Template for Electronic Submission to ACS Journals



The Anisotropic Responses of a Flexible Metal-Organic Framework Constructed from Asymmetric Flexible Linkers and Heptanuclear Zinc Carboxylate Secondary Building UnitsElliot J. Carrington, Rémi Pétuya, Rebecca K. Hylton, Yong Yan, Dmytro Antypov, George R. Darling, Matthew S. Dyer, Neil G. Berry, Alexandros P. Katsoulidis, and Matthew J. Rosseinsky*Department of Chemistry, University of Liverpool, Liverpool L69 7ZD, United KingdomABSTRACT A new porous and flexible metal-organic framework (MOF) has been synthesised from the flexible asymmetric linker N-(4-Carboxyphenyl)succinamate (CSA) and heptanuclear zinc oxo-clusters of formula [Zn7O2(Carboxylate)10DMF2] involving two coordinated terminal DMF ligands. The structural response of this MOF to the removal or exchange of its guest molecules has been probed using a combination of experimental and computational approaches. The topology of the material, involving double linker connections in the a and b directions and single linker connections along the c axis, is shown to be key in the material’s anisotropic response. The a and b directions remain locked during guest removal, while the c axis linker undergoes large changes significantly reducing the material’s void space. The changes to the c axis linker involve a combination of a hinge motion on the linker’s rigid side and conformational rearrangements on its flexible end, which were probed in detail during this process despite the presence of crystallographic disorder along this axis which prevented accurate characterisation by experimental methods alone. While inactive during guest removal, the flexible ends of the a and b axis linkers are observed to play a prominent role during DMF to DMSO solvent exchange, facilitating the exchange reaction arising in the cluster.IntroductionFlexible metal-organic frameworks (MOFs)ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/adma.201501523","ISBN":"1521-4095 (Electronic)\\r0935-9648 (Linking)","ISSN":"15214095","PMID":"26270630","abstract":"Flexible metal-organic frameworks (MOFs) receive much attention owing to their attractive properties that originate from their flexibility and dynamic behavior, and show great potential applications in many fields. Here, recent progress in the discovery, understanding, and property investigations of flexible MOFs are reviewed, and the examples of their potential applications in storage and separation, sensing, and guest capture and release are presented to highlight the developing trends in flexible MOFs.","author":[{"dropping-particle":"","family":"Chang","given":"Ze","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yang","given":"Dong Hui","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Jian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hu","given":"Tong Liang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bu","given":"Xian He","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Advanced Materials","id":"ITEM-1","issue":"36","issued":{"date-parts":[["2015"]]},"page":"5432-5441","title":"Flexible metal-organic frameworks: Recent advances and potential applications","type":"article-journal","volume":"27"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/ja411673b","ISBN":"0002-7863","ISSN":"00027863","PMID":"24460112","abstract":"Occasional, large amplitude flexibility in metal-organic frameworks (MOFs) is one of the most intriguing recent discoveries in chemistry and material science. Yet, there is at present no theoretical framework that permits the identification of flexible structures in the rapidly expanding universe of MOFs. Here, we propose a simple method to predict whether a MOF is flexible, based on treating it as a system of rigid elements, connected by hinges. This proposition is correct in application to MOFs based on rigid carboxylate linkers. We validate the method by correctly classifying known experimental MOFs into rigid and flexible groups. Applied to hypothetical MOFs, the method reveals an abundance of flexibility phenomena, and this seems to be at odds with the proportion of flexible structures among experimentally known MOFs. We speculate that the flexibility of a MOF may constitute an intrinsic impediment on its experimental realization. This highlights the importance of systematic prediction of large amplitude flexibility regimes in MOFs.","author":[{"dropping-particle":"","family":"Sarkisov","given":"Lev","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Martin","given":"Richard L.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Haranczyk","given":"Maciej","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Smit","given":"Berend","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-2","issue":"6","issued":{"date-parts":[["2014"]]},"page":"2228-2231","title":"On the flexibility of metal-organic frameworks","type":"article-journal","volume":"136"},"uris":["","",""]},{"id":"ITEM-3","itemData":{"DOI":"10.1039/C4CS00101J","ISSN":"0306-0012","author":[{"dropping-particle":"","family":"Schneemann","given":"a.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bon","given":"V.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schwedler","given":"I.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Senkovska","given":"I.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaskel","given":"S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fischer","given":"R. a.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chem. Soc. Rev.","id":"ITEM-3","issue":"16","issued":{"date-parts":[["2014"]]},"page":"6062-6096","title":"Flexible metal–organic frameworks","type":"article-journal","volume":"43"},"uris":[""]},{"id":"ITEM-4","itemData":{"DOI":"10.1039/C3CS60483G","ISBN":"0306-0012","ISSN":"0306-0012","PMID":"24699533","abstract":"This review presents the recent developments on FL-MOFs, including their structures and applications in gas adsorption, catalysis and proton conduction.","author":[{"dropping-particle":"","family":"Lin","given":"Zu-Jin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lü","given":"Jian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hong","given":"Maochun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cao","given":"Rong","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chem. Soc. Rev.","id":"ITEM-4","issue":"16","issued":{"date-parts":[["2014"]]},"page":"5867-5895","title":"Metal–organic frameworks based on flexible ligands (FL-MOFs): structures and applications","type":"article-journal","volume":"43"},"uris":["","",""]},{"id":"ITEM-5","itemData":{"DOI":"10.1038/nchem.444","ISSN":"17554330","abstract":"The field of host-guest complexation is intensely attractive from diverse perspectives, including materials science, chemistry and biology. The uptake and encapsulation of guest species by host frameworks are being investigated for a wide variety of purposes, including separation and storage using zeolites, and recognition and sensing by enzymes in solution. Here we focus on the concept of the cooperative integration of 'softness' and 'regularity'. Recent developments on porous coordination polymers (or metal-organic frameworks) have provided the inherent properties that combine these features. Such soft porous crystals exhibit dynamic frameworks that are able to respond to external stimuli such as light, electric fields or the presence of particular species, but they are also crystalline and can change their channels reversibly while retaining high regularity. We discuss the relationship between the structures and properties of these materials in view of their practical applications.","author":[{"dropping-particle":"","family":"Horike","given":"Satoshi","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Shimomura","given":"Satoru","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kitagawa","given":"Susumu","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-5","issue":"9","issued":{"date-parts":[["2009"]]},"page":"695-704","title":"Soft porous crystals","type":"article-journal","volume":"1"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>1–5</sup>","plainTextFormattedCitation":"1–5","previouslyFormattedCitation":"<sup>1–5</sup>"},"properties":{"noteIndex":0},"schema":""}1–5 are a relatively rare but interesting subset of extended framework materialsADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1126/science.1230444","ISBN":"1095-9203 (Electronic)\\n0036-8075 (Linking)","ISSN":"10959203","PMID":"23990564","abstract":"Crystalline metal-organic frameworks (MOFs) are formed by reticular synthesis, which creates strong bonds between inorganic and organic units. Careful selection of MOF constituents can yield crystals of ultrahigh porosity and high thermal and chemical stability. These characteristics allow the interior of MOFs to be chemically altered for use in gas separation, gas storage, and catalysis, among other applications. The precision commonly exercised in their chemical modification and the ability to expand their metrics without changing the underlying topology have not been achieved with other solids. MOFs whose chemical composition and shape of building units can be multiply varied within a particular structure already exist and may lead to materials that offer a synergistic combination of properties.","author":[{"dropping-particle":"","family":"Furukawa","given":"Hiroyasu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cordova","given":"Kyle E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Science","id":"ITEM-1","issued":{"date-parts":[["2013"]]},"page":"1230444","title":"The chemistry and applications of metal-organic frameworks","type":"article-journal","volume":"341"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1039/C4CS90059F","ISBN":"0306-0012\\r1460-4744","ISSN":"0306-0012","PMID":"25011480","abstract":"A graphical abstract is available for this content","author":[{"dropping-particle":"","family":"Zhou","given":"Hong-Cai “Joe”","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kitagawa","given":"Susumu","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chem. Soc. Rev.","id":"ITEM-2","issue":"16","issued":{"date-parts":[["2014"]]},"page":"5415-5418","title":"Metal–Organic Frameworks (MOFs)","type":"article-journal","volume":"43"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>6,7</sup>","plainTextFormattedCitation":"6,7","previouslyFormattedCitation":"<sup>6,7</sup>"},"properties":{"noteIndex":0},"schema":""}6,7 which can undergo structural changes in the presence of external stimuli.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/acs.chemmater.5b00046","ISSN":"0897-4756","author":[{"dropping-particle":"","family":"Coudert","given":"Fran?ois-Xavier","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemistry of Materials","id":"ITEM-1","issued":{"date-parts":[["2015"]]},"page":"150211155239004","title":"Responsive Metal–Organic Frameworks and Framework Materials: Under Pressure, Taking the Heat, In the Spotlight, With Friends","type":"article-journal"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>8</sup>","plainTextFormattedCitation":"8","previouslyFormattedCitation":"<sup>8</sup>"},"properties":{"noteIndex":0},"schema":""}8 Their potential to provide a highly adaptable pore environment, which changes its size and shape to suit the requirements of specific chemistry, makes them particularly attractive as the next generation of materials targeted at difficult separation, catalysis, adsorption or multifunctional applications (e.g., ferroelectric and non-linear optical properties).ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/acs.accounts.7b00151","ISSN":"0001-4842","author":[{"dropping-particle":"","family":"Karmakar","given":"Avishek","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Samanta","given":"Partha","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"V","family":"Desai","given":"Aamod","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ghosh","given":"Sujit K","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Accounts of Chemical Research","id":"ITEM-1","issue":"10","issued":{"date-parts":[["2017","10","17"]]},"page":"2457-2469","title":"Guest-Responsive Metal–Organic Frameworks as Scaffolds for Separation and Sensing Applications","type":"article-journal","volume":"50"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1002/adma.201501523","ISBN":"1521-4095 (Electronic)\\r0935-9648 (Linking)","ISSN":"15214095","PMID":"26270630","abstract":"Flexible metal-organic frameworks (MOFs) receive much attention owing to their attractive properties that originate from their flexibility and dynamic behavior, and show great potential applications in many fields. Here, recent progress in the discovery, understanding, and property investigations of flexible MOFs are reviewed, and the examples of their potential applications in storage and separation, sensing, and guest capture and release are presented to highlight the developing trends in flexible MOFs.","author":[{"dropping-particle":"","family":"Chang","given":"Ze","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yang","given":"Dong Hui","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Jian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hu","given":"Tong Liang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bu","given":"Xian He","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Advanced Materials","id":"ITEM-2","issue":"36","issued":{"date-parts":[["2015"]]},"page":"5432-5441","title":"Flexible metal-organic frameworks: Recent advances and potential applications","type":"article-journal","volume":"27"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/ja9000129","ISSN":"00027863","abstract":"The tetracarboxylate organic linker and Zn(II) ions assemble into chiral building blocks for a porous metal-organic framework with ferroelectric and second-order nonlinear optical properties.","author":[{"dropping-particle":"","family":"Guo","given":"Zhengang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cao","given":"Rong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Xin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Hongfang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yuan","given":"Wenbing","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Guojian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wu","given":"Haohan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Jing","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-3","issue":"20","issued":{"date-parts":[["2009"]]},"page":"6894-6895","title":"A multifunctional 3D ferroelectric and NLO-active porous metal-organic framework","type":"article-journal","volume":"131"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>1,9,10</sup>","plainTextFormattedCitation":"1,9,10","previouslyFormattedCitation":"<sup>1,9,10</sup>"},"properties":{"noteIndex":0},"schema":""}1,9,10 Recently, several attractive features for the development of future porous materials, in particular anisotropy and asymmetry, have been identified by Kitagawa.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/acs.accounts.6b00500","ISSN":"15204898","abstract":"Developing science and technology of porous materials provides fuels and useful substances from ubiquitous gaseous substances such as air.","author":[{"dropping-particle":"","family":"Kitagawa","given":"Susumu","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Accounts of Chemical Research","id":"ITEM-1","issue":"3","issued":{"date-parts":[["2017"]]},"page":"514-516","title":"Future porous materials","type":"article-journal","volume":"50"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>11</sup>","plainTextFormattedCitation":"11","previouslyFormattedCitation":"<sup>11</sup>"},"properties":{"noteIndex":0},"schema":""}11 However, in order to develop these new materials, we first need to improve our understanding of how their flexible responses arise.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1038/s41467-017-02666-y","ISSN":"20411723","abstract":"Knowledge of the thermodynamic potential in terms of the independent variables allows to characterize the macroscopic state of the system. However, in practice, it is difficult to access this potential experimentally due to irreversible transitions that occur between equilibrium states. A showcase example of sudden transitions between (meta)stable equilibrium states is observed for soft porous crystals possessing a network with long-range structural order, which can transform between various states upon external stimuli such as pressure, temperature and guest adsorption. Such phase transformations are typically characterized by large volume changes and may be followed experimentally by monitoring the volume change in terms of certain external triggers. Herein, we present a generalized thermodynamic approach to construct the underlying Helmholtz free energy as a function of the state variables that governs the observed behaviour based on microscopic simulations. This concept allows a unique identification of the conditions under which a material becomes flexible.","author":[{"dropping-particle":"","family":"Vanduyfhuys","given":"L.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rogge","given":"S. M.J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wieme","given":"J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vandenbrande","given":"S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Maurin","given":"G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Waroquier","given":"M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Speybroeck","given":"V.","non-dropping-particle":"Van","parse-names":false,"suffix":""}],"container-title":"Nature Communications","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2018"]]},"page":"1-9","publisher":"Springer US","title":"Thermodynamic insight into stimuli-responsive behaviour of soft porous crystals","type":"article-journal","volume":"9"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>12</sup>","plainTextFormattedCitation":"12","previouslyFormattedCitation":"<sup>12</sup>"},"properties":{"noteIndex":0},"schema":""}12 This requires in-depth structural characterisation of the MOFs before and after their stimulus-driven transitions, which is often hindered by their inherent porosity, structural disorder, or lack of crystallinity, making structural solution by traditional direct methods particularly challenging. To overcome these issues, it is possible to combine experimental and computational methods. For example, in the case of MIL-88ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ja054900x","ISBN":"0002-7863","ISSN":"0002-7863","PMID":"16287320","abstract":"Using a combination of simulations and powder diffraction, we report here the study of the very large swelling of a three-dimensional nanoporous iron(III) carboxylate (MIL-88) which exhibits almost a reversible doubling (approximately 85%) of its cell volume while fully retaining its open-framework topology. The crystal structure of the open form of MIL-88 has been successfully refined and indicates that atomic displacements larger than 4 angstroms are observed when water or various alcohols are adsorbed in the porous structure, revealing an unusually flexible crystallized framework. X-ray thermodiffractometry shows that only a displacive transition occurs during the swelling phenomenon, ruling out any bond breaking.","author":[{"dropping-particle":"","family":"Mellot-Draznieks","given":"Caroline","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Surblé","given":"Suzy","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Audebrand","given":"Nathalie","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-1","issue":"46","issued":{"date-parts":[["2005","11"]]},"page":"16273-16278","title":"Very Large Swelling in Hybrid Frameworks: A Combined Computational and Powder Diffraction Study","type":"article-journal","volume":"127"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>13</sup>","plainTextFormattedCitation":"13","previouslyFormattedCitation":"<sup>13</sup>"},"properties":{"noteIndex":0},"schema":""}13 and its isoreticular compounds,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/c8cy02329h","ISSN":"2044-4753","abstract":" Hierarchical Ni@C hollow spheres composed of dispersed Ni nanoparticles confined in carbon shells were readily synthesized for efficient CO 2 methanation. The hydrogenation of CO 2 to CH 4 is attracting growing attention due to the universal energy and climate challenges. In this work, we present the facile synthesis of hierarchical Ni@C spheres composed of dispersed Ni nanoparticles confined in carbon shells as an efficient catalyst for low-temperature CO 2 methanation at ambient pressure. With Ni-based metal–organic frameworks (MOFs) as a precursor, the Ni@C composite was produced by thermal annealing in a N 2 atmosphere. The MOF-derived Ni@C hybrid with unique hollow and porous structures could afford high surface area and rich isolated active sites for CO 2 adsorption/activation and redox processes. Therefore, the Ni@C catalyst exhibited high activity, excellent selectivity and superior stability for CO 2 methanation reaction. The possible intermediates and reaction pathway of the CO 2 conversion catalysis were carefully investigated by CO 2 -TPD measurements and in situ FTIR characterization. Furthermore, a possible CO 2 methanation mechanism over the Ni@C catalyst was proposed. ","author":[{"dropping-particle":"","family":"Wang","given":"Sibo","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lin","given":"Xiahui","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hou","given":"Yidong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ding","given":"Zhengxin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dai","given":"Wenxin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tu","given":"Wenguang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hu","given":"Zhibiao","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Rong","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Catalysis Science & Technology","id":"ITEM-1","issue":"3","issued":{"date-parts":[["2019"]]},"page":"731-738","publisher":"Royal Society of Chemistry","title":" MOF-derived hierarchical hollow spheres composed of carbon-confined Ni nanoparticles for efficient CO 2 methanation ","type":"article-journal","volume":"9"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1039/b512169h","ISBN":"1359-7345","ISSN":"1359-7345","PMID":"16391735","abstract":"We report here a new family of isoreticular MOFs, comprising three larger analogues of the nanoporous metallocarboxylate MIL-88; these solids were synthesized using a controlled SBU approach and the three crystal structures were solved using an original simulation-assisted structure determination method in direct space.","author":[{"dropping-particle":"","family":"Surblé","given":"Suzy","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mellot-Draznieks","given":"Caroline","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Millange","given":"Franck","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical communications (Cambridge, England)","id":"ITEM-2","issue":"3","issued":{"date-parts":[["2006"]]},"page":"284-286","title":"A new isoreticular class of metal-organic-frameworks with the MIL-88 topology.","type":"article-journal"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>14,15</sup>","plainTextFormattedCitation":"14,15","previouslyFormattedCitation":"<sup>14,15</sup>"},"properties":{"noteIndex":0},"schema":""}14,15 which display a large “swelling motion” enabled via a hinge motion, the relatively low quality of the diffraction patterns prevented the characterization of the framework's dynamics via crystallographic methods alone. Therefore, a combination of simulations (force field constrained optimizations) and powder diffraction refinements were employed to obtain a qualitative structural picture of the dynamical response.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/c8cy02329h","ISSN":"2044-4753","abstract":" Hierarchical Ni@C hollow spheres composed of dispersed Ni nanoparticles confined in carbon shells were readily synthesized for efficient CO 2 methanation. The hydrogenation of CO 2 to CH 4 is attracting growing attention due to the universal energy and climate challenges. In this work, we present the facile synthesis of hierarchical Ni@C spheres composed of dispersed Ni nanoparticles confined in carbon shells as an efficient catalyst for low-temperature CO 2 methanation at ambient pressure. With Ni-based metal–organic frameworks (MOFs) as a precursor, the Ni@C composite was produced by thermal annealing in a N 2 atmosphere. The MOF-derived Ni@C hybrid with unique hollow and porous structures could afford high surface area and rich isolated active sites for CO 2 adsorption/activation and redox processes. Therefore, the Ni@C catalyst exhibited high activity, excellent selectivity and superior stability for CO 2 methanation reaction. The possible intermediates and reaction pathway of the CO 2 conversion catalysis were carefully investigated by CO 2 -TPD measurements and in situ FTIR characterization. Furthermore, a possible CO 2 methanation mechanism over the Ni@C catalyst was proposed. ","author":[{"dropping-particle":"","family":"Wang","given":"Sibo","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lin","given":"Xiahui","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hou","given":"Yidong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ding","given":"Zhengxin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dai","given":"Wenxin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tu","given":"Wenguang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hu","given":"Zhibiao","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Rong","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Catalysis Science & Technology","id":"ITEM-1","issue":"3","issued":{"date-parts":[["2019"]]},"page":"731-738","publisher":"Royal Society of Chemistry","title":" MOF-derived hierarchical hollow spheres composed of carbon-confined Ni nanoparticles for efficient CO 2 methanation ","type":"article-journal","volume":"9"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1039/b512169h","ISBN":"1359-7345","ISSN":"1359-7345","PMID":"16391735","abstract":"We report here a new family of isoreticular MOFs, comprising three larger analogues of the nanoporous metallocarboxylate MIL-88; these solids were synthesized using a controlled SBU approach and the three crystal structures were solved using an original simulation-assisted structure determination method in direct space.","author":[{"dropping-particle":"","family":"Surblé","given":"Suzy","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mellot-Draznieks","given":"Caroline","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Millange","given":"Franck","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical communications (Cambridge, England)","id":"ITEM-2","issue":"3","issued":{"date-parts":[["2006"]]},"page":"284-286","title":"A new isoreticular class of metal-organic-frameworks with the MIL-88 topology.","type":"article-journal"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/ja054900x","ISBN":"0002-7863","ISSN":"0002-7863","PMID":"16287320","abstract":"Using a combination of simulations and powder diffraction, we report here the study of the very large swelling of a three-dimensional nanoporous iron(III) carboxylate (MIL-88) which exhibits almost a reversible doubling (approximately 85%) of its cell volume while fully retaining its open-framework topology. The crystal structure of the open form of MIL-88 has been successfully refined and indicates that atomic displacements larger than 4 angstroms are observed when water or various alcohols are adsorbed in the porous structure, revealing an unusually flexible crystallized framework. X-ray thermodiffractometry shows that only a displacive transition occurs during the swelling phenomenon, ruling out any bond breaking.","author":[{"dropping-particle":"","family":"Mellot-Draznieks","given":"Caroline","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Surblé","given":"Suzy","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Audebrand","given":"Nathalie","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-3","issue":"46","issued":{"date-parts":[["2005","11"]]},"page":"16273-16278","title":"Very Large Swelling in Hybrid Frameworks: A Combined Computational and Powder Diffraction Study","type":"article-journal","volume":"127"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>13–15</sup>","plainTextFormattedCitation":"13–15","previouslyFormattedCitation":"<sup>13–15</sup>"},"properties":{"noteIndex":0},"schema":""}13–15 In this paper we present a combined computational and crystallographic study reporting the behaviour of the new framework ZnCSA, where CSA (Figure 1(a)) is the asymmetric flexible linker N-(4-Carboxyphenyl)succinamate. CSA was selected because it combines flexible and rigid characteristics, as it can be considered to have one rigid end, consisting of an aromatic ring, and one flexible end, built from amide and aliphatic functionalities. The inherent flexibility involves sp3 carbons and rotatable bonds, which have previously been shown to produce flexible responses in a range of MOFs.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/anie.201307074","ISBN":"1521-3773","ISSN":"14337851","PMID":"24302659","abstract":"The peptide-based porous 3D framework, ZnCar, has been synthesized from Zn(2+) and the natural dipeptide carnosine (β-alanyl-L-histidine). Unlike previous extended peptide networks, the imidazole side chain of the histidine residue is deprotonated to afford Zn-imidazolate chains, with bonding similar to the zeolitic imidazolate framework (ZIF) family of porous materials. ZnCar exhibits permanent microporosity with a surface area of 448 m(2) g(-1) , and its pores are 1D channels with 5 ? openings and a characteristic chiral shape. This compound is chemically stable in organic solvents and water. Single-crystal X-ray diffraction (XRD) showed that the ZnCar framework adapts to MeOH and H2 O guests because of the torsional flexibility of the main His-β-Ala chain, while retaining the rigidity conferred by the Zn-imidazolate chains. The conformation adopted by carnosine is driven by the H bonds formed both to other dipeptides and to the guests, permitting the observed structural transformations.","author":[{"dropping-particle":"","family":"Katsoulidis","given":"Alexandros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Park","given":"Kyo Sung","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"Dmytro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Martí-Gastaldo","given":"Carlos","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Miller","given":"Gary J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"John E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Robertson","given":"Craig M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Blanc","given":"Frédéric","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"George R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"Neil G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Purton","given":"John A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Adams","given":"Dave J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"Matthew J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2014","1","3"]]},"page":"193-198","title":"Guest-Adaptable and Water-Stable Peptide-Based Porous Materials by Imidazolate Side Chain Control","type":"article-journal","volume":"53"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1038/nchem.1871","ISBN":"1755-4349","ISSN":"1755-4330","PMID":"24651203","abstract":"Porous materials are attractive for separation and catalysis-these applications rely on selective interactions between host materials and guests. In metal-organic frameworks (MOFs), these interactions can be controlled through a flexible structural response to the presence of guests. Here we report a MOF that consists of glycyl-serine dipeptides coordinated to metal centres, and has a structure that evolves from a solvated porous state to a desolvated non-porous state as a result of ordered cooperative, displacive and conformational changes of the peptide. This behaviour is driven by hydrogen bonding that involves the side-chain hydroxyl groups of the serine. A similar cooperative closure (reminiscent of the folding of proteins) is also displayed with multipeptide solid solutions. For these, the combination of different sequences of amino acids controls the framework's response to the presence of guests in a nonlinear way. This functional control can be compared to the effect of single-point mutations in proteins, in which exchange of single amino acids can radically alter structure and function.","author":[{"dropping-particle":"","family":"Martí-Gastaldo","given":"C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"D.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"J. E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Briggs","given":"M. E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chater","given":"P. A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"V.","family":"Wiper","given":"P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Miller","given":"G. J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Khimyak","given":"Y. Z.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"G. R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"N. G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"M. J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-2","issue":"4","issued":{"date-parts":[["2014","4","23"]]},"page":"343-351","title":"Side-chain control of porosity closure in single- and multiple-peptide-based porous materials by cooperative folding","type":"article-journal","volume":"6"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1038/s41586-018-0820-9","ISSN":"0028-0836","abstract":"Metal–organic frameworks (MOFs) are crystalline synthetic porous materials formed by binding organic linkers to metal nodes: they can be either rigid1,2 or flexible3. Zeolites and rigid MOFs have widespread applications in sorption, separation and catalysis that arise from their ability to control the arrangement and chemistry of guest molecules in their pores via the shape and functionality of their internal surface, defined by their chemistry and structure4,5. Their structures correspond to an energy landscape with a single, albeit highly functional, energy minimum. By contrast, proteins function by navigating between multiple metastable structures using bond rotations of the polypeptide6,7, where each structure lies in one of the minima of a conformational energy landscape and can be selected according to the chemistry of the molecules that interact with the protein. These structural changes are realized through the mechanisms of conformational selection (where a higher-energy minimum characteristic of the protein is stabilized by small-molecule binding) and induced fit (where a small molecule imposes a structure on the protein that is not a minimum in the absence of that molecule)8. Here we show that rotation about covalent bonds in a peptide linker can change a flexible MOF to afford nine distinct crystal structures, revealing a conformational energy landscape that is characterized by multiple structural minima. The uptake of small-molecule guests by the MOF can be chemically triggered by inducing peptide conformational change. This change transforms the material from a minimum on the landscape that is inactive for guest sorption to an active one. Chemical control of the conformation of a flexible organic linker offers a route to modifying the pore geometry and internal surface chemistry and thus the function of open-framework materials.","author":[{"dropping-particle":"","family":"Katsoulidis","given":"Alexandros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"Dmytro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Whitehead","given":"George F. S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Carrington","given":"Elliot J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Adams","given":"Dave J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"Neil G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"George R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dyer","given":"Matthew S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"Matthew J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature","id":"ITEM-3","issue":"7738","issued":{"date-parts":[["2019","1","9"]]},"page":"213-217","publisher":"Springer US","title":"Chemical control of structure and guest uptake by a conformationally mobile porous material","type":"article-journal","volume":"565"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>16–18</sup>","plainTextFormattedCitation":"16–18","previouslyFormattedCitation":"<sup>16–18</sup>"},"properties":{"noteIndex":0},"schema":""}16–18 Although reported as an organic linker in several extended framework materials, the potential of CSA for introducing flexible responses associated with the distinct ends of the linker has not yet been explored.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/C7FD00085E","ISSN":"1359-6640","PMID":"28612863","abstract":"Two new amide functionalised metal–organic frameworks, In(OH)CSA and In(OH)PDG, were synthesized using two flexible linkers, N -(4-carboxyphenyl)succinamic acid (CSA) and N , N ′-(1,4-phenylenedicarbonyl)diglycine (PDG), respectively. Both structures consist of corner-sharing {InO 4 (OH) 2 } octahedra in the form of trans indium hydroxide chains, which are interconnected by the dicarboxylate linkers to form stacked 2-dimensional layers. The different symmetries and configurations of the flexible and rigid features on the linkers results in different supramolecular interactions dominating between linkers, resulting in different shaped pores and functional group orientation. In(OH)CSA lacks hydrogen bonding between linkers, which results in close packing between the layers and very small solvent accessible pores running perpendicular to the plane of the layers. In(OH)PDG exhibits strong intra- and interlayer hydrogen bonding, which prevents the layers from close packing and results in larger cylindrical pores running parallel to the indium hydroxide chains, producing a total accessible volume of 25% of the unit cell volume.","author":[{"dropping-particle":"","family":"Haddad","given":"J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Whitehead","given":"G. F. S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Katsoulidis","given":"A. P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"M. J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Faraday Discussions","id":"ITEM-1","issued":{"date-parts":[["2017"]]},"page":"327-335","publisher":"Royal Society of Chemistry","title":"In-MOFs based on amide functionalised flexible linkers","type":"article-journal","volume":"201"},"uris":[""]},{"id":"ITEM-2","itemData":{"author":[{"dropping-particle":"","family":"Li, Qian-hai, Jiang, Shun-feng","given":"Cheng Peng-xiang","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Synthetic Crystals","id":"ITEM-2","issue":"7","issued":{"date-parts":[["2016"]]},"page":"1875","title":"No Title","type":"article-journal"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>19,20</sup>","plainTextFormattedCitation":"19,20","previouslyFormattedCitation":"<sup>19,20</sup>"},"properties":{"noteIndex":0},"schema":""}19,20 The secondary building unit (SBU) in ZnCSA is an heptanuclear zinc cluster with two coordinated terminal DMF ligands of formula [Zn7O2(Carboxylate)10DMF2]. While MOFs built from the same SBU, but incorporating rigid bidentate carboxylate based organic linkers, have previously been reported with the same overall topology,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/chem.200500963","ISBN":"0947-6539","ISSN":"09476539","abstract":"A novel, three-dimensional, noninterpenetrating microporous metal-organic framework (MOF), [Zn7O2(pda)5(H2O)2]5 DMF4 EtOH 6 H2O (1) (H2PDA=p-phenylenediacrylic acid, DMF=N,N-dimethylformamide, EtOH=ethanol), was synthesized by constructing heptanuclear zinc carboxylate secondary building units (SBUs) and by using rigid and linear aromatic carboxylate ligands, PDA. The X-ray crystallographic data reveals that the seven zinc centers of 1 are held together with ten carboxylate groups of the PDA ligands and four water molecules to form a heptametallic SBU, Zn7O4(CO2)10, with dimensions of 9.8 x 9.8 x 13.8 A3. Furthermore, the heptametallic SBUs are interconnected by PDA acting as linkers, thereby generating an extended network with a three-dimensional, noninterpenetrating, intersecting large-channel system with spacing of about 17.3 A. As a microporous framework, polymer 1 shows adsorption behavior that is favorable towards H2O and CH3OH, and substantial H2 uptake. In terms of the heptanuclear zinc carboxylate SBUs, polymer 1 exhibits interesting photoelectronic properties, which would facilitate the exploration of new types of semiconducting materials, especially among MOFs containing multinuclear metal carboxylate SBUs.","author":[{"dropping-particle":"","family":"Fang","given":"Qian Rong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zhu","given":"Guang Shan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xue","given":"Ming","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zhang","given":"Qing Lin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Sun","given":"Jin Yu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Guo","given":"Xiao Dan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Qiu","given":"Shi Lun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Shi Tao","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Ping","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"De Jun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wei","given":"Yen","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemistry - A European Journal","id":"ITEM-1","issue":"14","issued":{"date-parts":[["2006"]]},"page":"3754-3758","title":"Microporous metal-organic framework constructed from heptanuclear zinc carboxylate secondary building units","type":"article-journal","volume":"12"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1002/anie.201202925","ISSN":"14337851","author":[{"dropping-particle":"","family":"Choi","given":"Sang Beom","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Furukawa","given":"Hiroyasu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Nam","given":"Hye Jin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jung","given":"Duk-young","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jhon","given":"Young Ho","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Walton","given":"Allan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Book","given":"David","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kim","given":"Jaheon","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-2","issue":"35","issued":{"date-parts":[["2012","8","27"]]},"page":"8791-8795","title":"Reversible Interpenetration in a Metal-Organic Framework Triggered by Ligand Removal and Addition","type":"article-journal","volume":"51"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/ic302127y","ISSN":"0020-1669","abstract":"A new metal-organic framework, CALF-22 comprising Zn7O2(COO)10 secondary building units and 2-nitro-1,4-benzenedicarboxylate, is reported. The porosity and gas adsorption of N2, H2, CO2, and CH4 are studied, and CALF-22 has a surface area in excess of 1000 m(2)/g. The stability of the larger zinc cluster and the effect of the nitro group on gas sorption are also studied.","author":[{"dropping-particle":"","family":"Iremonger","given":"Simon S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vaidhyanathan","given":"Ramanathan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mah","given":"Roger K.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Shimizu","given":"George K. H.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-3","issue":"8","issued":{"date-parts":[["2013","4","15"]]},"page":"4124-4126","title":"Zn 7 O 2 (RCOO) 10 Clusters and Nitro Aromatic Linkers in a Porous Metal–Organic Framework","type":"article-journal","volume":"52"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>21–23</sup>","plainTextFormattedCitation":"21–23","previouslyFormattedCitation":"<sup>21–23</sup>"},"properties":{"noteIndex":0},"schema":""}21–23 the incorporation of the flexible CSA linkers enables ZnCSA to display new dynamical behaviours. The three-dimensional connectivity of ZnCSA could be determined by crystallographic methods, enabling us to describe the connections between the inorganic cluster and organic linker components, however, a subset of the CSA linkers displayed crystallographic disorder of the linker orientation, and therefore the relative binding of the rigid and flexible ends, making the exact linker environment hard to accurately characterise. In addition, ZnCSA loses single crystal crystallinity upon complete removal of its guest species. These obstacles required additional information beyond crystallography to develop an understanding of the flexible response mechanism of the material. Periodic density functional theory (DFT) calculations were therefore used in conjunction with crystallographic data to guide our understanding of the system. These methods revealed that the topology plays a crucial role in the behaviour of the framework, with the framework displaying an anisotropic response depending on the environments of the individual linkers. Post-synthetic single crystal coordinated solvent exchange (SCCSE)ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ic3005085","ISSN":"0020-1669","abstract":"Single-crystal-to-single-crystal (SCSC) transformations represent some of the most fascinating phenomena in chemistry. They are not only intriguing from a basic science point of view but also provide a means to modify or tune the properties of the materials via the postsynthetic introduction of suitable guest molecules or organic functional groups into their structures. Here, we describe UCY-2, a new flexible Nd3+ metal-organic framework (MOF), which exhibits a unique capability to undergo a plethora of SCSC transformations with some of them being very uncommon. These structural alterations involve the replacement of coordinating solvent molecules of UCY-2 by terminally ligating solvents and organic ligands with multiple functional groups including -OH, -SH, -NH-, and -NH2 or their combinations, chelating ligands, anions, and two different organic compounds. The SCSC coordinating solvent exchange is thus demonstrated as a powerful method for the functionalization of MOFs.","author":[{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kyprianidou","given":"Eleni J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Papaefstathiou","given":"Giannis S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-1","issue":"11","issued":{"date-parts":[["2012","6","4"]]},"page":"6308-6314","title":"Insertion of Functional Groups into a Nd 3+ Metal–Organic Framework via Single-Crystal-to-Single-Crystal Coordinating Solvent Exchange","type":"article-journal","volume":"51"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>24</sup>","plainTextFormattedCitation":"24","previouslyFormattedCitation":"<sup>24</sup>"},"properties":{"noteIndex":0},"schema":""}24 of the DMF ligands to DMSO was also performed, resulting in a new material that displayed a significantly different predicted response during its guest removal. While postsynthetic modification in general has shown great potential in the functionalization of MOFs, most examples involve ligandADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/C0CS00031K","ISSN":"0306-0012","abstract":"Metal-organic frameworks (MOFs) are an important class of hybrid inorganic-organic materials. In this tutorial review, a progress report on the postsynthetic modification (PSM) of MOFs is provided. PSM refers to the chemical modification of the MOF lattice in a heterogeneous fashion. This powerful synthetic approach has grown in popularity and resulted in a number of advances in the functionalization and application of MOFs. The use of PSM to develop MOFs with improved gas sorption, catalytic activity, bioactivity, and more robust physical properties is discussed. The results reported to date clearly show that PSM is an important approach for the development and advancement of these hybrid solids.","author":[{"dropping-particle":"","family":"Tanabe","given":"Kristine K.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cohen","given":"Seth M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chem. Soc. Rev.","id":"ITEM-1","issue":"2","issued":{"date-parts":[["2011"]]},"page":"498-519","title":"Postsynthetic modification of metal–organic frameworks—a progress report","type":"article-journal","volume":"40"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/cr200179u","ISSN":"1520-6890","PMID":"21916418","author":[{"dropping-particle":"","family":"Cohen","given":"Seth M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical reviews","id":"ITEM-2","issue":"2","issued":{"date-parts":[["2012","2","8"]]},"page":"970-1000","title":"Postsynthetic methods for the functionalization of metal-organic frameworks.","type":"article-journal","volume":"112"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/jacs.6b11259","ISSN":"0002-7863","abstract":"Metal?organic frameworks (MOFs) have rapidly grown into a major area of chemical research over the last two decades. MOFs represent the development of covalent chemistry \"beyond the molecule\" and into extended structures. MOFs also present an unprecedented scaffold for performing heterogeneous organic transformations in the solid state, allowing for deliberate and precise preparation of new materials. The development of these transformations has given rise to the \"postsynthetic renaissance\", a suite of methods by which these materials can be transformed in a single-crystal-to-single-crystal manner. Postsynthetic modification, postsynthetic depro-tection, postsynthetic exchange, postsynthetic insertion, and postsynthetic polymerization have exploited the unique features of both the organic and inorganic components of MOFs to create crystalline, porous solids of unique complexity and functionality.","author":[{"dropping-particle":"","family":"Cohen","given":"Seth M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-3","issue":"8","issued":{"date-parts":[["2017","3","10"]]},"page":"2855-2863","title":"The Postsynthetic Renaissance in Porous Solids","type":"article-journal","volume":"139"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>25–27</sup>","plainTextFormattedCitation":"25–27","previouslyFormattedCitation":"<sup>25–27</sup>"},"properties":{"noteIndex":0},"schema":""}25–27 or cation exchange ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/c4cs00002a","ISSN":"14604744","abstract":"This review examines the governing factors of cation exchange at the secondary building units of MOFs and reviews its applications.Cation exchange is an emerging synthetic route for modifying the secondary building units (SBUs) of metal–organic frameworks (MOFs). This technique has been used extensively to enhance the properties of nanocrystals and molecules, but the extent of its applications for MOFs is still expanding. To harness cation exchange as a rational tool, we need to elucidate its governing factors. Not nearly enough experimental observations exist for drawing these conclusions, so we provide a conceptual framework for approaching this task. We address which SBUs undergo exchange, why certain ions replace others, how the framework influences the process, the role of the solvent, and current applications. Using these guidelines, certain trends emerge from the available data and missing experiments become obvious. If future studies follow this framework, then a more comprehensive body of observations will furnish a deeper understanding of cation exchange and inspire future applications.","author":[{"dropping-particle":"","family":"Brozek","given":"C. K.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dincǎ","given":"M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical Society Reviews","id":"ITEM-1","issue":"16","issued":{"date-parts":[["2014"]]},"page":"5456-5467","title":"Cation exchange at the secondary building units of metal-organic frameworks","type":"article-journal","volume":"43"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1002/chem.201200137","ISSN":"09476539","abstract":"Two anionic metal-organic frameworks were successfully prepared based on pre-designed flexible multicarboxylate ligands and indium cations. Owing to the flexibility of the bridging organic linkers, which could not themselves sustain the frameworks, both of the frameworks showed thermal instability and shrinkage after removal of guest solvent molecules. Inspired by bamboo, we used a guest-dependent approach to tune the permanent porosity of the MOFs. In this approach, several tetraalkyammonium cations of different sizes were introduced into the channels by cation exchange to act as partitions and to support the main frameworks. This approach significantly enhanced the stability of the framework and its permanent porosity. Moreover, the gas-adsorption properties (such as gate sorption, hysteresis, and selectivity) of the MOFs were also modulated by the judicious choice of guest cations.","author":[{"dropping-particle":"","family":"Lin","given":"Zu Jin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Liu","given":"Tian Fu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Huang","given":"Yuan Biao","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jian Lü","given":"","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cao","given":"Rong","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemistry - A European Journal","id":"ITEM-2","issue":"25","issued":{"date-parts":[["2012"]]},"page":"7896-7902","title":"A guest-dependent approach to retain permanent pores in flexible metal- Organic frameworks by cation exchange","type":"article-journal","volume":"18"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/ic1019229","ISSN":"00201669","abstract":"A series of coordination polymers with anionic, cationic, and neutral metal-carboxylate frameworks have been synthesized by using a flexible tetrapodal ligand tetrakis[4-(carboxyphenyl)oxamethyl] methane acid (H(4)X). The reactions between divalent transition-metal ions and H(4)X ligands gave [M(3)X(2)]·[NH(2)(CH(3))(2)](2)·8DMA (M = Co (1), Mn (2), Cd(3)) which have anionic metal-carboxylate frameworks with NH(2)(CH(3))(2)(+) cations filled in channels. The reactions of trivalent metal ions Y(III), Dy(III), and In(III) with H(4)X ligands afforded cationic metal-carboxylate frameworks [M(3)X(2)·(NO(3))·(DMA)(2)·(H(2)O)]·5DMA·2H(2)O (M = Y(4), Dy(5)) and [In(2)X·(OH)(2)]·3DMA·6H(2)O (6) with the NO(3)(-) and OH(-) serving as counterions, respectively. Moreover, a neutral metal-carboxylate framework [Pb(2)X·(DMA)(2)]·2DMA (7) can also be isolated from reaction of Pb(II) and H(4)X ligands. The charged metal-carboxylate frameworks 1-5 have selectivity for specific counterions in the reaction system, and compounds 1 and 2 display ion-exchange behavior. Moreover, magnetic property measurements on compounds 1, 2, and 5 indicate that there exists weak antiferromagnetic interactions between magnetic centers in the three compounds.","author":[{"dropping-particle":"","family":"Liu","given":"Tian Fu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lü","given":"Jian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tian","given":"Chongbin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cao","given":"Minna","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lin","given":"Zujin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cao","given":"Rong","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-3","issue":"6","issued":{"date-parts":[["2011"]]},"page":"2264-2271","title":"Porous anionic, cationic, and neutral metal-carboxylate frameworks constructed from flexible tetrapodal ligands: Syntheses, structures, ion-exchanges, and magnetic properties","type":"article-journal","volume":"50"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>28–30</sup>","plainTextFormattedCitation":"28–30","previouslyFormattedCitation":"<sup>28–30</sup>"},"properties":{"noteIndex":0},"schema":""}28–30 and such SCCSE are relatively rarely reported.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ic502639v","ISSN":"0020-1669","abstract":"Solvent-assisted ligand incorporation (SALI) is useful for functionalizing the channels of metal–organic framework (MOF) materials such as NU-1000 that offer substitutionally labile zirconium(IV) coordination sites for nonbridging ligands. Each of the 30 or so previous examples relied upon coordination of a carboxylate ligand to achieve incorporation. Here we show that, with appropriate attention to ligand/node stoichiometry, SALI can also be achieved with phosphonate-terminated ligands. Consistent with stronger M(IV) coordination of phosphonates versus carboxylates, this change extends the pH range for retention of incorporated ligands. The difference in coordination strength can be exploited to achieve stepwise incorporation of pairs of ligands—specifically, phosphonates species followed by carboxylate species—without danger of displacement of the first ligand type by the second. Diffuse reflectance infrared Fourier-transform spectroscopy suggests that the phosphonate ligands are connected to the MOF node as RPO2(OH)? species in a moiety that leaves a base-accessible ?OH moiety on each bound phosphonate.","author":[{"dropping-particle":"","family":"Deria","given":"Pravas","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bury","given":"Wojciech","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hod","given":"Idan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kung","given":"Chung-Wei","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Karagiaridi","given":"Olga","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hupp","given":"Joseph T.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Farha","given":"Omar K.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-1","issue":"5","issued":{"date-parts":[["2015","3","2"]]},"page":"2185-2192","title":"MOF Functionalization via Solvent-Assisted Ligand Incorporation: Phosphonates vs Carboxylates","type":"article-journal","volume":"54"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/cg501141m","ISSN":"1528-7483","abstract":"The synthesis and characterization of {[Co9(INA)18(H2O)6]·11DMF·15H2O}∞ (Co9-INA·11DMF·15H2O) (INA- = the anion of isonicotinic acid) is reported. It exhibits a rigid 3D-porous structure with a Co9 repeating unit consisting of four [CoII 2(μ-O2CR)2(μ-H2O)] subunits (two unique) linked through bridging INA- ligands to an isolated CoII ion (half unique). The [CoII 2] dimers and the isolated CoII ion have assembled to create a trinodal (6,7,8)-coordinated network with point symbol (32.411.56.62)2(32.418.54.64)2(34.44.54.63). Gas sorption studies revealed that Co9-INA exhibits 910 m2 g-1 BET area, 4.2 mmol g-1 CO2 uptake at 273 K/1 bar, and 6.7 CO2/CH4 selectivity at zero coverage. Furthermore, Co9-INA displays capability for exchange of the guest solvent molecules by various organic molecules in a single-crystal to single-crystal fashion. Direct and alternating current magnetic susceptibility studies revealed the existence of dominant antiferromagnetic interactions between the Co2+ ions that result in a paramagnetic ST = 3/2 spin ground state value. Overall, this work emphasizes the potential of relatively simple and inexpensive polytopic ligands, such as isonicotic acid, to stabilize microporous MOFs with significant CO2 sorption capacity. ? 2014 American Chemical Society.","author":[{"dropping-particle":"","family":"Moushi","given":"Eleni E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kourtellaris","given":"Andreas","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Spanopoulos","given":"Ioannis","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Papaefstathiou","given":"Giannis S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Trikalitis","given":"Pantelis N.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Crystal Growth & Design","id":"ITEM-2","issue":"1","issued":{"date-parts":[["2015","1","7"]]},"page":"185-193","title":"A Microporous Co 2+ Metal Organic Framework with Single-Crystal to Single-Crystal Transformation Properties and High CO 2 Uptake","type":"article-journal","volume":"15"},"uris":["","",""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/acs.cgd.6b01366","ISSN":"1528-7483","abstract":"Four similar Mn(II) metal–organic frameworks (MOFs), {[Mn2(nbtc)(H2O)2(S)]·S·0.5H2O}n [S = DMF (1), DMA (2), NMP (3), DEF (4)] (DMF = N,N′-dimethylformamide, DMA = N,N′-dimethylacetamide, NMP = N-methyl-2-pyrrolidinone, DEF = N,N′-diethylformamide), have been assembled solvothermally from the nitro and carboxyl doubly functionalized ligand 6,6′-dinitro-2,2′,4,4′-biphenyl tetracarboxylic acid (H4nbtc) and characterized by single-crystal X-ray diffraction, elemental analyses, infrared spectroscopy, thermogravimetric analyses, and powder X-ray diffraction. All MOFs exhibit unique three-dimensional double-walled open-frameworks with one-dimensional parallelogram channels and have guest/coordinated water and carbonyl solvent molecules, reasonably providing a good example of the competitive behavior of water and carbonyl solvent molecules and an excellent candidate for studying the single crystal coordinated solvent exchange transformations. Interestingly, because of the different steric hindrance of the series...","author":[{"dropping-particle":"","family":"Zhang","given":"Wen-Qian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zhang","given":"Wen-Yan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Rui-Dong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ren","given":"Chun-Yan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Quan-Quan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fan","given":"Yan-Ping","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Liu","given":"Bin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Liu","given":"Ping","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Yao-Yu","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Crystal Growth & Design","id":"ITEM-3","issue":"2","issued":{"date-parts":[["2017","2","11"]]},"page":"517-526","title":"Effect of Coordinated Solvent Molecules on Metal Coordination Sphere and Solvent-Induced Transformations","type":"article-journal","volume":"17"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>31–33</sup>","plainTextFormattedCitation":"31–33","previouslyFormattedCitation":"<sup>31–33</sup>"},"properties":{"noteIndex":0},"schema":""}31–33 These exchanges, however, offer a distinct route to control MOF behaviour, including the modulation of MOFs luminescence properties, ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/c3ta14489e","ISSN":"2050-7488","abstract":"The discovery of new methods for the post-synthesis modification of materials is essential in order to establish suitable strategies for the tuning of their properties in a rational manner. Here we present a series of single-crystal-to-single-crystal (SCSC) transformations for the flexible [Eu2(CIP)2(DMF)2(H2O)2] (UCY-8) [H3CIP = 5-(4-carboxybenzylideneamino)isophthalic acid] and rigid [Eu2(N-BDC)3(DMF)4] (EuN-BDC) (H 2N-BDC = 2-amino-1,4-benzene dicarboxylic acid) Metal-Organic Frameworks (MOFs) that involve the replacement of their coordinating solvent molecules by terminally ligating organic molecules with multiple functional groups including -OH, -SH, -NH- and -NH2 or their combinations, chelating ligands, and two different organic compounds. The capability of the flexible MOF, which contains small pores and channels (<4 ?), to exchange its coordinating solvent molecules by relatively bulky molecules (such as pyridine, 2-hydroxymethyl-phenol, etc.) is shown to be the result of its breathing capacity. Remarkably, the rigid MOF is also highly capable of replacing its coordinating solvent molecules by bulky ligands, despite its small pores (2-5 ?) and lack of structural flexibility. Interestingly, the insertion of some organic ligands into the rigid MOF results in a significant modification of its framework structure and substantial expansion of its potential void space. Not only a plethora of exchanged analogues of these MOFs have been isolated and crystallographically characterized, but also, in some cases, a tremendous enhancement of their Eu3+-based photoluminescence (PL) signals, lifetimes and quantum yields (up to ~16 times) compared to those of the pristine materials has been observed due to the replacement of terminal solvents by organic ligands being efficient sensitizers for the Eu 3+ ion. Overall this work indicates that the Single Crystal Coordinating Solvent Exchange (SCCSE) can be applied as a general post-synthetic modification method for LnMOFs and also constitutes a highly efficient strategy for the enhancement of the Ln3+-based PL. ? 2014 the Partner Organisations.","author":[{"dropping-particle":"","family":"Kyprianidou","given":"Eleni J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lazarides","given":"Theodore","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaziannis","given":"Spyridon","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kosmidis","given":"Constantine","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Itskos","given":"Grigorios","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Materials Chemistry A","id":"ITEM-1","issue":"15","issued":{"date-parts":[["2014"]]},"page":"5258","title":"Single crystal coordinating solvent exchange as a general method for the enhancement of the photoluminescence properties of lanthanide MOFs","type":"article-journal","volume":"2"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/ic202635a","ISSN":"0020-1669","abstract":"In this work, for the first time, we have systematically demonstrated that solvent plays crucial roles in both controllable synthesis of metal-organic frameworks (MOFs) and their structural transformation process. With solvent as the only variable, five new MOFs with different structures have been constructed, in which one MOF undergoes solvent-induced single-crystal to single-crystal (SCSC) transformation that involves not only solvent exchange but also the cleavage and formation of coordination bonds. Particularly, a significant crystallographic change has been realized through an unprecedented three-step SCSC transformation process. Furthermore, we have demonstrated that the obtained MOF could be an excellent host for chromophores such as Alq3 for modulated luminescent properties.","author":[{"dropping-particle":"","family":"Lan","given":"Ya-Qian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jiang","given":"Hai-Long","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Shun-Li","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Qiang","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-2","issue":"14","issued":{"date-parts":[["2012","7","16"]]},"page":"7484-7491","title":"Solvent-Induced Controllable Synthesis, Single-Crystal to Single-Crystal Transformation and Encapsulation of Alq3 for Modulated Luminescence in (4,8)-Connected Metal–Organic Frameworks","type":"article-journal","volume":"51"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>34,35</sup>","plainTextFormattedCitation":"34,35","previouslyFormattedCitation":"<sup>34,35</sup>"},"properties":{"noteIndex":0},"schema":""}34,35 and the tuning of the flexible responses in rigid linker MOFs.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/C5DT03504J","ISBN":"1477-9234 (Electronic) 1477-9226 (Linking)","ISSN":"1477-9226","PMID":"26876816","abstract":"A new approach for the fine tuning of flexibility in MOFs, involving functionalization of the secondary building unit, is presented.","author":[{"dropping-particle":"","family":"Bon","given":"Volodymyr","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kavoosi","given":"Negar","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Senkovska","given":"Irena","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Müller","given":"Philipp","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schaber","given":"Jana","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wallacher","given":"Dirk","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"T?bbens","given":"Daniel M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mueller","given":"Uwe","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaskel","given":"Stefan","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Dalton Transactions","id":"ITEM-1","issue":"10","issued":{"date-parts":[["2016"]]},"page":"4407-4415","publisher":"Royal Society of Chemistry","title":"Tuning the flexibility in MOFs by SBU functionalization","type":"article-journal","volume":"45"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>36</sup>","plainTextFormattedCitation":"36","previouslyFormattedCitation":"<sup>36</sup>"},"properties":{"noteIndex":0},"schema":""}36 Work by Manos et al.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ic3005085","ISSN":"0020-1669","abstract":"Single-crystal-to-single-crystal (SCSC) transformations represent some of the most fascinating phenomena in chemistry. They are not only intriguing from a basic science point of view but also provide a means to modify or tune the properties of the materials via the postsynthetic introduction of suitable guest molecules or organic functional groups into their structures. Here, we describe UCY-2, a new flexible Nd3+ metal-organic framework (MOF), which exhibits a unique capability to undergo a plethora of SCSC transformations with some of them being very uncommon. These structural alterations involve the replacement of coordinating solvent molecules of UCY-2 by terminally ligating solvents and organic ligands with multiple functional groups including -OH, -SH, -NH-, and -NH2 or their combinations, chelating ligands, anions, and two different organic compounds. The SCSC coordinating solvent exchange is thus demonstrated as a powerful method for the functionalization of MOFs.","author":[{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kyprianidou","given":"Eleni J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Papaefstathiou","given":"Giannis S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-1","issue":"11","issued":{"date-parts":[["2012","6","4"]]},"page":"6308-6314","title":"Insertion of Functional Groups into a Nd 3+ Metal–Organic Framework via Single-Crystal-to-Single-Crystal Coordinating Solvent Exchange","type":"article-journal","volume":"51"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>24</sup>","plainTextFormattedCitation":"24","previouslyFormattedCitation":"<sup>24</sup>"},"properties":{"noteIndex":0},"schema":""}24 has also shown the potential for conducting SCCSE on a flexible material, reporting a wide range of solvent exchanges that result in changes to the overall framework size. Distinctively from these previous SCCSE examples, the mechanism for the exchange reaction at the SBU of ZnCSA appears to be directly linked to the flexibility of the CSA linker. These important structural rearrangements, occurring in the previously locked double linker connections, lead to a significantly different predicted response during guest removal.ExperimentalGeneralAll reagents were purchased from Sigma Aldrich Ltd and were used as received without further purification. SynthesisZnCSA.DMF. A 0.2M solution of Zn(NO3)2.6H2O in N,N’-Dimethylformamide (DMF) (375 μl, 0.075 mmols) was added to a (10 ml) screw capped Pyrex vial containing N-(4-Carboxyphenyl)succinamic acid (CSA-H2) (6.9 mg, 0.03 mmols). An additional volume of DMF (3.33 ml) was then added to the mixture and the vial was then capped and placed in an oven. The temperature inside the oven was raised slowly to 100 °C (1 oC/min) and the reaction was left to proceed for 24hrs before cooling back to room temperature at a ramp rate of 0.1 °C/min. The reaction afforded a white crystalline precipitate which was washed and stored in fresh DMF. ZnCSA.DMSO. Crystals of ZnCSA.DMF (20 mg) were transferred from DMF into a vial containing dimethylsulphoxide (DMSO) (1ml). The crystals were left to sit in the vial for 5 days, with the solvent being exchanged twice daily.Thermogravimetric analysisThermogravimetric analysis was carried out on a TA 500 instrument in a temperature range of 25 °C to 600 °C with a scan rate of 3 °C / min and air flow rate of 60 ml / min.X-ray CrystallographySingle crystal X-ray diffraction data were collected for ZnCSA.DMF using a Rigaku MicroMax-007 HF X-ray generator with a Mo-Kα rotating anode microfocus source and a Saturn 724+ CCD detector. Crystals were transferred from DMF into Fomblin oil and then mounted onto the diffractometer using a MiteGen MicroMount. The sample temperature was maintained at 100K using a 5ml/min N2 flow from an Oxford Cryosystems Cyrostream Plus device. Intensity data were indexed, integrated and corrected for absorption using CrysAlisPro software.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"author":[{"dropping-particle":"","family":"Crys Alis Pro","given":"","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Oxford Diffraction","id":"ITEM-1","issued":{"date-parts":[["0"]]},"title":"No Title","type":"article-journal"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>37</sup>","plainTextFormattedCitation":"37","previouslyFormattedCitation":"<sup>37</sup>"},"properties":{"noteIndex":0},"schema":""}37 Single Crystal X-ray diffraction data for ZnCSA.DMSO meanwhile were collected using synchrotron radiation at beamline I19, Diamond Light Source.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S0909049512008801","ISSN":"0909-0495","abstract":"The dedicated small-molecule single-crystal X-ray diffraction beamline (I19) at Diamond Light Source has been operational and supporting users for over three years. I19 is a high-flux tunable-wavelength beamline and its key details are described in this article. Much of the work performed on the beamline involves structure determination from small and weakly diffracting crystals. Other experiments that have been supported to date include structural studies at high pressure, studies of metastable species, variable-temperature crystallography, studies involving gas exchange in porous materials and structural characterizations that require analysis of the diffuse scattering between Bragg reflections. A range of sample environments to facilitate crystallographic studies under non-ambient conditions are available as well as a number of options for automation. An indication of the scope of the science carried out on the beamline is provided by the range of highlights selected for this paper.","author":[{"dropping-particle":"","family":"Nowell","given":"Harriott","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Barnett","given":"Sarah a.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Christensen","given":"Kirsten E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Teat","given":"Simon J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Allan","given":"David R.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Synchrotron Radiation","id":"ITEM-1","issue":"3","issued":{"date-parts":[["2012","5","1"]]},"page":"435-441","publisher":"International Union of Crystallography","title":"I19, the small-molecule single-crystal diffraction beamline at Diamond Light Source","type":"article-journal","volume":"19"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>38</sup>","plainTextFormattedCitation":"38","previouslyFormattedCitation":"<sup>38</sup>"},"properties":{"noteIndex":0},"schema":""}38 Intensity data, also collected at 100K, were indexed, integrated and corrected for absorption using Xia2.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S0021889809045701","ISBN":"0021-8898","ISSN":"0021-8898","abstract":"An expert system for macromolecular crystallography data reduction is presented, which builds on existing software to automate the complete data reduction process from images to merged structure factor amplitudes. This can automatically identify multi-wedge, multi-pass and multiwavelength data sets and includes explicit procedures to test for crystallographic special cases. With the push towards high-thoughput crystallography at synchrotron beamlines and automation of structure solution, the ability to reduce data with no user input fills an important gap in the pipeline.","author":[{"dropping-particle":"","family":"Winter","given":"G.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Applied Crystallography","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2010","2","1"]]},"page":"186-190","publisher":"International Union of Crystallography","title":"xia2 : an expert system for macromolecular crystallography data reduction","type":"article-journal","volume":"43"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1107/S0108767394005726","ISSN":"0108-7673","author":[{"dropping-particle":"","family":"Blessing","given":"R. H.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Acta Crystallographica Section A Foundations of Crystallography","id":"ITEM-2","issue":"1","issued":{"date-parts":[["1995","1","1"]]},"page":"33-38","title":"An empirical correction for absorption anisotropy","type":"article-journal","volume":"51"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>39,40</sup>","plainTextFormattedCitation":"39,40","previouslyFormattedCitation":"<sup>39,40</sup>"},"properties":{"noteIndex":0},"schema":""}39,40 All structures were initially solved by direct methods in the conventional setting of the unit cell.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S2053273314026370","ISSN":"2053-2733","abstract":"The new computer program SHELXT employs a novel dual-space algorithm to solve the phase problem for single-crystal reflection data expanded to the space group P 1. Missing data are taken into account and the resolution extended if necessary. All space groups in the specified Laue group are tested to find which are consistent with the P 1 phases. After applying the resulting origin shifts and space-group symmetry, the solutions are subject to further dual-space recycling followed by a peak search and summation of the electron density around each peak. Elements are assigned to give the best fit to the integrated peak densities and if necessary additional elements are considered. An isotropic refinement is followed for non-centrosymmetric space groups by the calculation of a Flack parameter and, if appropriate, inversion of the structure. The structure is assembled to maximize its connectivity and centred optimally in the unit cell. SHELXT has already solved many thousand structures with a high success rate, and is optimized for multiprocessor computers. It is, however, unsuitable for severely disordered and twinned structures because it is based on the assumption that the structure consists of atoms.","author":[{"dropping-particle":"","family":"Sheldrick","given":"George M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Acta Crystallographica Section A Foundations and Advances","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2015","1","1"]]},"page":"3-8","publisher":"International Union of Crystallography","title":"SHELXT – Integrated space-group and crystal-structure determination","type":"article-journal","volume":"71"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>41</sup>","plainTextFormattedCitation":"41","previouslyFormattedCitation":"<sup>41</sup>"},"properties":{"noteIndex":0},"schema":""}41 The atomic positions were analysed and compared to the other structures collected for the material. Lattice transformations were then applied using the program WinGX to ensure the structures were consistently described throughout the manuscript.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S0021889812029111","ISSN":"0021-8898","abstract":"The WinGX suite provides a complete set of programs for the treatment of small-molecule single-crystal diffraction data, from data reduction and processing, structure solution, model refinement and visualization, and metric analysis of molecular geometry and crystal packing, to final report preparation in the form of a CIF. It includes several well known pieces of software and provides a repository for programs when the original authors no longer wish to, or are unable to, maintain them. It also provides menu items to execute external software, such as the SIR and SHELX suites of programs. The program ORTEP for Windows provides a graphical user interface (GUI) for the classic ORTEP program, which is the original software for the illustration of anisotropic displacement ellipsoids. The GUI code provides input capabilities for a wide variety of file formats, and extra functionality such as geometry calculations and ray-traced outputs. The programs WinGX and ORTEP for Windows have been distributed over the internet for about 15 years, and this article describes some of the more modern features of the programs.","author":[{"dropping-particle":"","family":"Farrugia","given":"Louis J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Applied Crystallography","id":"ITEM-1","issue":"4","issued":{"date-parts":[["2012","7","14"]]},"page":"849-854","publisher":"International Union of Crystallography","title":"WinGX and ORTEP for Windows : an update","type":"article-journal","volume":"45"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>42</sup>","plainTextFormattedCitation":"42","previouslyFormattedCitation":"<sup>42</sup>"},"properties":{"noteIndex":0},"schema":""}42 Details can be found in Supporting Information S2. The crystal structures in the new unit cell were refined by full-matrix least squares using SHELXL accessed via the program Olex2.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S2053229614024218","ISBN":"2053-2296 (Electronic)","ISSN":"2053-2296","PMID":"25567568","abstract":"The improvements in the crystal structure refinement program SHELXL have been closely coupled with the development and increasing importance of the CIF (Crystallographic Information Framework) format for validating and archiving crystal structures. An important simplification is that now only one file in CIF format (for convenience, referred to simply as `a CIF') containing embedded reflection data and SHELXL instructions is needed for a complete structure archive; the program SHREDCIF can be used to extract the .hkl and .ins files required for further refinement with SHELXL . Recent developments in SHELXL facilitate refinement against neutron diffraction data, the treatment of H atoms, the determination of absolute structure, the input of partial structure factors and the refinement of twinned and disordered structures. SHELXL is available free to academics for the Windows, Linux and Mac OS X operating systems, and is particularly suitable for multiple-core processors.","author":[{"dropping-particle":"","family":"Sheldrick","given":"George M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Acta Crystallographica Section C Structural Chemistry","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2015","1","1"]]},"page":"3-8","publisher":"International Union of Crystallography","title":"Crystal structure refinement with SHELXL","type":"article-journal","volume":"71"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1107/S0021889808042726","ISSN":"0021-8898","author":[{"dropping-particle":"V.","family":"Dolomanov","given":"Oleg","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bourhis","given":"Luc J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Gildea","given":"Richard J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Howard","given":"Judith a. K.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Puschmann","given":"Horst","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Applied Crystallography","id":"ITEM-2","issue":"2","issued":{"date-parts":[["2009","1","24"]]},"page":"339-341","publisher":"International Union of Crystallography","title":"OLEX2 : a complete structure solution, refinement and analysis program","type":"article-journal","volume":"42"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>43,44</sup>","plainTextFormattedCitation":"43,44","previouslyFormattedCitation":"<sup>43,44</sup>"},"properties":{"noteIndex":0},"schema":""}43,44 Due to porous and flexible nature of the material, high resolution data could not always be obtained. Non-hydrogen atoms were therefore only refined anisotropically if there was sufficient data to parameter ratio available. Hydrogen atoms were placed in idealized positions and refined using a riding model with isotropic thermal parameters dependant on the connected atom. In the crystal structures, one of the organic linkers is situated on an inversion centre and is therefore highly disordered. Restrained isotropic thermal parameters combined with idealised distance restraints were used to ensure a chemically sensible model for this linker. The routine SQUEEZE from the program PLATON was used to account for the scattering contribution of disordered solvent molecules contained within the large accessible void space of the framework.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S2053229614024929","ISSN":"2053-2296","abstract":"The completion of a crystal structure determination is often hampered by the presence of embedded solvent molecules or ions that are seriously disordered. Their contribution to the calculated structure factors in the least-squares refinement of a crystal structure has to be included in some way. Traditionally, an atomistic solvent disorder model is attempted. Such an approach is generally to be preferred, but it does not always lead to a satisfactory result and may even be impossible in cases where channels in the structure are filled with continuous electron density. This paper documents the SQUEEZE method as an alternative means of addressing the solvent disorder issue. It conveniently interfaces with the 2014 version of the least-squares refinement program SHELXL [Sheldrick (2015). Acta Cryst. C 71 . In the press] and other refinement programs that accept externally provided fixed contributions to the calculated structure factors. The PLATON SQUEEZE tool calculates the solvent contribution to the structure factors by back-Fourier transformation of the electron density found in the solvent-accessible region of a phase-optimized difference electron-density map. The actual least-squares structure refinement is delegated to, for example, SHELXL . The current versions of PLATON SQUEEZE and SHELXL now address several of the unnecessary complications with the earlier implementation of the SQUEEZE procedure that were a necessity because least-squares refinement with the now superseded SHELXL97 program did not allow for the input of fixed externally provided contributions to the structure-factor calculation. It is no longer necessary to subtract the solvent contribution temporarily from the observed intensities to be able to use SHELXL for the least-squares refinement, since that program now accepts the solvent contribution from an external file (.fab file) if the ABIN instruction is used. In addition, many twinned structures containing disordered solvents are now also treatable by SQUEEZE. The details of a SQUEEZE calculation are now automatically included in the CIF archive file, along with the unmerged reflection data. The current implementation of the SQUEEZE procedure is described, and discussed and illustrated with three examples. Two of them are based on the reflection data of published structures and one on synthetic reflection data generated for a published structure.","author":[{"dropping-particle":"","family":"Spek","given":"Anthony L.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Acta Crystallographica Section C Structural Chemistry","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2015","1","1"]]},"page":"9-18","publisher":"International Union of Crystallography","title":"PLATON SQUEEZE: a tool for the calculation of the disordered solvent contribution to the calculated structure factors","type":"article-journal","volume":"71"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>45</sup>","plainTextFormattedCitation":"45","previouslyFormattedCitation":"<sup>45</sup>"},"properties":{"noteIndex":0},"schema":""}45 Full details of the crystals, data collections and refinement parameters are given in Supporting Information S1. Powder X-ray diffraction data were collected on a Bruker D8 advance diffractometer in transmission geometry using monochromated Cu-Kα radiation and 0.7 mm borosilicate capillary tubes. Due to the sensitivity of the material to loss of solvent, data for ZnCSA.DMF was initially collected while the sample was immersed in DMF within a sealed 0.7 mm borosilicate capillary tube. Crystal structures were visualised and images produced using a combination of Mercury 4.0 (CCDC)ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S002188980600731X","ISSN":"0021-8898","author":[{"dropping-particle":"","family":"Macrae","given":"Clare F.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Edgington","given":"Paul R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"McCabe","given":"Patrick","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Pidcock","given":"Elna","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Shields","given":"Greg P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Taylor","given":"Robin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Towler","given":"Matthew","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Streek","given":"Jacco","non-dropping-particle":"van de","parse-names":false,"suffix":""}],"container-title":"Journal of Applied Crystallography","id":"ITEM-1","issue":"3","issued":{"date-parts":[["2006","5","10"]]},"page":"453-457","title":"Mercury : visualization and analysis of crystal structures","type":"article-journal","volume":"39"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>46</sup>","plainTextFormattedCitation":"46","previouslyFormattedCitation":"<sup>46</sup>"},"properties":{"noteIndex":0},"schema":""}46 and VESTA 3ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1107/S0021889811038970","ISSN":"0021-8898","abstract":"VESTA is a three-dimensional visualization system for crystallographic studies and electronic state calculations. It has been upgraded to the latest version, VESTA 3 , implementing new features including drawing the external morphology of crystals; superimposing multiple structural models, volumetric data and crystal faces; calculation of electron and nuclear densities from structure parameters; calculation of Patterson functions from structure parameters or volumetric data; integration of electron and nuclear densities by Voronoi tessellation; visualization of isosurfaces with multiple levels; determination of the best plane for selected atoms; an extended bond-search algorithm to enable more sophisticated searches in complex molecules and cage-like structures; undo and redo in graphical user interface operations; and significant performance improvements in rendering isosurfaces and calculating slices.","author":[{"dropping-particle":"","family":"Momma","given":"Koichi","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Izumi","given":"Fujio","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Applied Crystallography","id":"ITEM-1","issue":"6","issued":{"date-parts":[["2011","12","1"]]},"page":"1272-1276","publisher":"International Union of Crystallography","title":"VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data","type":"article-journal","volume":"44"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>47</sup>","plainTextFormattedCitation":"47","previouslyFormattedCitation":"<sup>47</sup>"},"properties":{"noteIndex":0},"schema":""}putational MethodsDFT calculations were performed using the VASPADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"author":[{"dropping-particle":"","family":"Kresse","given":"G","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hafner","given":"J","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Phys. Rev. B","id":"ITEM-1","issued":{"date-parts":[["1993"]]},"page":"558","title":"No Title","type":"article-journal","volume":"47"},"uris":["",""]}],"mendeley":{"formattedCitation":"<sup>48</sup>","plainTextFormattedCitation":"48","previouslyFormattedCitation":"<sup>48</sup>"},"properties":{"noteIndex":0},"schema":""}48 code. Input geometries were generated from experimental structures in which all atomic positions except for the disordered linker along the c axis had been determined. To account for the experimental random coordination of the c axis linker 2×2×1 periodic supercells were built alternating the binding through the rigid and flexible ends of the linkers in the c axis direction. Each supercell contained 20 CSA linkers, 28 zinc cations, 8 cation-bridging oxygens and 8 DMF or DMSO molecules coordinated to the SBU, a total of 652 or 628 ions for ZnCSA.DMF and ZnCSA.DMSO respectively. The experimental structures (1, 2, 3) were initially optimized with the unit cell parameters fixed (ISIF = 2) at the measured values, yielding structures C1opt , C2opt and C3opt in the respective cases of DMF, partially removed DMF, and DMSO. The flexible behaviour of the desolvated materials was then computationally addressed via geometry optimizations by allowing both ion positions and unit cell parameters (volume and shape, ISIF = 3) to relax, thus producing the structures C1relax and C3relax. All calculations were conducted with their coordinated guest molecules, but not any guest molecules contained within the pore due to computational cost. Allowing both the unit cell parameters and ion positions to relax therefore probes the structural effect of removal of the solvent from the pores of the material. The ion-electron interaction was described with the Projector Augmented-Wave (PAW) method.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"author":[{"dropping-particle":"","family":"Bl?chl","given":"P E","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Physical Review B","id":"ITEM-1","issued":{"date-parts":[["1994"]]},"page":"17953","title":"Projector augmented-wave method","type":"article-journal","volume":"50"},"uris":["",""]}],"mendeley":{"formattedCitation":"<sup>49</sup>","plainTextFormattedCitation":"49","previouslyFormattedCitation":"<sup>49</sup>"},"properties":{"noteIndex":0},"schema":""}49 To account for van der Waals dispersion forces, the non-local correlation functional method was usedADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1103/PhysRevLett.92.246401","ISBN":"0031-9007 (Print)\\n0031-9007 (Linking)","ISSN":"00319007","PMID":"15245113","abstract":"To understand sparse systems, we must account for both strong local atom bonds and weak nonlocal van der Waals forces between atoms separated by empty space. A fully nonlocal functional form [Phys. Rev. B 62, 6997 (2000)]] of density-functional theory (DFT) is applied here to the layered systems graphite, boron nitride, and molybdenum sulfide to compute bond lengths, binding energies, and compressibilities. These key examples show that the DFT with the generalized-gradient approximation does not apply for calculating properties of sparse matter, while use of the fully nonlocal version appears to be one way to proceed.","author":[{"dropping-particle":"","family":"Dion","given":"M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rydberg","given":"H.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schr?der","given":"E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Langreth","given":"D. C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lundqvist","given":"B. I.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Physical Review Letters","id":"ITEM-1","issue":"24","issued":{"date-parts":[["2004"]]},"page":"246401","title":"Van der Waals Density Functional for General Geometries","type":"article-journal","volume":"92"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>50</sup>","plainTextFormattedCitation":"50","previouslyFormattedCitation":"<sup>50</sup>"},"properties":{"noteIndex":0},"schema":""}50 with optB86b-vdW exchange functional.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1088/0953-8984/22/2/022201","ISBN":"0953-8984","ISSN":"09538984","PMID":"21386245","abstract":"The non-local van der Waals density functional (vdW-DF) of Dion et al(2004 Phys. Rev. Lett. 92 246401) is a very promising scheme for the efficient treatment of dispersion bonded systems. We show here that the accuracy of vdW-DF can be dramatically improved both for dispersion and hydrogen bonded complexes through the judicious selection of its underlying exchange functional. New and published exchange functionals are identified that deliver much better than chemical accuracy from vdW-DF for the S22 benchmark set of weakly interacting dimers and for water clusters. Improved performance for the adsorption of water on salt is also obtained.","author":[{"dropping-particle":"","family":"Klime?","given":"Jii?í","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bowler","given":"David R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Michaelides","given":"Angelos","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Physics Condensed Matter","id":"ITEM-1","issue":"2","issued":{"date-parts":[["2010"]]},"page":"022201","title":"Chemical accuracy for the van der Waals density functional","type":"article-journal","volume":"22"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>51</sup>","plainTextFormattedCitation":"51","previouslyFormattedCitation":"<sup>51</sup>"},"properties":{"noteIndex":0},"schema":""}51 An energy cut-off of 520 eV was employed for the plane-wave expansion, as well as a k-point mesh of 1×1×2 to sample the Brillouin zone in the reciprocal space. For all calculations, the “Normal” precision setting with convergence criteria of 1×10-6 eV for the electronic energy convergence and 1×10-5 eV for the ionic convergence were used.ResultsThe MOF ZnCSA.DMF was synthesised using solvothermal methods from Zn(NO3)2.6H2O, CSA-H2 (Figure 1a) and DMF. This afforded plate-shaped single crystals of formula [Zn7O2(CSA)5DMF2].14DMF which were analysed by X-ray diffraction at 100K. The crystal structure of ZnCSA.DMF (1) revealed that the material crystallises in the triclinic space group P1 (No. 2) with unit cell dimensions a = 15.7350(7) ?, b = 17.4602(8) ?, c = 19.3022(9) ?, α = 95.955(4)°, β = 109.926(4)°, γ = 105.846(4)°, V = 4681.2(4) ?3. The framework is constructed from a heptanuclear secondary building unit of formula [Zn7O2(Carboxylate)10DMF2]. The cluster contains seven zinc atoms bridged by 10 bidentate carboxylate groups (8 μ2 connections & 2 μ3 connections), two μ4-O atoms and two terminal DMF ligands (see Figure 1b). The carboxylate groups belong to 10 different CSA linkers, bound through a mixture of rigid and flexible ends, which connect to 6 other clusters to give a non-interpenetrated fqr net.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/c0ce00636j","ISSN":"1466-8033","author":[{"dropping-particle":"V","family":"Alexandrov","given":"E","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Blatov","given":"V A","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"V","family":"Kochetkov","given":"A","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Proserpio","given":"D M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"CrystEngComm","id":"ITEM-1","issue":"12","issued":{"date-parts":[["2011"]]},"page":"3947","title":"Underlying nets in three-periodic coordination polymers: topology, taxonomy and prediction from a computer-aided analysis of the Cambridge Structural Database","type":"article-journal","volume":"13"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/cr200205j","ISSN":"1520-6890","PMID":"21916513","author":[{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical reviews","id":"ITEM-2","issue":"2","issued":{"date-parts":[["2012","2","8"]]},"page":"675-702","title":"Deconstructing the crystal structures of metal-organic frameworks and related materials into their underlying nets.","type":"article-journal","volume":"112"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>52,53</sup>","plainTextFormattedCitation":"52,53","previouslyFormattedCitation":"<sup>52,53</sup>"},"properties":{"noteIndex":0},"schema":""}52,53 The metal clusters are located at the vertices of the crystallographic unit cell with the linker connections running along the crystallographic axes, four corresponding to the crystallographic a axis, four to the b axis and two to the c axis. Within the cluster there are one octahedral, two square pyramidal and four tetrahedral Zn environments. The octahedral Zn is located at the centre of the cluster and is apically coordinated to the two μ4-O2- anions and equatorially coordinated to four different carboxylate groups. This Zn is then connected through the two μ4-O2- anions to two equivalent trinuclear units, each centred around one of the O2- anions and consisting of one square pyramidal Zn (with the coordinated DMF ligand) and two tetrahedral Zn atoms. The overall cluster could also be considered as two symmetry related Zn4O tetrahedra which share the central octahedral Zn as a common vertex. This metal cluster unit has been previously reported in molecular complexes (although often with different terminal ligands),ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1007/s00706-012-0824-3","ISBN":"0070601208","ISSN":"0026-9247","abstract":"The heptanuclear complex [(Zn 7 O 2 (OAc) 10 (dma) 2 ] (dma = N,N-dimethylacetamide) was prepared sonochemically in 60 % yield by the reaction of zinc acetate dihydrate with N,N-dimethylacetamide in an appropriate molar ratio using dry methanol as a solvent, and characterised by m.p., IR, TG, FTIR, and single-crystal X-ray analysis. The X-ray crystallography revealed that the complex belongs to a monoclinic crystal system with space group C2/c with cell dimensions a = 18.5951(6), b = 10.6718(4), and c = 24.0520(8) ?, β = 102.948(2). Highly conformal thin films of zincite (zinc oxide) and manganese-doped zincite were fabricated on a gold interdigitated ceramic substrate from the heptanuclear complex, and a mixture of manganese acetate dihydrate and the complex in an appropriate molar ratio using the ultrasonic aerosol-assisted chemical vapour deposition technique. The size, shape, surface morphology, microstructure, chemical composition, and crystallinity of the resulting films were analysed by SEM/EDX and powder X-ray diffraction. Investigation of sensing properties of the deposited films reveals that both zincite and manganese-doped zincite thin films are sensitive to alcohol vapours at a concentration of 100 ppm. ? 2012 Springer-Verlag.","author":[{"dropping-particle":"","family":"Hussain","given":"Muzammil","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tahir","given":"Muhammad Nawaz","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mansoor","given":"Muhammad Adil","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Arifin","given":"Zainudin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mazhar","given":"Muhammad","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Monatshefte für Chemie - Chemical Monthly","id":"ITEM-1","issue":"3","issued":{"date-parts":[["2013","3","22"]]},"page":"285-294","title":"Heptanuclear zinc cluster for growth of zincite and manganese-doped zincite thin films for sensor applications","type":"article-journal","volume":"144"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/ic0107983","ISSN":"0020-1669","abstract":"Zinc complexes derived from benzoic acids containing electron-withdrawing substituents have been synthesized from Zn(II)(bis-trimethylsilyl amide)(2) and the corresponding carboxylic acid (2,6-X(2)C(6)H(3)COOH, where X = F, Cl, or OMe) in THF and structurally characterized via X-ray crystallography. The 2,6-difluorobenzoate complex crystallizes from THF or CH(3)CN as a seven membered zinc aggregate, where the metal atoms are interconnected by a combination of 10 mu-benzoates and mu(4)-oxo ligands, that is, [(2,6-difluorobenzoate)(10)O(2)Zn(7)](solvent)(2), solvent = THF (1) and CH(3)CN (1a). On the other hand, the 2,6-dichlorobenzoate zinc derivative crystallizes from THF as a dimer, [(2,6-dichlorobenzoate)(4)Zn(2)](THF)(3) (2), where the two zinc centers are bridged by three benzoate ligand. One of the zinc centers possesses a tetrahedral ligand environment where the fourth ligand is a unidentate benzoate, and the other zinc center has an octahedral arrangement of ligands which is accomplished by the additional binding of three THF molecules. Upon dissolution of complex 1 or 2 in the strongly binding pyridine solvent, disruption of these zinc carboxylates occurs with concomitant formation of mononuclear zinc bis-benzoates with three pyridine ligands in the metal coordination sphere. Complexes 1 and 2 were found to be effective catalysts for the copolymerization of cyclohexene oxide and carbon dioxide to afford polycarbonates devoid of polyether linkages, that is, completely alternating copolymers. Although these catalysts or catalyst precursors in the presence of CO(2)/propylene oxide afforded mostly propylene carbonate, they did serve as efficient catalysts for the terpolymerization of carbon dioxide/cyclohexene oxide/propylene oxide. The reactivities of these zinc carboxylates were very similar to those previously reported analogous complexes which have not been structurally characterized. Hence, it is suggested here that all of these zinc carboxylates provide similar catalytic sites for CO(2)/epoxide coupling processes.","author":[{"dropping-particle":"","family":"Darensbourg","given":"Donald J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wildeson","given":"Jacob R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yarbrough","given":"Jason C.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-2","issue":"4","issued":{"date-parts":[["2002","2"]]},"page":"973-980","title":"Solid-State Structures of Zinc(II) Benzoate Complexes. Catalyst Precursors for the Coupling of Carbon Dioxide and Epoxides","type":"article-journal","volume":"41"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1039/a802844c","ISSN":"13597345","author":[{"dropping-particle":"","family":"Lalioti","given":"Nikolia","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Perlepes","given":"Spyros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Manessi-Zoupa","given":"Evy","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Raptopoulou","given":"Catherine P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Terzis","given":"Aris","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Aliev","given":"Abil E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Gerothanassis","given":"Ioannis P.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical Communications","id":"ITEM-3","issue":"15","issued":{"date-parts":[["1998"]]},"page":"1513-1514","title":"Rare M7O2 double tetrahedral core in molecular species: preparation, structure and properties of [Zn7O2(O2CMe)10(1-Meim)2] (1-Meim = 1-methylimidazole)","type":"article-journal","volume":"2"},"uris":[""]},{"id":"ITEM-4","itemData":{"DOI":"10.1134/S0036023610010092","ISSN":"0036-0236","author":[{"dropping-particle":"V.","family":"Anan’ev","given":"I.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"V.","family":"Perova","given":"E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Nefedov","given":"S. E.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Russian Journal of Inorganic Chemistry","id":"ITEM-4","issue":"1","issued":{"date-parts":[["2010","1","21"]]},"page":"40-52","title":"Pyrazolate-bridged binuclear zinc complexes Zn2(μ-dmpz)2(Hdmpz)2(OOCR)2 (R = Me, Ph; Hdmpz = 3,5-dimethylpyrazole)","type":"article-journal","volume":"55"},"uris":[""]},{"id":"ITEM-5","itemData":{"DOI":"10.1039/B414983C","ISBN":"9789501265675","ISSN":"1477-9226","author":[{"dropping-particle":"","family":"Waheed","given":"Abdul","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jones","given":"Richard A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"McCarty","given":"Jeffrey","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yang","given":"Xiaoping","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Dalton Trans.","id":"ITEM-5","issue":"22","issued":{"date-parts":[["2004"]]},"page":"3840-3841","title":"Synthesis and structure of Zn 7 (? 4 -O) 2 (OAc) 10 (Pz) 2 (OAc = acetate; Pz = pyrazine)","type":"article-journal","volume":"7"},"uris":[""]},{"id":"ITEM-6","itemData":{"DOI":"10.1007/s11243-017-0127-y","ISSN":"0340-4285","abstract":"? 2017, Springer International Publishing Switzerland. Reaction of [Zn 4 (? 4 -O)(O 2 CCH 3 ) 6 ] with one equivalent of 4-tertiary-butylpyridine leads to a nearly quantitative formation of the heptanuclear cluster [Zn 7 O 2 (O 2 CCH 3 ) 10 ( t bupy) 2 ] (1). Here, we present the crystal structure of 1 and first results of the examination of its catalytic activity in transesterification reactions and polymerization of lactide compared to the activity of [Zn 4 (? 4 -O)(O 2 CCH 3 ) 6 ] in the presence of 4-tertiary-butylpyridine.","author":[{"dropping-particle":"","family":"Dittrich","given":"Dennis","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tewes","given":"Heiko","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"W?lper","given":"Christoph","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bl?ser","given":"Dieter","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schulz","given":"Stephan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Roll","given":"Joachim","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Transition Metal Chemistry","id":"ITEM-6","issue":"3","issued":{"date-parts":[["2017","4","16"]]},"page":"237-241","title":"Preparation, catalytical activity and crystal structure of a heptanuclear zinc acetate cluster","type":"article-journal","volume":"42"},"uris":[""]},{"id":"ITEM-7","itemData":{"DOI":"10.1039/DT9790000028","ISSN":"0300-9246","author":[{"dropping-particle":"","family":"Attanasio","given":"Donato","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dessy","given":"Giulia","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fares","given":"Vincenzo","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"J. Chem. Soc., Dalton Trans.","id":"ITEM-7","issue":"1","issued":{"date-parts":[["1979"]]},"page":"28-32","title":"Crystal and molecular structure of deca-?-acetato-dioxobis(pyridine)-heptazinc( <scp>II</scp> ) and the electron paramagnetic resonance spectrum of its copper-doped crystals","type":"article-journal","volume":"84"},"uris":[""]},{"id":"ITEM-8","itemData":{"DOI":"10.1016/j.poly.2012.06.011","ISSN":"02775387","author":[{"dropping-particle":"","family":"Reger","given":"Daniel L.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Debreczeni","given":"Agota","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Pascui","given":"Andrea E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Smith","given":"Mark D.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Polyhedron","id":"ITEM-8","issued":{"date-parts":[["2013","3"]]},"page":"1317-1322","title":"Heptanuclear zinc carboxylate complex: New supramolecular building unit and unique supramolecular architecture","type":"article-journal","volume":"52"},"uris":[""]},{"id":"ITEM-9","itemData":{"DOI":"10.1016/j.inoche.2007.04.007","ISSN":"13877003","author":[{"dropping-particle":"","family":"Feazell","given":"Rodney P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Carson","given":"Cody E","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Klausmeyer","given":"Kevin K.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry Communications","id":"ITEM-9","issue":"8","issued":{"date-parts":[["2007","8"]]},"page":"873-875","title":"Synthesis of a funtionalized monomer of the rare double tetrahedron Zn cluster: [Zn7O2(O2C2H3)10(3-{CH2OH}C5H4N)2]","type":"article-journal","volume":"10"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>54–62</sup>","plainTextFormattedCitation":"54–62","previouslyFormattedCitation":"<sup>54–62</sup>"},"properties":{"noteIndex":0},"schema":""}54–62 one-dimensional coordination polymers,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1016/j.poly.2013.07.002","ISSN":"02775387","author":[{"dropping-particle":"","family":"Constable","given":"Edwin C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Housecroft","given":"Catherine E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Sch?nle","given":"Jonas","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vujovic","given":"Srboljub","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zampese","given":"Jennifer A.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Polyhedron","id":"ITEM-1","issued":{"date-parts":[["2013","10"]]},"page":"260-267","publisher":"Elsevier Ltd","title":"Coordination polymers with 4′-(4-(anthracen-9-yl)phenyl)- and 4′-(4-(naphthalen-1-yl)phenyl)-4,2′:6′,4″-terpyridines: Mono-, di- and heptazinc(II) nodes","type":"article-journal","volume":"62"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1016/j.poly.2006.01.031","ISSN":"02775387","author":[{"dropping-particle":"","family":"Suen","given":"Maw-Cherng","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chan","given":"Zhi-Kai","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chen","given":"Jhy-Der","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Ju-Chun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hung","given":"Chen-Hsiung","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Polyhedron","id":"ITEM-2","issue":"11","issued":{"date-parts":[["2006","7"]]},"page":"2325-2332","title":"Syntheses and structures of three new coordination polymers generated from the flexible 1,3-bis(4-pyridyl)propane ligand and zinc salts","type":"article-journal","volume":"25"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1016/j.poly.2006.01.015","ISSN":"02775387","abstract":"The long and rigid spacers 1,4-bis(3-pyridyl)-2,3-diaza-1,3-butadiene (3pdb) and 1,4-bis(4-pyridyl)-2,3-diaza-1,3-butadiene (4pdb) exhibit a self-assembly reaction with zinc acetate to produce the new coordination polymers [Zn7(μ4-O)2(OAc)10(3pdb)] n (1) and [Zn7(μ4-O)2(OAc)10(4pdb)] n (2). The X-ray crystallography results show one-dimensional polymeric chains with the cluster moiety {Zn7(μ4-O)2(OAc)10} as a node. The formation of the heptanuclear nodes is discussed on the basis of the key importance of the length of the spacer. ? 2006 Elsevier Ltd. All rights reserved.","author":[{"dropping-particle":"","family":"Granifo","given":"Juan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Garland","given":"María T.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Baggio","given":"Ricardo","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Polyhedron","id":"ITEM-3","issue":"11","issued":{"date-parts":[["2006","7"]]},"page":"2277-2283","title":"The effect in the assembly and node nuclearity of the long and rigid character of bis-pyridyl exo-bidentate spacers when react with zinc acetate: Crystal structures of the high nuclearity coordination polymers [Zn7(μ4-O)2(OAc)10(3pdb)]n (3pdb=1,4-bis(3-py","type":"article-journal","volume":"25"},"uris":[""]},{"id":"ITEM-4","itemData":{"DOI":"10.1016/S0020-1693(02)01488-3","ISBN":"6567791691","ISSN":"00201693","abstract":"Two one-dimensional coordination polymeric chains namely, [Zn7(μ4-O)2(CH3CO 2)10(bpe)] (1) and ([Zn7(μ4-O)2(CH3CO 2)10(dpds)] (2) (bpe=1,2-bis(4-pyridyl)ethane and dpds=4,4′-dipyridyldisulfide) have been obtained by self assembly method from Zn(O2CCH3)2 and spacer ligands dpe and dpds in the ratio 7:1, and characterized by X-ray crystallography. Compound 1 is a quasi-linear one-dimensional coordination polymer whereas 2 has the zigzag polymeric chain structure. The conformations of the backbone of these flexible angular ligands appear to influence the way in which the heptanuclear metal aggregates are bonded in the solid state. In 1, the py-C-C-py fragment has an anti conformation with the torsion angle, 180°. The S-C-C-S torsion angle for dpds in 2 is 61.8(5)° with the interplanar angle between the pyridine rings of 89.4(4)°. ? 2002 Elsevier Science B.V. All rights reserved.","author":[{"dropping-particle":"","family":"Ng","given":"Meng Tack","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Deivaraj","given":"Theivanayagam C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"J. Vittal","given":"Jagadese","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganica Chimica Acta","id":"ITEM-4","issued":{"date-parts":[["2003","5"]]},"page":"173-178","title":"Self assembly of heptanuclear zinc(II) clusters linked by angular spacer ligands","type":"article-journal","volume":"348"},"uris":[""]},{"id":"ITEM-5","itemData":{"DOI":"10.1016/j.poly.2012.06.011","ISSN":"02775387","author":[{"dropping-particle":"","family":"Reger","given":"Daniel L.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Debreczeni","given":"Agota","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Pascui","given":"Andrea E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Smith","given":"Mark D.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Polyhedron","id":"ITEM-5","issued":{"date-parts":[["2013","3"]]},"page":"1317-1322","title":"Heptanuclear zinc carboxylate complex: New supramolecular building unit and unique supramolecular architecture","type":"article-journal","volume":"52"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>61,63–66</sup>","plainTextFormattedCitation":"61,63–66","previouslyFormattedCitation":"<sup>61,63–66</sup>"},"properties":{"noteIndex":0},"schema":""}61,63–66 and three-dimensional metal-organic frameworks.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/chem.200500963","ISBN":"0947-6539","ISSN":"09476539","abstract":"A novel, three-dimensional, noninterpenetrating microporous metal-organic framework (MOF), [Zn7O2(pda)5(H2O)2]5 DMF4 EtOH 6 H2O (1) (H2PDA=p-phenylenediacrylic acid, DMF=N,N-dimethylformamide, EtOH=ethanol), was synthesized by constructing heptanuclear zinc carboxylate secondary building units (SBUs) and by using rigid and linear aromatic carboxylate ligands, PDA. The X-ray crystallographic data reveals that the seven zinc centers of 1 are held together with ten carboxylate groups of the PDA ligands and four water molecules to form a heptametallic SBU, Zn7O4(CO2)10, with dimensions of 9.8 x 9.8 x 13.8 A3. Furthermore, the heptametallic SBUs are interconnected by PDA acting as linkers, thereby generating an extended network with a three-dimensional, noninterpenetrating, intersecting large-channel system with spacing of about 17.3 A. As a microporous framework, polymer 1 shows adsorption behavior that is favorable towards H2O and CH3OH, and substantial H2 uptake. In terms of the heptanuclear zinc carboxylate SBUs, polymer 1 exhibits interesting photoelectronic properties, which would facilitate the exploration of new types of semiconducting materials, especially among MOFs containing multinuclear metal carboxylate SBUs.","author":[{"dropping-particle":"","family":"Fang","given":"Qian Rong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zhu","given":"Guang Shan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xue","given":"Ming","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zhang","given":"Qing Lin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Sun","given":"Jin Yu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Guo","given":"Xiao Dan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Qiu","given":"Shi Lun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Shi Tao","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Ping","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"De Jun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wei","given":"Yen","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemistry - A European Journal","id":"ITEM-1","issue":"14","issued":{"date-parts":[["2006"]]},"page":"3754-3758","title":"Microporous metal-organic framework constructed from heptanuclear zinc carboxylate secondary building units","type":"article-journal","volume":"12"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/ic302127y","ISSN":"0020-1669","abstract":"A new metal-organic framework, CALF-22 comprising Zn7O2(COO)10 secondary building units and 2-nitro-1,4-benzenedicarboxylate, is reported. The porosity and gas adsorption of N2, H2, CO2, and CH4 are studied, and CALF-22 has a surface area in excess of 1000 m(2)/g. The stability of the larger zinc cluster and the effect of the nitro group on gas sorption are also studied.","author":[{"dropping-particle":"","family":"Iremonger","given":"Simon S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vaidhyanathan","given":"Ramanathan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mah","given":"Roger K.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Shimizu","given":"George K. H.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-2","issue":"8","issued":{"date-parts":[["2013","4","15"]]},"page":"4124-4126","title":"Zn 7 O 2 (RCOO) 10 Clusters and Nitro Aromatic Linkers in a Porous Metal–Organic Framework","type":"article-journal","volume":"52"},"uris":["","","","",""]},{"id":"ITEM-3","itemData":{"DOI":"10.1002/anie.201202925","ISSN":"14337851","author":[{"dropping-particle":"","family":"Choi","given":"Sang Beom","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Furukawa","given":"Hiroyasu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Nam","given":"Hye Jin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jung","given":"Duk-young","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jhon","given":"Young Ho","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Walton","given":"Allan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Book","given":"David","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kim","given":"Jaheon","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-3","issue":"35","issued":{"date-parts":[["2012","8","27"]]},"page":"8791-8795","title":"Reversible Interpenetration in a Metal-Organic Framework Triggered by Ligand Removal and Addition","type":"article-journal","volume":"51"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>21–23</sup>","plainTextFormattedCitation":"21–23","previouslyFormattedCitation":"<sup>21–23</sup>"},"properties":{"noteIndex":0},"schema":""}21–23 In ZnCSA, the carboxylates located around the centre of the cluster, coordinated to the central octahedral Zn, are all connections made through the flexible ends of the asymmetric CSA linker, and correspond to two a axis linkers and two b axis linkers. The carboxylate groups on outer parts of the cluster meanwhile are connections to the rigid ends of the remaining two a and two b axis linkers, displayed in green and purple in Figure 1b respectively, or connections to the two c axis linkers, displayed in orange. Figure 1. a) The CSA linker with distinct rigid and flexible ends. For the single c axis linker, these ends of the molecule are disordered such that it coordinates randomly via either end. Atoms labelled from A to F are involved in the definition of torsion angles φ (Ocarbox(A)-Ccarbox(B) -Csp3(C)-Csp3(D)), ψ (Ccarbox(B)-Csp3(C)-Csp3(D)-Camide(E)) and δ (Csp3(C)-Csp3(D)-Camide(E)-Namide(F)). Note that values of these torsions are tabulated in Supporting Information S7 for the structures discussed in the text. b) Secondary building unit of ZnCSA.DMF (1) showing the seven Zn atoms, ten carboxylate connections and two terminal DMF ligands. Carboxylate carbons coloured green are connections running along the a axis, coloured purple are connections running along the b axis, and coloured orange are connections running along the c axis. Striped coloured connections are connections through the flexible end of the linker, solid colours are connections through the rigid end, and half striped connections correspond to 50/50 distribution through the flexible and rigid ends of the linkers. The carboxylates showing mixed connections correspond to the single c axis linkers. Note that two b axis connections are hidden behind the cluster. c) Schematic showing the linker connections between the metal clusters situated at the corners of each crystallographic unit cell. Blue ellipses represent the metal clusters, double green lines represent CSA linker connections along the a axis, double purple lines represent CSA linker connections along the b axis and single orange lines represent CSA linker connections along the c axis. Disorder has also been removed from the c axis linker for clarity and coordinated DMF molecules are present on the complete clusters, although their orientations make them hard to distinguish from the Zn polyhedra. d) Overlay of the 3 different linker conformations present in 1, coloured based on the axis it defines. e) The two half-occupancy inversion symmetry-related orientations of the single c axis linker modelled crystallographically. Hydrogen atoms have been removed for clarity.The four a and four b axis CSA linkers can be divided into pairs, each corresponding to double linker connections between two different clusters. Each pair of asymmetric linkers is oriented anti-parallel to each other (i.e., the rigid ends of each linker pointing in opposite directions and crystallographically ordered) above and below the crystallographic axis, related by inversion symmetry. The c axis, meanwhile, only shows one linker connection on either side of the cluster, running directly along its axis. This is a key component of the structure and from here will be referred to as the single c axis linker. As with the a and b axes linkers, the single c axis linker binds through both its rigid and flexible ends, however, it is crystallographically disordered around the inversion centre due to a random distribution of the two possible orientations within the whole structure. The crystallographic model of these two orientations, which required a number of bond distance restraints, is shown in Figure 1e. The three linkers modelled in the asymmetric unit, two full occupancy corresponding to the a and b axes likers, and one half occupancy corresponding to disordered c axis linker, were all observed to adopt different conformations from each other based on rotations around the torsional angles of the flexible end of the CSA linker. The rigid ends of the three linkers, however, remained the same. This can be seen visually in Figure 1d, displaying an overlay of the three crystallographically modelled conformations. The torsional angles involved, displayed in Figure 1a, can be defined as φ (Ocarbox(A)-Ccarbox(B) –Csp3(C)-Csp3(D)), ψ (Ccarbox(B)-Csp3(C)-Csp3(D)-Camide€) and δ (Csp3(C)-Csp3(D)-Camide€-Namide(F)). In particular the a axis linkers are highly bent, with a ψ value of 60°, while the b and c axis linkers are almost planar with ψ values of -176° and 163° respectively. The values of these torsions in all structures discussed are tabulated in Supporting Information S7. Overall, the framework affords a large (2032 ?3) 3-dimensional pore structure with absolute window apertures of 12.7 × 12.6 ?, 12.3 × 9.0 ? and 4.5 × 4.3 ?. The pore limiting diameter, defined as the smallest opening along the pore, calculated with the Zeo++ software packageADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1016/j.micromeso.2011.08.020","ISSN":"13871811","author":[{"dropping-particle":"","family":"Willems","given":"Thomas F.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rycroft","given":"Chris H.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kazi","given":"Michaeel","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Meza","given":"Juan C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Haranczyk","given":"Maciej","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Microporous and Mesoporous Materials","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2011"]]},"page":"134-141","publisher":"Elsevier Inc.","title":"Algorithms and tools for high-throughput geometry-based analysis of crystalline porous materials","type":"article-journal","volume":"149"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>67</sup>","plainTextFormattedCitation":"67","previouslyFormattedCitation":"<sup>67</sup>"},"properties":{"noteIndex":0},"schema":""}67 is 9.7 ?. The solvent accessible void (measured with 1.2 ? probe radius) accounts for 43.4% of the unit cell volume. Based on thermogravimetric analysis, there are 14 guest DMF molecules contained within the pore, in addition to those coordinated to the cluster (see Supplementary Information S3), which corresponds well to the crystallographic data suggesting 13.6 guest DMF molecules. This value was obtained from a combination of the 1 guest site in the pore which could be resolved in the asymmetric unit, giving a refined formula of [Zn7O2(CSA)5DMF2].1.79(1) DMF, and the residual electron density calculated by the routine SQUEEZE. As a comparison, the absolute pore volume could accommodate 15.8 DMFs, assuming a packing density identical to liquid DMF. A schematic of ZnCSA, designed to emphasise the relationship of the structure to its crystallographic unit cell, is given in Figure 1c and coloured based on the connections along the different axes. More detailed views of the structure along with its Connolly surface can be found in Figure 2. The bulk phase purity of the material was confirmed by PXRD at room temperature measured while the material was immersed in DMF. The diffraction pattern was indexed to a unit cell of dimensions a = 15.877 (1) ?, b = 17.500(1) ?, c = 19.842(2) ? , α = 97.702(5)°, β = 103.279(4)°, γ = 104.588(4)°, V = 5083.8(6) ?3, which is a 9% anisotropic expansion compared to the single crystal data collected at 100K, with the majority difference being in the β angle.Figure 2. a) View of the channel running along the c axis in ZnCSA.DMF (left), the double linker connection running along the a axis (centre) and the Connolly surface (calculated using a 1.2 ? probe radius) (right). b) View of the channel running along the a axis (left), the double linker connection running along the b axis (centre) and the Connolly surface (calculated using a 1.2 ? probe radius) (right). c) View of the channel running along b axis (left), the single linker connection running along the c axis (centre) and the Connolly surface (calculated using a 1.2 ? probe radius) (right). Angles ω1 and ω2 are shown here, but are given in more detail in Figure 4b. Note that disorder has been removed from the single c axis linker for clarity.A PXRD pattern of ZnCSA.DMF was also collected once the material was filtered, dried and heated at 120 oC for 30 mins. TGA data (collected before and after heating) suggests that this temperature should be sufficient to remove the majority of the contained guests for the material (10 out of 14 guests) (see Figure 3a), and therefore probes the behaviour of the framework during removal of the DMF molecules in the pore but not from the SBU. The large difference observed in the two obtained PXRD patterns (see Figure 3) indicates clear changes to the unit cell parameters, and therefore the MOF crystal structure, during the desolvation process. Unfortunately, due to limitations in the data quality and the triclinic nature of the material, this new pattern could not be indexed. However, the general shift to larger 2θ values is suggestive of a reduction in d-spacing and therefore a reduction in overall unit cell volume. Heating to higher temperatures to completely remove the guests (beyond point B in Figure 3a) resulted in a loss of crystallinity. This fully activated the material also displayed a negligible volumetric uptake of nitrogen (Supporting Information S6).Figure 3. a) Thermogravimetric analysis of ZnCSA.DMF (1) ramping at 3 oC/min. b) PXRD patterns of ZnCSA.DMF while immersed in DMF (Blue) and after drying and heating at 120oC to remove some of its guests (Green). The labels A and B in the figure are to show the position of the obtained PXRD patterns in the solvent loss process. A corresponds to the as-synthesised material 1, while B is a partially desolvated material, but is significantly more desolvated than the partially-desolvated single crystal structure 2.To gain further structural insight into the framework changes during guest removal, a single crystal structure was obtained after the material was left to dry in air for 1hr, affording the partially desolvated material 2. The level of desolvation of this crystal, however, is unknown because the significantly reduced data quality from undergoing the phase transition makes analysis of the residual electron density unreliable. 2 showed an identical building unit (metal-oxo cluster and bound DMF molecules), connectivity and topology to 1, but displayed noticeable differences in the relative angles between the different metal cluster connections, resulting in changes to the unit cell a = 15.644(1) ?, b = 17.386(1) ?, c = 18.668(3) ?, α = 97.05(1)°, β = 115.49 (1)°, γ = 105.851(7)°, V = 4236.8(9) ?3 (a 9.5% volume reduction). In particular, the single c axis linker changes significantly, which is directly reflected in the change of β from 1 (Δβ = 5.5°). This resulted in a reduced volume for the unit cell and therefore a reduction in void space from 43.4% to 38.3%. The presence of crystallographic disorder on the single c axis linker, requiring a number of bond distance restraints to model, means information on the changes to this part of the material could not be accurately interpreted from the refinement. To approach this problem and provide a chemically sensible model for the disordered components of the two structures we employed computational methods, optimising non-disordered versions of crystal structures 1 and 2 using Density Functional Theory (DFT) within a fixed 2×2×1 supercell, C1opt and C2 opt respectively (where C stands for computational and 1 and 2 stand for the crystal structure used as input geometry). Although computationally expensive, this choice of supercell dimensions (double the experimental cell along a and b axes) allows us to model the crystallographic disorder in an alternating manner, such that for any given SBU the binding of the single c axis linker (e.g., through its rigid end) is different to its four neighbouring SBU’s (e.g., through its flexible end) in the ab plane. It is worth mentioning that computational structures were always modelled with uncoordinated solvent molecules removed (the terminal DMF coordinated to the SBU are preserved), therefore fixed unit cell (to the experimental parameter values) geometry optimizations (the subscript “opt” refers to optimized) were intended to represent the solvated crystal, but the solvent was not actually modelled using the experimentally determined compositions due to computational cost. The overall validity of the modelling method was assessed by directly overlaying 1 and C1opt (using the central atoms of four different metal SBUs) and comparing the positions of the computationally optimised atoms with the non-disordered atomic sites observed experimentally (i.e., the framework without the single c axis linker). As 1 displayed significantly higher resolution experimental data than 2, it was deemed most reliable for this analysis. Figure 4a shows the extracted sections of the overlay for all the ordered and crystallographically unique components of the structure, including the SBU, one of the two symmetry related linkers lying along the a axis and one of the two symmetry related linkers lying along the b axis. A good agreement between the experimental and computational models was observed, with the positions of all the calculated atoms, represented using a green stick model, lying within the anisotropic thermal displacement ellipsoids determined experimentally. Small root-mean-square displacement (RMSD) values (see Figure 4a) of the individual section overlays further support the computational-experimental agreement on the non-disordered parts of the structure. While the disordered components of the structure could not be directly compared in a similar way, the crystallographic model of the disordered single c axis linker, and the calculated linker environment after inversion symmetry has been applied post calculation, is shown overlaid in Figure 4b which suggests no major differences. Accordingly, the computational models were deemed suitable to provide reliable information and could be used to study the changing single c axis linkers in the structure.Figure 4. a) Extracted sections of the overlay of the experimentally determined structure 1, shown using ellipsoids drawn at 50% probability, and the DFT optimized structure C1opt, represented with a green stick model. Root mean square displacements (RMSD) are given below each section. b) Left: The two half occupancy inversion symmetry related orientations of the single c axis linker modelled crystallographically in 1. Right: Two orientations of the single c axis linker in C1opt obtained by artificially introducing inversion symmetry into the non-disordered model post calculation. c) Environment of the single c axis linker in C1opt showing the defined coordination angles θ1, θ2 and torsional angles φ, ψ, δ. d) Overlay of computationally optimised single c axis linkers in C1opt (Dark colours) & the partially desolvated C2opt (Light colours). The linkers are overlaid using the two carboxylate oxygens, and the two Zn atoms which they coordinate to, of the flexible end of the linker. Hydrogen atoms have been removed for clarity.To understand the starting point for this change, the environment of the asymmetric linker in the computed ordered structure of C1opt should be described initially. Significantly, this single c axis linker, shown in Figure 4c, is predicted to lie slightly out of the plane of the connection between its two SBUs. This distortion can be quantified by the two angles ω1 and ω2, one measured on either side of the linker, using the centroid of the two Zn atoms from the connecting SBU, the centroid of the two coordinating carboxylate oxygens and the centre of the linker. These angles are almost identical, with ω1 = 154° at the rigid end and ω2 = 153° at the flexible end of the linker. At the rigid end, ω1 can be seen to primarily arise from the direct coordination angle of the carboxylate group (θ1) at 158°, defined in Figure 4c using the centroid of the two Zn atoms, the centroid of the two carboxylate oxygens, and the first carbon after the carboxylate group. Meanwhile, at the flexible end, ω2 cannot be described solely by the much wider θ2 at 170°, defined in Figure 4c consistently with θ1. The same ω (bend out of plane) is instead achieved through additional conformational adjustments driven by rotations of three torsional angles associated with the linker’s sp3 carbons. These angles (φ, ψ & δ) become -158°, 163° and 152° respectively. Comparing C1opt and C2 opt, to understand the structural transformation upon guest removal, the major change arises due to a shift in the degree that the c axis linker lies out of the plane of the connection of the SBUs, with ω1 and ω2 both changing to become 144°. At the rigid end of the linker this occurs only through decreasing θ1 to 149°, while at the flexible end of the linker θ2 decreases to 162°, combined with further torsional adjustments to ψ and δ, which change to 158° and 149° respectively. φ meanwhile is predicted to remain almost the same at -157° (torsions values tabulated in Supporting Information S7). The resulting effect is shown in Figure 4d, which shows an overlay of the predicted c axis linker in C1opt and C2opt. While these angles cannot be directly compared to experimental results, as the modelling of disorder makes exact torsional measurements on the individual components unreliable, an average value for the bend of the linker out of plane at both ends can be obtained. This shows very similar overall response to that predicted by calculation, with the averaged ω changing from 156° to 146°, in very good agreement with the predicted 154° to 144° change. Therefore, although guest species adsorbed within the pores are not explicitly modelled in the calculations, fixing the unit cell to its experimental parameters allows us to reliably reproduce the change in the c axis linker responsible for material flexible response. Further treatment of single crystals, including heating at 120 oC to mimic the preparation conditions of the sample from which the powder diffraction pattern was collected, resulted in the complete loss of single crystal crystallinity. Therefore, to try to predict the structural behaviour of the MOF during full solvent removal, our DFT calculations were extended by optimising the empty structure while allowing the dimensions of the 2×2×1 supercell to vary. This “full structural relaxation”, C1relax (“relax” stands for relaxed), in which both the ions and unit cell parameters are allowed to relax, aims to predict the energy minimum of the solvent free framework. This is particularly relevant because the lattice parameters could not be determined from the PXRD, and therefore even the size of the repeating unit is completely unknown. The optimisation, which gave similar relaxed geometries for both 1 and 2, suggests the material undergoes a drastic change to its unit cell, equivalent to a = 15.30 ?, b = 17.19 ?, c = 15.15 ?, α = 96.8°, β = 138.43°, γ = 109.58°, V = 2010.4 ?3 (C1relax) when related back to a 1x1x1 cell. This new cell is roughly 40% the size of 1 and, similar to the differences between C1opt and C2opt - a result which are have been experimentally validated, the change between C1opt and C1relax is primarily driven by movement of the single c axis linker, producing large differences in c axis length and β angle. While this cell does not directly index the powder pattern obtained after heating at 120 °C, the d-spacings of the first two predicted peaks (14.1? & 13.1?) are similar to those observed experimentally (14.4? & 13.4?), therefore suggesting that it may provide a reasonable approximation to the largest changes in the structure. A comparison of the experimental powder pattern after heating at 120 °C and the predicted powder pattern of C1relax is shown in Supporting Figure S5. The changes to the structure between C1opt and C1relax occur broadly through the same mechanism observed in the optimised single crystal structures (C1opt and C2opt) i.e., changes in the conformation and binding of the c axis linker, but to a larger extent. ω1 reduces to 109°, driven primarily by a reduction in θ1 to 126°, but now also involves a bending out of plane of the aromatic ring. ω2 meanwhile reduces to 114°, θ2 changing to 143°, and the torsional angles φ, ψ and δ becoming -115°, 164° and 168° (torsion values tabulated in Supporting Information S7). Compared to the change between C1opt and C2opt, C1relax shows a much more significant response in the torsion φ during the optimisation, accounting for the largest change in the flexible end of the linker. The overall mechanism, shown in Figure 5, causes a large reduction in the accessible void space to 5.9% and in the pore limiting diameter to 3.7 ?. C1relax displays a complete removal of the channels running along the a and b axes resulting from the single c axis linker folding back to lie almost in plane with the a axis linkers (see Figure 5d). This anisotropic single axis response observed in ZnCSA during the “structural relaxation” is closely related to the MOF’s topology. The presence of the double linker connections along the a and b axes prevents any significant changes along these directions, thus effectively providing two-dimensional rigid platforms. The single CSA linker connections between these platforms, however, grants the necessary freedom to respond along the c axis, with changes occurring at both its rigid end (change in coordination angle θ1 in a hinge motion) and at its flexible end (conformational adjustments of the torsional angles φ, ψ and δ).Figure 5. a) Computed optimized structure with fixed experimentally determined cell parameters, C1opt, viewed down the a axis. b) Computed relaxed structure after full relaxation of both positions and cell parameters, C1relax, viewed down the a axis. c) 2 SBUs of C1opt connected via the single c axis linker connection. φ and θ1 are shown in expanded sections. d) 2 SBUs of C1relax connected via the single c axis linker connection. φ and θ1 are shown in expanded sections to illustrate the conformational rearrangements and the coordination change observed at the flexible and at the rigid ends of the linker respectively.The behaviour of ZnCSA was further explored through exchange of the DMF solvent in ZnCSA.DMF (1) with DMSO giving ZnCSA.DMSO (3), formula [Zn7O2(CSA)5DMSO2].18DMSO. The crystal structure was obtained by X-ray diffraction at 100K and transformed from its standard setting to ensure consistency in the structural description. 3 shows a drastic change in the unit cell dimensions, a = 17.605(4) ?, b = 17.548(4) ?, c = 20.650(6) ?, α = 110.09(2)°, β = 79.19(2)°, γ = 87.58(2)°, V = 5845(3) ?3, but displayed the same overall topology as 1. The pore contains 18 guest DMSO molecules based on thermogravimetric analysis, however, the crystallographic data suggests a slightly higher value of 22.5 guest DMSO molecules. Similar to ZnCSA.DMF this was calculated using a combination of the two guest sites crystallographically resolved in the asymmetric unit, giving a refined composition of [Zn7O2(CSA)5DMSO2].2.64(3)DMSO, and the residual electron density calculated by the routine SQUEEZE. This value is reasonably close to upper limit of 27.7 molecules, based on the absolute pore volume (3150 ?3) and a packing density identical to liquid DMSO. The changes in structure from 1 to 3 arise primarily due to the exchange of the two coordinated DMF molecules in the SBU for two DMSO molecules. This process involves a significant rearrangement of the metal cluster; two of the tetrahedral Zn atoms (symmetry equivalents) become octahedral, coordinating to the incoming DMSO molecules and the flexible end of a b axis CSA linker, while the two square pyramidal Zn atoms (coordinated to the outgoing DMFs) becomes tetrahedral, losing their DMF and swapping a connection to the flexible end of a b axis linker for a flexible end of an a axis linker. This change is coupled to conformational adjustments in the flexible ends of the connected a and b axes linkers, which display rotations around the torsions φ, ψ and δ (values tabulated in Supporting Information S7). In the a axis linkers these can be seen to change as follows: φ -151° to 107°, ψ 60° to -177° and δ 173° to 170°. The b axis linker, meanwhile, shows the changes φ -116° to -73°, ψ -176° to -174° and δ -150° to 167°. In addition, the bend out of plane of the c axis linker is also observed to show a response, with the averaged experimental value changing from 156° to 193°, essentially switching from bending in one direction to bending in the other (see Fig 6b). A possible mechanism, depicted in Figure 6a, starts with the binding of the new DMSO molecule to a tetrahedral Zn site without any coordinated solvent (Zn2). This causes a coordinated carboxylate from an a axis linker to rotate, transferring an oxygen (O1) from bonding to Zn2 to bonding to Zn1, the site coordinated to DMF. There is also a shift of a second carboxylate (b axis linker) changing an oxygen (O3) from bonding to Zn1 to bonding to Zn2’ (the symmetry equivalent to Zn2). Finally, the mechanism results in the departure of the original DMF molecule. The carboxylates involved in the cluster reorganisation are the central a and b axes linkers connected through the flexible end of the CSA linker, these are ordered and could be accurately characterised by crystallographic techniques, thus enabling the analysis of the coordination change without relying on computational structures. It is worth mentioning that any conformational changes in these linkers during DMF removal were not observed to significantly affect the stability of the a and b directions, thus retaining the pore aperture along the c axis. It should also be noted that the mechanism is mimicked on the opposing side of the SBU with the symmetry generated atoms Zn1’, O1’, O3’ and Zn2’. The large rearrangement of the a axis linker results in a new much straighter linker conformation, which, combined with the changes to the c axis linker, gives a significant increase in unit cell volume (20% in comparison to 1), and in the void space of the channel running along the crystallographic b axis (see Figure 6c-d). The solvent accessible void increases to 53.9% and the pore limiting diameter (now running along the b axis) increases to 11.3 ?. Exchanges with other solvents, MeOH and THF, also showed evidence of structural rearrangements by PXRD (Supporting Figure S7). However, unlike exchange with DMSO, the material didn’t retain its single crystal crystallinity.Figure 6. a) SBU and one linker (situated along the crystallographic a axis) of 1. A proposed mechanism for DMSO exchange to structure 3 is shown. Analogous to Figure 1 carboxylate carbons coloured green are connections running along the a axis, coloured purple are connections running along the b axis, and coloured orange are connections running along the c axis. Striped coloured connections are connections through the flexible end of the linker, solid colours are connections through the rigid end and half striped connections correspond to 50/50 distribution through the flexible and rigid ends of the linkers. b) SBU and one linker (situated along the crystallographic a axis) of 3. c) View along the b axis of 1, Connolly surface for this orientation (calculated using a 1.2 ? probe radius) and crystallographically modelled a axis linker showing a bent conformation (ψ = 60°). d) View along the b axis of 3, Connolly surface for this orientation (calculated using a 1.2 ? probe radius) and crystallographically modelled a axis linker showing a straight conformation (ψ = -177°). In an attempt to experimentally study the behavior of ZnCSA.DMSO during removal of the DMSO contained within the pores, but not coordinated to the SBU, a sample was heated in analogous manner to ZnCSA.DMF. The TGA of ZnCSA.DMSO (Supporting Figure S2) showed that the majority of the solvent is removed at 175 oC (higher than the DMF analogue), however, heating at a conservative 150 oC was observed to result in a loss of crystallinity, with the powder pattern (Supporting Figure S8) losing all diffraction peaks except for the strongest peak in the pattern. This remaining peak also showed a reduction in intensity and a significant broadening. It should be noted that in both ZnCSA.DMF and ZnCSA.DMSO the TGA data suggest that the coordinated solvent is not lost until higher temperatures, roughly 350 oC. The full structural relaxation was therefore instead investigated via DFT optimization, an attempt to predict the behavior, yielding an equilibrium structure (C3relax) of unit cell parameters equivalent to a = 16.37 ?, b = 17.34 ?, c = 16.60 ?, α = 140.49°, β = 61.60°, γ = 86.61°, V = 1702.3 ?3. The extent of the overall change, upon removal of only the uncoordinated guests, is predicted to be even larger than for ZnCSA.DMF, the cell of C3relax being only 30% of the size of 3. The overall mechanism leads to a large reduction in the accessible void space, now only 2% for C3relax, and in the pore limiting diameter to 1.2 ?, such that any remaining void space can essentially be considered as pockets rather channels. Since the MOF topology is preserved during the solvent exchange process, in particular the single and double CSA linker connections, the structural change in ZnCSA.DMSO occurs via qualitatively the same mechanism as in ZnCSA.DMF. Major changes involve out of plane distortions of the single c axis linker through adjustment of the coordination angles θ1 & θ2 and the torsional angles φ, ψ & δ. However, in ZnCSA.DMSO this bending of the linker occurs in the opposite direction, relative to its SBU, to that predicted for ZnCSA.DMF (shown in Figure 7). This is related to the change in bending direction of the linker that occurs during the solvent exchange process and means the linker folds over the SBU’s coordinated solvent, which resides on opposite sides in the two materials. The folding in ZnCSA.DMF is also mainly along the crystallographic a axis, while the folding in ZnCSA.DMSO is in-between the a and b axes, resulting in C3relax showing a collapse into the centre of the ab plane. This behavior is evidenced from the unit cell changes where ZnCSA.DMF (1 to C1relax) shows major changes to the β angle (suggesting a mechanism mainly involving the a and c directions), while ZnCSA.DMSO (3 to C3relax) shows a large increase of α (similar to β in ZnCSA.DMF) coupled with a decrease of β (a summary table of unit cell parameters is available in Supporting Information S9). This difference is responsible for the predicted larger reduction in void space, since the collapse of the DMSO structure into the centre of the ab plane eliminates the channel running along the original c axis (Figure 7h) that is still present in ZnCSA.DMF (1relax) (Figure 7d). Quantified changes to the single c axis linker, measured between C3opt (the optimized structure with fixed experimentally determined cell parameters) and C3relax, can be found in Supporting Information S8 where they are compared to changes between C1opt and C1relax.Figure 7. a) View along the b axis of ZnCSA.DMF (1). CSA linkers coloured green are connections running along the a axis and CSA linkers coloured orange are connections running along the c axis. b) View along the b axis of C1relax, the predicted DMF structure after guest removal, using the same colour scheme. c) Two SBUs in C1relax connected by a single c axis linker. d) Connolly surface of C1relax (calculated using a 1.2 ? probe radius) showing the 1D channel remaining after structure collapse. The channel runs in the direction of the c axis defined for the original structure 1. e) View along the b axis of ZnCSA.DMSO (3). CSA linkers coloured green are connections running along the a axis and CSA linkers coloured orange are connections running along the c axis. f) View along the b axis of C3relax, the predicted DMSO structure after guest removal, using the same colour scheme. g) Two SBUs in C3relax connected by a single c axis linker. h) Connolly surface of C3relax (calculated using a 1.2 ? probe radius) showing the removal of all channels.DiscussionThe diffraction data collected for ZnCSA reveals that the material exhibits a significant degree of flexibility. This behaviour can be divided into two separate responses: A large contraction of the void space during removal of its contained guests, and a structural rearrangement caused by the SCCSE of DMF to DMSO. The structural change during guest removal is most obviously seen in the PXRD pattern collected after heating, where the general shift of the Bragg reflections to higher 2θ values is suggestive of a significant reduction in the unit cell dimensions. Indications of the mechanism behind this contraction emerge from the analysis of the single crystal-derived structures of 1 & 2, which show a small scale (Δ9.5 % in volume) structural transformation after partial loss of the contained guests. The transformation is largely related to the single c axis linker, with the double linker connections lying along the a and b axes being observed to show very little motion. However, the exact role of the c axis linker in the transformation of 1 to 2 could not be accurately interpreted from crystallographic refinements alone. Unlike the double linker connections along the a and b axes, which are highly ordered throughout the structure sitting antiparallel and related by inversion symmetry, the single c axis linkers coordinate to their SBUs in one of two possible orientations of their rigid and flexible ends. These orientations are randomly distributed throughout the structure and therefore result in high levels of crystallographic disorder, requiring idealised bond distance restraints to generate as a refineable model, which should not be over-interpreted. The interplay between disorder and flexibility, here a consequence of the material topology, is highly intriguing and has been discussed before by Bennet et al.,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1038/nchem.2691","ISBN":"1755-4349 (Electronic)\\r1755-4330 (Linking)","ISSN":"1755-4330","PMID":"27995920","abstract":"A graphical abstract is available for this content","author":[{"dropping-particle":"","family":"Bennett","given":"Thomas D.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cheetham","given":"Anthony K.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fuchs","given":"Alain H.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Coudert","given":"Fran?ois-Xavier","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2016","12","20"]]},"page":"11-16","publisher":"Nature Publishing Group","title":"Interplay between defects, disorder and flexibility in metal-organic frameworks","type":"article-journal","volume":"9"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>68</sup>","plainTextFormattedCitation":"68","previouslyFormattedCitation":"<sup>68</sup>"},"properties":{"noteIndex":0},"schema":""}68 particularly on the dynamical disorder of UiO-abdc, an analogue of UiO-67 made of azobenzene-4,4?-dicarboxylate linkers.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/anie.201509352","ISBN":"1433-7851","ISSN":"14337851","PMID":"26797762","abstract":"Whilst many metal-organic frameworks possess the chemical stability needed to be used as functional materials, they often lack the physical strength required for industrial applications. Herein, we have investigated the mechanical properties of two UiO-topology Zr-MOFs, the planar UiO-67 ([Zr6O4(OH)4(bpdc)6], bpdc: 4,4′-biphenyl dicarboxylate) and UiO-abdc ([Zr6O4(OH)4(abdc)6], abdc: 4,4′-azobenzene dicarboxylate) by single-crystal nanoindentation, high-pressure X-ray diffraction, density functional theory calculations, and first-principles molecular dynamics. On increasing pressure, both UiO-67 and UiO-abdc were found to be incompressible when filled with methanol molecules within a diamond anvil cell. Stabilization in both cases is attributed to dynamical linker disorder. The diazo-linker of UiO-abdc possesses local site disorder, which, in conjunction with its longer nature, also decreases the capacity of the framework to compress and stabilizes it against direct compression, compared to UiO-67, characterized by a large elastic modulus. The use of non-linear linkers in the synthesis of UiO-MOFs therefore creates MOFs that have more rigid mechanical properties over a larger pressure range. ? 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.","author":[{"dropping-particle":"","family":"Hobday","given":"Claire L.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Marshall","given":"Ross J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Murphie","given":"Colin F.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Sotelo","given":"Jorge","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Richards","given":"Tom","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Allan","given":"David R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Düren","given":"Tina","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Coudert","given":"Fran?ois-Xavier","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Forgan","given":"Ross S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Morrison","given":"Carole A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Moggach","given":"Stephen A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bennett","given":"Thomas D.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-1","issue":"7","issued":{"date-parts":[["2016","2","12"]]},"page":"2401-2405","title":"A Computational and Experimental Approach Linking Disorder, High-Pressure Behavior, and Mechanical Properties in UiO Frameworks","type":"article-journal","volume":"55"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>69</sup>","plainTextFormattedCitation":"69","previouslyFormattedCitation":"<sup>69</sup>"},"properties":{"noteIndex":0},"schema":""}69Periodic DFT calculations, benchmarked by comparing the output to the experimentally determined non-disordered a and b axes linkers, were therefore employed to gain insights into the environment of the changing single c axis linkers. These calculations predict that the transition is a result of a joint mechanism, where the rigid end of the organic linker undergoes changes in carboxylate coordination angle, while the flexible aliphatic end shows similar coordination changes combined with conformational adjustments around its sp3 carbons. The coordination angle changes predicted are commonly observed in flexible MOFs built from rigid linkers, particularly wine-rack style materials, e.g., MIL-53ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/adma.200602645","ISSN":"09359648","author":[{"dropping-particle":"","family":"Serre","given":"C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bourrelly","given":"S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vimont","given":"A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ramsahye","given":"N. A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Maurin","given":"G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Llewellyn","given":"P. L.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Daturi","given":"M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Filinchuk","given":"Y.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Leynaud","given":"O.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Barnes","given":"P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"G.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Advanced Materials","id":"ITEM-1","issue":"17","issued":{"date-parts":[["2007","9","3"]]},"page":"2246-2251","title":"An Explanation for the Very Large Breathing Effect of a Metal–Organic Framework during CO2 Adsorption","type":"article-journal","volume":"19"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>70</sup>","plainTextFormattedCitation":"70","previouslyFormattedCitation":"<sup>70</sup>"},"properties":{"noteIndex":0},"schema":""}70 or DMOFADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/anie.200460712","ISSN":"1433-7851","PMID":"15384114","author":[{"dropping-particle":"","family":"Dybtsev","given":"Danil N","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chun","given":"Hyungphil","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kim","given":"Kimoon","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie (International ed. in English)","id":"ITEM-1","issue":"38","issued":{"date-parts":[["2004","9","27"]]},"page":"5033-6","title":"Rigid and flexible: a highly porous metal-organic framework with unusual guest-dependent dynamic behavior.","type":"article-journal","volume":"43"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>71</sup>","plainTextFormattedCitation":"71","previouslyFormattedCitation":"<sup>71</sup>"},"properties":{"noteIndex":0},"schema":""}71, and are often referred to as a knee cap or hinge motion.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/b804302g","ISBN":"0306-0012","ISSN":"0306-0012","PMID":"19384443","abstract":"This critical review focuses on a strange behaviour of crystallized solid matter: its reversible swelling with large magnitude. This will be of interest for experts in porous solids but also for solid state chemists and physicists. Some examples, classified according to the dimensionality of the inorganic subnetwork, present the general requirements and the structural rules which govern the existence of this phenomenon. Its consequences concern specific applications related to sensors, energy savings, sustainable development and health (100 references).","author":[{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical Society reviews","id":"ITEM-1","issue":"5","issued":{"date-parts":[["2009"]]},"page":"1380-1399","title":"Large breathing effects in three-dimensional porous hybrid matter: facts, analyses, rules and consequences.","type":"article-journal","volume":"38"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1038/nchem.2747","ISSN":"1755-4330","abstract":"Understanding the behaviour of flexible metal–organic frameworks (MOFs)—porous crystalline materials that undergo a structural change upon exposure to an external stimulus—underpins their design as responsive materials for specific applications, such as gas separation, molecular sensing, catalysis and drug delivery. Reversible transformations of a MOF between open- and closed-pore forms—a behaviour known as ‘breathing’—typically occur through well-defined crystallographic transitions. By contrast, continuous breathing is rare, and detailed characterization has remained very limited. Here we report a continuous-breathing mechanism that was studied by single-crystal diffraction in a MOF with a diamondoid network, (Me2NH2)[In(ABDC)2] (ABDC, 2-aminobenzene-1,4-dicarboxylate). Desolvation of the MOF in two different solvents leads to two polymorphic activated forms with very different pore openings, markedly different gas-adsorption capacities and different CO2 versus CH4 selectivities. Partial desolvation introduces a gating pressure associated with CO2 adsorption, which shows that the framework can also undergo a combination of stepped and continuous breathing.","author":[{"dropping-particle":"","family":"Carrington","given":"Elliot J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"McAnally","given":"Craig A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fletcher","given":"Ashleigh J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Thompson","given":"Stephen P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"Mark","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Brammer","given":"Lee","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-2","issue":"9","issued":{"date-parts":[["2017","3","14"]]},"page":"882-889","publisher":"Nature Publishing Group","title":"Solvent-switchable continuous-breathing behaviour in a diamondoid metal–organic framework and its influence on CO2 versus CH4 selectivity","type":"article-journal","volume":"9"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>72,73</sup>","plainTextFormattedCitation":"72,73","previouslyFormattedCitation":"<sup>72,73</sup>"},"properties":{"noteIndex":0},"schema":""}72,73 Similarly, conformational changes around sp3 carbons are often the key mechanism to many flexible motions reported in frameworks built from solely aliphatic or peptide-based linkers.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/anie.201307074","ISBN":"1521-3773","ISSN":"14337851","PMID":"24302659","abstract":"The peptide-based porous 3D framework, ZnCar, has been synthesized from Zn(2+) and the natural dipeptide carnosine (β-alanyl-L-histidine). Unlike previous extended peptide networks, the imidazole side chain of the histidine residue is deprotonated to afford Zn-imidazolate chains, with bonding similar to the zeolitic imidazolate framework (ZIF) family of porous materials. ZnCar exhibits permanent microporosity with a surface area of 448 m(2) g(-1) , and its pores are 1D channels with 5 ? openings and a characteristic chiral shape. This compound is chemically stable in organic solvents and water. Single-crystal X-ray diffraction (XRD) showed that the ZnCar framework adapts to MeOH and H2 O guests because of the torsional flexibility of the main His-β-Ala chain, while retaining the rigidity conferred by the Zn-imidazolate chains. The conformation adopted by carnosine is driven by the H bonds formed both to other dipeptides and to the guests, permitting the observed structural transformations.","author":[{"dropping-particle":"","family":"Katsoulidis","given":"Alexandros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Park","given":"Kyo Sung","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"Dmytro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Martí-Gastaldo","given":"Carlos","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Miller","given":"Gary J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"John E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Robertson","given":"Craig M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Blanc","given":"Frédéric","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"George R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"Neil G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Purton","given":"John A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Adams","given":"Dave J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"Matthew J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2014","1","3"]]},"page":"193-198","title":"Guest-Adaptable and Water-Stable Peptide-Based Porous Materials by Imidazolate Side Chain Control","type":"article-journal","volume":"53"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1038/nchem.1871","ISBN":"1755-4349","ISSN":"1755-4330","PMID":"24651203","abstract":"Porous materials are attractive for separation and catalysis-these applications rely on selective interactions between host materials and guests. In metal-organic frameworks (MOFs), these interactions can be controlled through a flexible structural response to the presence of guests. Here we report a MOF that consists of glycyl-serine dipeptides coordinated to metal centres, and has a structure that evolves from a solvated porous state to a desolvated non-porous state as a result of ordered cooperative, displacive and conformational changes of the peptide. This behaviour is driven by hydrogen bonding that involves the side-chain hydroxyl groups of the serine. A similar cooperative closure (reminiscent of the folding of proteins) is also displayed with multipeptide solid solutions. For these, the combination of different sequences of amino acids controls the framework's response to the presence of guests in a nonlinear way. This functional control can be compared to the effect of single-point mutations in proteins, in which exchange of single amino acids can radically alter structure and function.","author":[{"dropping-particle":"","family":"Martí-Gastaldo","given":"C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"D.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"J. E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Briggs","given":"M. E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chater","given":"P. A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"V.","family":"Wiper","given":"P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Miller","given":"G. J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Khimyak","given":"Y. Z.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"G. R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"N. G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"M. J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-2","issue":"4","issued":{"date-parts":[["2014","4","23"]]},"page":"343-351","title":"Side-chain control of porosity closure in single- and multiple-peptide-based porous materials by cooperative folding","type":"article-journal","volume":"6"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1038/s41586-018-0820-9","ISSN":"0028-0836","abstract":"Metal–organic frameworks (MOFs) are crystalline synthetic porous materials formed by binding organic linkers to metal nodes: they can be either rigid1,2 or flexible3. Zeolites and rigid MOFs have widespread applications in sorption, separation and catalysis that arise from their ability to control the arrangement and chemistry of guest molecules in their pores via the shape and functionality of their internal surface, defined by their chemistry and structure4,5. Their structures correspond to an energy landscape with a single, albeit highly functional, energy minimum. By contrast, proteins function by navigating between multiple metastable structures using bond rotations of the polypeptide6,7, where each structure lies in one of the minima of a conformational energy landscape and can be selected according to the chemistry of the molecules that interact with the protein. These structural changes are realized through the mechanisms of conformational selection (where a higher-energy minimum characteristic of the protein is stabilized by small-molecule binding) and induced fit (where a small molecule imposes a structure on the protein that is not a minimum in the absence of that molecule)8. Here we show that rotation about covalent bonds in a peptide linker can change a flexible MOF to afford nine distinct crystal structures, revealing a conformational energy landscape that is characterized by multiple structural minima. The uptake of small-molecule guests by the MOF can be chemically triggered by inducing peptide conformational change. This change transforms the material from a minimum on the landscape that is inactive for guest sorption to an active one. Chemical control of the conformation of a flexible organic linker offers a route to modifying the pore geometry and internal surface chemistry and thus the function of open-framework materials.","author":[{"dropping-particle":"","family":"Katsoulidis","given":"Alexandros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"Dmytro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Whitehead","given":"George F. S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Carrington","given":"Elliot J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Adams","given":"Dave J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"Neil G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"George R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dyer","given":"Matthew S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"Matthew J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature","id":"ITEM-3","issue":"7738","issued":{"date-parts":[["2019","1","9"]]},"page":"213-217","publisher":"Springer US","title":"Chemical control of structure and guest uptake by a conformationally mobile porous material","type":"article-journal","volume":"565"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>16–18</sup>","plainTextFormattedCitation":"16–18","previouslyFormattedCitation":"<sup>16–18</sup>"},"properties":{"noteIndex":0},"schema":""}16–18 The two types of flexibility appear to lead to very similar distortion to the linker. This is evidenced by the 10o change in ω at both the rigid and flexible ends of the linker, which closely matches the averaged distortion modelled crystallographically. The random distribution of the orientation of this single c axis linker is likely the cause for the similarity, requiring symmetrical responses in order to maintain the more rigid 2D network constructed from the double linker connections along the a and b axes. Importantly, the change to the average bend of the single c axis linker out of the plane of the two SBUs it connects during the transformation of 1 to 2 is reliably captured by the differences between the computationally optimised structures C1opt and C2opt. Therefore, by extending our calculations further we have also probed the behaviour under full removal of the solvent, predicting extended changes on the same joint mechanism, Δω≈ 50o, resulting in a significantly contracted form with a considerably reduced void space, as suggested by PXRD. The extent of the knee cap motion observed in these calculations is comparable with other highly flexible systems. The coordination angle at the rigid end (θ1) is observed to change by 32°, which compares to the 31° response seen in MIL-88-B.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ja206936e","ISBN":"1520-5126 (Electronic)\\n0002-7863 (Linking)","ISSN":"0002-7863","PMID":"21950795","abstract":"A series of organically modified iron(III) terephthalate MIL-88B and iron(III) 4,4'-biphenyl dicarboxylate MIL-88D flexible solids have been synthesized and characterized through a combination of X-ray diffraction, IR spectroscopy, and thermal analysis (MIL stands for Material from Institut Lavoisier). The swelling amplitude of the highly flexible MOFs tuned by introducing functional groups onto the phenyl rings shows a clear dependence on the steric hindrance and on the number of groups per aromatic ring. For instance, while the introduction of four methyl groups per spacer in dried MIL-88B results in a large permanent porosity, introducing two or four methyl groups in MIL-88D allows an easier pore opening in the presence of liquids without drastically decreasing the swelling magnitude. The influence of the degree of saturation of the metal center and the nature of the solvent on the swelling is also discussed. Finally, a computationally assisted structure determination has led to a proposal of plausible structures for the closed (dried) and open forms of modified MIL-88B and MIL-88D and to evaluation of their framework energies subject to the nature of the functional groups.","author":[{"dropping-particle":"","family":"Horcajada","given":"Patricia","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Salles","given":"Fabrice","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wuttke","given":"Stefan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Devic","given":"Thomas","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Heurtaux","given":"Daniela","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Maurin","given":"Guillaume","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vimont","given":"Alexandre","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Daturi","given":"Marco","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"David","given":"Olivier","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Magnier","given":"Emmanuel","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Stock","given":"Norbert","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Filinchuk","given":"Yaroslav","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Popov","given":"Dmitry","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Riekel","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-1","issue":"44","issued":{"date-parts":[["2011","11","9"]]},"page":"17839-17847","title":"How Linker’s Modification Controls Swelling Properties of Highly Flexible Iron(III) Dicarboxylates MIL-88","type":"article-journal","volume":"133"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1039/b512169h","ISBN":"1359-7345","ISSN":"1359-7345","PMID":"16391735","abstract":"We report here a new family of isoreticular MOFs, comprising three larger analogues of the nanoporous metallocarboxylate MIL-88; these solids were synthesized using a controlled SBU approach and the three crystal structures were solved using an original simulation-assisted structure determination method in direct space.","author":[{"dropping-particle":"","family":"Surblé","given":"Suzy","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mellot-Draznieks","given":"Caroline","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Millange","given":"Franck","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical communications (Cambridge, England)","id":"ITEM-2","issue":"3","issued":{"date-parts":[["2006"]]},"page":"284-286","title":"A new isoreticular class of metal-organic-frameworks with the MIL-88 topology.","type":"article-journal"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1038/nchem.2747","ISSN":"1755-4330","abstract":"Understanding the behaviour of flexible metal–organic frameworks (MOFs)—porous crystalline materials that undergo a structural change upon exposure to an external stimulus—underpins their design as responsive materials for specific applications, such as gas separation, molecular sensing, catalysis and drug delivery. Reversible transformations of a MOF between open- and closed-pore forms—a behaviour known as ‘breathing’—typically occur through well-defined crystallographic transitions. By contrast, continuous breathing is rare, and detailed characterization has remained very limited. Here we report a continuous-breathing mechanism that was studied by single-crystal diffraction in a MOF with a diamondoid network, (Me2NH2)[In(ABDC)2] (ABDC, 2-aminobenzene-1,4-dicarboxylate). Desolvation of the MOF in two different solvents leads to two polymorphic activated forms with very different pore openings, markedly different gas-adsorption capacities and different CO2 versus CH4 selectivities. Partial desolvation introduces a gating pressure associated with CO2 adsorption, which shows that the framework can also undergo a combination of stepped and continuous breathing.","author":[{"dropping-particle":"","family":"Carrington","given":"Elliot J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"McAnally","given":"Craig A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fletcher","given":"Ashleigh J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Thompson","given":"Stephen P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"Mark","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Brammer","given":"Lee","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-3","issue":"9","issued":{"date-parts":[["2017","3","14"]]},"page":"882-889","publisher":"Nature Publishing Group","title":"Solvent-switchable continuous-breathing behaviour in a diamondoid metal–organic framework and its influence on CO2 versus CH4 selectivity","type":"article-journal","volume":"9"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>15,73,74</sup>","plainTextFormattedCitation":"15,73,74","previouslyFormattedCitation":"<sup>15,73,74</sup>"},"properties":{"noteIndex":0},"schema":""}15,73,74 Meanwhile, the torsional changes at the flexible end of the linker observed in ZnCSA are in the range of 100o – 120o, similar to the rotations observed in peptide-based MOFs displaying structural transformations that are entirely attributed to conformational freedom.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/anie.201307074","ISBN":"1521-3773","ISSN":"14337851","PMID":"24302659","abstract":"The peptide-based porous 3D framework, ZnCar, has been synthesized from Zn(2+) and the natural dipeptide carnosine (β-alanyl-L-histidine). Unlike previous extended peptide networks, the imidazole side chain of the histidine residue is deprotonated to afford Zn-imidazolate chains, with bonding similar to the zeolitic imidazolate framework (ZIF) family of porous materials. ZnCar exhibits permanent microporosity with a surface area of 448 m(2) g(-1) , and its pores are 1D channels with 5 ? openings and a characteristic chiral shape. This compound is chemically stable in organic solvents and water. Single-crystal X-ray diffraction (XRD) showed that the ZnCar framework adapts to MeOH and H2 O guests because of the torsional flexibility of the main His-β-Ala chain, while retaining the rigidity conferred by the Zn-imidazolate chains. The conformation adopted by carnosine is driven by the H bonds formed both to other dipeptides and to the guests, permitting the observed structural transformations.","author":[{"dropping-particle":"","family":"Katsoulidis","given":"Alexandros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Park","given":"Kyo Sung","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"Dmytro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Martí-Gastaldo","given":"Carlos","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Miller","given":"Gary J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"John E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Robertson","given":"Craig M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Blanc","given":"Frédéric","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"George R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"Neil G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Purton","given":"John A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Adams","given":"Dave J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"Matthew J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2014","1","3"]]},"page":"193-198","title":"Guest-Adaptable and Water-Stable Peptide-Based Porous Materials by Imidazolate Side Chain Control","type":"article-journal","volume":"53"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1038/nchem.1871","ISBN":"1755-4349","ISSN":"1755-4330","PMID":"24651203","abstract":"Porous materials are attractive for separation and catalysis-these applications rely on selective interactions between host materials and guests. In metal-organic frameworks (MOFs), these interactions can be controlled through a flexible structural response to the presence of guests. Here we report a MOF that consists of glycyl-serine dipeptides coordinated to metal centres, and has a structure that evolves from a solvated porous state to a desolvated non-porous state as a result of ordered cooperative, displacive and conformational changes of the peptide. This behaviour is driven by hydrogen bonding that involves the side-chain hydroxyl groups of the serine. A similar cooperative closure (reminiscent of the folding of proteins) is also displayed with multipeptide solid solutions. For these, the combination of different sequences of amino acids controls the framework's response to the presence of guests in a nonlinear way. This functional control can be compared to the effect of single-point mutations in proteins, in which exchange of single amino acids can radically alter structure and function.","author":[{"dropping-particle":"","family":"Martí-Gastaldo","given":"C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"D.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Warren","given":"J. E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Briggs","given":"M. E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chater","given":"P. A.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"V.","family":"Wiper","given":"P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Miller","given":"G. J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Khimyak","given":"Y. Z.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"G. R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"N. G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"M. J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature Chemistry","id":"ITEM-2","issue":"4","issued":{"date-parts":[["2014","4","23"]]},"page":"343-351","title":"Side-chain control of porosity closure in single- and multiple-peptide-based porous materials by cooperative folding","type":"article-journal","volume":"6"},"uris":[""]},{"id":"ITEM-3","itemData":{"DOI":"10.1038/s41586-018-0820-9","ISSN":"0028-0836","abstract":"Metal–organic frameworks (MOFs) are crystalline synthetic porous materials formed by binding organic linkers to metal nodes: they can be either rigid1,2 or flexible3. Zeolites and rigid MOFs have widespread applications in sorption, separation and catalysis that arise from their ability to control the arrangement and chemistry of guest molecules in their pores via the shape and functionality of their internal surface, defined by their chemistry and structure4,5. Their structures correspond to an energy landscape with a single, albeit highly functional, energy minimum. By contrast, proteins function by navigating between multiple metastable structures using bond rotations of the polypeptide6,7, where each structure lies in one of the minima of a conformational energy landscape and can be selected according to the chemistry of the molecules that interact with the protein. These structural changes are realized through the mechanisms of conformational selection (where a higher-energy minimum characteristic of the protein is stabilized by small-molecule binding) and induced fit (where a small molecule imposes a structure on the protein that is not a minimum in the absence of that molecule)8. Here we show that rotation about covalent bonds in a peptide linker can change a flexible MOF to afford nine distinct crystal structures, revealing a conformational energy landscape that is characterized by multiple structural minima. The uptake of small-molecule guests by the MOF can be chemically triggered by inducing peptide conformational change. This change transforms the material from a minimum on the landscape that is inactive for guest sorption to an active one. Chemical control of the conformation of a flexible organic linker offers a route to modifying the pore geometry and internal surface chemistry and thus the function of open-framework materials.","author":[{"dropping-particle":"","family":"Katsoulidis","given":"Alexandros P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Antypov","given":"Dmytro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Whitehead","given":"George F. S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Carrington","given":"Elliot J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Adams","given":"Dave J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Berry","given":"Neil G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Darling","given":"George R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Dyer","given":"Matthew S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"Matthew J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature","id":"ITEM-3","issue":"7738","issued":{"date-parts":[["2019","1","9"]]},"page":"213-217","publisher":"Springer US","title":"Chemical control of structure and guest uptake by a conformationally mobile porous material","type":"article-journal","volume":"565"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>16–18</sup>","plainTextFormattedCitation":"16–18","previouslyFormattedCitation":"<sup>16–18</sup>"},"properties":{"noteIndex":0},"schema":""}16–18 The CSA linker’s flexibility, expressed through torsional changes, appears to be highly important to the mechanism because the same flexible behaviour was not observed in MOF-123, a framework displaying the same SBU and topology but with a symmetric rigid linker capable of just the knee cap motion. Instead, MOF-123 displays a completely different behaviour involving a transition from its non-interpenetrated structure to a two-fold interpenetrated structure (MOF-246) upon heating. This occurs at high temperatures (270 °C) capable of removing the coordinated DMF molecules in addition to the guest DMF molecules in the pores, which is outside the scope of the current study on ZnCSA.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1002/anie.201202925","ISSN":"14337851","author":[{"dropping-particle":"","family":"Choi","given":"Sang Beom","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Furukawa","given":"Hiroyasu","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Nam","given":"Hye Jin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jung","given":"Duk-young","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jhon","given":"Young Ho","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Walton","given":"Allan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Book","given":"David","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kim","given":"Jaheon","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Angewandte Chemie International Edition","id":"ITEM-1","issue":"35","issued":{"date-parts":[["2012","8","27"]]},"page":"8791-8795","title":"Reversible Interpenetration in a Metal-Organic Framework Triggered by Ligand Removal and Addition","type":"article-journal","volume":"51"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>22</sup>","plainTextFormattedCitation":"22","previouslyFormattedCitation":"<sup>22</sup>"},"properties":{"noteIndex":0},"schema":""}22 Furthermore, the previously reported indium-based CSA MOF,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/C7FD00085E","ISSN":"1359-6640","PMID":"28612863","abstract":"Two new amide functionalised metal–organic frameworks, In(OH)CSA and In(OH)PDG, were synthesized using two flexible linkers, N -(4-carboxyphenyl)succinamic acid (CSA) and N , N ′-(1,4-phenylenedicarbonyl)diglycine (PDG), respectively. Both structures consist of corner-sharing {InO 4 (OH) 2 } octahedra in the form of trans indium hydroxide chains, which are interconnected by the dicarboxylate linkers to form stacked 2-dimensional layers. The different symmetries and configurations of the flexible and rigid features on the linkers results in different supramolecular interactions dominating between linkers, resulting in different shaped pores and functional group orientation. In(OH)CSA lacks hydrogen bonding between linkers, which results in close packing between the layers and very small solvent accessible pores running perpendicular to the plane of the layers. In(OH)PDG exhibits strong intra- and interlayer hydrogen bonding, which prevents the layers from close packing and results in larger cylindrical pores running parallel to the indium hydroxide chains, producing a total accessible volume of 25% of the unit cell volume.","author":[{"dropping-particle":"","family":"Haddad","given":"J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Whitehead","given":"G. F. S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Katsoulidis","given":"A. P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosseinsky","given":"M. J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Faraday Discussions","id":"ITEM-1","issued":{"date-parts":[["2017"]]},"page":"327-335","publisher":"Royal Society of Chemistry","title":"In-MOFs based on amide functionalised flexible linkers","type":"article-journal","volume":"201"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>19</sup>","plainTextFormattedCitation":"19","previouslyFormattedCitation":"<sup>19</sup>"},"properties":{"noteIndex":0},"schema":""}19 displaying a 2D sheet structure built from similar inversion related anti-parallel double linker connections, also does not show the hinge motion observed in ZnCSA. This illustrates the intimate relationship between the linker flexibility and the 3D topology of the MOF, i.e., the presence of both single connections and a flexible linker, in forming this dynamical responsive structure. The overall response of ZnCSA during removal of its guest species therefore arises from the combined effects of the inherent nature of the CSA linker, which is both asymmetric and flexible, and the topology of the MOF. The double connection of anti-parallel oriented asymmetric linkers, i.e., rigid and flexible ends pointing in opposite directions, provides structural stability, impeding distortions in the a and b directions, while the randomly oriented single c axis linker remains free and therefore governs the flexible response. The mechanism would be classified as “2D rigid breathing” according to the review by Murdock et al..ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1016/r.2013.09.006","ISBN":"0010-8545","ISSN":"00108545","abstract":"A subclass of metal-organic frameworks (MOFs) undergoes a \"breathing\" phenomenon which is a reversible flexing of the framework as a function of adsorbed guest. The existence of multiple stable states in a single framework leads to the modification of the shape and size of the pores as a function of guest, and has encouraged researchers to consider innovative applications for these flexible materials. To assist in improving the designs of these breathing MOFs, a thorough understanding of the different mechanisms by which frameworks breathe, must be established. This review will classify current breathing mechanisms by designating which part of the framework will not flex. By classifying the dimensional rigidity within a framework, a more in-depth discussion of each breathing mechanism can be established. Additionally, strategies for the design and synthesis of the next generation of breathing MOFs based on the reviewed research are discussed. ?? 2013 Elsevier B.V.","author":[{"dropping-particle":"","family":"Murdock","given":"Christopher R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hughes","given":"Brianna C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lu","given":"Zheng","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jenkins","given":"David M.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Coordination Chemistry Reviews","id":"ITEM-1","issue":"1","issued":{"date-parts":[["2014"]]},"page":"119-136","publisher":"Elsevier B.V.","title":"Approaches for synthesizing breathing MOFs by exploiting dimensional rigidity","type":"article-journal","volume":"258-259"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>75</sup>","plainTextFormattedCitation":"75","previouslyFormattedCitation":"<sup>75</sup>"},"properties":{"noteIndex":0},"schema":""}75 MOFs displaying this mode tend to consist of rigid 2D sheets made flexible by additional linkers running in the flexible direction.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/c2dt11992g","ISBN":"1477-9226","ISSN":"1477-9226","PMID":"22278089","abstract":"A new flexible ultramicroporous solid, La(H(5)DTMP)·7H(2)O (1), has been crystallized at room temperature using the tetraphosphonic acid H(8)DTMP, hexamethylenediamine-N,N,N',N'-tetrakis(methylenephosphonic acid). Its crystal structure, solved by synchrotron powder X-ray diffraction, is characterised by a 3D pillared open-framework containing 1D channels filled with water. Upon dehydration, a new related crystalline phase, La(H(5)DTMP) (2) is formed. Partial rehydration of 2 led to La(H(5)DTMP)·2H(2)O (3). These new phases contain highly corrugated layers showing different degrees of conformational flexibility of the long organic chain. The combination of the structural study and the gas adsorption characterization (N(2) and CO(2)) suggests an ultramicroporous flexible framework. NO isotherms are indicative of a strong irreversible adsorption of NO within the pores. Impedance data indicates that 1 is a proton-conductor with a conductivity of 8 × 10(-3) S cm(-1) at 297 K and 98% of relative humidity, and an activation energy of 0.25 eV.","author":[{"dropping-particle":"","family":"Colodrero","given":"Rosario M. P.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Olivera-Pastor","given":"Pascual","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Losilla","given":"Enrique R.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Aranda","given":"Miguel a. G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Leon-Reina","given":"Laura","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Papadaki","given":"Maria","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"McKinlay","given":"Alistair C.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Morris","given":"Russell E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Demadis","given":"Konstantinos D.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Cabeza","given":"Aurelio","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Dalton Transactions","id":"ITEM-1","issue":"14","issued":{"date-parts":[["2012"]]},"page":"4045","title":"Multifunctional lanthanum tetraphosphonates: Flexible, ultramicroporous and proton-conducting hybrid frameworks","type":"article-journal","volume":"41"},"uris":[""]},{"id":"ITEM-2","itemData":{"DOI":"10.1039/c2ce25715g","ISSN":"1466-8033","author":[{"dropping-particle":"","family":"Li","given":"Xiao-Ling","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Liu","given":"Guang-Zhen","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xin","given":"Ling-Yun","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Li-Ya","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"CrystEngComm","id":"ITEM-2","issue":"18","issued":{"date-parts":[["2012"]]},"page":"5757","title":"A novel metal–organic framework displaying reversibly shrinking and expanding pore modulation","type":"article-journal","volume":"14"},"uris":["","",""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/cg301760a","ISSN":"15287483","abstract":"A novel 2-fold interpenetrated, pillared, cadmium metal?organic framework, namely, [Cd(HBTC)-BPE]n·nDMF, has been synthesized using 1,3,5-benzene tricarboxylic acid and 1,2-bis(4-pyridyl)ethane (BPE). Thiscompound has been desolvated and subjected to various liquids and gases for sorption studies. Structures of the as-synthesized (1), desolvated (2), and resolvated in benzene (3) have been determined by single-crystal X-ray diffraction analysis and further characterized by elemental analysis, IR spectra, and thermogravimetric/differential scanning calorimetry analysis. Single crystal X-ray analysis revealed a 2-fold interpenetrated, three-dimensional (3D) framework which exhibits a 3,5-connected network with the Schlafl? i symbol of [(63)(69.8) and hms topology. Compound 1 exhibits a temperature-induced single-crystal-to-single-crystal (SC?SC) transformation upon the release of N,N′-dimethylformamide molecules forming compound 2 (stable up to 300 \\,^{\\circ}C). SC?SC transformation is also observed when it is immersed in benzene, chloroform, 1,4- dioxane, and tetrahydrofuran. The uptake of different solvent molecules was analyzed, and desolvated samples selectively adsorb benzene, chloroform, 1,4-dioxane, and THF molecules over other selected polar solvents. Gas (N2, CO2, and N2O) sorption experiments were also performed and the structure showed 2.5% N2, 4.5% CO2, and 3.4% N2O absorption by mass at room temperature and moderate gas pressures (~10 bar).","author":[{"dropping-particle":"","family":"Husain","given":"Ahmad","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ellwart","given":"Mario","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bourne","given":"Susan a.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"?hrstr?m","given":"Lars","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Oliver","given":"Clive L.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Crystal Growth and Design","id":"ITEM-3","issue":"4","issued":{"date-parts":[["2013"]]},"page":"1526-1534","title":"Single-crystal-to-single-crystal transformation of a novel 2-fold interpenetrated cadmium-organic framework with trimesate and 1,2-bis(4-pyridyl)ethane into the thermally desolvated form which exhibits liquid and gas sorption properties","type":"article-journal","volume":"13"},"uris":["","",""]},{"id":"ITEM-4","itemData":{"DOI":"10.1002/chem.201203267","ISBN":"1521-3765 (Electronic)\\r0947-6539 (Linking)","ISSN":"09476539","PMID":"23303617","abstract":"Solvothermal reaction of Zn(NO(3))(2).4H(2)O, 1,4-bis[2-(4-pyridyl)ethenyl]benzene (bpeb) and 4,4'-oxybisbenzoic acid (H(2)obc) in the presence of dimethylacetamide (DMA) as one of the solvents yielded a threefold interpenetrated pillared-layer porous coordination polymer with pcu topology, [Zn(2)(bpeb)(obc)(2)].5H(2)O (1), which comprised an unusual isomer of the well-known paddle-wheel building block and the trans-trans-trans isomer of the bpeb pillar ligand. When dimethylformamide (DMF) was used instead of DMA, a supramolecular isomer [Zn(2)(bpeb)(obc)(2)].2DMF.H(2)O (2), with the trans-cis-trans isomer of the bpeb ligand with a slightly different variation of the paddle-wheel repeating unit, was isolated. In MeOH, single crystals of 2 were transformed by solvent exchange in a single-crystal-to-single-crystal (SCSC) manner to yield [Zn(2)(bpeb)(obc)(2)].2H(2)O (3), which is a polymorph of 1. SCSC conversion of 3 to 2 was achieved by soaking 3 in DMF. Compounds 1 and 2 as well as 2 and 3 are supramolecular isomers.","author":[{"dropping-particle":"","family":"Park","given":"In-Hyeok","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lee","given":"Shim Sung","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vittal","given":"Jagadese J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemistry - A European Journal","id":"ITEM-4","issue":"8","issued":{"date-parts":[["2013","2","18"]]},"page":"2695-2702","title":"Guest-Triggered Supramolecular Isomerism in a Pillared-Layer Structure with Unusual Isomers of Paddle-Wheel Secondary Building Units by Reversible Single-Crystal-to-Single-Crystal Transformation","type":"article-journal","volume":"19"},"uris":[""]},{"id":"ITEM-5","itemData":{"DOI":"10.1021/ic801794e","ISBN":"0020-1669","ISSN":"0020-1669","PMID":"19588928","abstract":"A 2D pillared bilayer coordination polymer, [Co(5-NH(2)-bdc)(bpy)(0.5)(H(2)O)] x 2 H(2)O (1; 5-NH(2)-bdc = 5-aminoisophthalate; bpy = 4,4'-bipyridine) has been hydrothermally synthesized and shows a novel microporous host framework with 1D channels and high thermal stability (approximately 400 degrees C). The framework of 1 exhibits reversible single-crystal-to-single-crystal transformations upon removing and rebinding the coordinated waters as well as replacing them with MeOH and EtOH from the solvent. X-ray crystallography reveals that the coordination geometry of Co(II) changes from octahedron to square pyramid, as well as the shrinkage/expansion of pore deformation in respect to the subsequent shear motion of bpy pillars and vice versa. The dehydrated form 2 exhibits a shape recognition ability, which can accommodate linear molecules, such as MeCN and 2-propynyl alcohol, and interesting storage capabilities for oversized MeOH, EtOH, and benzene molecules, concomitant with spongelike dynamic transformation. The microcalorimetric study indicates that the crystalline state-liquid guest exchange and guest inclusion processes (1 superset MeOH or EtOH, 2 superset MeOH, EtOH or MeCN) are feasibly endothermic reactions with the values of molar enthalpy, DeltaH(theta)(m), of +21.38(96), +12.68(85), +25.92(86), +17.03(57), and +14.93(75) kJ mol(-1), respectively.","author":[{"dropping-particle":"","family":"Zeng","given":"Ming-Hua","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hu","given":"Sheng","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chen","given":"Qing","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xie","given":"Gang","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Shuai","given":"Qi","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Gao","given":"Sheng-Li","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tang","given":"Li-Yuan","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-5","issue":"15","issued":{"date-parts":[["2009","8","3"]]},"page":"7070-7079","title":"Apical Ligand Substitution, Shape Recognition, Vapor-Adsorption Phenomenon, and Microcalorimetry for a Pillared Bilayer Porous Framework That Shrinks or Expands in Crystal-to-Crystal Manners upon Change in the Cobalt(II) Coordination Environment","type":"article-journal","volume":"48"},"uris":["","",""]},{"id":"ITEM-6","itemData":{"DOI":"10.1021/ja904363b","ISBN":"0002-7863","ISSN":"0002-7863","PMID":"19681608","abstract":"The design of pore properties utilizing flexible motifs and functional groups is of importance to obtain porous coordination polymers with desirable functions. We have prepared a 3D pillared-layer coordination polymer, {[Cd(2)(pzdc)(2)L(H(2)O)(2)].5(H(2)O).(CH(3)CH(2)OH)}(n) (1, H(2)pzdc = 2,3-pyrazinedicarboxylic acid; L = 2,5-bis(2-hydroxyethoxy)-1,4-bis(4-pyridyl)benzene) showing (i) a rotatable pillar bearing ethylene glycol side chains acting as a molecular gate with locking/unlocking interactions triggered by guest inclusion between the side chains, (ii) framework flexibility with slippage of the layers, and (iii) coordinatively unsaturated metal centers as guest accessible sites through the removal of the water coligands. The framework clearly shows reversible single-crystal-to-single-crystal transformations in response to the removal and rebinding of guest molecules, the observation of these processes has provided fundamental clues to the understanding of the sorption profiles. The X-ray structures indicate that the 3D host framework is retained during the transformations, involving mainly rotation of the pillars and slippage of the layers. The structure of dried form 2, [Cd(2)(pzdc)(2)L](n), has no void volume and no water coligands. Interestingly, the adsorption isotherm of water for 2 at 298 K exhibits three distinct steps coinciding with the framework functions. Compound 2 favors the uptake of CO(2) (195 K) over N(2) (77 K) and O(2) (77 K). Above all, we report on a molecular gate with a rotational module exhibiting a locking/unlocking system which accounts for gate-opening type sorption profiles.","author":[{"dropping-particle":"","family":"Seo","given":"Joobeom","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Matsuda","given":"Ryotaro","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Sakamoto","given":"Hirotoshi","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bonneau","given":"Charlotte","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kitagawa","given":"Susumu","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-6","issue":"35","issued":{"date-parts":[["2009","9","9"]]},"page":"12792-12800","title":"A Pillared-Layer Coordination Polymer with a Rotatable Pillar Acting as a Molecular Gate for Guest Molecules","type":"article-journal","volume":"131"},"uris":[""]},{"id":"ITEM-7","itemData":{"DOI":"10.1021/ja981565m","ISSN":"00027863","author":[{"dropping-particle":"","family":"Alberti","given":"G.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Murcia-Mascarós","given":"S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vivani","given":"R.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-7","issue":"36","issued":{"date-parts":[["1998"]]},"page":"9291-9295","title":"Pillared derivatives of γ-zirconium phosphate containing nonrigid alkyl chain pillars","type":"article-journal","volume":"120"},"uris":[""]},{"id":"ITEM-8","itemData":{"DOI":"10.1039/b804302g","ISBN":"0306-0012","ISSN":"0306-0012","PMID":"19384443","abstract":"This critical review focuses on a strange behaviour of crystallized solid matter: its reversible swelling with large magnitude. This will be of interest for experts in porous solids but also for solid state chemists and physicists. Some examples, classified according to the dimensionality of the inorganic subnetwork, present the general requirements and the structural rules which govern the existence of this phenomenon. Its consequences concern specific applications related to sensors, energy savings, sustainable development and health (100 references).","author":[{"dropping-particle":"","family":"Férey","given":"Gérard","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Serre","given":"Christian","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Chemical Society reviews","id":"ITEM-8","issue":"5","issued":{"date-parts":[["2009"]]},"page":"1380-1399","title":"Large breathing effects in three-dimensional porous hybrid matter: facts, analyses, rules and consequences.","type":"article-journal","volume":"38"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>72,76–82</sup>","manualFormatting":"72,76–82","plainTextFormattedCitation":"72,76–82","previouslyFormattedCitation":"<sup>76–83</sup>"},"properties":{"noteIndex":0},"schema":""}72,76–82 However, unlike these materials, ZnCSA is built only using one type of linker and the rigidity in the ab plane is caused by the framework’s topology. Figure 8. Overview of the responses of ZnCSA. The behaviour during DMF to DMSO single crystal coordinated solvent exchange is shown in green. The behaviour during uncoordinated guest (DMF or DMSO) removal is shown in yellow. The framework also shows an interesting structural rearrangement during exchange of its DMF guest species for DMSO. This occurs because exchanging with DMSO not only replaces the solvent located in the pores, but also the terminal DMF ligands bound to the SBU. This appears to involve a complicated mechanism where DMSO molecules coordinate to completely different Zn atoms than the original DMF, triggering a rearrangement of the cluster involving changes in the coordination geometry of four out of seven Zn cations, and the reorganisation of four carboxylate groups. As these carboxylate groups belong to the highly ordered a and b axes linkers we are able to observe these changes crystallographically. The changing carboxylates are all connected to the flexible end of the linker, suggesting that the solvent exchange can be directly linked to the linker’s inherent flexibility. The process results in changes to the pore size and shape, but conserves the topology of the MOF, maintaining the single and double linker connections. Similar single crystal coordinated solvent exchanges (SCCSE) have been applied to a small number of rigid linkers MOFs,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ic502639v","ISSN":"0020-1669","abstract":"Solvent-assisted ligand incorporation (SALI) is useful for functionalizing the channels of metal–organic framework (MOF) materials such as NU-1000 that offer substitutionally labile zirconium(IV) coordination sites for nonbridging ligands. Each of the 30 or so previous examples relied upon coordination of a carboxylate ligand to achieve incorporation. Here we show that, with appropriate attention to ligand/node stoichiometry, SALI can also be achieved with phosphonate-terminated ligands. Consistent with stronger M(IV) coordination of phosphonates versus carboxylates, this change extends the pH range for retention of incorporated ligands. The difference in coordination strength can be exploited to achieve stepwise incorporation of pairs of ligands—specifically, phosphonates species followed by carboxylate species—without danger of displacement of the first ligand type by the second. Diffuse reflectance infrared Fourier-transform spectroscopy suggests that the phosphonate ligands are connected to the MOF node as RPO2(OH)? species in a moiety that leaves a base-accessible ?OH moiety on each bound phosphonate.","author":[{"dropping-particle":"","family":"Deria","given":"Pravas","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Bury","given":"Wojciech","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hod","given":"Idan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kung","given":"Chung-Wei","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Karagiaridi","given":"Olga","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Hupp","given":"Joseph T.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Farha","given":"Omar K.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-1","issue":"5","issued":{"date-parts":[["2015","3","2"]]},"page":"2185-2192","title":"MOF Functionalization via Solvent-Assisted Ligand Incorporation: Phosphonates vs Carboxylates","type":"article-journal","volume":"54"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/cg501141m","ISSN":"1528-7483","abstract":"The synthesis and characterization of {[Co9(INA)18(H2O)6]·11DMF·15H2O}∞ (Co9-INA·11DMF·15H2O) (INA- = the anion of isonicotinic acid) is reported. It exhibits a rigid 3D-porous structure with a Co9 repeating unit consisting of four [CoII 2(μ-O2CR)2(μ-H2O)] subunits (two unique) linked through bridging INA- ligands to an isolated CoII ion (half unique). The [CoII 2] dimers and the isolated CoII ion have assembled to create a trinodal (6,7,8)-coordinated network with point symbol (32.411.56.62)2(32.418.54.64)2(34.44.54.63). Gas sorption studies revealed that Co9-INA exhibits 910 m2 g-1 BET area, 4.2 mmol g-1 CO2 uptake at 273 K/1 bar, and 6.7 CO2/CH4 selectivity at zero coverage. Furthermore, Co9-INA displays capability for exchange of the guest solvent molecules by various organic molecules in a single-crystal to single-crystal fashion. Direct and alternating current magnetic susceptibility studies revealed the existence of dominant antiferromagnetic interactions between the Co2+ ions that result in a paramagnetic ST = 3/2 spin ground state value. Overall, this work emphasizes the potential of relatively simple and inexpensive polytopic ligands, such as isonicotic acid, to stabilize microporous MOFs with significant CO2 sorption capacity. ? 2014 American Chemical Society.","author":[{"dropping-particle":"","family":"Moushi","given":"Eleni E.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kourtellaris","given":"Andreas","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Spanopoulos","given":"Ioannis","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Papaefstathiou","given":"Giannis S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Trikalitis","given":"Pantelis N.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Crystal Growth & Design","id":"ITEM-2","issue":"1","issued":{"date-parts":[["2015","1","7"]]},"page":"185-193","title":"A Microporous Co 2+ Metal Organic Framework with Single-Crystal to Single-Crystal Transformation Properties and High CO 2 Uptake","type":"article-journal","volume":"15"},"uris":["","",""]},{"id":"ITEM-3","itemData":{"DOI":"10.1021/acs.cgd.6b01366","ISSN":"1528-7483","abstract":"Four similar Mn(II) metal–organic frameworks (MOFs), {[Mn2(nbtc)(H2O)2(S)]·S·0.5H2O}n [S = DMF (1), DMA (2), NMP (3), DEF (4)] (DMF = N,N′-dimethylformamide, DMA = N,N′-dimethylacetamide, NMP = N-methyl-2-pyrrolidinone, DEF = N,N′-diethylformamide), have been assembled solvothermally from the nitro and carboxyl doubly functionalized ligand 6,6′-dinitro-2,2′,4,4′-biphenyl tetracarboxylic acid (H4nbtc) and characterized by single-crystal X-ray diffraction, elemental analyses, infrared spectroscopy, thermogravimetric analyses, and powder X-ray diffraction. All MOFs exhibit unique three-dimensional double-walled open-frameworks with one-dimensional parallelogram channels and have guest/coordinated water and carbonyl solvent molecules, reasonably providing a good example of the competitive behavior of water and carbonyl solvent molecules and an excellent candidate for studying the single crystal coordinated solvent exchange transformations. Interestingly, because of the different steric hindrance of the series...","author":[{"dropping-particle":"","family":"Zhang","given":"Wen-Qian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Zhang","given":"Wen-Yan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Rui-Dong","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Ren","given":"Chun-Yan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Quan-Quan","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Fan","given":"Yan-Ping","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Liu","given":"Bin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Liu","given":"Ping","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wang","given":"Yao-Yu","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Crystal Growth & Design","id":"ITEM-3","issue":"2","issued":{"date-parts":[["2017","2","11"]]},"page":"517-526","title":"Effect of Coordinated Solvent Molecules on Metal Coordination Sphere and Solvent-Induced Transformations","type":"article-journal","volume":"17"},"uris":["","",""]},{"id":"ITEM-4","itemData":{"DOI":"10.1039/c3ta14489e","ISSN":"2050-7488","abstract":"The discovery of new methods for the post-synthesis modification of materials is essential in order to establish suitable strategies for the tuning of their properties in a rational manner. Here we present a series of single-crystal-to-single-crystal (SCSC) transformations for the flexible [Eu2(CIP)2(DMF)2(H2O)2] (UCY-8) [H3CIP = 5-(4-carboxybenzylideneamino)isophthalic acid] and rigid [Eu2(N-BDC)3(DMF)4] (EuN-BDC) (H 2N-BDC = 2-amino-1,4-benzene dicarboxylic acid) Metal-Organic Frameworks (MOFs) that involve the replacement of their coordinating solvent molecules by terminally ligating organic molecules with multiple functional groups including -OH, -SH, -NH- and -NH2 or their combinations, chelating ligands, and two different organic compounds. The capability of the flexible MOF, which contains small pores and channels (<4 ?), to exchange its coordinating solvent molecules by relatively bulky molecules (such as pyridine, 2-hydroxymethyl-phenol, etc.) is shown to be the result of its breathing capacity. Remarkably, the rigid MOF is also highly capable of replacing its coordinating solvent molecules by bulky ligands, despite its small pores (2-5 ?) and lack of structural flexibility. Interestingly, the insertion of some organic ligands into the rigid MOF results in a significant modification of its framework structure and substantial expansion of its potential void space. Not only a plethora of exchanged analogues of these MOFs have been isolated and crystallographically characterized, but also, in some cases, a tremendous enhancement of their Eu3+-based photoluminescence (PL) signals, lifetimes and quantum yields (up to ~16 times) compared to those of the pristine materials has been observed due to the replacement of terminal solvents by organic ligands being efficient sensitizers for the Eu 3+ ion. Overall this work indicates that the Single Crystal Coordinating Solvent Exchange (SCCSE) can be applied as a general post-synthetic modification method for LnMOFs and also constitutes a highly efficient strategy for the enhancement of the Ln3+-based PL. ? 2014 the Partner Organisations.","author":[{"dropping-particle":"","family":"Kyprianidou","given":"Eleni J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Lazarides","given":"Theodore","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaziannis","given":"Spyridon","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kosmidis","given":"Constantine","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Itskos","given":"Grigorios","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of Materials Chemistry A","id":"ITEM-4","issue":"15","issued":{"date-parts":[["2014"]]},"page":"5258","title":"Single crystal coordinating solvent exchange as a general method for the enhancement of the photoluminescence properties of lanthanide MOFs","type":"article-journal","volume":"2"},"uris":["","",""]},{"id":"ITEM-5","itemData":{"DOI":"10.1021/ic202635a","ISSN":"0020-1669","abstract":"In this work, for the first time, we have systematically demonstrated that solvent plays crucial roles in both controllable synthesis of metal-organic frameworks (MOFs) and their structural transformation process. With solvent as the only variable, five new MOFs with different structures have been constructed, in which one MOF undergoes solvent-induced single-crystal to single-crystal (SCSC) transformation that involves not only solvent exchange but also the cleavage and formation of coordination bonds. Particularly, a significant crystallographic change has been realized through an unprecedented three-step SCSC transformation process. Furthermore, we have demonstrated that the obtained MOF could be an excellent host for chromophores such as Alq3 for modulated luminescent properties.","author":[{"dropping-particle":"","family":"Lan","given":"Ya-Qian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Jiang","given":"Hai-Long","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Shun-Li","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Xu","given":"Qiang","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-5","issue":"14","issued":{"date-parts":[["2012","7","16"]]},"page":"7484-7491","title":"Solvent-Induced Controllable Synthesis, Single-Crystal to Single-Crystal Transformation and Encapsulation of Alq3 for Modulated Luminescence in (4,8)-Connected Metal–Organic Frameworks","type":"article-journal","volume":"51"},"uris":["","",""]},{"id":"ITEM-6","itemData":{"DOI":"10.1039/C5DT03504J","ISBN":"1477-9234 (Electronic) 1477-9226 (Linking)","ISSN":"1477-9226","PMID":"26876816","abstract":"A new approach for the fine tuning of flexibility in MOFs, involving functionalization of the secondary building unit, is presented.","author":[{"dropping-particle":"","family":"Bon","given":"Volodymyr","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kavoosi","given":"Negar","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Senkovska","given":"Irena","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Müller","given":"Philipp","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schaber","given":"Jana","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wallacher","given":"Dirk","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"T?bbens","given":"Daniel M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mueller","given":"Uwe","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaskel","given":"Stefan","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Dalton Transactions","id":"ITEM-6","issue":"10","issued":{"date-parts":[["2016"]]},"page":"4407-4415","publisher":"Royal Society of Chemistry","title":"Tuning the flexibility in MOFs by SBU functionalization","type":"article-journal","volume":"45"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>31–36</sup>","plainTextFormattedCitation":"31–36","previouslyFormattedCitation":"<sup>31–36</sup>"},"properties":{"noteIndex":0},"schema":""}31–36 but SCCSE on MOFs capable of flexible responses has to our knowledge only been reported once by Manos et al.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ic3005085","ISSN":"0020-1669","abstract":"Single-crystal-to-single-crystal (SCSC) transformations represent some of the most fascinating phenomena in chemistry. They are not only intriguing from a basic science point of view but also provide a means to modify or tune the properties of the materials via the postsynthetic introduction of suitable guest molecules or organic functional groups into their structures. Here, we describe UCY-2, a new flexible Nd3+ metal-organic framework (MOF), which exhibits a unique capability to undergo a plethora of SCSC transformations with some of them being very uncommon. These structural alterations involve the replacement of coordinating solvent molecules of UCY-2 by terminally ligating solvents and organic ligands with multiple functional groups including -OH, -SH, -NH-, and -NH2 or their combinations, chelating ligands, anions, and two different organic compounds. The SCSC coordinating solvent exchange is thus demonstrated as a powerful method for the functionalization of MOFs.","author":[{"dropping-particle":"","family":"Manos","given":"Manolis J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kyprianidou","given":"Eleni J.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Papaefstathiou","given":"Giannis S.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Tasiopoulos","given":"Anastasios J.","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Inorganic Chemistry","id":"ITEM-1","issue":"11","issued":{"date-parts":[["2012","6","4"]]},"page":"6308-6314","title":"Insertion of Functional Groups into a Nd 3+ Metal–Organic Framework via Single-Crystal-to-Single-Crystal Coordinating Solvent Exchange","type":"article-journal","volume":"51"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>24</sup>","plainTextFormattedCitation":"24","previouslyFormattedCitation":"<sup>24</sup>"},"properties":{"noteIndex":0},"schema":""}24 This work describes the structural response of a MOF built from a “semirigid” tricarboxylic linker, one consisting of rigid components connected though a “semirigid” imine (CH=N) linkage which displays limited conformational freedom, during a wide range of topotactic solvent exchanges. However, this framework does not display the large flexibility observed for ZnCSA during removal of its guests, and the flexible response of the framework does not show the cluster rearrangements seen in ZnCSA, which are enabled through the flexibility of the linker.In order to study the effect this structural rearrangement might then have on the removal of solvent, we returned to using periodic DFT calculations. While ZnCSA.DMSO showed a qualitatively similar response in the absolute changes of the single c axis linker response, the direction of distortion changed significantly. This is thought to originate from the previous reorganisation of the SBU from 1 to 3 upon DMSO coordination, and the resulting overall structural changes of the MOF. Consequently, a much more compact structure with a reduced void space is predicted in comparison to the predicted empty structure derived from ZnCSA.DMF. The tuning of MOF dynamical response by SBU functionalization has been reported for a rigid linker MOF, by Bon et al. for the MOF [Zn3(bpydc)2(HCOO2)] (JLU-Liu4), which exhibits a defined gate pressure response.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/C5DT03504J","ISBN":"1477-9234 (Electronic) 1477-9226 (Linking)","ISSN":"1477-9226","PMID":"26876816","abstract":"A new approach for the fine tuning of flexibility in MOFs, involving functionalization of the secondary building unit, is presented.","author":[{"dropping-particle":"","family":"Bon","given":"Volodymyr","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kavoosi","given":"Negar","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Senkovska","given":"Irena","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Müller","given":"Philipp","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schaber","given":"Jana","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wallacher","given":"Dirk","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"T?bbens","given":"Daniel M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mueller","given":"Uwe","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaskel","given":"Stefan","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Dalton Transactions","id":"ITEM-1","issue":"10","issued":{"date-parts":[["2016"]]},"page":"4407-4415","publisher":"Royal Society of Chemistry","title":"Tuning the flexibility in MOFs by SBU functionalization","type":"article-journal","volume":"45"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>36</sup>","plainTextFormattedCitation":"36","previouslyFormattedCitation":"<sup>36</sup>"},"properties":{"noteIndex":0},"schema":""}36 These workers conducted a systematic substitution of the monocarboxylates in the SBU via solvent exchange, replacing formic acid by acetic acid, benzoic acid or cinnamic acid and, due to steric effect of the monocarboxylates, the resulting isostructural materials displayed different dynamics during the wine-rack closing. However, this example is for a rigid linker MOF which, unlike ZnCSA, does not adapt its structure to accommodate the new SBU species. In ZnCSA, the torsions of the flexible ends of the a and b axes linkers actively respond to facilitate the coordinated solvent exchange reaction occurring at the cluster. DMF and DMSO are also comparable in size, whereas the size of the SBU terminal ligand was the major factor in the control exhibited by Bon et al.ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1039/C5DT03504J","ISBN":"1477-9234 (Electronic) 1477-9226 (Linking)","ISSN":"1477-9226","PMID":"26876816","abstract":"A new approach for the fine tuning of flexibility in MOFs, involving functionalization of the secondary building unit, is presented.","author":[{"dropping-particle":"","family":"Bon","given":"Volodymyr","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kavoosi","given":"Negar","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Senkovska","given":"Irena","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Müller","given":"Philipp","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Schaber","given":"Jana","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wallacher","given":"Dirk","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"T?bbens","given":"Daniel M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Mueller","given":"Uwe","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kaskel","given":"Stefan","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Dalton Transactions","id":"ITEM-1","issue":"10","issued":{"date-parts":[["2016"]]},"page":"4407-4415","publisher":"Royal Society of Chemistry","title":"Tuning the flexibility in MOFs by SBU functionalization","type":"article-journal","volume":"45"},"uris":[""]}],"mendeley":{"formattedCitation":"<sup>36</sup>","plainTextFormattedCitation":"36","previouslyFormattedCitation":"<sup>36</sup>"},"properties":{"noteIndex":0},"schema":""}36 The predicted changes in response of ZnCSA.DMSO compared to ZnCSA.DMF therefore illustrates the control possibilities provided by fine tuning of the SBU functionalization in flexible linker MOFs.ZnCSA demonstrates the attractive potential of flexible asymmetric linkers, composed of rigid and flexible ends, to control the guest response of MOFs, relying on the delicate balance between topology and flexibility. Further exploration of such linkers, for example with the extensive Zn-carboxylate cluster chemistryADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"DOI":"10.1021/ja010825o","ISSN":"0002-7863","PMID":"11516275","abstract":"The secondary building unit (SBU) has been identified as a useful tool in the analysis of complex metal-organic frameworks (MOFs). We illustrate its applicability to rationalizing MOF crystal structures by analysis of nine new MOFs which have been characterized by single-crystal X-ray diffraction. Tetrahedral SBUs in Zn(ADC)(2).(HTEA)(2) (MOF-31), Cd(ATC).[Cd(H(2)O)(6)](H2O)(5) (MOF-32), and Zn(2)(ATB)(H2O).(H2O)(3)(DMF)(3) (MOF-33) are linked into diamond networks, while those of Ni(2)(ATC)(H(2)O)(4).(H2O)(4) (MOF-34) have the structure of the Al network in SrAl(2). Frameworks constructed from less symmetric tetrahedral SBUs have the Ga network of CaGa(2)O(4) as illustrated by Zn(2)(ATC).(C(2)H(5)OH)(2)(H2O)(2) (MOF-35) structure. Squares and tetrahedral SBUs in Zn(2)(MTB)(H2O)(2).(DMF)(6)(H2O)(5) (MOF-36) are linked into the PtS network, which is the simplest structure type known for the assembly of these shapes. The octahedral SBUs found in Zn(2)(NDC)(3).[(HTEA)(DEF)(ClBz)](2) (MOF-37) form the most common structure for linking octahedral shapes, namely, the boron network in CaB(6). New structure types for linking triangular and trigonal prismatic SBUs are found in Zn(3)O(BTC)(2).(HTEA)(2) (MOF-38) and Zn(3)O(HBTB)(2)(H2O).(DMF)(0.5)(H2O)(3) (MOF-39). The synthesis, crystal structure, and structure analysis using the SBU approach are presented for each MOF.","author":[{"dropping-particle":"","family":"Kim","given":"Jaheon","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chen","given":"Banglin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Reineke","given":"Theresa M","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Hailian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Eddaoudi","given":"Mohamed","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Moler","given":"David B","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Journal of the American Chemical Society","id":"ITEM-1","issue":"34","issued":{"date-parts":[["2001","8"]]},"page":"8239-8247","title":"Assembly of Metal?Organic Frameworks from Large Organic and Inorganic Secondary Building Units: New Examples and Simplifying Principles for Complex Structures ?","type":"article-journal","volume":"123"},"uris":["","",""]},{"id":"ITEM-2","itemData":{"DOI":"10.1021/ar000034b","ISSN":"0001-4842","PMID":"11308306","abstract":"Secondary building units (SBUs) are molecular complexes and cluster entities in which ligand coordination modes and metal coordination environments can be utilized in the transformation of these fragments into extended porous networks using polytopic linkers (1,4-benzenedicarboxylate, 1,3,5,7-adamantanetetracarboxylate, etc.). Consideration of the geometric and chemical attributes of the SBUs and linkers leads to prediction of the framework topology, and in turn to the design and synthesis of a new class of porous materials with robust structures and high porosity.","author":[{"dropping-particle":"","family":"Eddaoudi","given":"Mohamed","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Moler","given":"David B","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Li","given":"Hailian","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Chen","given":"Banglin","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Reineke","given":"Theresa M","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"Michael","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Accounts of Chemical Research","id":"ITEM-2","issue":"4","issued":{"date-parts":[["2001","4"]]},"page":"319-330","title":"Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal?Organic Carboxylate Frameworks","type":"article-journal","volume":"34"},"uris":["","",""]},{"id":"ITEM-3","itemData":{"author":[{"dropping-particle":"","family":"Eddaoudi","given":"Mohamed","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kim","given":"Jaheon","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Rosi","given":"Nathaniel","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Vodak","given":"David","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Wachter","given":"Joseph","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Keeffe","given":"Michael O","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Kimrn","given":"Jaheon","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghil","given":"Omar M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Science","id":"ITEM-3","issue":"5554","issued":{"date-parts":[["2002"]]},"page":"469-472","title":"Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage All use subject to JSTOR Terms and Conditions Systematic Design and Functionality MOFs and Size Isoreticular in Application Storage","type":"article-journal","volume":"295"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>83–85</sup>","manualFormatting":"83–85","plainTextFormattedCitation":"83–85","previouslyFormattedCitation":"<sup>84–86</sup>"},"properties":{"noteIndex":0},"schema":""}83–85 as exemplified by the prototypical MOF-5,ADDIN CSL_CITATION {"citationItems":[{"id":"ITEM-1","itemData":{"author":[{"dropping-particle":"","family":"Li","given":"H.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Yaghi","given":"O. M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"O'Keeffe","given":"M.","non-dropping-particle":"","parse-names":false,"suffix":""},{"dropping-particle":"","family":"Eddaoudi","given":"M","non-dropping-particle":"","parse-names":false,"suffix":""}],"container-title":"Nature","id":"ITEM-1","issue":"November","issued":{"date-parts":[["1999"]]},"page":"276-279","title":"Design and synthesis of an exceptionally stable and highly","type":"article-journal","volume":"402"},"uris":["","",""]}],"mendeley":{"formattedCitation":"<sup>86</sup>","plainTextFormattedCitation":"86","previouslyFormattedCitation":"<sup>87</sup>"},"properties":{"noteIndex":0},"schema":""}86 provides interesting new directions in the design of future flexible MOFs.ConclusionsWe report the synthesis and characterisation of a new highly porous flexible MOF constructed from heptanuclear zinc carboxylate secondary building units and flexible asymmetric CSA linkers. Using a combination of crystallographic and computational approaches we are able to explore the dynamic response of this material during guest removal, which is driven by the changes in the disordered single linker connection along the c axis, while the double antiparallel linker connections along the a and b axes remain locked. The response of the CSA linker reflects its asymmetric nature: its rigid end experiences only a hinge motion, while its flexible end shows a combination of a hinge motion and conformational adjustments around its sp3 carbons. The overall anisotropic nature can also be related to the three-dimensional lattice topology within which it is embedded – the same linker in a two-dimensional structure with only one unique linker position does not respond in the same way. Further structural rearrangements of the material are also induced through DMF to DMSO solvent exchange, replacing the two terminal solvent molecules coordinated to the SBU based on a significant rearrangement of the cluster. This occurs through large conformational adjustments to the a and b axes CSA linkers, connected to the cluster through their flexible ends, resulting in coordination changes to the individual Zn atoms. The reorganisation significantly affects the size and shape of the solvent accessible volume of the material and is predicted to lead to a different structural response during guest removal, which highlights the potential of controlling the dynamic responses of flexible linker MOFs through SBU functionalization.ASSOCIATED CONTENTSupporting Information. SCXRD information. Lattice transformation used for ZnCSA.DMF (2) and ZnCSA.DMSO (3). Thermogravimetric analysis. PXRD fitting and comparison. Infrared spectroscopy. Volumetric nitrogen adsorption. Summary table of torsion angle changes. Summary table of changes between ZnCSA.DMF (1) and ZnCSA.DMSO (3). Summary table of experimental and computational structures unit cell parameters.The following files are available free of charge:Zip archive containing the computational structures in cif format. AUTHOR INFORMATIONCorresponding Author*Email: m.j.rosseinsky@liv.ac.uk (M. J. R.)Author ContributionsThe manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. NotesThe authors declare no competing financial interest.ACKNOWLEDGMENTSThis project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement number 692685). We acknowledge the HEC Materials Chemistry Consortium funded by EPSRC (EP/L000202, EP/R029431) for provision of computer time on the ARCHER UK National Supercomputing Service, and on the UK Materials and Molecular Modelling Hub, MMM Hub, which is partially funded by EPSRC (EP/P020194). We thank the Diamond Light Source for provision of beamtime on the I11 and I19 beamlines.ABBREVIATIONSCSA, N-(4-Carboxyphenyl)succinamate; DFT, density functional theory; DMF, dimethylformamide; DMSO, dimethylsulphoxide. MOF, metal-organic framework; PXRD, powder X-ray diffraction; SBU, secondary building unit.REFERENCESADDIN Mendeley Bibliography CSL_BIBLIOGRAPHY (1) Chang, Z.; Yang, D. H.; Xu, J.; Hu, T. L.; Bu, X. H. Flexible Metal-Organic Frameworks: Recent Advances and Potential Applications. Adv. Mater. 2015, 27 (36), 5432–5441. (2) Sarkisov, L.; Martin, R. L.; Haranczyk, M.; Smit, B. On the Flexibility of Metal-Organic Frameworks. J. Am. Chem. Soc. 2014, 136 (6), 2228–2231. (3) Schneemann, A.; Bon, V.; Schwedler, I.; Senkovska, I.; Kaskel, S.; Fischer, R. A. Flexible Metal–organic Frameworks. Chem. Soc. Rev. 2014, 43 (16), 6062–6096. (4) Lin, Z.-J.; Lü, J.; Hong, M.; Cao, R. Metal–Organic Frameworks Based on Flexible Ligands (FL-MOFs): Structures and Applications. Chem. Soc. Rev. 2014, 43 (16), 5867–5895.(5) Horike, S.; Shimomura, S.; Kitagawa, S. Soft Porous Crystals. Nat. Chem. 2009, 1 (9), 695–704. (6) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science. 2013, 341, 1230444. (7) Zhou, H.-C. “Joe”; Kitagawa, S. Metal–Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43 (16), 5415–5418. (8) Coudert, F.-X. Responsive Metal–Organic Frameworks and Framework Materials: Under Pressure, Taking the Heat, In the Spotlight, With Friends. Chem. Mater. 2015, 27 (6), 1905-1916(9) Karmakar, A.; Samanta, P.; Desai, A. V; Ghosh, S. K. Guest-Responsive Metal–Organic Frameworks as Scaffolds for Separation and Sensing Applications. Acc. Chem. Res. 2017, 50 (10), 2457–2469. (10) Guo, Z.; Cao, R.; Wang, X.; Li, H.; Yuan, W.; Wang, G.; Wu, H.; Li, J. A Multifunctional 3D Ferroelectric and NLO-Active Porous Metal-Organic Framework. J. Am. Chem. Soc. 2009, 131 (20), 6894–6895.(11) Kitagawa, S. Future Porous Materials. Acc. Chem. Res. 2017, 50 (3), 514–516.(12) Vanduyfhuys, L.; Rogge, S. M. J.; Wieme, J.; Vandenbrande, S.; Maurin, G.; Waroquier, M.; Van Speybroeck, V. Thermodynamic Insight into Stimuli-Responsive Behaviour of Soft Porous Crystals. Nat. Commun. 2018, 9, 204. (13) Mellot-Draznieks, C.; Serre, C.; Surblé, S.; Audebrand, N.; Férey, G. Very Large Swelling in Hybrid Frameworks: A Combined Computational and Powder Diffraction Study. J. Am. Chem. Soc. 2005, 127 (46), 16273–16278. (14) Lin, X.; Wang, S.; Hou, Y.; Ding, Z.; Dai, W.; Tu, W.; Hu, Z.; Xu, R. MOF-Derived Hierarchical Hollow Spheres Composed of Carbon-Confined Ni Nanoparticles for Efficient CO2 Methanation . Catal. Sci. Technol. 2019, 9 (3), 731–738. (15) Surblé, S.; Serre, C.; Mellot-Draznieks, C.; Millange, F.; Férey, G. A New Isoreticular Class of Metal-Organic-Frameworks with the MIL-88 Topology. Chem. Commun. 2006, No. 3, 284–286. (16) Katsoulidis, A. P.; Park, K. S.; Antypov, D.; Martí-Gastaldo, C.; Miller, G. J.; Warren, J. E.; Robertson, C. M.; Blanc, F.; Darling, G. R.; Berry, N. G.; et al. Guest-Adaptable and Water-Stable Peptide-Based Porous Materials by Imidazolate Side Chain Control. Angew. Chemie Int. Ed. 2014, 53 (1), 193–198.(17) Martí-Gastaldo, C.; Antypov, D.; Warren, J. E.; Briggs, M. E.; Chater, P. A.; Wiper, P. V.; Miller, G. J.; Khimyak, Y. Z.; Darling, G. R.; Berry, N. G.; et al. Side-Chain Control of Porosity Closure in Single- and Multiple-Peptide-Based Porous Materials by Cooperative Folding. Nat. Chem. 2014, 6 (4), 343–351. (18) Katsoulidis, A. P.; Antypov, D.; Whitehead, G. F. S.; Carrington, E. J.; Adams, D. J.; Berry, N. G.; Darling, G. R.; Dyer, M. S.; Rosseinsky, M. J. Chemical Control of Structure and Guest Uptake by a Conformationally Mobile Porous Material. Nature 2019, 565 (7738), 213–217. (19) Haddad, J.; Whitehead, G. F. S.; Katsoulidis, A. P.; Rosseinsky, M. J. In-MOFs Based on Amide Functionalised Flexible Linkers. Faraday Discuss. 2017, 201, 327–335. (20) Li, Qian-hai, Jiang, Shun-feng, C. P. Synthesis and Crystal Structure of Copper(II) Complex Constructed by Flexible Dicarboxylate Ligands. J. Synth. Cryst. 2016, No. 7, 1875.(21) Fang, Q. R.; Zhu, G. S.; Xue, M.; Zhang, Q. L.; Sun, J. Y.; Guo, X. D.; Qiu, S. L.; Xu, S. T.; Wang, P.; Wang, D. J.; et al. Microporous Metal-Organic Framework Constructed from Heptanuclear Zinc Carboxylate Secondary Building Units. Chem. - A Eur. J. 2006, 12 (14), 3754–3758.(22) Choi, S. B.; Furukawa, H.; Nam, H. J.; Jung, D.; Jhon, Y. H.; Walton, A.; Book, D.; O’Keeffe, M.; Yaghi, O. M.; Kim, J. Reversible Interpenetration in a Metal-Organic Framework Triggered by Ligand Removal and Addition. Angew. Chemie Int. Ed. 2012, 51 (35), 8791–8795.(23) Iremonger, S. S.; Vaidhyanathan, R.; Mah, R. K.; Shimizu, G. K. H. Zn7O2(RCOO)10 Clusters and Nitro Aromatic Linkers in a Porous Metal–Organic Framework. Inorg. Chem. 2013, 52 (8), 4124–4126. (24) Manos, M. J.; Kyprianidou, E. J.; Papaefstathiou, G. S.; Tasiopoulos, A. J. Insertion of Functional Groups into a Nd3+ Metal–Organic Framework via Single-Crystal-to-Single-Crystal Coordinating Solvent Exchange. Inorg. Chem. 2012, 51 (11), 6308–6314. (25) Tanabe, K. K.; Cohen, S. M. Postsynthetic Modification of Metal–Organic Frameworks—a Progress Report. Chem. Soc. Rev. 2011, 40 (2), 498–519. (26) Cohen, S. M. Postsynthetic Methods for the Functionalization of Metal-Organic Frameworks. Chem. Rev. 2012, 112 (2), 970–1000.(27) Cohen, S. M. The Postsynthetic Renaissance in Porous Solids. J. Am. Chem. Soc. 2017, 139 (8), 2855–2863. (28) Brozek, C. K.; Dincǎ, M. Cation Exchange at the Secondary Building Units of Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43 (16), 5456–5467. (29) Lin, Z. J.; Liu, T. F.; Huang, Y. B.; Jian Lü; Cao, R. A Guest-Dependent Approach to Retain Permanent Pores in Flexible Metal-Organic Frameworks by Cation Exchange. Chem. - A Eur. J. 2012, 18 (25), 7896–7902.(30) Liu, T. F.; Lü, J.; Tian, C.; Cao, M.; Lin, Z.; Cao, R. Porous Anionic, Cationic, and Neutral Metal-Carboxylate Frameworks Constructed from Flexible Tetrapodal Ligands: Syntheses, Structures, Ion-Exchanges, and Magnetic Properties. Inorg. Chem. 2011, 50 (6), 2264–2271. (31) Deria, P.; Bury, W.; Hod, I.; Kung, C.-W.; Karagiaridi, O.; Hupp, J. T.; Farha, O. K. MOF Functionalization via Solvent-Assisted Ligand Incorporation: Phosphonates vs Carboxylates. Inorg. Chem. 2015, 54 (5), 2185–2192. (32) Moushi, E. E.; Kourtellaris, A.; Spanopoulos, I.; Manos, M. J.; Papaefstathiou, G. S.; Trikalitis, P. N.; Tasiopoulos, A. J. A Microporous Co2+ Metal Organic Framework with Single-Crystal to Single-Crystal Transformation Properties and High CO2 Uptake. Cryst. Growth Des. 2015, 15 (1), 185–193. (33) Zhang, W.-Q.; Zhang, W.-Y.; Wang, R.-D.; Ren, C.-Y.; Li, Q.-Q.; Fan, Y.-P.; Liu, B.; Liu, P.; Wang, Y.-Y. Effect of Coordinated Solvent Molecules on Metal Coordination Sphere and Solvent-Induced Transformations. Cryst. Growth Des. 2017, 17 (2), 517–526. (34) Kyprianidou, E. J.; Lazarides, T.; Kaziannis, S.; Kosmidis, C.; Itskos, G.; Manos, M. J.; Tasiopoulos, A. J. Single Crystal Coordinating Solvent Exchange as a General Method for the Enhancement of the Photoluminescence Properties of Lanthanide MOFs. J. Mater. Chem. A 2014, 2 (15), 5258. (35) Lan, Y.-Q.; Jiang, H.-L.; Li, S.-L.; Xu, Q. Solvent-Induced Controllable Synthesis, Single-Crystal to Single-Crystal Transformation and Encapsulation of Alq3 for Modulated Luminescence in (4,8)-Connected Metal–Organic Frameworks. Inorg. Chem. 2012, 51 (14), 7484–7491. (36) Bon, V.; Kavoosi, N.; Senkovska, I.; Müller, P.; Schaber, J.; Wallacher, D.; T?bbens, D. M.; Mueller, U.; Kaskel, S. Tuning the Flexibility in MOFs by SBU Functionalization. Dalt. Trans. 2016, 45 (10), 4407–4415.(37) Crys Alis Pro. Oxford Diffraction.(38) Nowell, H.; Barnett, S. A.; Christensen, K. E.; Teat, S. J.; Allan, D. R. I19, the Small-Molecule Single-Crystal Diffraction Beamline at Diamond Light Source. J. Synchrotron Radiat. 2012, 19 (3), 435–441. (39) Winter, G. Xia2?: An Expert System for Macromolecular Crystallography Data Reduction. J. Appl. Crystallogr. 2010, 43 (1), 186–190. (40) Blessing, R. H. An Empirical Correction for Absorption Anisotropy. Acta Crystallogr. Sect. A Found. Crystallogr. 1995, 51 (1), 33–38.(41) Sheldrick, G. M. SHELXT – Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. Sect. A Found. Adv. 2015, 71 (1), 3–8. (42) Farrugia, L. J. WinGX and ORTEP for Windows?: An Update. J. Appl. Crystallogr. 2012, 45 (4), 849–854.(43) Sheldrick, G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71 (1), 3–8.(44) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2?: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42 (2), 339–341. (45) Spek, A. L. PLATON SQUEEZE: A Tool for the Calculation of the Disordered Solvent Contribution to the Calculated Structure Factors. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71 (1), 9–18.(46) Macrae, C. F.; Edgington, P. R.; McCabe, P.; Pidcock, E.; Shields, G. P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury?: Visualization and Analysis of Crystal Structures. J. Appl. Crystallogr. 2006, 39 (3), 453–457. (47) Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44 (6), 1272–1276. (48) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558.(49) Bl?chl, P. E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953.(50) Dion, M.; Rydberg, H.; Schr?der, E.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals Density Functional for General Geometries. Phys. Rev. Lett. 2004, 92 (24), 246401. (51) Klime?, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van Der Waals Density Functional. J. Phys. Condens. Matter 2010, 22 (2), 022201. (52) Alexandrov, E. V; Blatov, V. A.; Kochetkov, A. V; Proserpio, D. M. Underlying Nets in Three-Periodic Coordination Polymers: Topology, Taxonomy and Prediction from a Computer-Aided Analysis of the Cambridge Structural Database. CrystEngComm 2011, 13 (12), 3947. (53) O’Keeffe, M.; Yaghi, O. M. Deconstructing the Crystal Structures of Metal-Organic Frameworks and Related Materials into Their Underlying Nets. Chem. Rev. 2012, 112 (2), 675–702. (54) Hussain, M.; Tahir, M. N.; Mansoor, M. A.; Arifin, Z.; Mazhar, M. Heptanuclear Zinc Cluster for Growth of Zincite and Manganese-Doped Zincite Thin Films for Sensor Applications. Monatshefte für Chemie - Chem. Mon. 2013, 144 (3), 285–294. (55) Darensbourg, D. J.; Wildeson, J. R.; Yarbrough, J. C. Solid-State Structures of Zinc(II) Benzoate Complexes. Catalyst Precursors for the Coupling of Carbon Dioxide and Epoxides. Inorg. Chem. 2002, 41 (4), 973–980. (56) Lalioti, N.; Perlepes, S. P.; Manessi-Zoupa, E.; Raptopoulou, C. P.; Terzis, A.; Aliev, A. E.; Gerothanassis, I. P. Rare M7O2 Double Tetrahedral Core in Molecular Species: Preparation, Structure and Properties of [Zn7O2(O2CMe)10(1-Meim)2] (1-Meim = 1-Methylimidazole). Chem. Commun. 1998, 2 (15), 1513–1514.(57) Anan’ev, I. V.; Perova, E. V.; Nefedov, S. E. Pyrazolate-Bridged Binuclear Zinc Complexes Zn2(μ-Dmpz)2(Hdmpz)2(OOCR)2 (R = Me, Ph; Hdmpz = 3,5-Dimethylpyrazole). Russ. J. Inorg. Chem. 2010, 55 (1), 40–52.(58) Waheed, A.; Jones, R. A.; McCarty, J.; Yang, X. Synthesis and Structure of Zn 7 (? 4 -O) 2 (OAc) 10 (Pz) 2 (OAc = Acetate; Pz = Pyrazine). Dalt. Trans. 2004, 7 (22), 3840–3841. (59) Dittrich, D.; Tewes, H.; W?lper, C.; Bl?ser, D.; Schulz, S.; Roll, J. Preparation, Catalytical Activity and Crystal Structure of a Heptanuclear Zinc Acetate Cluster. Transit. Met. Chem. 2017, 42 (3), 237–241.(60) Attanasio, D.; Dessy, G.; Fares, V. Crystal and Molecular Structure of Deca-?-Acetato-Dioxobis(Pyridine)-Heptazinc(II) and the Electron Paramagnetic Resonance Spectrum of Its Copper-Doped Crystals. J. Chem. Soc., Dalt. Trans. 1979, 84 (1), 28–32. (61) Reger, D. L.; Debreczeni, A.; Pascui, A. E.; Smith, M. D. Heptanuclear Zinc Carboxylate Complex: New Supramolecular Building Unit and Unique Supramolecular Architecture. Polyhedron 2013, 52, 1317–1322. (62) Feazell, R. P.; Carson, C. E.; Klausmeyer, K. K. Synthesis of a Funtionalized Monomer of the Rare M7O210+ Double Tetrahedron Zn Cluster: [Zn7O2(O2C2H3)10(3-{CH2OH}C5H4N)2]. Inorg. Chem. Commun. 2007, 10 (8), 873–875. (63) Constable, E. C.; Housecroft, C. E.; Sch?nle, J.; Vujovic, S.; Zampese, J. A. Coordination Polymers with 4′-(4-(Anthracen-9-Yl)Phenyl)- and 4′-(4-(Naphthalen-1-Yl)Phenyl)-4,2′:6′,4″-Terpyridines: Mono-, Di- and Heptazinc(II) Nodes. Polyhedron 2013, 62, 260–267. (64) Suen, M.-C.; Chan, Z.-K.; Chen, J.-D.; Wang, J.-C.; Hung, C.-H. Syntheses and Structures of Three New Coordination Polymers Generated from the Flexible 1,3-Bis(4-Pyridyl)Propane Ligand and Zinc Salts. Polyhedron 2006, 25 (11), 2325–2332. (65) Granifo, J.; Garland, M. T.; Baggio, R. The Effect in the Assembly and Node Nuclearity of the Long and Rigid Character of Bis-Pyridyl exo-Bidentate Spacers When React with Zinc Acetate: Crystal Structures of the High Nuclearity Coordination Polymers [Zn7(μ4-O)2(OAc)10(3pdb)]n (3pdb?=?1,4-bis(3-pyridyl)-2,3-diaza-1,3-butadiene) and [Zn7(μ4-O)2(OAc)10(4pdb)]n (4pdb?=?1,4-bis(4-pyridyl)-2,3-diaza-1,3-butadiene). Polyhedron 2006, 25 (11), 2277–2283. (66) Ng, M. T.; Deivaraj, T. C.; J. Vittal, J. Self Assembly of Heptanuclear Zinc(II) Clusters Linked by Angular Spacer Ligands. Inorganica Chim. Acta 2003, 348, 173–178. (67) Willems, T. F.; Rycroft, C. H.; Kazi, M.; Meza, J. C.; Haranczyk, M. Algorithms and Tools for High-Throughput Geometry-Based Analysis of Crystalline Porous Materials. Microporous Mesoporous Mater. 2011, 149 (1), 134–141. (68) Bennett, T. D.; Cheetham, A. K.; Fuchs, A. H.; Coudert, F.-X. Interplay between Defects, Disorder and Flexibility in Metal-Organic Frameworks. Nat. Chem. 2016, 9 (1), 11–16. (69) Hobday, C. L.; Marshall, R. J.; Murphie, C. F.; Sotelo, J.; Richards, T.; Allan, D. R.; Düren, T.; Coudert, F.-X.; Forgan, R. S.; Morrison, C. A.; et al. A Computational and Experimental Approach Linking Disorder, High-Pressure Behavior, and Mechanical Properties in UiO Frameworks. Angew. Chemie Int. Ed. 2016, 55 (7), 2401–2405. (70) Serre, C.; Bourrelly, S.; Vimont, A.; Ramsahye, N. A.; Maurin, G.; Llewellyn, P. L.; Daturi, M.; Filinchuk, Y.; Leynaud, O.; Barnes, P.; et al. An Explanation for the Very Large Breathing Effect of a Metal–Organic Framework during CO2 Adsorption. Adv. Mater. 2007, 19 (17), 2246–2251. (71) Dybtsev, D. N.; Chun, H.; Kim, K. Rigid and Flexible: A Highly Porous Metal-Organic Framework with Unusual Guest-Dependent Dynamic Behavior. Angew. Chem. Int. Ed. Engl. 2004, 43 (38), 5033–5036.(72) Férey, G.; Serre, C. Large Breathing Effects in Three-Dimensional Porous Hybrid Matter: Facts, Analyses, Rules and Consequences. Chem. Soc. Rev. 2009, 38 (5), 1380–1399. (73) Carrington, E. J.; McAnally, C. A.; Fletcher, A. J.; Thompson, S. P.; Warren, M.; Brammer, L. Solvent-Switchable Continuous-Breathing Behaviour in a Diamondoid Metal–organic Framework and Its Influence on CO2 versus CH4 Selectivity. Nat. Chem. 2017, 9 (9), 882–889. (74) Horcajada, P.; Salles, F.; Wuttke, S.; Devic, T.; Heurtaux, D.; Maurin, G.; Vimont, A.; Daturi, M.; David, O.; Magnier, E.; et al. How Linker’s Modification Controls Swelling Properties of Highly Flexible Iron(III) Dicarboxylates MIL-88. J. Am. Chem. Soc. 2011, 133 (44), 17839–17847. (75) Murdock, C. R.; Hughes, B. C.; Lu, Z.; Jenkins, D. M. Approaches for Synthesizing Breathing MOFs by Exploiting Dimensional Rigidity. Coord. Chem. Rev. 2014, 258–259 (1), 119–136. (76) Colodrero, R. M. P.; Olivera-Pastor, P.; Losilla, E. R.; Aranda, M. a. G.; Leon-Reina, L.; Papadaki, M.; McKinlay, A. C.; Morris, R. E.; Demadis, K. D.; Cabeza, A. Multifunctional Lanthanum Tetraphosphonates: Flexible, Ultramicroporous and Proton-Conducting Hybrid Frameworks. Dalt. Trans. 2012, 41 (14), 4045.(77) Li, X.-L.; Liu, G.-Z.; Xin, L.-Y.; Wang, L.-Y. A Novel Metal–organic Framework Displaying Reversibly Shrinking and Expanding Pore Modulation. CrystEngComm 2012, 14 (18), 5757.(78) Husain, A.; Ellwart, M.; Bourne, S. a.; ?hrstr?m, L.; Oliver, C. L. Single-Crystal-to-Single-Crystal Transformation of a Novel 2-Fold Interpenetrated Cadmium-Organic Framework with Trimesate and 1,2-Bis(4-Pyridyl)Ethane into the Thermally Desolvated Form Which Exhibits Liquid and Gas Sorption Properties. Cryst. Growth Des. 2013, 13 (4), 1526–1534.(79) Park, I.-H.; Lee, S. S.; Vittal, J. J. Guest-Triggered Supramolecular Isomerism in a Pillared-Layer Structure with Unusual Isomers of Paddle-Wheel Secondary Building Units by Reversible Single-Crystal-to-Single-Crystal Transformation. Chem. - A Eur. J. 2013, 19 (8), 2695–2702.(80) Zeng, M.-H.; Hu, S.; Chen, Q.; Xie, G.; Shuai, Q.; Gao, S.-L.; Tang, L.-Y. Apical Ligand Substitution, Shape Recognition, Vapor-Adsorption Phenomenon, and Microcalorimetry for a Pillared Bilayer Porous Framework That Shrinks or Expands in Crystal-to-Crystal Manners upon Change in the Cobalt(II) Coordination Environment. Inorg. Chem. 2009, 48 (15), 7070–7079.(81) Seo, J.; Matsuda, R.; Sakamoto, H.; Bonneau, C.; Kitagawa, S. A Pillared-Layer Coordination Polymer with a Rotatable Pillar Acting as a Molecular Gate for Guest Molecules. J. Am. Chem. Soc. 2009, 131 (35), 12792–12800. (82) Alberti, G.; Murcia-Mascarós, S.; Vivani, R. Pillared Derivatives of γ-Zirconium Phosphate Containing Nonrigid Alkyl Chain Pillars. J. Am. Chem. Soc. 1998, 120 (36), 9291–9295. (83) Kim, J.; Chen, B.; Reineke, T. M.; Li, H.; Eddaoudi, M.; Moler, D. B.; O’Keeffe, M.; Yaghi, O. M. Assembly of Metal?Organic Frameworks from Large Organic and Inorganic Secondary Building Units: New Examples and Simplifying Principles for Complex Structures ?. J. Am. Chem. Soc. 2001, 123 (34), 8239–8247. (84) Eddaoudi, M.; Moler, D. B.; Li, H.; Chen, B.; Reineke, T. M.; O’Keeffe, M.; Yaghi, O. M. Modular Chemistry: Secondary Building Units as a Basis for the Design of Highly Porous and Robust Metal?Organic Carboxylate Frameworks. Acc. Chem. Res. 2001, 34 (4), 319–330. (85) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; Keeffe, M. O.; Yaghi, O. M.; Kimrn, J.; Yaghil, O. M. Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science. 2002, 295 (5554), 469–472.(86) Li, H.; Yaghi, O. M.; O’Keeffe, M.; Eddaoudi, M. Design and Synthesis of an Exceptionally Stable and Highly Porous Metal-Organic Framework. Nature 1999, 402, 276–279.TOC Graphic ................
................

In order to avoid copyright disputes, this page is only a partial summary.

Google Online Preview   Download