in the synthesis of Zr-MOFs. In 2010, Garibay et al. reported the isoreticular synthesis of UiO-66-1,4–Naph with 1,4-NDC2 as the ligand.148 Later, Zr6(m3-O)4(m3-OH)4(NDC)6 (DUT-52) and Zr6(m3-O)8(NDC)3(AcO)2 (DUT-84) with 2,6-NDC2 as the ligand were reported by Bon et al.149 DUT-52 is isoreticular to UiO-66. However, DUT-84 crystallizes in the orthorhombic space group of Cmma and has a different topology from that of UiO-66. In DUT-84, each Zr6 SBU is connected with only six ligands. Four ligands link the Zr6 clusters forming a layer, and the other two ligands connect two such layers, leading to a double-layer structure of DUT-84. This is the first reported Zr-MOF with a 2D structure. Though the structure is 2D, pores exist with an inner diameter of 6.5 Å. Meanwhile, Pu et al. used H2AN and synthesized UiO-66 analog Zr6(m3-O)4(m3-OH)4(AN)6 (UiO-66(AN)).150 Furthermore, Wang et al. obtained an ‘‘X-ray scintillating’’ Zr-MOF, Zr6(m3-O)4(m3-OH)4(ADC)6 through using anthracene-based emitter compound 9,10anthracenyl-bis(benzoate) (ADC2) as the bridging ligand.151 In addition, linear heterocyclic dicarboxylate ligands were also introduced into the design and preparation of Zr-MOFs. Wang et al. used S-heterocyclic dimethyl substituted thiophene DTDC2 as the ligand to synthesize Zr6(m3-O)4(m3-OH)4(DTDC)6 (Zr-DTDC) with the same structural topology of UiO-66.152 Due to the joint action of the electronic effect and steric hindrance from methyl groups as well as strong Zr–O bonds, Zr-DTDC shows ultrahigh hydrothermal stability. Yoon et al. adopted BTDC2 to construct a UiO-66 type structure Zr6(m3-O)4(m3-OH)4(BTDC)6 (Zr-BTDC).153 The sizes of the octahedral and tetrahedral cages in Zr-BTDC are about 12 and 8 Å, respectively. In addition, in order to incorporate Lewis basic sites into the pore surfaces of Zr-MOFs, Li et al. exploited N-heterocyclic ligand BPYDC2 and also obtained Zr6(m3-O)4(m3-OH)4(BPYDC)6 (UiO(BPYDC)) with fcu topology.154 Aside from the above-discussed linear ligands, angular dicarboxylate ligands have also been explored to construct Zr-MOFs. For example, DUT-51 reported by Bon et al. was built using the DTTDC2 ligand with a bend angle of 148.611 and a [Zr6(m3-O)6(m3-OH)2]10+ octahedral cluster (Fig. 16).49 In its structure, each Zr-based cluster is 8-linked by DTTDC2 ligands through the carboxylate coordination on eight edges of the Zr6 octahedral cluster. The remaining coordination sites in the equatorial plane of the Zr6 cluster are occupied by six DMF and two BC ligands. As a result, the assembly between the This journal is © The Royal Society of Chemistry 2016 View Article Online Review Article Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Scheme 1 Some ligands used in the construction of Zr-MOFs. [Zr6(m3-O)6(m3-OH)2]10+ cluster and the bent ligand generates a 3D framework of DUT-51 with the reo topology. The framework contains two types of cages: octahedral and cubo-octahedral This journal is © The Royal Society of Chemistry 2016 ones with diameters of 15.6 and 18.8 Å, respectively. The cubooctahedral cage possesses six square and eight triangular windows with the size of 9 and 4.2 Å, respectively. Each triangular Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Chem Soc Rev Fig. 14 Structure of UiO-66: (a) tetrahedral cage; (b) octahedral cage; and (c) packing of the two types of cages.27 Fig. 16 Structure of DUT-51: (a) [Zr6(m3-O)6(m3-OH)2]10+ cluster based SBU; (b) the remaining four coordination sites occupied by DMF and benzoic acid molecules in the equatorial plane; (c) octahedral and cubooctahedral pores; and (d) polyhedral packing. Reprinted with permission from ref. 49. Copyright 2012 The Royal Society of Chemistry. Fig. 15 (a) One of the two frameworks of PIZOF-2 showing the arrangement of concave and convex tetrahedral cavities; (b) topological representation of the interpenetrated structure in PIZOF-2; (c) convex cavity; (d) concave cavity; and (e) the whole structure of PIZOF-2. Reprinted with permission from ref. 73. Copyright 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. window is shared with one octahedral cage, and each square window is shared with one cubo-octahedral cage. Following this work, the same group synthesized a series of Zr-MOFs, DUT-67–69, by using the bent TDC2 ligand under different reaction conditions (Fig. 17).50 Carboxylate groups of TDC2 with a bridging angle of 147.91 are spaced at a distance of 5.31 Å, which can be viewed as a shortened version of DTTDC2. The structure of DUT-67 has a reo topological net, similar to that of DUT-51, containing 8-connected [Zr6(m3-O)6(m3-OH)2]10+ octahedral clusters and cuboctahedral and octahedral cage pores. DUT-68 crystallizes in the Im3% m space group and has two different crystallographically independent 8-connected [Zr6(m3-O)6(m3-OH)2]10+ clusters with a binodal net. The structure of DUT-68 represents Chem. Soc. Rev. a complicated hierarchical pore system containing four types of cage pores of the rhombicuboctahedral mesoporous cage, the square antiprismatic cage, the cuboctahedral cage, and the smallest octahedral cage. Their inner diameters are 27.7, 13.9, 12.5, and 8 Å, respectively. The largest rhombicuboctahedral cage shares its square windows with six cuboctahedral cages, while the remaining windows are shared by twelve square antiprismatic cages. The octahedral cage is formed as a result of stacking six square antiprismatic cages and two rhombicuboctahedral cages. DUT-69 with the bct topology is the first example of a Zr-MOF containing a uninodal 10-connected Zr6(m3-O)4(m3-OH)4(CO2)10(AcO)2 SBU. The structure of DUT-69 involves octahedral cages with a diameter of 5 Å and rectangular channels of 9.15 2.66 Å in size. Similar to the structure of DUT-69, Furukawa et al. utilized PZDC2 to synthesize Zr6(m3-O)4(m3-OH)4(PZDC)5(HCO2)2(H2O)2 (MOF-802), in which each SBU is 10-connected with PZDC2 linkers, while formate and DMF ligands occupy the remaining coordination sites.155 As discussed above, using rigid aromatic ditopic carboxylates has resulted in a variety of Zr-MOFs. In contrast, few Zr-MOFs based on flexible aliphatic ditopic carboxylates have been reported until now. Compared with the framework of rigid aromatic counterparts, a skeleton of a Zr-MOF with flexible aliphatic carboxylate is highly possible to exhibit dynamic behavior and unique properties, which is worth exploring. For the first time, using flexible AP2 and AP-CH32 ligands Reinsch et al. synthesized Zr6(m3-OH)8(OH)8(AP-CH3)4, Zr6(m3-OH)8(OH)4(SO4)4(AP)2 and Zr6(m3-OH)8(OH)2(Cr2O7)3(AP)4, respectively (Fig. 18).156 For all the three Zr-MOFs, the inorganic nodes are based on the well-known Zr6O8 cluster. In the structure of Zr6(m3-OH)8(OH)8(AP-CH3)4, eight carboxylate groups coordinate This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article Fig. 18 (a) The SBU and framework connectivity in Zr6(m3-OH)8(OH)8(AP-CH3)4; (b) core architecture and framework structure of Zr6(m3OH)8(OH)4(SO4)4(AP)2; and (c) core architecture and framework structure of Zr6(m3-OH)8(OH)2(Cr2O7)3(AP)4. Reprinted with permission from ref. 156. Copyright 2015 The Royal Society of Chemistry. Fig. 17 (a) Linker orientation and local environment of Zr clusters in DUT-67–69; (b) the framework structure of DUT-67 showing two types of cages; (c) the framework structure of DUT-68 showing four types of cages; and (d) the framework structure of DUT-69 showing one type of cage. Reprinted with permission from ref. 50. Copyright 2013 American Chemical Society. with a Zr6(m3-OH)8(OH)8 cluster, resulting in a uninodal net with the bcu topology. Different from those in Zr6(m3-OH)8(OH)8(AP-CH3)4, four SO42 ions in Zr6(m3-OH)8(OH)4(SO4)4(AP)2 coordinate to the cluster all in a bidentate fashion. With these binding modes, the substructure with the dia topology is formed by hydrogen bonding of Zr-sulfate clusters, while the Zr(IV) atoms in the equatorial plane of the cluster are coordinated by the organic linkers, leading to a square grid layer. In Zr6(m3-OH)8(OH)2(Cr2O7)3(AP)4 the Cr2O72 ions served as connectors between Zr-based clusters in a monodentate coordination fashion. The remaining coordination sites in the Zr clusters are occupied by eight carboxylate groups of organic ligands, which extend the structure to form a bcu framework. It is clear that these inorganic anions played an important role in tuning the connectivity of the Zr-based clusters in this system, thereby forming different network structures of Zr-MOFs. 3.2.2 Tritopic carboxylate ligands. Among tritopic ligands, BTC3 was first used in Zr-MOF construction. Zr6(m3-O)4(m3-OH)4(BTC)2(HCO2)6 (MOF-808) reported by Furukawa et al. crystallizes in the cubic space group of Fd3% m, in which each Zr6(m3-O)4(m3-OH)4(CO2)12 SBU is connected to six BTC3 ligands and each BTC3 ligand coordinates with three SBUs, leading to the formation of a (3,6)-connected 3D framework with spn topology.155 HCO2 ions act as the terminal ligand which This journal is © The Royal Society of Chemistry 2016 coordinate with the [Zr6(m3-O)4(m3-OH)4]12+ cluster. The framework of MOF-808 features tetrahedral cages with the internal pore diameter of 4.8 Å, where the Zr6(m3-O)4(m3-OH)4(CO2)12 SBUs are at the vertices and the BTC3 ligands at the faces of the tetrahedron. Furthermore, these tetrahedral cages share their vertices to form a large adamantine-like cage with the internal pore diameter of 18.4 Å. Then, using another tritopic ligand TATB3 Feng et al. synthesized zeotype PCN-777 with a b-cristobalite-type structure similar to that of MOF-808 (Fig. 19).147 However, because the size of organic linker TATB3 is almost twice that of BTC3, the internal pore diameter of the adamantine-like cage reaches up to 38 Å, which ranks PCN-777 among mesoporous MOFs. Soon after, Wang et al. synthesized Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(BTB)2 (Zr-BTB) by using BTB3 as the ligand.157 Fig. 19 (a) The Zr6(m3-O)4(m3-OH)4(CO2)12 SBU and trigonal-planar organic linker TATB3 in constructing PCN-777; (b) structure of supertetrahedra in PCN-777; and (c) the overall framework structure of PCN-777 with a large mesoporous cage.147 Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Zr-BTB crystallizes in the space group of I41/amd. In its structure, the Zr6 clusters as 6-connected nodes were connected by BTB3 ligands as the 3-connected linkers to give rise to a (3,6)-connected 2D network with the kgd topology. The 2D framework possesses elliptical windows with the size of 7.16 12.61 Å. Moreover, the parallel 2D layers of one set interpenetrate perpendicularly with another set to generate a 3D open framework with 1D square channels (14.37 14.37 Å). Zr-BTB thus represented the first Zr-MOF with a 2D - 3D interpenetration structure. 3.2.3 Tetratopic carboxylic ligands. Compared with di- and tricarboxylate ligands, tetracarboxylate ligands have more coordination sites and rich coordination patterns. More importantly, tetracarboxylic ligands with large space steric hindrance could effectively avoid framework interpenetration, being favorable for constructing mesoporous Zr-MOFs. Among all tetracarboxylic ligands for constructing reported Zr-MOFs, H4TCPP is quite common. Some related studies can be found in a published review on metal-metalloporphyrin frameworks reported by Gao et al.158 The variation of its connecting number and symmetry provides a large number of topological possibilities to form various frameworks with different pore sizes and shapes. In addition, most reported porphyrinic Zr-MOFs exhibited excellent thermal and chemistry stability, providing good platforms for exploring the applications of MOFs under harsh conditions. Currently, eight Zr-MOFs constructed using H4TCPP have been reported by Zhou’s group from PCN-221 to PCN-225,135,143–146 Yaghi’s group for Zr6(m3-O)4(m3-OH)4(TCPP)3 (MOF-525) and Zr6(m3-O)8(H2O)8(TCPP)2 (MOF-545),159 and Ma’s group for Zr6(m3-O)8(H2O)8(CO2)8 (MMPF-6).160 In PCN-221, four peripheral benzene rings of the ligand are perpendicular to the central porphyrin ring.135 Each TCPP4 linker connects with four Zr8 clusters in a 4-connected mode, and each Zr8 cluster is linked by twelve TCPP4 linkers in a 12-connected node, resulting in a (4,12)-connected ftw topological network. In addition, the structure features two types of polyhedral cages with distorted octahedral and cubic shapes. The slightly distorted octahedral cage, with a cavity diameter of B11 Å, involves two Zr8 clusters at the vertices along the axial direction and four TCPP4 ligands in the equatorial plane. The vertices and faces of the cubic cage with the edge length of B20 Å are delimited by eight Zr8 clusters and six TCPP4 ligands, respectively. For PCN-223, following the same topological simplification way of PCN-221, each Zr6 cluster can be regarded as a 12-connected node and the TCPP4 linker can be simplified to a 4-connected node.143 The framework represents the first (4,12)-connected MOF with the shp topology. In addition, PCN-223 contains uniform triangular 1D channels of 12 Å in diameter delimited by three SBUs and three ligands. MOF-545 prepared by Morris et al. crystallizes in the hexagonal space group P6/mmm.159 In its structure, eight edges of the Zr6 octahedral cluster are bridged by carboxylate groups and the remaining positions are occupied by terminating water, leading to the formation of a SBU with a formula of Zr6(m3-O)8(CO2)8(H2O)8. Combination of these SBUs and the tetratopic linker TCPP4 resulted in a network with the csq topology. The structure of MOF-545 features triangular and hexagonal 1D channels with Chem. Soc. Rev. Chem Soc Rev Fig. 20 Structures of MOF-525 and -545: (a) Zr6(m3-O)4(m3-OH)4(CO2)12 SBU in MOF-525; (b) TCPP4 linker used in MOF-525 and -545; (c) ftw topology; (d) the structure of MOF-525; (e) Zr6(m3-O)8(CO2)8(H2O)8 SBU in MOF-545; (f) csq topology; and (g) the structure of MOF-545. Reprinted with permission from ref. 159. Copyright 2012 American Chemical Society. diameters of 8 and 36 Å, respectively. In addition, MOF-525 with the ftw topology was also obtained in the same work. It crystallizes in the space group Pm3% m and features large cubic cages with an edge length of 20 Å, which are packed in a primitive cubic lattice. The cubic cage is comprised of eight corner-sharing Zr6(m3-O)4(m3-OH)4 units and six face-sharing porphyrin-based units (Fig. 20). Parallel to this work, PCN-222 and MMPF-6 with nearly the same structure as MOF-545 were reported at almost the same time. In PCN-225, the Zr6 cluster and the TCPP4 ligand can also be viewed as 8- and 4-connected nodes, respectively.145 But different from the (4,8)-connected csq topology of MOF-545, the whole network of PCN-225 can be simplified as the (4,8)-connected sqc net. There are two types of channels in PCN-225 running along the crystallographic a and b direction, respectively. The small quadrangle-shaped channel with the dimension of 8 15 Å is comprised of two Zr6 clusters and two TCPP4 ligands, whereas the big pear-like one with a size of 9 22 Å is delimited by three Zr6 clusters and three TCPP4 ligands. As for PCN-224, the Zr-based cluster and the TCPP4 ligand are 6- and 4-connected, respectively.146 The framework can thus be regarded as a (4,6)-connected she topological net. As the coordination number of the cluster is low, large 3D open channels of 19 Å are found in PCN-224. Besides TCPP4, other tetratopic carboxylate ligands used in building Zr-MOFs can be classified into four categories: augmented porphyrinic derivatives63,65,161 (in constructing CPM-99, NU-1102, NU-1104, PCN-228, -229, and -230), planar pyrene-based derivatives63,83,118–121,162 (in constructing NU-1000, -1100, -1101, -1103, and Zr-PTBA), flexible planar tetratopic ligands159,162,163 (in constructing PCN-94, MOF-535, and Zr-BTBA), and tetrahedral ligands34,155 (in constructing PCN-521, MOF-812, and MOF-841). All reported Zr-MOFs constructed by tetratopic augmented porphyrinic derivatives are the ftw network structures. Lin et al. reported Zr6(m3-O)4(m3-OH)4(TCBPP)3 (CPM-99) with TCBPP4 as the ligand.161 The CPM-99 framework includes large cubic This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev cavities with an edge length as large as 25 Å, which are packed in a primitive cubic lattice. Each 12-connected Zr6 cluster resides on one cube vertex while the faces of the cube are occupied by TCBPP4 ligands. Simultaneously, the distorted octahedron cage with B11 Å size in diameter is created by two Zr6 clusters in the axial sites and four TCBPP4 ligands located in the equatorial plane. With the same ligand, Wang et al. prepared isostructural NU-1102. Meanwhile, NU-1104 was prepared with an extended version of the TCBPP4 ligand, designed by inserting triple-bonds between the phenyl groups of each arm.63 The diameters of the large cubic cage and octahedral cage in NU-1104 are 24.2 and 13.5 Å, respectively. Furthermore, by further increasing the arms’ length, PCN-228 to -230 were synthesized with TCP-14, TCP-24, and TCP-34, respectively.65 Due to the extra-large size of these three porphyrinic linkers, the resulting Zr-MOFs are mesoporous with pore sizes ranging from 25 to 38 Å. Most reported Zr-MOFs built with pyrene-based ligands, such as the above-mentioned NU-1100, -1101, and -1103, possess ftw topological structures as well, with the exceptional case of NU-1000. NU-1000 crystallizes in the space group of P6/mmm, similar to PCN-222.118–121 In the structure of NU-1000 eight of the twelve edges in the octahedral Zr6 cluster are occupied by eight carboxylate groups of TBAPy4 ligands and the remaining positions are occupied by terminal –OH groups pointing into the mesoporous channels. The resulting 3D structure can thus be viewed as 2D Kagome sheets linked by TBAPy4 ligands. With rigid PTBA4 as the ligand, Kalidindi et al. prepared [Zr6(m3-O)4(m3-OH)4(PTBA)2.75(OH)](C7H6O2)0.25(H2O)2 (Zr-PTBA), which is isostructural to MOF-525 built from TCPP4.162 According to the CHN elementary analysis and TGA results, it was suggested that about 8% of the PTBA4 linkers are replaced by a hydroxide anion per formula in Zr-PTBA. Despite the defects observed, Zr-PTBA showed permanent porosity, but was not stable in water. Compared with rigid ligands, the tetratopic ligands with conformational flexibility should be more compatible with Zr clusters. Tetrabenzoic acid H4BTBA with a configurationally flexible biphenyl core was used by the same group to synthesize Zr6(m3-O)4(m3-OH)4(BTBA)3 (Zr-BTBA).162 It should be noted that BTBA4 has the same four-fold connectivity being essential identical metrics, but distinct torsional flexibility compared with PTBA4. Zr-BTBA crystallizes in the Pm3% m space group. The edges of the octahedral Zr6(m3-O)4(m3-OH)4(CO)12 SBUs are interconnected by twelve BTBA4 ligands, which in turn are bridging four neighboring SBUs, resulting in an ftw framework as in MOF-525. Zr-BTBA displays a 3D porous system with an average pore diameter of about 19 Å. Unlike its free conformation, a significant deviation from planarity of the BTBA4 ligand was observed in Zr-BTBA. Interestingly, it was found that contrary to Zr-PTBA, Zr-BTBA showed excellent water stability, although its mechanical stability was poor. Combining the stability features of Zr-PTBA and Zr-BTBA, a solid solution material Zr6(m3-O)4(m3-OH)4(PTBA)2.6(BTBA)0.4 with a BET surface area up to 4000 m2 g1 was then prepared using a mixed ligand strategy (Fig. 21). Remarkably, the synergy of flexible BTBA4 and rigid PTBA4 linkers makes this material hydrolytically This journal is © The Royal Society of Chemistry 2016 Review Article Fig. 21 The designed synthesis of Zr6(m3-O)4(m3-OH)4L1XL23X with the mixed tetratopic ligand acids H4BTBA (H4L1) and H4PTBA (H4L2), highlighting its stability. Reprinted with permission from ref. 162. Copyright 2015 WileyVCH Verlag GmbH & Co. KGaA, Weinheim. and structurally stable, which is difficult for Zr-BTBA and Zr-PTBA, individually. Similarly, Wei et al. synthesized Zr6(m3-O)4(m3-OH)4(ETTC)3 (PCN-94) by using the ETTC4 ligand, which is isostructural to MOF-525.163 The size of the cubic cage in PCN-94 is about 17.5 Å across an edge. Such cages are packed in a primitive cubic lattice to form additional channels with a window size of 14 14 Å2. Note that there exists a prominent conformational change for the twisted ETTC4 linker in the Zr-MOF framework compared with that in its free state. The minimum angle between adjacent carboxylate groups, the average dihedral angles between phenyl rings, as well as those between phenyl rings and the central plane in the ETTC4 ligand are all larger than those in its free conformation. With the exception of planar tetratopic ligands as discussed above, tetrahedral carboxylate-based ligands with the Td symmetry are also quite intriguing for MOF construction.164–172 However, due to inherent 3D directive conformation of this type of ligand, geometrical symmetry incompatibility emerges when they combine with SBUs of high symmetry. As a result, reported Zr-MOFs with tetrahedral ligands are relatively rare. Currently, only four Zr-MOFs, PCN-521 as discussed above with the MTBC4 ligand,34 Zr6(m3-O)4(m3-OH)4(H2O)2(MTB)3 (MOF-812),155 Zr6(m3-O)4(m3-OH)4(MTB)2(HCO2)4(H2O)4 (MOF-841) with the MTB4 ligand,155 and Zr6(m3-O)4(m3-OH)4(OH)4(H2O)4(TCPS)2 (Zr-TCPS) with the TCPS4 ligand,36 have been reported. Except for the ith type network of MOF-812, the other three Zr-MOFs all are (4,8)-connected networks with the flu topology. PCN-521 crystallizes in the tetragonal space group of I4/m and its framework consists of 8-connected Zr6(m3-O)4(m3-OH)4(OH)4(H2O)4(CO2)8 SBUs linked by tetrahedral MTBC4 linkers.34 Similar to those in PCN-222 and MOF-545, the symmetry of the Zr6(m3-O)4(m3-OH)4(H2O)4(CO2)8 SBU is D4h being symmetrically compatible with the tetrahedral MTBC4 with the D2d symmetry. As a result, a (4,8)-connected network with the flu topology was generated. The structure of PCN-521 contains interstitial cavities with the size of 20.5 20.5 37.4 Å. Hereafter, Furukawa et al. obtained MOF-841 and MOF-812 with the MTB4 ligand.155 The framework structure of MOF-841 is similar to that of PCN-521, but with small cavities of 11.6 Å due to its shortened ligand. Different from that in MOF-841, the SBUs in MOF-812 are connected to twelve MTB4 ligands, with four monodentate carboxylate groups. Each SBU in MOF-812 can thus be simplified Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article to an icosahedron, being different from the cuboctahedral 12-connected SBUs in fcu Zr-MOFs. Meanwhile, each MTB4 ligand is bridging four SBUs, leading to a 3D ith network with 5.6 Å cavities in diameter. In addition, Fang et al. synthesized Zr-TCPS by using a silane-cored TCPS4 ligand. Zr-TCPS is isostructural to PCN-521 with a (4,8)-connected 3D flu network.36 In this MOF, there exist channels with a window size of 10.2 7.98 Å running along the crystallographic a-axis direction, as well as 13.3 Å octahedron cages in diameter. Interestingly, it was found that monocarboxylates could also be used to build Zr-MOFs, which were commonly used as modulators to control crystal growth during the synthesis of Zr-MOFs as discussed above. Liang et al. obtained a novel Zr-MOF, Zr6(m3-O)4(m3-OH)4(FA)12 (ZrFA) with formate as the ligand.173 ZrFA crystallizes in the orthorhombic space group of Cmcm. The structure is also constructed by the common Zr6(m3-O)4(m3-OH)4 clusters. The eight edges up and down the equatorial plane of the Zr6 octahedron are occupied by eight formate ions in a chelating way, forming a Zr6(m3-O)4(m3-OH)4(HCO2)8 cluster. In the ac plane, Zr6(m3-O)4(m3-OH)4(HCO2)8 clusters are bridged by other inter-cluster formate linkers to give infinite 2D layers which are further stacked into 3D structure along the b-axis. 3.2.4 Phosphate ligands. Aside from carboxylates, phosphate ligands have also been used to construct Zr-MOFs. Taddei et al. synthesized Zr3(H3BTBP)4 with a honeycomb-like layered structure (Fig. 22).136 In its structure, each Zr atom is coordinated with six phosphate O atoms forming a ZrO6 entity. Three ZrO6 assembled together through sharing O atoms to form a trimeric SBU, in which the center ZrO6 entity connects with the other two via six bidentate PO3C and six monodentate PO3C groups. Each SBU is further connected with the nearest six equivalent SBUs via H3BTBP3 linkers, leading to a layered structure. The 3D structure of Zr3(H3BTBP)4 with cavities is formed by these layers stacking along the crystallographic c-axis direction in an ABAB mode. However, the generated cavities are capped by the SBUs of adjacent layers, thus making them inaccessible. In addition, the ligand H6TTBMP with slight conformational freedom was applied to synthesize another Zr-phosphonate MOF, UPG-1.137 In the structure of UPG-1, three phosphate groups of each H4TTBMP2 ligand display different coordination fashions with Zr atoms: two are connected to adjacent Zr–O chains while the third is non-coordinated and protonated. The ligands link Zr–O Chem Soc Rev Fig. 23 Structure of MIL-163: (a) view of a single chain SBU showing the Zr atoms coordinated by 1,2,3-trioxobenzene groups; and (b) view of the structure along the [001] direction showing square-shaped channels. Reprinted with permission from ref. 142. Copyright 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. chains to finally form an open framework with two types of channels with the sizes of about 10 and 5 Å, respectively. In contrast to Zr3(H3BTBP)4, UPG-1 displays permanent porosity and strong absorption affinity towards n-butane and CO2. 3.2.5 Phenolate ligands. Recently, it was demonstrated that phenolate ligands could also be used to construct Zr-MOFs. These MOFs should be much more stable because of the high pKa of the phenolate group, which could strengthen the Zr–O bonds. In early stages, Cooper et al. synthesized two Zr–hydroxycarboxylate compounds, (MIL-153) and (MIL-154) with a phenolate ligand.141 The two compounds have a 1D inorganic chain structure composed of edge-sharing ZrO8 entities. On the basis of this work, with ditopic phenolic acid H6TzGal as the starting ligand resource, Mouchaham et al. recently successfully obtained a highly stable porous Zr-MOF, Zr(H2TzGal) (MIL-163).142 Similar to that in MIL-153, each 1,2,3-trioxobenzene group chelates two Zr atoms and two pairs of these groups coordinate with one Zr atom through all O coordination to form a chain SBU (Fig. 23a). These chains are further connected with each other through H2TzGal2 linkers to yield a 3D open framework with squareshaped channels of about 12 12 Å in diameter (Fig. 23b). 4. Properties and applications Due to their diverse structures, large and accessible specific surface areas, uniform and tunable pore sizes, outstanding stability, and specific properties, intensive research has been directed towards the exploration of Zr-MOF applications. In this section, we discuss the studies of Zr-MOFs in applications ranging from catalysis, molecule adsorption and separation, drug delivery, and fluorescence sensing, and as porous carriers. 4.1 Fig. 22 (a) The single layer of ZrBTBP along the c-axis; and (b) the stacking of layers in ZrBTBP in an ABAB mode. Reprinted with permission from ref. 136. Copyright 2014 The Royal Society of Chemistry. Chem. Soc. Rev. Catalysis Catalysis is one important application field of MOFs. In this regard, several comprehensive reviews have been published, in which state-of-the-art in MOF catalysis was highlighted and analyzed. In the following section, the catalytic properties of Zr-MOFs and their functionalized derivatives are discussed based on different types of catalytic reactions, including Lewis This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Fig. 24 The schematic representation of Meerwein reduction of 4-tertbutylcyclohexanone catalyzed by UiO-66 with unsaturated metal sites.56 acid catalysis, oxidation catalysis, biomimetic catalysis, electrocatalysis, hydrogenation catalysis, and photocatalysis. 4.1.1 Lewis acid catalysis. Zr-MOFs with unsaturated metal sites were mainly applied in Lewis acid catalysis reactions, in which the open metal sites act as electron pair acceptors capable of accelerating the reaction process. Vermoortele et al. revealed that coordinatively unsaturated metal sites in UiO-66 can be drastically increased by using specific modulators in the synthesis.56 Thus, the Lewis acid catalytic activity of the MOF can be improved correspondingly. With UiO-66 as the catalyst the Meerwein reduction of 4-tert-butylcyclohexanone (TCH) with isopropanol (IPA) was studied, which requires activating the alcohol and ketone simultaneously to facilitate the transfer of hydride (Fig. 24). The results illustrated that the non-modulated UiO-66 shows nearly no activity, while the materials modulated with ten equivalents of TFA display high catalytic activity, leading to appreciable yields of tertbutylcyclohexanol. Moreover, it was also demonstrated that the space around the Zr cluster in UiO-66-NO2 increased when HCl and TFA were used as modulators, which could facilitate the simultaneous activation of reactants. As a result, 93% conversion of tert-butylcyclohexanol was achieved. Similarly, Wang et al. successfully removed coordinated water of Zr6(m3-O)4(m3-OH)4(OH)6(H2O)6(CO2)6 SBUs in Zr-BTB (Fig. 25).157 Fig. 25 The single layer structure of Zr-BTB and its Lewis acid catalyzed reaction of carbonyl compounds with cyanide.157 This journal is © The Royal Society of Chemistry 2016 Review Article Thus, the coordinative unsaturated metal sites were generated, which are favorable for the Lewis acid catalytic reaction of carbonyl compounds with cyanide. The resulting material showed high catalytic activity, especially for the cyanosilylation. The conversion reached 100% in 24 hours at room temperature, which was comparable to the performance of other reported MOF catalysts in the cyanosilylation of benzaldehyde, but much higher in the cyanosilylation of naphthaldehyde.174–176 It is worth noting that postsynthetic modification can be used as a method to obtain coordinatively unsaturated metal sites in MOFs for improving their catalytic activity. Mondloch et al. used the AIM strategy to obtain metallized NU-1000, denoted as Zn-AIM and Al-AIM.121 As a proof-of-concept to verify whether these AIM materials are able to elicit new catalytic behavior, Knoevenagel condensation reactions between ethyl cyanoacetate and benzaldehyde were studied. As expected, the Zr sites in NU-1000 proved inactive towards the Knoevenagel condensation. Interestingly, Zn-AIM and Al-AIM were active catalysts, due to the presence of Lewis acidic Zn(II) and Al(III) sites. Moreover, no Zn or Al elements were determined in the catalytic reaction systems after removing the functionalized MOFs, suggesting that Zn- and Al-AIM acted as the stable catalysts. In addition, Rasero-Almansa et al. prepared multifunctional heterogeneous catalyst Ir–Zr-MOFs (Zr–[IrL]) using a PSM process for the hydrogenation of aromatic compounds.105 This reaction involves the activation of aromatic compounds by Lewis acids and a sequential hydrogenation step. This bifunctional Zr–[IrL] catalyst with Zr as the Lewis acid and Ir as the active metal allows the activation process to work cooperatively. Using aniline as a model compound, its selective conversion to cyclohexanamine catalyzed by Zr–[IrL] resulted in a 100% yield in only 10 h, indicating the high catalytic activity of this MOF based catalyst. In addition, the Zr–[IrL] catalyst can be reused after the reaction via a simple centrifugation treatment. 4.1.2 Oxidation catalysis. Currently, oxidation reactions with Zr-MOFs as catalysts mainly include the oxidation of alkanes, alkenes, and water, as well as regioselective C–H oxidation of benzo[h]quinolone. In industry, the selective oxidation of cyclohexane to produce cyclohexanone (K) and cyclohexanol (A) is conducted over cobalt-based homogeneous catalysts with about only 4% conversion and 70–85% selectivity for the K/A product when the oxidant is air. Therefore it is highly desirable to develop new catalysts which will improve the conversion and selectivity of this reaction. Due to the existence of accessible active porphyrinic iron(III) centers, Zr-PCN-221(Fe) was employed by Feng et al. as the catalyst for the selective oxidation of cyclohexane with tert-butyl hydroperoxide (TBHP) as the oxidant.135 The results demonstrated that a quick oxidation reaction proceeded for only about 11 h. Moreover, the utilization efficiency of the oxidant TBHP reached nearly 100%. More importantly, Zr-PCN-221(Fe) as the catalyst displayed high selectivity of cyclohexanone up to 86.9%, and for cyclohexanol selectivity was only 5.4%. This observed cyclohexanone selectivity of Zr-PCN-221(Fe) was higher than those of other MOF-based or conventional molecular sieve catalysts.177–179 On the other hand, the epoxidation of alkenes is one of the most important reactions in organic synthesis Chem. Soc. Rev. View Article Online Review Article Chem Soc Rev Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Table 2 Results of the selective oxidation of cyclooctene to cyclooctene oxide with TBHP as the oxidant for 12 h using different catalysts. Reprinted with permission from ref. 99. Copyright 2015 American Chemical Society Catalyst Conversion (%) Yield (%) Selectivity (%) Zr-MOF-BPYDC-CuBr2 BPYDC-CuBr2 Zr-MOF-BPYDC No catalyst 88.5 28.5 23.0 13.4 84.3 21.2 20.1 8.5 95.3 74.4 87.3 63.2 since the epoxide group is a key intermediate in the industrial synthesis of polymers and bioactive molecules.108,180,181 However, traditional transition metal compounds such as homogeneous catalysts generally suffer from product contamination and poor recyclability. In order to solve these problems, Toyao et al. immobilized CuBr2 into Zr-MOF-BPYDC to obtain Zr-MOF-BPYDC-CuBr2 through a PSM method, which was then used as the catalyst for the selective oxidation of cyclooctene to produce cyclooctene oxide using TBHP as the oxidant.99 Zr-MOF-BPYDC-CuBr2 was found to show higher catalytic activity and selectivity than Zr-MOF-BPYDC and (BPYDC)CuBr2 in this reaction. The yield of the cyclooctene oxide after 12 h reaction period reached 84.3% with 95.3% selectivity (Table 2). Moreover, Zr-MOF-BPYDC-CuBr2 exhibited an excellent recyclability without a significant loss of catalytic activity over three cycles due to the highly stable framework of the Zr-MOF. It should be pointed out that the observed high catalytic activity can be ascribed to the larger pore size (7 Å) in the MOF compared with the size of the cyclooctene molecule (5.5 Å), which facilitates the free diffusion of cyclooctene inside the pores. Furthermore, the formation of Cu-peroxo species with TBHP might also be favorable for accelerating the catalytic process. In addition, regioselective C–H oxidations with Zr-MOF catalysts were also investigated. Fei et al. used UiO-66PdTCAT with site-isolated Pd to catalyze the C–H oxidation of benzo[h]quinoline with iodobenzene diacetate [PhI(OAc)2] as the oxidant (Fig. 26).108 It was found that the metalated Pd(thiocatecholato) sites exhibited highly efficient, reusable, and selective catalysis for the oxidation of the aromatic C–H bond. Nearly 99% yield of methoxy-functionalized benzo[h]quinoline Fig. 26 Schematic representation of the regioselective C–H oxidation of benzo[h]quinolone catalyzed by UiO-66-PdTCAT. Reprinted with permission from ref. 108. Copyright 2015 American Chemical Society. Chem. Soc. Rev. was achieved using UiO-66-PdTCAT (5 mol% in Pd) as the catalyst. In contrast, no obvious catalytic activity was observed for both pristine UiO-66 and UiO-66-TCAT. Moreover, UiO-66-PdTCAT could be recycled at least five times without a significant decrease in the product yields (92–99%). In addition, such chelate-directed oxidation could also be expanded to halogenation of benzo[h]quinoline, in which N-halosuccinimides served as both oxidants and halogenating reagents. The yield of monochlorinated benzo[h]quinoline could reach up to 95%. The high catalytic activity and excellent recyclability of the UiO-66-PdTCAT catalyst in converting C–H bonds to ethers and aryl-halides could be attributed to the strong covalent metalthiocatecholato binding. It is also interesting that Zr-MOF catalysts can be used in water oxidation reactions. Wang et al. adopted a mix-and-match synthetic strategy to obtain Ir(III) functionalized Zr-MOFs by replacing the ligands in UiO-67 with Cp*Ir(DCPPY)Cl, [Cp*Ir(BPY)Cl]Cl, and [Ir(DCPPY)2(H2O)2]OTf (Cp* = pentamethylcyclopentadienyl and H2DCPPY = 2-phenylpyridine-5,4 0 -dicarboxylic acid).182 The obtained MOFs were used for water oxidation with cerium ammonium nitrate (CAN) as the oxidant. All of these Zr-MOFs were proven to be effective for the water oxidation reaction with high turnover frequencies (TOFs) up to 4.8 h1. In addition, these Zr-MOF catalysts retain their integrity after reactions and can be recovered from the reaction systems. However, compared with the homogeneous catalysts, those Zr-MOFs displayed lower TOFs because CAN is apparently too large to enter the MOF pores, where catalytic active sites are located. 4.1.3 Biomimetic catalysis. MOFs have been attracting intense interest for biomimetic catalysis, where the self-destruction of the catalyst and the dilution of active site density can be effectively avoided. However, many MOFs are not stable enough to be applied in biomimetic catalysis, especially in an aqueous medium. In addition, most MOFs are microporous, limiting the access of large substrate molecules to catalytic active sites and slowing down the catalysis rate. In contrast, ultrahigh stable Zr-MOFs with large mesopores and rich accessible redox sites are promising for biomimetic catalysis. PCN-222(Fe) with large 1D open channels (37 Å) prepared by Feng et al. exhibited exceptional stability, even in concentrated hydrochloric acid, which is favorable for biomimetic catalysis.144 Three standard oxidation assays with pyrogallol, 3,3,5,5-tetramethylbenzidine, and o-phenylenediamine as substrates were carried out to check the catalytic performances of PCN-222(Fe) (Fig. 27). For the pyrogallol oxidation reaction, the derived kcat (catalytic activity) value using PCN-222(Fe) (16.1) was seven times higher than that of free hemin (2.4). The derived Km (binding affinity) value of 0.33 was lower than that of natural horseradish peroxidase (HRP) enzyme (0.81), which indicated strong affinity between the substrate and the PCN-222(Fe) framework. For the other two substrates, PCN-222(Fe) also showed superior catalytic activity to free hemin. These outstanding catalytic performances could be attributed to the high density of porphyrin metal active centers in PCN-222(Fe). Simultaneously, Chen et al. reported MMPF-6 as a biomimetic catalyst for the oxidation of pyrogallol (THB) and 2,2 0 -azinodi(3-ethylbenzothiazoline)-6-sulfonate (ABTS) with oxygen and electron transfer capabilities, respectively.160 This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Fig. 27 (a) Structure of PCN-222(Fe); (b) schematic representation of peroxidase; (c) the reactions catalyzed by the peroxidase-like catalyst PCN-222(Fe).144 MMPF-6 exhibited a very fast initial rate of 1.39 103 mM s1 for purpurogallin formation and 8.18 104 mM s1 for ABTS + formation, which are about one-third and one-fourth as fast as that of free heme protein Mb (oxygen storage protein), respectively. Thus, it was indicated that MMPF-6 was highly active for oxidation catalysis both involving oxygen transfer and electron transfer. Moreover, this Zr-MOF also showed excellent adaptability for organic solvents and stable reusability. On the other hand, the catalytic decomposition of phosphate ester bonds is of great importance for the destruction of chemical warfare agents (CWAs).183–186 Considering that Zr–OH–Zr bonds are similar to the bimetallic Zn–OH–Zn active site found in phosphotriesterase enzymes, Katz et al. used UiO-66 as a biomimetic catalyst for the methanolysis and hydrolysis of phosphatebased nerve agent simulants, dimethyl 4-nitrophenyl phosphate (DMNP) and p-nitrophenyl diphenyl phosphate (PNPDPP).187 It was found that UiO-66 had fast half-lives of nearly 45 and 10 min for the two solvolysis reactions, respectively, benefitting from the functions of strong Lewis-acidic Zr(IV) and bridging hydroxide anions on its pore surface. Following this, Mondloch et al. also used NU-1000 as the catalyst for the catalytic hydrolysis of DMNP and CWA O-pinacolyl methylphosphonofluoridate (GD).188 77% conversion was achieved over the course of 60 min with a measured half-life of 15 min. In particular, dehydrated NU-1000 was found to be remarkably active for the hydrolysis of DMNP, giving a half-life of only 1.5 min and exhibiting 100% conversion after approximately 10 min. For the catalytic hydrolysis of GD, the measured half-life was as low as 3 min in the presence of N-ethylmorpholine buffer. Extending this work, a series of Zr-MOFs with different structural features were applied in the degradation of DMNP by the same group.189 Among them, MOF-808 showed the highest hydrolysis rates. The observed TOF was up to 350-fold compared with those of other checked Zr-MOFs, while the half-life was less than 0.5 min. This journal is © The Royal Society of Chemistry 2016 Review Article In addition, López-Maya et al. reported the introduction of lithium alkoxide into the UiO-66 framework (UiO-66@LiOtBu) to improve the phosphotriesterase activity for the capture and hydrolytic degradation of CWA analogues.123 The observed halflife and TOF for the hydrolysis of P–F, P–O, and C–Cl bonds were 5 min and 0.13 min1, 25 min and 0.02 min1, and 3 min and 0.17 min1, respectively. Inspired by such a boosting of the phosphotriesterase catalytic activity, this material was integrated onto textiles as self-detoxifying filters for CWAs. 4.1.4 Electrocatalysis. Due to their high stability, another emerging application of Zr-MOFs is electrocatalysis for oxygen reduction reaction (ORR) as precursors. Lin et al. used Fe-porphyrinic carbon CPM-99X/C (X = H2, Zn, Co, Fe) composites, derived from the carbonization of CPM-99X at 700 1C as the electrocatalysts for ORR.161 As shown in Fig. 28, among the four materials, CPM-99Fe/C exhibited the best electrocatalytic activity, with the ORR peak at the most positive potential of 0.836 V and the onset and half-wave potentials of 0.950 and 0.802 V, respectively, being comparable to the commercial Pt/C catalyst. Moreover, it exhibited long-term durability and was methanol-tolerant, superior to Pt/C. The Koutecky–Levich (K–L) plots from rotating ring electrode (RDE) polarization curves of CPM-99Fe/C ranging from 0.1 to 0.6 V presented a good linear feature, with the calculated electron-transfer number of 4.1 (0.1), indicating the complete reduction of O2. In addition to the basic medium in use, the CPM-99Fe/C catalyst also exhibited high activity in O2-saturated acidic medium, illustrating its high adaptability for solvent medium. The prominent performances of CPM-99Fe/C could be attributed to the unique structural features of the parent Zr-MOF framework, including high porosity, thermal stability, and uniform distribution of FeN4 active sites, as well as the hard-templating effect of its rigid skeleton. 4.1.5 Hydrogenation catalysis. Hydrogenation of olefins is one important processing technology in the petrochemical industry, which can improve the quality of oil products and the yield of light oil. Currently, industrial hydrogenation reactions still rely on noble-metal catalysts, which suffer from high costs and inherent toxicity. Exploring the substitution of precious metal catalysts with earth-abundant and less toxic base metals thus occupies the forefront of contemporary molecular catalysis research.190–197 However, base metals in the homogeneous catalytic system are inclined to deactivation. To avoid this problem, Manna et al. afforded robust and highly active sal-M-MOFs (M = Fe or Co) as single-site solid catalysts for olefin hydrogenation.190 Sal-M-MOF materials were designed by the incorporation of an orthogonal secondary functional group salicylaldimine (sal) into the UiO-67 backbone along with postsynthetic metalation with iron and cobalt salts (Fig. 29). The obtained sal-Fe-MOF displayed excellent catalytic activity in the hydrogenation reaction of a wide variety of aliphatic and aromatic terminal alkenes. With only a 0.05–0.01 mol% catalyst loading, styrene and its alkoxy and halide derivatives could be completely hydrogenated. The yields were up to 93–100%. Moreover, the Sal-Fe-MOF also displayed high turnover numbers (TONs) up to 145 000 for 1-octene hydrogenation in 8 h with only 6.5 104 mol% of the catalyst. Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Chem Soc Rev Fig. 28 (a) Structure of CPM-99; (b) cyclic voltammetry (CV) curves and (c) RDE polarization curves of CPM-99X/C and 20% Pt/C in 0.1 M KOH solution; (d) RDE polarization curves of CPM-99Fe/C at different rotation rates. Inset: K–L plot of J1 vs. o1 at different potentials; (e) RDE polarization curves of CPM-99X/C and 20% Pt/C in 0.1 M HClO4 solution. Reprinted with permission from ref. 161. Copyright 2015 American Chemical Society. Fig. 29 The alkene hydrogenation catalyzed by sal-M-MOFs obtained by the postsynthetic metalation of sal-MOF. Reprinted with permission from ref. 190. Copyright 2014 American Chemical Society. In addition, the sal-Fe-MOF could be recycled and reused at least eight times without the loss of catalytic activity. Interestingly, the catalyst could be regenerated after deactivation and further reused for another six cycles with a simple treatment of ten equivalents of NaBEt3H followed by THF washing. 4.1.6 Brønsted acid catalysis. Brønsted acid catalytic reactions are mainly utilized in hydrocarbon cracking and restructuring of the petroleum refining industry. The catalysts used to activate hydrocarbons are mainly superacids (Hammett acidity functions H0 r 12). Jiang et al. treated MOF-808 with aqueous sulfuric acid for 1 day to generate a sulfated analogue, MOF-808-2.5SO4 as a superacid (Fig. 30).198 Based on the standard catalytic acid/base test reactions of the cyclization of citronellal to isopulegol and a-pinene isomerization, it was shown that Brønsted acidity was introduced into MOF-808-2.5SO4, which was catalytically active in various Brønsted acid-catalyzed reactions including Friedel–Crafts acylation, esterification, and isomerization. Chem. Soc. Rev. Fig. 30 Schematic representation for constructing superacid MOF-8082.5SO4. Reprinted with permission from ref. 198. Copyright 2014 American Chemical Society. 4.1.7 Photocatalysis. The harvesting of solar energy and conversion into renewable energy has attracted increasing attention due to global energy consumption and increasing concern with environmental contamination. Numerous photocatalysts have been investigated for harvesting solar energy. With respect to MOFs, pristine UiO-66 as a photocatalyst was first applied for hydrogen evolution, which showed poor photocatalytic activity due to the absence of efficient catalytic sites for substrate activation and inefficient electron transfer between the catalyst and substrates. In order to enhance photocatalytic performance, an effective way is to build elaborate hybrid photocatalytic systems by introducing functional entities. In 2011, Wang et al. doped [Re(CO)3(5,5 0 -BPYDC)Cl] into the UiO-67 framework using a mix-and-match synthetic strategy for photochemical CO2 reduction (Fig. 31).182 The activity of the resulting Zr-MOF catalyst was examined in CO2-saturated This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article degraded by UiO-66(AN). The high performances could be attributed to the extended visible-light response generated by the introduction of a greater conjugation system of AN. Moreover, the structural integrity of this MOF after the photocatalytic reactions was retained, which guarantees its potential practical application. Meanwhile, Wang et al. used UiO-66(Ti) nanocomposites prepared by a modified post-grafting method in the removal of methylene blue (MB).201 The removal efficiency could reach 87.1%, which may be attributed to the synergistic effect of the adsorption and photo-degradation. 4.2 Fig. 31 Structure model of the doped UiO-67 and the illustration of its photocatalytic application in CO2 reduction and organic transformations.182 acetonitrile (MeCN). The results showed that this catalyst leads to the selective reduction of CO2 to CO under light with the TON of 10.9, which is three times higher than that of complex Re(CO)3(5,5 0 -BPYDC)Cl. Extending this work, phosphorescent [IrIII(PPY)2(DCBPY)]Cl (H2L5) and [RuII(BPY)2(BPYDC)]Cl2 (H2L6) (PPY = 2-phenylpyridine, BPY = 2,2 0 -bipyridine) were also incorporated into the UiO-67 framework for photocatalytic organic transformations (aza-Henry reaction, aerobic amine coupling, and aerobic oxidation of thioanisole) (Fig. 31). Both generated Zr-MOFs, Zr6(m3-O)4(m3-OH)4(L5)6 and Zr6(m3-O)4(m3-OH)4(L6)6, showed high catalytic efficiency towards aza-Henry reactions between nitromethane and methoxy-substituted tetrahydroisoquinoline, with conversion as high as 96%. For the aerobic amine coupling, the catalytic performances of the latter MOF Zr6(m3-O)4(m3-OH)4(L6)6 were comparable to those of homogeneous molecular catalysts. Furthermore, Zr6(m3-O)4(m3-OH)4(L6)6 also represented excellent selectivity towards the aerobic oxidation of thioanisole, giving a conversion as high as 73% after 22 h. Meanwhile, Sun et al. used NH2-UiO-66(Zr) as a photocatalyst for the CO2 reduction, which exhibited high catalytic activity.199 A proposed mechanism was that under visible-light irradiation, the photoinduced electron from the excited ligand can be transferred to the Zr oxo clusters generating Zr(III), which reduce CO2 to HCOO with TEOA (triethanolamine) serving as a hydrogen source. Moreover, Lee et al. prepared mixed metal (Zr/Ti), mixed ligand UiO-66 derivative Zr4.3Ti1.7O4(OH)4(BDC-NH2)5.17(BDC-(NH2)2)0.83 using the PSE strategy, which could also act as an effective photocatalyst for CO2 reduction to HCOOH under visible light irradiation in the presence of TEOA as a sacrificial base and 1-benzyl-1,4-dihydronicotinamide (BNAH) as a sacrificial reductant.200 Such a mixed ligand strategy introduced new energy levels into the band structure of the MOF and thus enhanced the photocatalytic activity. In addition, Zr-MOFs have also served as the photocatalyst for the degradation of dyes. Pu et al. performed photocatalytic reactions for the degradation of methyl orange with UiO66(AN).150 In this study, 65% of methyl orange was effectively This journal is © The Royal Society of Chemistry 2016 Molecule adsorption and separation The remarkable properties that distinguish MOFs from other porous solid materials are their permanent porosity, ultrahigh specific surface area, as well as feasible and tailorable functionalization. Combined with unique chemical stability, even in the presence of water, Zr-MOFs showed promising performances for molecule adsorption and separation applications.202–204 The porosity and surface area of Zr-MOFs are important factors for molecule adsorption and separation, which can be tuned by controlling the framework topology and ligand size. Based on the topology-guided design strategy, Gutov et al. prepared well-designed NU-1100 with a pore volume of 1.53 cm3 g1 and a BET surface area of 4020 m2 g1.83 No obvious variation of crystallinity and porosity was observed after soaking in liquid water for 24 h, indicating its excellent stability against water. NU-1100 was then checked for the H2 storage application. It was found that the total volumetric H2 adsorption of NU-1100 was 43 g L1 and gravimetric uptake is 0.092 g g1 at 65 bar and 77 K, which are comparable with those of most reported MOFs. In addition, the gravimetric and volumetric CH4 storage capacities at 65 bar and 298 K of NU-1100 were approximately 0.27 g g1 and 180 vSTP/v, respectively, which ranked this Zr-MOF among the most promising methane-storage materials. Functionalization of the pore surface is also an efficient approach for enhancing molecule adsorption capacity of MOFs. In order to achieve high guest molecule adsorption of Zr-MOFs, extensive efforts have been devoted to increasing the affinity between the framework pore surface and the guest molecules. Two of the most effective methods are the generation of active sites (such as open metal sites) by regulating the template agent and the encapsulation of functional groups on the pore surface. By varying the concentration of modulator agent, Wu et al. achieved a high concentration of missing-linker defects in the UiO-66 framework, yielding a dramatic increase of both pore volume and surface area up to B150% and B60%, respectively, which greatly affect gas adsorption behaviors in the resulting material.55 The modified UiO-66 with most defects showed a B50% higher uptake of CO2 than that of the one with the least defects at 35 bar. On the other hand, immobilization of functional groups to increase adsorption sites appears to be a promising strategy for enhancing gas capacity. However, the conventional introduction of functional groups often results in the reduction of specific surface area due to space occupation or blockage by functional moieties, which is a negative effect on adsorption capacity of resulting materials. In this respect, the Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article heterocyclic ligands without declining original pore spaces of parent frameworks are the best choices. Li et al. synthesized UiO(BPYDC), in which the bipyridyl moieties were incorporated as free Lewis basic sites in the framework.154 The high density of basic sites on the pore surface of UiO(BPYDC) endowed this Zr-MOF with high uptake capacities towards H2, CO2, and CH4. H2 uptake of this MOF was 5.7 wt% at 77 K and 20 bar, 26.6% higher than that in UiO-67 (4.5 wt% under the same conditions), being among the top rank of H2 uptakes in MOF materials. The CO2 uptakes at 0.15 and 1 bar were 1.5 and 8.0 wt%, respectively. At 293 K and 20 bar, the adsorbed amount of CH4 was 12.2 wt% (7.63 mmol g1) in this MOF. Moreover, the predicated CH4 storage capacity (10.52 mmol g1) reached the Department of Energy (DOE) target of 180 vSTP/v at ambient temperature and 35 bar, making UiO(BPYDC) one of the few MOFs that meet the DOE target. In addition, Yoon et al. reported Zr-BTDC, being composed of electronegative sulfur-containing ligands as strong gas-binding sites.153 In spite of the lower surface area of this MOF contrasted against UiO-67, it showed higher H2 and CO2 sorption capacities of 2.3 wt% at 77 K and 4.05 mmol g1 at 273 K, respectively. Apart from adsorption and storage, the selective adsorption and separation of molecules is also one of the most attractive research topics in Zr-MOFs. Recently, selective CO2 capture from gas mixtures (especially CO2/CH4 and CO2/N2) has been widely investigated due to its practical significance in halting the greenhouse effect.203 To improve the selectivity of CO2 adsorption, Wang et al. used carbonyl- and sulfone-functionalized ligands and isolated two UiO-67 analogues, Zr6(m3-O)4(m3-OH)4(FDCA)6 (BUT-10) and Zr6(m3-O)4(m3-OH)4(DTDAO)6 (BUT-11) (Fig. 32).205 The adsorption isotherms of pure CO2 of the Chem Soc Rev evacuated UiO-67, BUT-10 and BUT-11 demonstrated that the maximal CO2 uptakes were 22.9, 50.6, and 53.5 cm3 g1 at 298 K and 1 atm, respectively. It is clear that the adsorption capacities of CO2 in BUT-10 and -11 were more than double that of parent UiO-67. The functionalized BUT-10 and -11 presented enhanced heats of adsorption towards CO2 (21.8 and 25.9 kJ mol1, respectively), illustrating strong interactions between CO2 and their frameworks. Moreover, the evaluated selectivities for CO2/CH4 were 2.7–2.9, 5.1–5.2, and 9.0–9.2, while those for CO2/N2 were 9.4–9.9, 18.6–22.9, and 31.5–43.1 in UiO-67, BUT-10, and BUT-11, respectively. Clearly, the selectivities of CO2/CH4 and CO2/N2 were enhanced greatly in BUT-10 and -11 in contrast to those in UiO-67. Toxic Hg0 vapor is difficult to remove from some systems because of its insolublility. Based on hard-soft acid–base theory, the soft base S atom is apt to preferentially interact with soft metal Hg. The introduction of a thiophene group into Zr-MOFs can provide the ability to capture and sense Hg0 vapor.206–209 Using a thiophene-based ligand DTDC2, Wang et al. synthesized ultrastable Zr-DTDC, which can remain intact in aqueous solutions over a wide pH range from 0 to 12.152 The exceptional stability makes it an excellent adsorbent for Hg0 vapor in harsh environments. It was found that after the adsorption of Hg0 vapor the color of Zr-DTDC crystals changed from white to yellow, and about 70% of photoluminescence (PL) intensity of the MOF was suppressed. Clearly, the remarkable fluorescence quenching allows Zr-DTDC to function as an optical sensor for detecting Hg0 vapor. Although molecule adsorption in MOFs has been widely studied, water adsorption has been rarely reported because most MOFs are not sufficiently stable in the presence of water. Fig. 32 CO2 adsorption isotherms at 298 K and IAST-predicted selectivities toward CO2/N2 in BUT-11, BUT-10, and UiO-67. Reprinted with permission from ref. 205. Copyright 2014 American Chemical Society. Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev However, Zr-MOFs with their remarkable stability can exhibit high water uptake capacity, being useful in some practical applications, such as drying and sensing. Furukawa et al. proposed three criteria for consideration in the design of MOFs with excellent water absorption performance, involving condensation pressure of water in the pores, uptake capacity, recyclability and water stability of the material.155 Based on these criteria, they prepared a series of Zr-MOFs including MOF-802, Zr6(m3-O)4(m3-OH)4(NDC-(OH)2)6 (MOF-805), Zr6(m3-O)4(m3-OH)4(BPDC-(OH)2)6 (MOF-806), MOF-808, MOF-812, and MOF-841 by connecting Zr6(m3-O)4(m3-OH)4(CO2)n (n = 6, 8, 10, or 12) SBUs with different carboxylate ligands. The water sorption properties of those MOFs along with other reported Zr-MOFs were evaluated. Among them, MOF-801 and MOF-841 showed excellent performances: the former took up 22.5 wt% of water at P/P0 = 0.1 and water uptake of the latter was 44 wt% at P/P0 = 0.3 (Fig. 33). Interestingly, the two Zr-MOFs showed a step water uptake and maintained stable adsorption after five consecutive cycles. Thus, it was anticipated that with these excellent characteristics, MOF-801 might be potentially applied in advanced thermal batteries and MOF-841 was a good candidate to be used in the capture and release of atmospheric water in remote desert areas, respectively. Such good performance could be attributed to the fact that the appropriate pore size of MOF-801 and MOF-841 can lead to the aggregation of water molecules resulting from the formation of hydrogen bonds between water and Zr-SBUs. Selenium is a naturally occurring element which is essential for human health. Howarth et al. examined the adsorption ability of seven porous, water stable Zr-MOFs for the removal of selenate and selenite anions from aqueous solutions.210 Among them NU-1000 exhibited the fastest adsorption rate and the highest gravimetric adsorption capacity. Such high performance can be attributed to the large apertures and substantial Fig. 33 (a) Structure of MOF-841; (b) recycled performance of water uptake in MOF-841 at 25 1C; (c) structure of MOF-801; (d) recycled performance of water uptake in MOF-841 at 25 1C. Reprinted with permission from ref. 155. Copyright 2014 American Chemical Society. This journal is © The Royal Society of Chemistry 2016 Review Article numbers of substitutionally Zr(IV) coordination sites in NU-1000. Then, NU-1000 was further used for the extraction of SO42 from water solution.211 The overall maximum adsorption capacity could reach 56 mg g1, within only one minute. In addition, NU-1000 displayed high selectivity even in the presence of high concentration of competitive anions and could be recycled without a decrease of adsorption capacity. Highly toxic and persistent organic dyes in wastewater have led to serious environmental pollution problems.212 Recently, several MOFs have been explored for the adsorption and capture of pollutants including organic dyes.212,213 Lv et al. synthesized Zr6(m3-O)4(m3-OH)4(EDDB)6 (BUT-30) with a UiO-66 type structure, and used it to adsorb dyes.214 Interestingly, through checking thirteen different dyes it was found that BUT-30 was only able to adsorb cationic dyes such as rhodamine B but exclude others. This adsorption selectivity may be accounted by the negatively charged framework generated from the deprotonation of coordinated OH groups in involved [Zr6(m3-O)4(m3-OH)4]12+ clusters, which are prone to selectively absorb cationic molecules. Herbicides and pesticides in aqueous solution can also be adsorptively removed by Zr-MOFs. Zhu et al. reported the removal of glyphosate (GP) and glufosinate (GF) with UiO-67 based on its abundant Zr–OH groups which served as natural anchors for GP and GF molecules.215 Due to the strong affinity towards the phosphoric groups in these guest molecules and the moderate pore size in UiO-67, the adsorption capacities were as high as 3.18 mmol (537 mg) g1 for GP and 1.98 mmol (360 mg) g1 for GF, which outclassed other reported adsorbents. In addition, Seo et al. used UiO-66 for the adsorptive removal of methylchlorophenoxypropionic acid (MCPP) from water.216 The result illustrated that UiO-66 had a fast adsorption rate (nearly 30 times that of activated carbon) and a very high adsorption capacity, especially at low MCPP concentrations (B7.5 times at 1 ppm MCPP). Moreover, UiO-66 could be reused after washing with a solvent. Such rapid and high uptakes of MCPP may be attributed to its strong affinity with the MOF framework generated by p–p interactions and electrostatic interactions. Bioaccumulation of these persistent pharmaceuticals and personal care products (PPCPs) is a threat to aquatic life. Diclofenac sodium (DCF) is the most frequently detected PPCP, the aqueous-phase adsorption of which using UiO-66 and functionalized UiO-66s (with SO3H/NH2) was studied by Hasan et al.217 Those Zr-MOFs performed well in terms of adsorption kinetics and capacities. In particular, at low concentration the adsorption capacity for SO3H-functionalized UiO-66 is nearly 13 times higher than the traditional adsorbent activated carbon, which has a great significance in practical application. In addition, nuclear power has been proposed as an alternative energy source for many years.218 However, the negative effects such as release of legacy nuclear waste and environmental contamination currently are awaiting a disposal solution. Recently, Zr-MOFs were investigated to be used in nuclear power-related processes. Abney et al. treated UiO-66 with NaOH, Na3PO4, and H3PO4 to remove its BDC2 ligands, leading to ZrOx, ZrOxyPhos, and ZrPhos (Fig. 34).219 Those resulting porous inorganic materials preserved the original morphology of the parent UiO-66, with useful Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Fig. 34 Schematic representation of the ligand extraction process in UiO-66 to form the porous amorphous material, ZrPhos (top) and an efficient radionuclide sequestration process with ZrPhos (bottom). Reprinted with permission from ref. 219. Copyright 2014 American Chemical Society. surface functionalities (hydroxyl and phosphoryl groups), as well as high stability and porosity. Leveraging those excellent properties, these inorganic materials as sorbents were investigated for several nuclear power-related processes, including decontamination of high-level nuclear waste (HLW), lanthanide (Ln) extraction, and remediation of radioactive seawater. It was found that ZrOxyPhos was superior to others in decontaminating the HLW stimulant. In the Ln separation application, ZrPhos could extract approximately 99% Ln from a slightly acidic aqueous solution. Just as expected, the adsorption properties of these new sorbents in both capacity and kinetics were superior to the current state-of-the-art materials, which should be attributed to the high porosity, well-defined morphologies, imparted surface hydroxyl groups, as well as accessibility of coordinating metal sites in these porous inorganic materials. 4.3 Drug delivery In recent years, increasing attention has been paid to the application of MOFs as drug delivery vehicles.220,221 However, most MOFs are difficult to meet the prerequisites of this application due to the limitation of their poor chemical and thermal stabilities.222–225 In contrast, stable, non-toxic Zr-MOFs with ease of nanoparticle formation are considered ideal candidates for drug delivery.47 It has been well documented that zirconia oxides can be modified by phosphonates relying on the strong coordination between the Zr atom and the phosphonate O atom.226 As a result, Zr-MOFs with regularly arrayed Zr–O clusters are expected to be suitable for the controllable capture and release of alendronate (AL) molecules. Accordingly, Zhu et al. used UiO-66 nanoparticles (NPs) as carriers for AL delivery.227 It was found that about 42.7% of adsorbed AL could be released by UiO-66 NPs at pH = 7.4 Chem. Soc. Rev. Chem Soc Rev in 60 h, whereas over 59% of the drug was released at pH = 5.5 within the same time duration (Fig. 35a). It was proposed that the observed pH-responsive release character resulted from the protonation of phosphates in acidic medium.228 To further verify that the UiO-66 NPs could be effectively endocytosized by cancer cells, UiO-66-FMN, made by grafting fluorescent flavin mononucleotide (FMN) onto UiO-66, was incubated into HepG2 cells for 4 h. As shown in Fig. 35b, the UiO-66-FMN NPs were remarkably internalized within a short period and distributed intensively in the cytoplasm. In addition, the therapeutic effect of AL-UiO-66 was tested through in vitro cytotoxicity measurements against cancer cells. It was observed that in the first 24 h of incubation for both cancer cells HepG2 and MCF-7, the cytotoxicity of AL-UiO-66 was slightly lower than that of free AL with an equivalent dose. However, after 48 h of incubation, the AL-UiO-66 NPs caused a higher mortality rate of cancer cells than free AL (Fig. 35c and d). It was also interesting that the IC50 (a measure of drug effectiveness) values of AL-UiO-66 were far below those of free AL. 4.4 Fluorescence sensing Fluorescent materials have riveted significant attention due to their wide applications, especially as light-emitting diodes (LEDs) and solid state sensors.229–233 In spite of a wide variety of organic fluorescent molecules being explored, they often suffer from self-quenching and consequently low quantum yields.234,235 Recently, researchers conceived that fixture of fluorophores into MOFs could solve these problems effectively.236–240 Generally, the fluorescence of MOF materials originates from three aspects: the organic ligands, lanthanide metal cations in Ln-MOFs, as well as the charge transfers between metal ions and organic ligands.240–245 As for Zr-based fluorescent MOFs, in most cases fluorescence originates from organic ligands. As a well-known fluorophore, tetraphenylethylene (TPE) is attractive due to its aggregation-induced emission property.246 Wei et al. prepared the TPE-based Zr-MOF, PCN-94, and studied its fluorescence properties (Fig. 36).163 The twisted ligand conformation in PCN-94 led to bright blue fluorescence emission at 470 nm, which displayed a large blue shift compared to the 545 nm yellow emission of the free ligand H4ETTC. Moreover, the quantum yield of PCN-94 was as high as 99.9 0.5% under Ar, which could be primarily attributed to the strong coordination between the ETTC4 ligand and Zr atoms. In addition, the fluorescence lifetime and relative intensity increased anomalously when heating the solid PCN-94 from cryogenic to ambient temperatures. It was believed that the change in ligand conformation had a tremendous influence on the photophysical properties of the MOFs. This light-emitting Zr-MOF with unusual photoluminescence properties can be potentially applied in molecular electronics or sensor technologies. Fluorescent PCN-225 prepared by Jiang et al. exhibited exceptional chemical stability in aqueous solutions with a remarkably wide pH range.145 Interestingly, the pH-dependent fluorescence of PCN-225 was observed (Fig. 37). The emission intensity rose along with the increase of pH value. However, it was not a simple linear relationship. When the pH was below 5, an This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article Fig. 35 (a) Free AL release profile from the dialysis bag at pH 7.4 (black) and the cumulative release profiles of AL from UiO-66 NPs in 5 mM PBS at pH values of 5.5 (red) and 7.4 (green) at 37 1C; (b) the merged confocal image of HepG2 cells after their incubation with 100 mg mL1 of UiO-66-FMN for 4 h at 37 1C; cell viabilities of (c) HepG2, (d) MCF-7 cells incubated with free AL and AL-UiO-66 at different concentrations for 24 and 48 h. Reprinted with permission from ref. 227. Copyright 2014 The Royal Society of Chemistry. increase of pH value led to only a slightly increase of fluorescence intensity. The reason for this unique pH-sensitive fluorescence in PCN-225 was that the protons in acidic solution combined with pyridine-type N atoms of the porphyrin center, destroying the p-electron conjugated double-bond system of porphyrin, ultimately leading to significant fluorescence quenching. However, when the pH was above 7, the fluorescence intensity increased significantly in close association with the pH values, which might arise from the deprotonation of the imino group of porphyrin. Based on these findings, PCN-225 was regarded with potential applications in pH sensing, especially in the pH range of 7–10. Capitalizing on the intrinsic fluorescence of MOFs, researchers explored extensive sensing applications.247–259 For instance, high levels of phosphates may lead to the pollution of water and consequent massive death of creatures.260 As discussed above, Zr-MOFs presented high affinity towards phosphoric groups due to the formation of Zr–O–P bonds. Thus, Yang et al. exploited This journal is © The Royal Society of Chemistry 2016 intrinsically fluorescent UiO-66-NH2 as a probe for the selective recognition of phosphate anions in aqueous solution, in which Zr–O clusters played the role of phosphate recognition sites with the ligands as the signal reporters (Fig. 38).261 The fluorescence intensity showed a good linear correlation with the concentration of phosphate. As a result, UiO-66-NH2 as a fluorescent sensor displayed a wide phosphate detection range of 5–150 mM. Moreover, the detection limit was estimated to be 1.25 mM, which is far below the detection requirement of phosphate discharge criteria. Meanwhile, UiO-66-NH2 was also used by Ghosh’s group for the deamination reaction-based detection of aqueous phase nitric oxide (NO), which showed high sensory activity and selectivity, even with potentially interfering species.262 Based on a similar sensing mechanism, UiO-66-N3 and UiO-66-NO2 served as fluorescence turn-on probes, which can detect H2S rapidly and selectively even under physiological conditions.263,264 Extending those studies, Zr6(m3-O)4(m3-OH)4(L)6 (UiO-67@N, H2L = 2-phenylpyridine5,4 0 -dicarboxylic acid) was prepared to selectively detect nitro Chem. Soc. Rev. View Article Online Review Article Chem Soc Rev photoelectric effect. The fast electrons then transfer the energy to anthracene-based emitters by inelastic scattering. Finally, the excited anthracene-based ligand emits the visible photons for detection. Through the synergistic functionalization of Zr-SBUs and anthracene-based bridging ligands, the resultant MOFs exhibited excellent X-ray-to-light converting capabilities superior to the two individual components. Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. 4.5 Fig. 36 (a) ETTC4 linker in PCN-94; (b) PCN-94 framework; (c) solidstate absorption (dashed lines) and emission spectra (solid lines) of PCN94 and H4ETTC at room temperature; (d) photos of PCN-94 and H4ETTC are shown under ambient light and UV light. Reprinted with permission from ref. 163. Copyright 2014 American Chemical Society. explosives in aqueous media.265 Especially for 2,4,6-trinitrophenol (TNP) with concentrations as low as 2.6 mM, fast and high fluorescence quenching was observed, illustrating low detection limitation for UiO-67@N. In contrast, all of the other nitro analytes had minor effects on the fluorescence intensity, indicative of high selectivity towards TNP. Such unprecedented selective fluorescence quenching can be attributed to the occurrence of both electron and energy transfer processes between the MOF and TNP. MOFs are also applicable in metal sensing, mainly involving the detection of Eu(III), Fe(III), Tb(III), Cu(II), and Ag(I).247,252–259 As for Zr-MOFs, Carboni et al. prepared Zr6(m3-O)4(m3-OH)4(H2O)8(L1)4 (H2L1 = H2TPDC-NH2 derivative) with metal binding sites orthogonal to succinic acid, which possesses a broad emission band at 390 nm upon excitation at 330 nm.266 The ligand H2L1 was prepared by treating H2TPDC-NH2 with succinic acid in dry DMSO at room temperature for 12 h. Thus, the chelation of amide groups with transition metal cations leads to changes of the fluorescence intensity. Interestingly, those transition metal ions with unpaired d-electrons presented a significant quenching effect even at low metal concentrations, especially for Mn(II). An exceptionally high KSV value of (0.91 0.04) 106 m1, with the detection limit of o0.5 ppb for Mn(II), was obtained on Zr6O8(H2O)8(L1)4, which is three to four orders of magnitude greater sensitivity than those of previously reported MOFs. In addition to the fluorescence properties, scintillation is also an important emissive phenomenon found in luminescent materials. For example, luminescent organic crystals can work as radiation scintillators for detecting low-energy b-rays and neutrons.267,268 X-ray scintillating Zr-MOFs, Zr6(m3-O)4(m3-OH)4(ADC)6 with anthracene-based bridging ligands H2ADC as emitters, were designed and prepared by Wang et al. (Fig. 39).151 The high atomic number of Zr in the SBUs serves as the effective X-ray antenna. Upon X-ray absorption, the outer-shell electrons of the Zr(IV) ions are converted to fast electrons via the Chem. Soc. Rev. Electrochemistry Recently, there has been rapidly growing interest in electrochemistry applications of MOFs, such as in energy storage and conversion (supercapacitors) or electrode materials, which have been proven to be excellent candidates in this field.269–273 It is mainly because the chemical composition of MOFs can be tuned at the molecular level. Moreover, the high porosity of MOFs is favorable for fast mass transportation of related species. The studies of Zr-MOFs in electrochemistry application mostly involve three aspects: ionic and proton conduction, as well as supercapacitors. Ameloot et al. prepared LiOtBu-grafted UiO-66 upon dehydration of Zr-based clusters followed by lithium alkoxide grafting.122 The grafted tBuO anions shielded the negative charge, leading to a higher mobility of Li(I) ions. As a result, the ionic conductivity of the LiOtBu-grafted UiO-66 was promoted to be 1.8 105 S cm1, which is on par with current solid polymer electrolytes.274,275 The activation energies (Ea) declined as low as 0.18 eV, which is in the range for superionic conductors and even slightly lower than that of H+ conduction in Nafion.276,277 Moreover, the impedance spectra obtained from a setup prepared by placing a pellet of LiOtBu-grafted UiO-66 in a Swagelok cell showed that no Li(I) blocking passivation layer was formed. Furthermore, this cell could be cycled three times and the crystallinity of the MOF was intact after contacting with the Li electrode for three days, indicating a high electrode compatibility. With these properties, the solid electrolytes based on this LiOtBu-grafted UiO-66 are potentially usable in Li-based batteries. On the other hand, ligand defects in MOFs have been demonstrated to have important effects on their pore sizes and surface properties, which subsequently influence their performances, such as in gas storage and catalysis. Interestingly, Taylor et al. first showed that defect control could also enhance the proton conductivity of Zr-MOFs.278 Different from common methods for improving proton conductivity of MOFs by increasing overall acidity, they controlled the formation of ligand defects by introducing stearic acid to change the proton mobility within the UiO-66 framework.279–281 The results demonstrated that the ligand defects were favorable for improving proton mobility. As a result, proton conductivity increased by nearly three orders of magnitude to 6.93 103 S cm1. In addition, supercapacitors represent an important class of energy storage devices because of their high power density.270,283 Choi et al. recently prepared twenty-three different nanocrystals of MOFs (nMOFs) with various structures, geometries and functionalities, which were applied in the fabrication of MOF thin films doped with graphene and then incorporated into This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article Fig. 37 (a) Structure of PCN-225; (b) pH dependent fluorescence intensity of PCN-225 with pH ranging from 0 to 10.2 measured under excitation of 415 nm; (c) fluorescence emission of PCN-225 at different pH values matching the simulation fluorescence intensity; (d) protonation and deprotonation processes of porphyrin involved in the PCN-225 framework in experimental acidic and basic media (pH = 0–10.2). Reprinted with permission from ref. 145. Copyright 2013 American Chemical Society. Fig. 38 A schematic illustration of the phosphate coordination induced fluorescence enhancement effect in MOFs of UiO-66-NH2 for the selective and sensitive detection of phosphate with inorganic nodes as natural recognition sites and organic struts as the fluorescent reporter. Reprinted with permission from ref. 261. Copyright 2015 The Royal Society of Chemistry. This journal is © The Royal Society of Chemistry 2016 Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article Fig. 39 (a) Synthesis of Zr6(m3-O)4(m3-OH)4(ADC)6; (b) scheme showing the X-ray induced generation of fast photoelectrons from heavy metals followed by the scintillation of the anthracene-based linkers in the visible spectrum.151 devices to serve as supercapacitors.282 Among them, Zr6(m3-O)4(m3-OH)4(BPYDC)6 (nMOF-867) exhibited the highest capacitance (Fig. 40). The max stack capacitance (0.644 F cmstack3) and max areal capacitance (5.085 mF cmareal2) were over six and ten times those of activated carbon and graphene, respectively. The gravimetric capacitance, maximum energy, and power densities of nMOF-867 were estimated to be 726 F gnMOF-867/electrode, 6.04 104 Wh cmstack3, and 1.097 W cmstack3, respectively. It is worth noting that at the power density of 0.386 W cmstack3, the energy density (2.86 104 Wh cmstack3) of nMOF-867 was over three times that of activated carbon. Moreover, nMOF-867 based on this supercapacitor could consistently preserve its excellent performances for at least 10 000 cycles under the conditions of exposing to the maximum voltage (0–2.7 V). 4.6 Porous carrier Zr-MOFs are excellent candidates as carriers for various applications because of their porosity and high chemical and thermal stability. The integration of Zr-MOFs with encapsulated functional entities leads to the formation of new multifunctional composites, which could exhibit unique properties superior to the individual components.284,285 Moreover, the multifunctional nature and the remarkable properties of such composites will enrich functional applications of MOFs, typically in catalysis, molecule adsorption, as well as drug delivery. As a mesoporous Zr-MOF, PCN-777 has high chemical and thermal stability with the highest pore volume of 2.8 cm3 g1.147 Moreover, in this MOF each Zr6 cluster is coordinated by six carboxylates, leaving rich terminal OH groups pointing into the mesoporous cages, which renders PCN-777 a suitable candidate as a carrier. Feng et al. used 4-tert-butylbenzolate (tBu) and 4-carboxy-1-methylpyridinium iodide (Me–Py) to modify the internal surface of PCN-777. After that, the originally hydrophilic properties of the internal surface changed into highly hydrophobic. Three functionalized guest molecules, meso–tetra(4-sulfonatophenyl)porphyrin (TSPP), tris(2,20 -bipyridine)-dichlororuthenium(II) ([Ru(bpy)3]Cl2), and tetra-amido macrocyclic ligand catalytic Chem. Soc. Rev. Chem Soc Rev activator (TAML-NaFeB*), were then encapsulated into the modified PCN-777. Due to the formation of hydrogen bonding between peripheral sulfonic groups in TSPP and terminal OH/H2O groups in the Zr-based clusters, the pristine PCN-777 exhibited the highest loading amount and the slowest leaching towards TSPP relatively. Moreover, the Me–Py-modified PCN-777 with the lowest porosity showed the highest volumetric loading amount of anionic TAML-NaFeB*. As expected, the tBu-modified PCN-777 displayed the weakest interactions with both TSPP and TAML-NaFeB* because of the nonpolar feature of the internal pore surface of the modified PCN-777. In contrast, no preferential [Ru(bpy)3]Cl2 loading in any modified PCN-777 was observed since the functional part [Ru(bpy)3]2+ is cationic. Phosphorescent Zr-MOFs can serve as supports to encapsulate molecular photosensitizers for hydrogen evolution reactions (HERs). Initially, Zhang et al. loaded Pt nanoparticles with different sizes into two phosphorescent Zr-MOFs, Zr6(m3-O)4(m3-OH)4(BPDC)5.94(L1)0.06 and Zr6(m3-O)4(m3-OH)4(L2)6 via the MOF-mediated photoreduction of K2PtCl4 (Fig. 41a and b).286 The [Ir(ppy)2(bpy)]Clderived dicarboxylic acids H2L1 and H2L2 were firstly synthesized by treating [Ir(ppy)2Cl]2 with diethyl (2,20 -bipyridine)-5,50 -dicarboxylate (Et2L1) and dimethyl (2,2 0 -bipyridine)-5,5 0 -dibenzoate (Me2L2), respectively, followed by the base-catalyzed hydrolysis. Based on the synergistic photoexcitation of the MOF frameworks and electron injection into the Pt nanoparticles, the obtained Pt@MOFs can serve as effective photocatalysts for hydrogen evolution. It was found that the two Pt@MOFs gave a total Ir-TON of 3400 and 7000, respectively, which are 1.5 and 4.7 times as high as that of homogeneous control [Ir(ppy)2(bpy)]Cl/K2PtCl4 (2200 and 1500) under the same conditions. The enhanced photocatalytic activity of Pt@MOFs was thought to be mainly attributed to the efficient electron transfer between [Ir(ppy)2(bpy )] species and Pt NPs. That is not only favorable for improving hydrogen reduction rates but also avoiding the decomposition of Ir complexes. Then Wang et al. developed the charge-assisted self-assembly process to encapsulate [P2W18O62]6 into Zr6(m3-O)4(m3-OH)4(L)6 bearing [Ru(bpy)3]2+ to get the POM@UiO composite (Fig. 41c and d).287 The [Ru(bpy)3]2+derived dicarboxylate ligand (H2L) was prepared by heating the above Me2L2 followed by base-catalyzed hydrolysis. It was found that with methanol as the sacrificial electron donor, TONs for the hydrogen production could reach a maximum value of 79. However, with triethanolamine as the sacrificial electron donor, the TONs of the photocatalytic HERs reached 540, which was thirteen times higher than that of the homogeneous control. Moreover, POM@UiO could be recycled and reused at least three times with only a slight loss of catalytic activity and crystallinity of the MOF. In addition, Choi et al. embedded Pt nanoparticles (NPs) into the nanocrystalline UiO-66 samples functionalized with –SO3H (S) and –NH3+ (N) or both of them to form Zr-MOF supported composites: Pt C nUiO-66-S, Pt C nUiO-66-N, and Pt C nUiO66-SN.288 By treating three samples with NaCl, triethylamine, and sodium bicarbonate, another three samples Pt C nUiO-66S*, N*, and S*N* embedded with Pt-NPs were obtained. These six samples with different chemical functionalities were applied in the catalytic conversion of methylcyclopentane. It was found that these chemical functionalizations played an This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev Review Article Fig. 40 (a) Construction for nMOF supercapacitors; (b) the stack capacitance of nMOF-867 and comparison to activated carbon and graphene. Reprinted with permission from ref. 282. Copyright 2014 American Chemical Society. Fig. 41 Schematic representation of the synergistic photocatalytic hydrogen evolution process via the photoinjection of electrons from the light harvesting MOF frameworks into the Pt NPs (a) and POMs (c); (b) time-dependent hydrogen evolution curves of Pt@1 (green), Pt@2 (red), and homogeneous control [Ir(ppy)2(bpy)]Cl/K2PtCl4 (blue and black for different Pt/Ir ratios) under optimized conditions; (d) time-dependent HER TONs of POM@UiO with methanol as the sacrificial electron donor. Reprinted with permission from ref. 286 and 287. Copyright 2012 and 2015 American Chemical Society. important role in the product selectivity and the conversion of methylcyclopentane (MCP) to acyclic isomers, olefins, cyclohexane, and benzene. Pt C nUiO-66-S showed the highest selectivity to cyclohexane and benzene (62.4% and 28.6%, respectively), and This journal is © The Royal Society of Chemistry 2016 the catalytic activity was twofold compared with the nonfunctionalized Pt C nUiO-66. For Pt C nUiO-66-N, while the selectivity for C6-cyclic products decreased to o50%, the acyclic isomer selectivity increased to 38.6%. Interestingly, for Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article the Pt C nUiO-66-SN, benzene was the dominant product with olefins and acyclic isomers as byproducts, while no cyclohexane was produced. In addition, Xu et al. modified UiO-66 with predesigned Pt-NPs for the electrocatalysis of H2O2 oxidation.289 The Pt-NPs@UiO-66 composite presented excellent electrocatalytic activity reflected by remarkable anti-interference performance, the extended linear range (5 mM–14.75 mM), the low detection limit, as well as good stability and reproducibility. Except commonly used Pt-NPs and UiO-66, Zhuang et al. incorporated Pd-NPs into DUT-67 to obtain Pd-NP@DUT-67, which was then applied in a Suzuki coupling reaction.290 It was found that the conversion could reach 99%, while the selectivity was 89%. In addition, Pd-NPs@DUT-67 also showed good catalytic properties for the hydrogenation of nitrobenzene, with a conversion of 99% and a selectivity of 99%. Note that Zr-MOFs as carriers are not only used in catalysis but also in the encapsulation of drugs224,291–294 and cosmetics.295–297 D. Cunha et al. chose different functionalized UiO-66-X (X = H, CH3, 2OH, NH2, NO2, Cl, Br, and 2CF3) MOFs to encapsulate bioactive molecules, ibuprofen, and caffeine.298 The functional groups in the MOF played a crucial role in the encapsulation performance. It was demonstrated that the caffeine entrapping was enhanced when the functionalized linker had a large octanol– water partition coefficient and low hydrogen bond accepting ability. In contrast, the ibuprofen loading was optimized when the functional groups of the linker had a large solvent surface area and free volume, as well as low hydrogen bond accepting ability. Moreover, the solvents used for the biomolecule entrapping also significantly impacted the encapsulation performances due to the existence of competitive adsorption between active molecules and the solvents. More importantly, the drug payloads of these Zr-MOFs outperformed conventional porous solids or current drug formulations. Similarly, the studies of Zr-MOFs as carriers for the encapsulation of luminescent molecules have also been performed. The resulting composites displayed excellent sensing properties. For instance, He et al. reported the use of nanoscale UiO68-NH2 as a carrier of fluorescein isothiocyanate (FITC) for the intracellular pH sensing in live cells.299 The resulting composite was demonstrated to have high FITC loadings, effective fluorescence, and remarkable ratiometric pH-sensing properties. The Zr-MOF also remained structurally intact after the rapid and efficient endocytosis process. In addition, Hao et al. used UiO-66-(COOH)2 as a carrier to encapsulate Eu(III) cations to get Eu(III)@UiO-66-(COOH)2, which was applied as a fluorescent probe for Cd(II) detection in aqueous solutions. This luminescent probe showed excellent sensitivity with the detection limit of 0.06 mM and a fast response speed within B1 min. The luminescence intensity of Eu(III) was enhanced to about 8-fold, which could be ascribed to the effective energy transfer from the ligand to Eu(III) generated from the coordination of Cd(II) to Eu(III)@UiO-66-(COOH)2. With well-defined porous structures, MOFs can serve as versatile carriers for the growth of metallic nanostructures in their pores. Additionally, the high surface area and large opening windows of MOFs ensure that the surface of metallic Chem. Soc. Rev. Chem Soc Rev Fig. 42 Schematic representation of MOF-545 as the carrier for the growth of gold metallic nanowires. Reprinted with permission from ref. 300. Copyright 2015 American Chemical Society. nanostructures is readily accessible for appropriate functions for specific applications. Volosskiy et al. used a Zr-MOF, MOF545, with a 1D pore as a carrier to grow ultrafine gold metallic nanowires with a well-controlled shape and size (Fig. 42).300 The intrinsic surface could be preserved because no addition of bulky and long-chain surfactants was used for the nanowire growth inside the MOF. Such a strategy was also extended to the synthesis of a hybrid Pt@MOF. In addition, the synergy of high surface area, porphyrin active sites as well as highly active nanowires contributes to distinct properties and thus provides platforms of hybrid composites for catalytic applications. 5. Conclusion and prospect Although the class of Zr-MOFs is just a member of the large MOF family, starting from the seminal discovery of the UiO-66 series, Zr-MOFs have received widespread attention by virtue of their preeminent chemical and physical properties. In particular, Zr-MOFs inspired great interest of researchers in the last three years, which can be demonstrated through an everescalating number of structures and function explorations. From the point of intriguing properties, Zr-MOFs provide an ideal platform for tapping into the superior applications in the future. Through summarizing a majority of the reported studies of Zr-MOFs, this review provides a holistic generalization of state-of-the-art development of the synthesis, structure, and applications of Zr-MOF materials. A large number of MOF materials with excellent properties and potential applications were thus reviewed. In terms of the design strategies and synthetic methods for constructing Zr-MOFs, the examples of modulated synthesis, isoreticular expansion, topology-guided design, and postsynthetic modification are presented in detail. All of these methods and strategies are successfully used in building various Zr-MOFs, which open a new avenue for the rational design of targeted Zr-MOFs with the enhancement of chemical compatibility for expanding a wide variety of potential applications. In addition, these methods and strategies also provide good guidance and reference for the design and synthesis of other MOFs. However, these strategies are still developed by trial and error to some extent and in-depth understanding into the mechanism of formation is still lacking. So, further exploration is highly desirable in this regard. With regard to the structures, five types of Zr-clusters, Zr6O8, Zr8O6, ZrO6, ZrO7, and ZrO8, as the cores This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev of SBUs have appeared in the literature. We are confident that, with the gradual development of this field, Zr-MOFs based on other SBUs will spring up in the future years. In particular, relying on the use of newly selected/designed organic ligands, the structural types of Zr-MOFs will continuously increase, and more new members of Zr-MOFs with unique properties and functions will emerge. Based on their excellent porosity and structural stability, significant efforts have been put into versatile potential applications of Zr-MOFs, including catalysis, molecule adsorption and separation, drug delivery, fluorescence sensing, electrochemistry, porous carrier and so on. Even if these studies of Zr-MOFs cover a wide scope of applications, distribution of the existing studies in different application aspects is not balanced, which are mainly focused on catalysis right now. Moreover, the overall investigation on Zr-MOFs in practical applications still lacks a systematic approach and depth. In comparison to many other MOF subclasses, such as ZIFs, Cr-MOFs, Al-MOFs, Cu-MOFs, and Zn-MOFs, Zr-MOFs are still at their early stage. Therefore, more extensive investigations are highly desired to explore the great potential applications of various Zr-MOFs. Notwithstanding there are still great challenges to be overcome in the Zr-MOF field. This class of materials remains predominant in the MOF family due to their framework stability and unique properties. Moreover, green and economically viable methods for the scalable synthesis of some Zr-MOFs with repeatable quality have been successfully developed. The massive production of Zr-MOFs with excellent stability implies their unlimited potential in commercial applications. Once practical applications are in place, it will be certainly a leap in the development of MOF materials. With this review, we also hope to inspire researchers with new and smart ideas to stimulate the emergence of new concepts and applications in the field of MOFs, with numerous possibilities offered in this fascinating area. Acknowledgements We thank the financial support of the Natural Science Foundation of China (No. 21322601, 21271015, and 21576006) and the Program for New Century Excellent Talents in University (No. NCET-13-0647). References 1 H.-C. Zhou, J. R. Long and O. M. Yaghi, Chem. Rev., 2012, 112, 673–674. 2 H.-C. Zhou and S. Kitagawa, Chem. Soc. Rev., 2014, 43, 5415–5418. 3 H. Furukawa, K. E. Cordova, M. O’Keeffe and O. M. Yaghi, Science, 2013, 341, 974–986. 4 T. R. Cook, Y.-R. Zheng and P. J. Stang, Chem. Rev., 2013, 113, 734–777. 5 B. Li, M. Chrzanowski, Y. Zhang and S. Ma, Coord. Chem. Rev., 2016, 307, 106–129. This journal is © The Royal Society of Chemistry 2016 Review Article 6 J. Canivet, A. Fateeva, Y. Guo, B. Coasne and D. Farrusseng, Chem. Soc. Rev., 2014, 43, 5594–5617. 7 S. Keskin, T. M. van Heest and D. S. Sholl, ChemSusChem, 2010, 3, 879–891. 8 A. C. Kizzie, A. G. Wong-Foy and A. J. Matzger, Langmuir, 2011, 27, 6368–6373. 9 J. Gascon and F. Kapteijn, Angew. Chem., Int. Ed., 2010, 49, 1530–1532. 10 L. Zhang and Y. H. Hu, J. Phys. Chem. C, 2011, 115, 7967–7971. 11 N. C. Burtch, H. Jasuja and K. S. Walton, Chem. Rev., 2014, 114, 10575–10612. 12 G. Férey, C. Mellot-Draznieks, C. Serre, F. Millange, J. Dutour, S. Surblé and I. Margiolaki, Science, 2005, 309, 2040–2042. 13 I. J. Kang, N. A. Khan, E. Haque and S. H. Jhung, Chem. – Eur. J., 2011, 17, 6437–6442. 14 J. J. Low, A. I. Benin, P. Jakubczak, J. F. Abrahamian, S. A. Faheem and R. R. Willis, J. Am. Chem. Soc., 2009, 131, 15834–15842. 15 X. Liu, Y. Li, Y. Ban, Y. Peng, H. Jin, H. Bux, L. Xu, J. Caro and W. Yang, Chem. Commun., 2013, 49, 9140–9142. 16 C. Montoro, F. Linares, E. Quartapelle Procopio, I. Senkovska, S. Kaskel, S. Galli, N. Masciocchi, E. Barea and J. A. R. Navarro, J. Am. Chem. Soc., 2011, 133, 11888–11891. 17 P. Küsgens, M. Rose, I. Senkovska, H. Fröde, A. Henschel, S. Siegle and S. Kaskel, Microporous Mesoporous Mater., 2009, 120, 325–330. 18 S. S. Iremonger, J. Liang, R. Vaidhyanathan, I. Martens, G. K. H. Shimizu, T. D. Daff, M. Z. Aghaji, S. Yeganegi and T. K. Woo, J. Am. Chem. Soc., 2011, 133, 20048–20051. 19 H. Li, W. Shi, K. Zhao, H. Li, Y. Bing and P. Cheng, Inorg. Chem., 2012, 51, 9200–9207. 20 H. Jasuja, N. C. Burtch, Y.-g. Huang, Y. Cai and K. S. Walton, Langmuir, 2013, 29, 633–642. 21 H. Jasuja, Y.-g. Huang and K. S. Walton, Langmuir, 2012, 28, 16874–16880. 22 D. Ma, Y. Li and Z. Li, Chem. Commun., 2011, 47, 7377–7379. 23 J. G. Nguyen and S. M. Cohen, J. Am. Chem. Soc., 2010, 132, 4560–4561. 24 J. B. Decoste, G. W. Peterson, M. W. Smith, C. A. Stone and C. R. Willis, J. Am. Chem. Soc., 2012, 134, 1486–1489. 25 S. J. Yang and C. R. Park, Adv. Mater., 2012, 24, 4010–4013. 26 W. Zhang, Y. Hu, J. Ge, H.-L. Jiang and S.-H. Yu, J. Am. Chem. Soc., 2014, 136, 16978–16981. 27 J. H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga and K. P. Lillerud, J. Am. Chem. Soc., 2008, 130, 13850–13851. 28 M. Kandiah, M. H. Nilsen, S. Usseglio, S. Jakobsen, U. Olsbye, M. Tilset, C. Larabi, E. A. Quadrelli, F. Bonino and K. P. Lillerud, Chem. Mater., 2010, 22, 6632–6640. 29 J. B. DeCoste, G. W. Peterson, H. Jasuja, T. G. Glover, Y.-g. Huang and K. S. Walton, J. Mater. Chem. A, 2013, 1, 5642–5650. 30 X. Liu, N. K. Demir, Z. Wu and K. Li, J. Am. Chem. Soc., 2015, 137, 6999–7002. 31 L. Valenzano, B. Civalleri, S. Chavan, S. Bordiga, M. H. Nilsen, S. Jakobsen, K. P. Lillerud and C. Lamberti, Chem. Mater., 2011, 23, 1700–1718. Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article 32 T. Devic and C. Serre, Chem. Soc. Rev., 2014, 43, 6097–6115. 33 P. Lu, Y. Wu, H. Kang, H. Wei, H. Liu and M. Fang, J. Mater. Chem. A, 2014, 2, 16250–16267. 34 M. Zhang, Y.-P. Chen, M. Bosch, T. Gentle, K. Wang, D. Feng, Z. U. Wang and H.-C. Zhou, Angew. Chem., Int. Ed., 2014, 53, 815–818. 35 R. G. Pearson, J. Am. Chem. Soc., 1963, 85, 3533–3539. 36 S. Wang, J. Wang, W. Cheng, X. Yang, Z. Zhang, Y. Xu, H. Liu, Y. Wu and M. Fang, Dalton Trans., 2015, 44, 8049–8061. 37 A. Bétard and R. A. Fischer, Chem. Rev., 2012, 112, 1055–1083. 38 Y. He, W. Zhou, G. Qian and B. Chen, Chem. Soc. Rev., 2014, 43, 5657–5678. 39 C. G. Silva, A. Corma and H. Garcia, J. Mater. Chem., 2010, 20, 3141–3156. 40 F. Vermoortele, R. Ameloot, A. Vimont, C. Serre and D. De Vos, Chem. Commun., 2011, 47, 1521–1523. 41 P. S. Bárcia, D. Guimarães, P. A. P. Mendes, J. A. C. Silva, V. Guillerm, H. Chevreau, C. Serre and A. E. Rodrigues, Microporous Mesoporous Mater., 2011, 139, 67–73. 42 S. Chavan, J. G. Vitillo, D. Gianolio, O. Zavorotynska, B. Civalleri, S. Jakobsen, M. H. Nilsen, L. Valenzano, C. Lamberti, K. P. Lillerud and S. Bordiga, Phys. Chem. Chem. Phys., 2012, 14, 1614–1626. 43 Q. Yang, V. Guillerm, F. Ragon, A. D. Wiersum, P. L. Llewellyn, C. Zhong, T. Devic, C. Serre and G. Maurin, Chem. Commun., 2012, 48, 9831–9833. 44 D. B. N. Lee, M. Roberts, C. G. Bluchel and R. A. Odell, ASAIO J., 2010, 56, 550–556. 45 T. Tsuruoka, S. Furukawa, Y. Takashima, K. Yoshida, S. Isoda and S. Kitagawa, Angew. Chem., Int. Ed., 2009, 48, 4739–4743. 46 A. Umemura, S. Diring, S. Furukawa, H. Uehara, T. Tsuruoka and S. Kitagawa, J. Am. Chem. Soc., 2011, 133, 15506–15513. 47 A. Schaate, P. Roy, A. Godt, J. Lippke, F. Waltz, M. Wiebcke and P. Behrens, Chem. – Eur. J., 2011, 17, 6643–6651. 48 G. Wißmann, A. Schaate, S. Lilienthal, I. Bremer, A. M. Schneider and P. Behrens, Microporous Mesoporous Mater., 2012, 152, 64–70. 49 V. Bon, V. Senkovskyy, I. Senkovska and S. Kaskel, Chem. Commun., 2012, 48, 8407–8409. 50 V. Bon, I. Senkovska, I. A. Baburin and S. Kaskel, Cryst. Growth Des., 2013, 13, 1231–1237. 51 J. Ma, A. G. Wong-Foy and A. J. Matzger, Inorg. Chem., 2015, 54, 4591–4593. 52 M. Eddaoudi, D. B. Moler, H. Li, B. Chen, T. M. Reineke, M. O’Keeffe and O. M. Yaghi, Acc. Chem. Res., 2001, 34, 319–330. 53 C. Serre, F. Millange, S. Surblé and G. Férey, Angew. Chem., Int. Ed., 2004, 43, 6285–6289. 54 V. Guillerm, S. Gross, C. Serre, T. Devic, M. Bauer and G. Férey, Chem. Commun., 2010, 46, 767–769. 55 H. Wu, Y. S. Chua, V. Krungleviciute, M. Tyagi, P. Chen, T. Yildirim and W. Zhou, J. Am. Chem. Soc., 2013, 135, 10525–10532. 56 F. Vermoortele, B. Bueken, G. Le Bars, B. Van de Voorde, M. Vandichel, K. Houthoofd, A. Vimont, M. Daturi, M. Waroquier, V. Van Speybroeck, C. Kirschhock and D. E. De Vos, J. Am. Chem. Soc., 2013, 135, 11465–11468. Chem. Soc. Rev. Chem Soc Rev 57 O. K. Farha, A. Özgür Yazaydın, I. Eryazici, C. D. Malliakas, B. G. Hauser, M. G. Kanatzidis, S. T. Nguyen, R. Q. Snurr and J. T. Hupp, Nat. Chem., 2010, 2, 944–948. 58 H. Furukawa, N. Ko, Y. B. Go, N. Aratani, S. B. Choi, E. Choi, A. Ö. Yazaydin, R. Q. Snurr, M. O’Keeffe, J. Kim and O. M. Yaghi, Science, 2010, 329, 424–428. 59 O. K. Farha, I. Eryazici, N. C. Jeong, B. G. Hauser, C. E. Wilmer, A. A. Sarjeant, R. Q. Snurr, S. T. Nguyen, A. Ö. Yazaydın and J. T. Hupp, J. Am. Chem. Soc., 2012, 134, 15016–15021. 60 H. Furukawa, Y. B. Go, N. Ko, Y. K. Park, F. J. Uribe-Romo, J. Kim, M. O’Keeffe and O. M. Yaghi, Inorg. Chem., 2011, 50, 9147–9152. 61 O. M. Yaghi, M. O’Keeffe, N. W. Ockwig, H. K. Chae, M. Eddaoudi and J. Kim, Nature, 2003, 423, 705–714. 62 H. Deng, S. Grunder, K. E. Cordova, C. Valente, H. Furukawa, M. Hmadeh, F. Gándara, A. C. Whalley, Z. Liu, S. Asahina, H. Kazumori, M. O’Keeffe, O. Terasaki, J. F. Stoddart and O. M. Yaghi, Science, 2012, 336, 1018–1023. 63 T. C. Wang, W. Bury, D. A. Gómez-Gualdrón, N. A. Vermeulen, J. E. Mondloch, P. Deria, K. Zhang, P. Z. Moghadam, A. A. Sarjeant, R. Q. Snurr, J. F. Stoddart, J. T. Hupp and O. K. Farha, J. Am. Chem. Soc., 2015, 137, 3585–3591. 64 Y.-S. Bae, A. Ö. Yazaydın and R. Q. Snurr, Langmuir, 2010, 26, 5475–5483. 65 T.-F. Liu, D. Feng, Y.-P. Chen, L. Zou, M. Bosch, S. Yuan, Z. Wei, S. Fordham, K. Wang and H.-C. Zhou, J. Am. Chem. Soc., 2015, 137, 413–419. 66 R. Banerjee, H. Furukawa, D. Britt, C. Knobler, M. O’Keeffe and O. M. Yaghi, J. Am. Chem. Soc., 2009, 131, 3875–3877. 67 B. P. Biswal, S. Chandra, S. Kandambeth, B. Lukose, T. Heine and R. Banerjee, J. Am. Chem. Soc., 2013, 135, 5328–5331. 68 A. P. Nelson, O. K. Farha, K. L. Mulfort and J. T. Hupp, J. Am. Chem. Soc., 2009, 131, 458–460. 69 O. K. Farha and J. T. Hupp, Acc. Chem. Res., 2010, 43, 1166–1175. 70 J. E. Mondloch, O. Karagiaridi, O. K. Farha and J. T. Hupp, CrystEngComm, 2013, 15, 9258–9264. 71 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O’Keeffe and O. M. Yaghi, Science, 2002, 295, 469–472. 72 A. Schaate, S. Dühnen, G. Platz, S. Lilienthal, A. M. Schneider and P. Behrens, Eur. J. Inorg. Chem., 2012, 790–796. 73 A. Schaate, P. Roy, T. Preuße, S. J. Lohmeier, A. Godt and P. Behrens, Chem. – Eur. J., 2011, 17, 9320–9325. 74 S. Ma, D. Sun, M. Ambrogio, J. A. Fillinger, S. Parkin and H.-C. Zhou, J. Am. Chem. Soc., 2007, 129, 1858–1859. 75 J. Zhang, L. Wojtas, R. W. Larsen, M. Eddaoudi and M. J. Zaworotko, J. Am. Chem. Soc., 2009, 131, 17040–17041. 76 B. Chen, X. Wang, Q. Zhang, X. Xi, J. Cai, H. Qi, S. Shi, J. Wang, D. Yuan and M. Fang, J. Mater. Chem., 2010, 20, 3758–3767. 77 O. Shekhah, H. Wang, M. Paradinas, C. Ocal, B. Schupbach, A. Terfort, D. Zacher, R. A. Fischer and C. Woll, Nat. Mater., 2009, 8, 481–484. This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev 78 M. Li, D. Li, M. O’Keeffe and O. M. Yaghi, Chem. Rev., 2014, 114, 1343–1370. 79 D. Kim, X. Liu and M. S. Lah, Inorg. Chem. Front., 2015, 2, 336–360. 80 V. A. Blatov, L. Carlucci, G. Ciani and D. M. Proserpio, CrystEngComm, 2004, 6, 378–395. 81 M. O’Keeffe and O. M. Yaghi, Chem. Rev., 2012, 112, 675–702. 82 D. A. Gomez-Gualdron, O. V. Gutov, V. Krungleviciute, B. Borah, J. E. Mondloch, J. T. Hupp, T. Yildirim, O. K. Farha and R. Q. Snurr, Chem. Mater., 2014, 26, 5632–5639. 83 O. V. Gutov, W. Bury, D. A. Gomez-Gualdron, V. Krungleviciute, D. Fairen-Jimenez, J. E. Mondloch, A. A. Sarjeant, S. S. Al-Juaid, R. Q. Snurr, J. T. Hupp, T. Yildirim and O. K. Farha, Chem. – Eur. J., 2014, 20, 12389–12393. 84 H. Chun, D. Kim, D. N. Dybtsev and K. Kim, Angew. Chem., Int. Ed., 2004, 43, 971–974. 85 M. Dincă, A. Dailly and J. R. Long, Chem. – Eur. J., 2008, 14, 10280–10285. 86 J. M. Gotthardt, K. F. White, B. F. Abrahams, C. Ritchie and C. Boskovic, Cryst. Growth Des., 2012, 12, 4425–4430. 87 O. Delgado-Friedrichs, M. O’Keeffe and O. M. Yaghi, Phys. Chem. Chem. Phys., 2007, 9, 1035–1043. 88 O. Delgado Friedrichs, M. O’Keeffe and O. M. Yaghi, Solid State Sci., 2003, 5, 73–78. 89 S. Yuan, W. Lu, Y.-P. Chen, Q. Zhang, T.-F. Liu, D. Feng, X. Wang, J. Qin and H.-C. Zhou, J. Am. Chem. Soc., 2015, 137, 3177–3180. 90 S. M. Cohen, Chem. Rev., 2012, 112, 970–1000. 91 M. Kim and S. M. Cohen, CrystEngComm, 2012, 14, 4096–4104. 92 P. Roy, A. Schaate, P. Behrens and A. Godt, Chem. – Eur. J., 2012, 18, 6979–6985. 93 Y. Goto, H. Sato, S. Shinkai and K. Sada, J. Am. Chem. Soc., 2008, 130, 14354–14355. 94 T. Gadzikwa, G. Lu, C. L. Stern, S. R. Wilson, J. T. Hupp and S. T. Nguyen, Chem. Commun., 2008, 5493–5495. 95 M. Savonnet, D. Bazer-Bachi, N. Bats, J. Perez-Pellitero, E. Jeanneau, V. Lecocq, C. Pinel and D. Farrusseng, J. Am. Chem. Soc., 2010, 132, 4518–4519. 96 T. Gadzikwa, O. K. Farha, C. D. Malliakas, M. G. Kanatzidis, J. T. Hupp and S. T. Nguyen, J. Am. Chem. Soc., 2009, 131, 13613–13615. 97 H.-L. Jiang, D. Feng, T.-F. Liu, J.-R. Li and H.-C. Zhou, J. Am. Chem. Soc., 2012, 134, 14690–14693. 98 H. C. Kolb, M. G. Finn and K. B. Sharpless, Angew. Chem., Int. Ed., 2001, 40, 2004–2021. 99 T. Toyao, K. Miyahara, M. Fujiwaki, T.-H. Kim, S. Dohshi, Y. Horiuchi and M. Matsuoka, J. Phys. Chem. C, 2015, 119, 8131–8137. 100 K. Manna, T. Zhang, F. X. Greene and W. Lin, J. Am. Chem. Soc., 2015, 137, 2665–2673. 101 K. Manna, T. Zhang and W. Lin, J. Am. Chem. Soc., 2014, 136, 6566–6569. 102 M. J. Ingleson, J. Perez Barrio, J.-B. Guilbaud, Y. Z. Khimyak and M. J. Rosseinsky, Chem. Commun., 2008, 2680–2682. This journal is © The Royal Society of Chemistry 2016 Review Article 103 Z. Wang and S. M. Cohen, Angew. Chem., Int. Ed., 2008, 47, 4699–4702. 104 E. Dugan, Z. Wang, M. Okamura, A. Medina and S. M. Cohen, Chem. Commun., 2008, 3366–3368. 105 A. M. Rasero-Almansa, A. Corma, M. Iglesias and F. Sanchez, Green Chem., 2014, 16, 3522–3527. 106 Y. Han, J.-R. Li, Y. Xie and G. Guo, Chem. Soc. Rev., 2014, 43, 5952–5981. 107 H. Fei, J. Shin, Y. S. Meng, M. Adelhardt, J. Sutter, K. Meyer and S. M. Cohen, J. Am. Chem. Soc., 2014, 136, 4965–4973. 108 H. Fei and S. M. Cohen, J. Am. Chem. Soc., 2015, 137, 2191–2194. 109 W. Lu, Z. Wei, Z.-Y. Gu, T.-F. Liu, J. Park, J. Park, J. Tian, M. Zhang, Q. Zhang, T. Gentle III, M. Bosch and H.-C. Zhou, Chem. Soc. Rev., 2014, 43, 5561–5593. 110 H. Chun, D. N. Dybtsev, H. Kim and K. Kim, Chem. – Eur. J., 2005, 11, 3521–3529. 111 A. D. Burrows, CrystEngComm, 2011, 13, 3623–3642. 112 H. Deng, C. J. Doonan, H. Furukawa, R. B. Ferreira, J. Towne, C. B. Knobler, B. Wang and O. M. Yaghi, Science, 2010, 327, 846–850. 113 M. Dan-Hardi, C. Serre, T. Frot, L. Rozes, G. Maurin, C. Sanchez and G. Férey, J. Am. Chem. Soc., 2009, 131, 10857–10859. 114 C. H. Hendon, D. Tiana, M. Fontecave, C. Sanchez, L. D’arras, C. Sassoye, L. Rozes, C. Mellot-Draznieks and A. Walsh, J. Am. Chem. Soc., 2013, 135, 10942–10945. 115 J. Gao, J. Miao, P.-Z. Li, W. Y. Teng, L. Yang, Y. Zhao, B. Liu and Q. Zhang, Chem. Commun., 2014, 50, 3786–3788. 116 J. A. Mason, L. E. Darago, W. W. Lukens and J. R. Long, Inorg. Chem., 2015, 54, 10096–10104. 117 M. Kim, J. F. Cahill, H. Fei, K. A. Prather and S. M. Cohen, J. Am. Chem. Soc., 2012, 134, 18082–18088. 118 P. Deria, J. E. Mondloch, E. Tylianakis, P. Ghosh, W. Bury, R. Q. Snurr, J. T. Hupp and O. K. Farha, J. Am. Chem. Soc., 2013, 135, 16801–16804. 119 P. Deria, W. Bury, J. T. Hupp and O. K. Farha, Chem. Commun., 2014, 50, 1965–1968. 120 P. Deria, W. Bury, I. Hod, C.-W. Kung, O. Karagiaridi, J. T. Hupp and O. K. Farha, Inorg. Chem., 2015, 54, 2185–2192. 121 J. E. Mondloch, W. Bury, D. Fairen-Jimenez, S. Kwon, E. J. DeMarco, M. H. Weston, A. A. Sarjeant, S. T. Nguyen, P. C. Stair, R. Q. Snurr, O. K. Farha and J. T. Hupp, J. Am. Chem. Soc., 2013, 135, 10294–10297. 122 R. Ameloot, M. Aubrey, B. M. Wiers, A. P. GómoraFigueroa, S. N. Patel, N. P. Balsara and J. R. Long, Chem. – Eur. J., 2013, 19, 5533–5536. 123 E. López-Maya, C. Montoro, L. M. Rodrı́guez-Albelo, S. D. Aznar Cervantes, A. A. Lozano-Pérez, J. L. Cenı́s, E. Barea and J. A. R. Navarro, Angew. Chem., Int. Ed., 2015, 54, 6790–6794. 124 L.-J. Li, P.-Q. Liao, C.-T. He, Y.-S. Wei, H.-L. Zhou, J.-M. Lin, X.-Y. Li and J.-P. Zhang, J. Mater. Chem. A, 2015, 3, 21849–21855. 125 Q. Yang, S. Vaesen, F. Ragon, A. D. Wiersum, D. Wu, A. Lago, T. Devic, C. Martineau, F. Taulelle, P. L. Llewellyn, H. Jobic, Chem. Soc. Rev. View Article Online Review Article 126 127 Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 C. Zhong, C. Serre, G. De Weireld and G. Maurin, Angew. Chem., Int. Ed., 2013, 52, 10316–10320. H. Reinsch, B. Bueken, F. Vermoortele, I. Stassen, A. Lieb, K.-P. Lillerud and D. De Vos, CrystEngComm, 2015, 17, 4070–4074. Z. Hu, Y. Peng, Z. Kang, Y. Qian and D. Zhao, Inorg. Chem., 2015, 54, 4862–4868. M. Taddei, P. V. Dau, S. M. Cohen, M. Ranocchiari, J. A. van Bokhoven, F. Costantino, S. Sabatini and R. Vivani, Dalton Trans., 2015, 44, 14019–14026. D. J. Tranchemontagne, J. L. Mendoza-Cortes, M. O’Keeffe and O. M. Yaghi, Chem. Soc. Rev., 2009, 38, 1257–1283. J. J. Perry IV, J. A. Perman and M. J. Zaworotko, Chem. Soc. Rev., 2009, 38, 1400–1417. A. Otero, J. Fernández-Baeza, A. Antiñolo, J. Tejeda, A. Lara-Sánchez, L. Sánchez-Barba, M. Fernández-López and I. López-Solera, Inorg. Chem., 2004, 43, 1350–1358. S. O. Baumann, M. Puchberger and U. Schubert, Dalton Trans., 2011, 40, 1401–1406. G. Kickelbick, D. Holzinger, C. Brick, G. Trimmel and E. Moons, Chem. Mater., 2002, 14, 4382–4389. S. Gross, J. Mater. Chem., 2011, 21, 15853–15861. D. Feng, H.-L. Jiang, Y.-P. Chen, Z.-Y. Gu, Z. Wei and H.-C. Zhou, Inorg. Chem., 2013, 52, 12661–12667. M. Taddei, F. Costantino, R. Vivani, S. Sabatini, S.-H. Lim and S. M. Cohen, Chem. Commun., 2014, 50, 5737–5740. M. Taddei, F. Costantino, F. Marmottini, A. Comotti, P. Sozzani and R. Vivani, Chem. Commun., 2014, 50, 14831–14834. M. Taddei, F. Costantino and R. Vivani, Inorg. Chem., 2010, 49, 9664–9670. V. Guillerm, F. Ragon, M. Dan-Hardi, T. Devic, M. Vishnuvarthan, B. Campo, A. Vimont, G. Clet, Q. Yang, G. Maurin, G. Férey, A. Vittadini, S. Gross and C. Serre, Angew. Chem., Int. Ed., 2012, 51, 9267–9271. W. Liang, R. Babarao, T. L. Church and D. M. D’Alessandro, Chem. Commun., 2015, 51, 11286–11289. L. Cooper, N. Guillou, C. Martineau, E. Elkaim, F. Taulelle, C. Serre and T. Devic, Eur. J. Inorg. Chem., 2014, 6281–6289. G. Mouchaham, L. Cooper, N. Guillou, C. Martineau, E. Elkaı̈m, S. Bourrelly, P. L. Llewellyn, C. Allain, G. Clavier, C. Serre and T. Devic, Angew. Chem., Int. Ed., 2015, 54, 13297–13301. D. Feng, Z.-Y. Gu, Y.-P. Chen, J. Park, Z. Wei, Y. Sun, M. Bosch, S. Yuan and H.-C. Zhou, J. Am. Chem. Soc., 2014, 136, 17714–17717. D. Feng, Z.-Y. Gu, J.-R. Li, H.-L. Jiang, Z. Wei and H.-C. Zhou, Angew. Chem., Int. Ed., 2012, 51, 10307–10310. H.-L. Jiang, D. Feng, K. Wang, Z.-Y. Gu, Z. Wei, Y.-P. Chen and H.-C. Zhou, J. Am. Chem. Soc., 2013, 135, 13934–13938. D. Feng, W.-C. Chung, Z. Wei, Z.-Y. Gu, H.-L. Jiang, Y.-P. Chen, D. J. Darensbourg and H.-C. Zhou, J. Am. Chem. Soc., 2013, 135, 17105–17110. D. Feng, K. Wang, J. Su, T.-F. Liu, J. Park, Z. Wei, M. Bosch, A. Yakovenko, X. Zou and H.-C. Zhou, Angew. Chem., Int. Ed., 2015, 54, 149–154. Chem. Soc. Rev. Chem Soc Rev 148 S. J. Garibay and S. M. Cohen, Chem. Commun., 2010, 46, 7700–7702. 149 V. Bon, I. Senkovska, M. S. Weiss and S. Kaskel, CrystEngComm, 2013, 15, 9572–9577. 150 S. Pu, L. Xu, L. Sun and H. Du, Inorg. Chem. Commun., 2015, 52, 50–52. 151 C. Wang, O. Volotskova, K. Lu, M. Ahmad, C. Sun, L. Xing and W. Lin, J. Am. Chem. Soc., 2014, 136, 6171–6174. 152 K. Wang, H. Huang, W. Xue, D. Liu, X. Zhao, Y. Xiao, Z. Li, Q. Yang, L. Wang and C. Zhong, CrystEngComm, 2015, 17, 3586–3590. 153 M. Yoon and D. Moon, Microporous Mesoporous Mater., 2015, 215, 116–122. 154 L. Li, S. Tang, C. Wang, X. Lv, M. Jiang, H. Wu and X. Zhao, Chem. Commun., 2014, 50, 2304–2307. 155 H. Furukawa, F. Gándara, Y.-B. Zhang, J. Jiang, W. L. Queen, M. R. Hudson and O. M. Yaghi, J. Am. Chem. Soc., 2014, 136, 4369–4381. 156 H. Reinsch, I. Stassen, B. Bueken, A. Lieb, R. Ameloot and D. De Vos, CrystEngComm, 2015, 17, 331–337. 157 R. Wang, Z. Wang, Y. Xu, F. Dai, L. Zhang and D. Sun, Inorg. Chem., 2014, 53, 7086–7088. 158 W.-Y. Gao, M. Chrzanowski and S. Ma, Chem. Soc. Rev., 2014, 43, 5841–5866. 159 W. Morris, B. Volosskiy, S. Demir, F. Gándara, P. L. McGrier, H. Furukawa, D. Cascio, J. F. Stoddart and O. M. Yaghi, Inorg. Chem., 2012, 51, 6443–6445. 160 Y. Chen, T. Hoang and S. Ma, Inorg. Chem., 2012, 51, 12600–12602. 161 Q. Lin, X. Bu, A. Kong, C. Mao, X. Zhao, F. Bu and P. Feng, J. Am. Chem. Soc., 2015, 137, 2235–2238. 162 S. B. Kalidindi, S. Nayak, M. E. Briggs, S. Jansat, A. P. Katsoulidis, G. J. Miller, J. E. Warren, D. Antypov, F. Corà, B. Slater, M. R. Prestly, C. Martı́-Gastaldo and M. J. Rosseinsky, Angew. Chem., Int. Ed., 2015, 54, 221–226. 163 Z. Wei, Z.-Y. Gu, R. K. Arvapally, Y.-P. Chen, R. N. McDougald, J. F. Ivy, A. A. Yakovenko, D. Feng, M. A. Omary and H.-C. Zhou, J. Am. Chem. Soc., 2014, 136, 8269–8276. 164 J. M. Taylor, A. H. Mahmoudkhani and G. K. H. Shimizu, Angew. Chem., Int. Ed., 2007, 46, 795–798. 165 L. Ma, A. Jin, Z. Xie and W. Lin, Angew. Chem., Int. Ed., 2009, 48, 9905–9908. 166 Y. E. Cheon and M. P. Suh, Chem. Commun., 2009, 2296–2298. 167 C. Tan, S. Yang, N. R. Champness, X. Lin, A. J. Blake, W. Lewis and M. Schroder, Chem. Commun., 2011, 47, 4487–4489. 168 L. Wen, P. Cheng and W. Lin, Chem. Commun., 2012, 48, 2846–2848. 169 L. Wen, P. Cheng and W. Lin, Chem. Sci., 2012, 3, 2288–2292. 170 D. Liu, H. Wu, S. Wang, Z. Xie, J. Li and W. Lin, Chem. Sci., 2012, 3, 3032–3037. 171 M. Zhang, Y.-P. Chen and H.-C. Zhou, CrystEngComm, 2013, 15, 9544–9552. 172 B. Chen, M. Eddaoudi, T. M. Reineke, J. W. Kampf, M. O’Keeffe and O. M. Yaghi, J. Am. Chem. Soc., 2000, 122, 11559–11560. This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev 173 W. Liang, R. Babarao, M. J. Murphy and D. M. D’Alessandro, Dalton Trans., 2015, 44, 1516–1519. 174 M. Gustafsson, A. Bartoszewicz, B. Martı́n-Matute, J. Sun, J. Grins, T. Zhao, Z. Li, G. Zhu and X. Zou, Chem. Mater., 2010, 22, 3316–3322. 175 M. Fujita, Y. J. Kwon, S. Washizu and K. Ogura, J. Am. Chem. Soc., 1994, 116, 1151–1152. 176 S. Horike, M. Dincă, K. Tamaki and J. R. Long, J. Am. Chem. Soc., 2008, 130, 5854–5855. 177 N. V. Maksimchuk, K. A. Kovalenko, V. P. Fedin and O. A. Kholdeeva, Chem. Commun., 2012, 48, 6812–6814. 178 M. H. Alkordi, Y. Liu, R. W. Larsen, J. F. Eubank and M. Eddaoudi, J. Am. Chem. Soc., 2008, 130, 12639–12641. 179 N. Grootboom and T. Nyokong, J. Mol. Catal. A: Chem., 2002, 179, 113–123. 180 B. S. Lane and K. Burgess, Chem. Rev., 2003, 103, 2457–2474. 181 K. C. Gupta and A. K. Sutar, Coord. Chem. Rev., 2008, 252, 1420–1450. 182 C. Wang, Z. Xie, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2011, 133, 13445–13454. 183 E. Barea, C. Montoro and J. A. R. Navarro, Chem. Soc. Rev., 2014, 43, 5419–5430. 184 B. M. Smith, Chem. Soc. Rev., 2008, 37, 470–478. 185 F.-J. Ma, S.-X. Liu, C.-Y. Sun, D.-D. Liang, G.-J. Ren, F. Wei, Y.-G. Chen and Z.-M. Su, J. Am. Chem. Soc., 2011, 133, 4178–4181. 186 K. Kim, O. G. Tsay, D. A. Atwood and D. G. Churchill, Chem. Rev., 2011, 111, 5345–5403. 187 M. J. Katz, J. E. Mondloch, R. K. Totten, J. K. Park, S. T. Nguyen, O. K. Farha and J. T. Hupp, Angew. Chem., Int. Ed., 2014, 53, 497–501. 188 J. E. Mondloch, M. J. Katz, W. C. Isley Iii, P. Ghosh, P. Liao, W. Bury, G. W. Wagner, M. G. Hall, J. B. DeCoste, G. W. Peterson, R. Q. Snurr, C. J. Cramer, J. T. Hupp and O. K. Farha, Nat. Mater., 2015, 14, 512–516. 189 S.-Y. Moon, Y. Liu, J. T. Hupp and O. K. Farha, Angew. Chem., Int. Ed., 2015, 54, 6795–6799. 190 K. Manna, T. Zhang, M. Carboni, C. W. Abney and W. Lin, J. Am. Chem. Soc., 2014, 136, 13182–13185. 191 C. Bolm, J. Legros, J. Le Paih and L. Zani, Chem. Rev., 2004, 104, 6217–6254. 192 C.-L. Sun, B.-J. Li and Z.-J. Shi, Chem. Rev., 2011, 111, 1293–1314. 193 S. Enthaler, K. Junge and M. Beller, Angew. Chem., Int. Ed., 2008, 47, 3317–3321. 194 E. T. Hennessy and T. A. Betley, Science, 2013, 340, 591–595. 195 M. R. Friedfeld, M. Shevlin, J. M. Hoyt, S. W. Krska, M. T. Tudge and P. J. Chirik, Science, 2013, 342, 1076–1080. 196 R. T. Gephart and T. H. Warren, Organometallics, 2012, 31, 7728–7752. 197 C. Chen, T. R. Dugan, W. W. Brennessel, D. J. Weix and P. L. Holland, J. Am. Chem. Soc., 2014, 136, 945–955. 198 J. Jiang, F. Gándara, Y.-B. Zhang, K. Na, O. M. Yaghi and W. G. Klemperer, J. Am. Chem. Soc., 2014, 136, 12844–12847. 199 D. Sun, Y. Fu, W. Liu, L. Ye, D. Wang, L. Yang, X. Fu and Z. Li, Chem. – Eur. J., 2013, 19, 14279–14285. This journal is © The Royal Society of Chemistry 2016 Review Article 200 Y. Lee, S. Kim, J. K. Kang and S. M. Cohen, Chem. Commun., 2015, 51, 5735–5738. 201 A. Wang, Y. Zhou, Z. Wang, M. Chen, L. Sun and X. Liu, RSC Adv., 2016, 6, 3671–3679. 202 Z. Zhang, Z.-Z. Yao, S. Xiang and B. Chen, Energy Environ. Sci., 2014, 7, 2868–2899. 203 K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T.-H. Bae and J. R. Long, Chem. Rev., 2012, 112, 724–781. 204 Z. R. Herm, E. D. Bloch and J. R. Long, Chem. Mater., 2014, 26, 323–338. 205 B. Wang, H. Huang, X.-L. Lv, Y. Xie, M. Li and J.-R. Li, Inorg. Chem., 2014, 53, 9254–9259. 206 S. Yang, Y. Guo, N. Yan, D. Wu, H. He, J. Xie, Z. Qu, C. Yang and J. Jia, Chem. Commun., 2010, 46, 8377–8379. 207 M. H. Kim, S.-W. Ham and J.-B. Lee, Appl. Catal., B, 2010, 99, 272–278. 208 S. Yang, Y. Guo, N. Yan, D. Wu, H. He, J. Xie, Z. Qu and J. Jia, Appl. Catal., B, 2011, 101, 698–708. 209 R. Stolle, H. Koeser and H. Gutberlet, Appl. Catal., B, 2014, 144, 486–497. 210 A. J. Howarth, M. J. Katz, T. C. Wang, A. E. Platero-Prats, K. W. Chapman, J. T. Hupp and O. K. Farha, J. Am. Chem. Soc., 2015, 137, 7488–7494. 211 A. J. Howarth, T. C. Wang, S. S. Al-Juaid, S. G. Aziz, J. T. Hupp and O. K. Farha, Dalton Trans., 2016, 45, 93–97. 212 W.-T. Tsai, H.-C. Hsu, T.-Y. Su, K.-Y. Lin, C.-M. Lin and T.-H. Dai, J. Hazard. Mater., 2007, 147, 1056–1062. 213 J.-S. Qin, S.-R. Zhang, D.-Y. Du, P. Shen, S.-J. Bao, Y.-Q. Lan and Z.-M. Su, Chem. – Eur. J., 2014, 20, 5625–5630. 214 X.-L. Lv, M. Tong, H. Huang, B. Wang, L. Gan, Q. Yang, C. Zhong and J.-R. Li, J. Solid State Chem., 2015, 223, 104–108. 215 X. Zhu, B. Li, J. Yang, Y. Li, W. Zhao, J. Shi and J. Gu, ACS Appl. Mater. Interfaces, 2015, 7, 223–231. 216 Y. S. Seo, N. A. Khan and S. H. Jhung, Chem. Eng. J., 2015, 270, 22–27. 217 Z. Hasan, N. A. Khan and S. H. Jhung, Chem. Eng. J., 2016, 284, 1406–1413. 218 M. I. Hoffert, K. Caldeira, A. K. Jain, E. F. Haites, L. D. D. Harvey, S. D. Potter, M. E. Schlesinger, S. H. Schneider, R. G. Watts, T. M. L. Wigley and D. J. Wuebbles, Nature, 1998, 395, 881–884. 219 C. W. Abney, K. M. L. Taylor-Pashow, S. R. Russell, Y. Chen, R. Samantaray, J. V. Lockard and W. Lin, Chem. Mater., 2014, 26, 5231–5243. 220 P. Horcajada, R. Gref, T. Baati, P. K. Allan, G. Maurin, P. Couvreur, G. Férey, R. E. Morris and C. Serre, Chem. Rev., 2012, 112, 1232–1268. 221 Y. Chen, C. Tan, H. Zhang and L. Wang, Chem. Soc. Rev., 2015, 44, 2681–2701. 222 D. Cunha, M. Ben Yahia, S. Hall, S. R. Miller, H. Chevreau, E. Elkaı̈m, G. Maurin, P. Horcajada and C. Serre, Chem. Mater., 2013, 25, 2767–2776. 223 C.-Y. Sun, C. Qin, X.-L. Wang, G.-S. Yang, K.-Z. Shao, Y.-Q. Lan, Z.-M. Su, P. Huang, C.-G. Wang and E.-B. Wang, Dalton Trans., 2012, 41, 6906–6909. Chem. Soc. Rev. View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Review Article 224 K. M. L. Taylor-Pashow, J. D. Rocca, Z. Xie, S. Tran and W. Lin, J. Am. Chem. Soc., 2009, 131, 14261–14263. 225 C. Tamames-Tabar, D. Cunha, E. Imbuluzqueta, F. Ragon, C. Serre, M. J. Blanco-Prieto and P. Horcajada, J. Mater. Chem. B, 2014, 2, 262–271. 226 C. Queffélec, M. Petit, P. Janvier, D. A. Knight and B. Bujoli, Chem. Rev., 2012, 112, 3777–3807. 227 X. Zhu, J. Gu, Y. Wang, B. Li, Y. Li, W. Zhao and J. Shi, Chem. Commun., 2014, 50, 8779–8782. 228 J. Gu, M. Huang, J. Liu, Y. Li, W. Zhao and J. Shi, New J. Chem., 2012, 36, 1717–1720. 229 J.-S. Yang and T. M. Swager, J. Am. Chem. Soc., 1998, 120, 11864–11873. 230 R. D. Smiley and G. G. Hammes, Chem. Rev., 2006, 106, 3080–3094. 231 B. W. D’Andrade and S. R. Forrest, Adv. Mater., 2004, 16, 1585–1595. 232 M. A. Baldo, D. F. O’Brien, Y. You, A. Shoustikov, S. Sibley, M. E. Thompson and S. R. Forrest, Nature, 1998, 395, 151–154. 233 W. E. Moerner and M. Orrit, Science, 1999, 283, 1670–1676. 234 Y. Cui, Y. Yue, G. Qian and B. Chen, Chem. Rev., 2012, 112, 1126–1162. 235 M. D. Allendorf, C. A. Bauer, R. K. Bhakta and R. J. T. Houk, Chem. Soc. Rev., 2009, 38, 1330–1352. 236 N. B. Shustova, B. D. McCarthy and M. Dincă, J. Am. Chem. Soc., 2011, 133, 20126–20129. 237 N. B. Shustova, A. F. Cozzolino, S. Reineke, M. Baldo and M. Dincă, J. Am. Chem. Soc., 2013, 135, 13326–13329. 238 N. B. Shustova, T.-C. Ong, A. F. Cozzolino, V. K. Michaelis, R. G. Griffin and M. Dincă, J. Am. Chem. Soc., 2012, 134, 15061–15070. 239 N. B. Shustova, A. F. Cozzolino and M. Dincă, J. Am. Chem. Soc., 2012, 134, 19596–19599. 240 J.-P. Zou, Q. Peng, Z. Wen, G.-S. Zeng, Q.-J. Xing and G.-C. Guo, Cryst. Growth Des., 2010, 10, 2613–2619. 241 R. Feng, F.-L. Jiang, L. Chen, C.-F. Yan, M.-Y. Wu and M.-C. Hong, Chem. Commun., 2009, 5296–5298. 242 Y. Liu, M. Pan, Q.-Y. Yang, L. Fu, K. Li, S.-C. Wei and C.-Y. Su, Chem. Mater., 2012, 24, 1954–1960. 243 K. A. White, D. A. Chengelis, K. A. Gogick, J. Stehman, N. L. Rosi and S. Petoud, J. Am. Chem. Soc., 2009, 131, 18069–18071. 244 T.-F. Liu, W. Zhang, W.-H. Sun and R. Cao, Inorg. Chem., 2011, 50, 5242–5248. 245 H.-J. Son, S. Jin, S. Patwardhan, S. J. Wezenberg, N. C. Jeong, M. So, C. E. Wilmer, A. A. Sarjeant, G. C. Schatz, R. Q. Snurr, O. K. Farha, G. P. Wiederrecht and J. T. Hupp, J. Am. Chem. Soc., 2013, 135, 862–869. 246 V. S. Vyas and R. Rathore, Chem. Commun., 2010, 46, 1065–1067. 247 B. Chen, L. Wang, Y. Xiao, F. R. Fronczek, M. Xue, Y. Cui and G. Qian, Angew. Chem., Int. Ed., 2009, 48, 500–503. 248 M. M. Wanderley, C. Wang, C.-D. Wu and W. Lin, J. Am. Chem. Soc., 2012, 134, 9050–9053. 249 A. B. Braunschweig, A. J. Senesi and C. A. Mirkin, J. Am. Chem. Soc., 2009, 131, 922–923. Chem. Soc. Rev. Chem Soc Rev 250 R.-B. Lin, F. Li, S.-Y. Liu, X.-L. Qi, J.-P. Zhang and X.M. Chen, Angew. Chem., Int. Ed., 2013, 52, 13429–13433. 251 B. Chen, L. Wang, F. Zapata, G. Qian and E. B. Lobkovsky, J. Am. Chem. Soc., 2008, 130, 6718–6719. 252 S. Liu, J. Li and F. Luo, Inorg. Chem. Commun., 2010, 13, 870–872. 253 Y. Xiao, Y. Cui, Q. Zheng, S. Xiang, G. Qian and B. Chen, Chem. Commun., 2010, 46, 5503–5505. 254 Z. Xiang, C. Fang, S. Leng and D. Cao, J. Mater. Chem. A, 2014, 2, 7662–7665. 255 Z. Hao, X. Song, M. Zhu, X. Meng, S. Zhao, S. Su, W. Yang, S. Song and H. Zhang, J. Mater. Chem. A, 2013, 1, 11043–11050. 256 X.-H. Zhou, L. Li, H.-H. Li, A. Li, T. Yang and W. Huang, Dalton Trans., 2013, 42, 12403–12409. 257 J. Cao, Y. Gao, Y. Wang, C. Du and Z. Liu, Chem. Commun., 2013, 49, 6897–6899. 258 L. Zhang, Y. Jian, J. Wang, C. He, X. Li, T. Liu and C. Duan, Dalton Trans., 2012, 41, 10153–10155. 259 S. Dang, E. Ma, Z.-M. Sun and H. Zhang, J. Mater. Chem., 2012, 22, 16920–16926. 260 W. C. Mak, C. Chan, J. Barford and R. Renneberg, Biosens. Bioelectron., 2003, 19, 233–237. 261 J. Yang, Y. Dai, X. Zhu, Z. Wang, Y. Li, Q. Zhuang, J. Shi and J. Gu, J. Mater. Chem. A, 2015, 3, 7445–7452. 262 A. V. Desai, P. Samanta, B. Manna and S. K. Ghosh, Chem. Commun., 2015, 51, 6111–6114. 263 S. S. Nagarkar, T. Saha, A. V. Desai, P. Talukdar and S. K. Ghosh, Sci. Rep., 2014, 4, 7053. 264 S. S. Nagarkar, A. V. Desai and S. K. Ghosh, Chem. – Eur. J., 2015, 21, 9994–9997. 265 S. S. Nagarkar, A. V. Desai and S. K. Ghosh, Chem. Commun., 2014, 50, 8915–8918. 266 M. Carboni, Z. Lin, C. W. Abney, T. Zhang and W. Lin, Chem. – Eur. J., 2014, 20, 14965–14970. 267 Y. J. Kim, J. H. Kim, M. S. Kang, M. J. Lee, J. Won, J. C. Lee and Y. S. Kang, Adv. Mater., 2004, 16, 1753–1757. 268 B. Kesanli, K. Hong, K. Meyer, H.-J. Im and S. Dai, Appl. Phys. Lett., 2006, 89, 214104. 269 A. Morozan and F. Jaouen, Energy Environ. Sci., 2012, 5, 9269–9290. 270 W. Xia, A. Mahmood, R. Zou and Q. Xu, Energy Environ. Sci., 2015, 8, 1837–1866. 271 M. Yoon, K. Suh, S. Natarajan and K. Kim, Angew. Chem., Int. Ed., 2013, 52, 2688–2700. 272 C. Busche, L. Vila-Nadal, J. Yan, H. N. Miras, D.-L. Long, V. P. Georgiev, A. Asenov, R. H. Pedersen, N. Gadegaard, M. M. Mirza, D. J. Paul, J. M. Poblet and L. Cronin, Nature, 2014, 515, 545–549. 273 X. Xia, Y. Zhang, D. Chao, Q. Xiong, Z. Fan, X. Tong, J. Tu, H. Zhang and H. J. Fan, Energy Environ. Sci., 2015, 8, 1559–1568. 274 K. Xu, Chem. Rev., 2004, 104, 4303–4418. 275 E. Quartarone and P. Mustarelli, Chem. Soc. Rev., 2011, 40, 2525–2540. 276 R. G. Linford and S. Hackwood, Chem. Rev., 1981, 81, 327–364. This journal is © The Royal Society of Chemistry 2016 View Article Online Published on 17 February 2016. Downloaded by Texas A & M University on 30/03/2016 23:45:58. Chem Soc Rev 277 G. Kelemen, W. Lortz and G. Schön, J. Mater. Sci., 1989, 24, 333–338. 278 J. M. Taylor, S. Dekura, R. Ikeda and H. Kitagawa, Chem. Mater., 2015, 27, 2286–2289. 279 T. Yamada, K. Otsubo, R. Makiura and H. Kitagawa, Chem. Soc. Rev., 2013, 42, 6655–6669. 280 S. Horike, D. Umeyama and S. Kitagawa, Acc. Chem. Res., 2013, 46, 2376–2384. 281 J. Shi, Y. Jiang, X. Wang, H. Wu, D. Yang, F. Pan, Y. Su and Z. Jiang, Chem. Soc. Rev., 2014, 43, 5192–5210. 282 K. M. Choi, H. M. Jeong, J. H. Park, Y.-B. Zhang, J. K. Kang and O. M. Yaghi, ACS Nano, 2014, 8, 7451–7457. 283 P. Simon and Y. Gogotsi, Nat. Mater., 2008, 7, 845–854. 284 A. Macdonell, N. A. B. Johnson, A. J. Surman and L. Cronin, J. Am. Chem. Soc., 2015, 137, 5662–5665. 285 J.-J. Chen, M. D. Symes, S.-C. Fan, M.-S. Zheng, H. N. Miras, Q.-F. Dong and L. Cronin, Adv. Mater., 2015, 27, 4649–4654. 286 C. Wang, K. E. deKrafft and W. Lin, J. Am. Chem. Soc., 2012, 134, 7211–7214. 287 Z.-M. Zhang, T. Zhang, C. Wang, Z. Lin, L.-S. Long and W. Lin, J. Am. Chem. Soc., 2015, 137, 3197–3200. 288 K. M. Choi, K. Na, G. A. Somorjai and O. M. Yaghi, J. Am. Chem. Soc., 2015, 137, 7810–7816. 289 Z. Xu, L. Yang and C. Xu, Anal. Chem., 2015, 87, 3438–3444. 290 G.-l. Zhuang, J.-q. Bai, L. Tan, H.-l. Huang, Y.-f. Gao, X. Zhong, C.-l. Zhong and J.-g. Wang, RSC Adv., 2015, 5, 32714–32719. This journal is © The Royal Society of Chemistry 2016 Review Article 291 P. Horcajada, T. Chalati, C. Serre, B. Gillet, C. Sebrie, T. Baati, J. F. Eubank, D. Heurtaux, P. Clayette, C. Kreuz, J.-S. Chang, Y. K. Hwang, V. Marsaud, P.-N. Bories, L. Cynober, S. Gil, G. Ferey, P. Couvreur and R. Gref, Nat. Mater., 2010, 9, 172–178. 292 J. An, S. J. Geib and N. L. Rosi, J. Am. Chem. Soc., 2009, 131, 8376–8377. 293 R. Ananthoji, J. F. Eubank, F. Nouar, H. Mouttaki, M. Eddaoudi and J. P. Harmon, J. Mater. Chem., 2011, 21, 9587–9594. 294 W. J. Rieter, K. M. Pott, K. M. L. Taylor and W. Lin, J. Am. Chem. Soc., 2008, 130, 11584–11585. 295 N. J. Hinks, A. C. McKinlay, B. Xiao, P. S. Wheatley and R. E. Morris, Microporous Mesoporous Mater., 2010, 129, 330–334. 296 P. K. Allan, P. S. Wheatley, D. Aldous, M. I. Mohideen, C. Tang, J. A. Hriljac, I. L. Megson, K. W. Chapman, G. De Weireld, S. Vaesen and R. E. Morris, Dalton Trans., 2012, 41, 4060–4066. 297 C. Gaudin, D. Cunha, E. Ivanoff, P. Horcajada, G. Chevé, A. Yasri, O. Loget, C. Serre and G. Maurin, Microporous Mesoporous Mater., 2012, 157, 124–130. 298 D. Cunha, C. Gaudin, I. Colinet, P. Horcajada, G. Maurin and C. Serre, J. Mater. Chem. B, 2013, 1, 1101–1108. 299 C. He, K. Lu and W. Lin, J. Am. Chem. Soc., 2014, 136, 12253–12256. 300 B. Volosskiy, K. Niwa, Y. Chen, Z. Zhao, N. O. Weiss, X. Zhong, M. Ding, C. Lee, Y. Huang and X. Duan, ACS Nano, 2015, 9, 3044–3049. Chem. Soc. Rev.