Mesoporous Cages in Chemically Robust MOFs Created by a Large

Nov 19, 2018 - The reduction in connectivity is found to be an effective way to ... This [2-c, 3-c] connectivity represents the least connectivity req...
0 downloads 0 Views 4MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Mesoporous Cages in Chemically Robust MOFs Created by a Large Number of Vertices with Reduced Connectivity Qi Liu, Yinyin Song, Yanhang Ma, Yi Zhou, Hengjiang Cong, Chao Wang, Jorryn Wu, Gaoli Hu, Michael O'Keeffe, and Hexiang Deng J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b11230 • Publication Date (Web): 19 Nov 2018 Downloaded from http://pubs.acs.org on November 19, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Mesoporous Cages in Chemically Robust MOFs Created by a Large Number of Vertices with Reduced Connectivity Qi Liu,†, ‡, ▲ Yinyin Song,†, ▲ Yanhang Ma,¶ Yi Zhou,† Hengjiang Cong,† Chao Wang,† Jorryn Wu,§ Gaoli Hu,† Michael O’Keeffe# and Hexiang Deng*, †, ‡ †Key

Laboratory of Biomedical Polymers-Ministry of Education, College of Chemistry and Molecular Sciences, Wuhan University, Luojiashan, Wuhan 430072, China ¶School of Physical Science and Technology, ShanghaiTech University, Shanghai 201210, China §Dana Hall School, 45 Dana Road, Wellesley, Massachusetts, 02482, USA #School of Molecular Sciences, Arizona State University, Tempe, AZ 85287, USA ‡UC

Berkeley-Wuhan University Joint Innovative Center, The Institute of Advanced Studies, Wuhan University, Luojiashan, Wuhan 430072, China

Supporting Information Placeholder ABSTRACT: We report the design and synthesis of two metal-organic frameworks (MOFs) with permanent porosity, MOF-818 and MOF919, using a small di-topic organic linker, 1H-pyrazole-4-carboxylic acid (H2PyC), 0.4 nm in length. Three mesoporous cages of unprecedented polyhedra are identified in these MOFs, wuh cage in MOF-818, yys and liu cages in MOF-919, with the diameter of 3.8, 4.9 and 6.0 nm, respectively. The di-topic H2PyC linker functions as the edge in the structure, while two types of metal-containing second building units (SBUs) function as the vertices. 28 vertices present in wuh cage; 50 in yys cage; and 70 in liu cage. Systematic analysis of these cages along with other mesoporous cages in supramolecules and MOFs constructed by di-topic linkers reveals that the extension of cage size is dictated by both the number and connectivity of the vertices. The increase in cage size is proportional to the number of vertices, while the growth rate is determined by their connectivity. The reduction in connectivity is found to be an effective way to create large cages. All three cages in this report are constructed by three-connecting (3-c) vertices and two-connecting (2-c) vertices. This [2-c, 3-c] connectivity represents the least connectivity required for the construction of cage and the most effective one for cage size expansion. The largest cage liu, exhibits a cage size to linker size ratio of 15, outstanding in supramolecules and MOFs. MOF-818 is stable in water with a wide pH range (pH = 2-12) and the wuh cage is big enough for the inclusion of biomolecules such as vitamin-B12 and insulin. INTRODUCTION Large open cages constructed by molecular building blocks are highly desirable in supramolecular and reticular chemistry.1-4 Recently, the increase in cage size from micropore (< 2 nm) to mesopore (2-50 nm) in supramolecules and metal-organic frameworks (MOFs) allows for the introduction of large guests into the cages, thus extends their application from the storage and separation of gases5,6 to catalysis,7 drug delivery8 and inclusion of biomolecules.8,9 The development of topological analysis and prediction also promotes the design and synthesis of structures with mesoporous cages.10 In general, these molecular based cages can be described either as individual zero-dimensional (0D) polyhedra in the case of supramolecules, or extended polyhedra in threedimensional networks (3D nets) in MOFs, where the basic units are edges and vertices.4,10,11 In the cages constructed by the simplest di-topic organic linkers, the linkers function as the edges, where the metal-containing SBUs function as the vertices. The increase of cage size can be achieved by extending the edge or increasing the number of vertices in the polyhedra.4,12-14 The former approach was carried out by extending the length of organic linkers,22-24 thus leads to the precise control of cage size without altering the

underlying polyhedra, while the later one introduces more metalcontaining SBUs in the cages, resulted in enrichment in the variety of polyhedra.4 The connectivity of vertices in the polyhedra, however, has received much less attention in previous studies, where systematic analysis is lack. Here, we show that the connectivity of vertices is a critical parameter for the construction of large molecular cages. We find that, when the connectivity is reduced, larger polyhedra is formed with the same number of vertices. Thus, it is possible to use even a single type of small organic linker for the construction of extremely large molecular cages. This is demonstrated here by the creation of three mesoporous cages of unprecedented polyhedra wuh, yys and liu in two new MOFs, MOF-818 and MOF-919 with spn and moo 3D nets, respectively, by a small linker 1H-pyrazole-4-carboxylic acid (H2PyC), 0.4 nm in length. The largest cage, liu, is composed of 70 vertices, with permanent porosity and an internal pore size of 6.0 nm, 15 times that of the linker, which outstands in both supramolecules and MOFs. The three new cages are studied alongside with other cages in supramolecules and 3D MOFs (Figure 1). All cages summarized here are composed of vertices linked by di-topic organic linker,4,15-20 where the length of the linker is

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

normalized to outline the impact of connectivity on cage size. For the first time, the connectivity of vertices in the polyhedra is systematically studied, and leads to the discovery of two

Page 2 of 16

general principles. (1) the size of cage is proportional to the number of vertices, when the connectivity remains the same (Figure 1A and

Figure 1. (A) Correlation between the polyhedra size with the number of vertices in polyhedra. (B) topological illustration of polyhedra classified by their connectivity in both 0D molecular cages in supramolecular chemistry and 3D extended cages in MOFs. Red dots represent polyhedra composed of 2-c and 3-c vertices, blue dots 3-c vertices, green dots 3-c and 4-c vertices, and black dots 4-c vertices, respectively. There are larger polyhedra exist with a large number of 3-c vertices. In order to provide clear comparison, these larger polyhedra are listed in SI. These results don’t alter the clear trend that larger polyhedra are achieved with less connectivity of the vertices, when the number is almost the same. Table S1, supporting information, SI). (2) the growth of cage size is dictated by the connectivity of the compositional vertices in the polyhedra (Figure 1B), where larger cages can be achieved by reducing the connectivity of the vertices in the polyhedra with the same number of vertices. The synthesis of three new cages here, wuh, yys and liu, represents a good example, where the compositional vertices exhibit the minimum connectivity [2-c, 3-c] for the construction of polyhedra. This explains the fastest growth rate of their sizes among all molecular cages synthesized so far, therefore, the combination of 3- and 2- connectivity stands as a promising method for the further extension in MOFs and supramolecules. EXPERIMENTAL SECTION Synthesis and activation of MOF-818. ZrOCl2∙8H2O (42.5 mg), Cu(NO3)2∙3H2O (124.0 mg), H2PyC (32.5 mg) and trifluoroacetic acid (120 μL) were completely dissolved in 10 mL of N, N-dimethylformamide (DMF) by ultrasonic in a 20 mL Pyrex vial. The mixture was heated at 100 °C for 10 hours to yield blue crystals after cooling down to room temperature. The as-synthesized MOF-818 crystals were immersed in DMF for 3 days, during which the solvent exchange was performed for five times using fresh DMF. This was followed by solvent

exchange with acetone for 3 days, 3 times per day. The acetone in the pores of the MOF sample was removed by supercritical CO2 to give activated MOF-818. In the supercritical CO2 process, the sample was exchanged with fresh liquid CO2 for seven times, with a duration of 15 minutes each time. Then the system was heated to 35 °C to reach a supercritical condition (typically 1200 psi) and held for 1 hour, followed by slowly release of CO2. The activated MOF-818 samples were degassed in vacuum at 150 °C for 12 hours and refilled with N2 for storage. Synthesis and activation of MOF-919. For the synthesis of MOF-919-Fe, FeCl3∙6H2O (35.5 mg), Cu(NO3)2∙3H2O (114.5 mg) and H2PyC (54.0 mg) were completely dissolved in 10 mL of DMF by ultrasonic in a 20 mL Pyrex vial. The mixture was heated at 100 °C for 12 hours to yield green crystals after cooling down to room temperature. Similarly, for the synthesis of MOF-919-Sc, ScCl3∙6H2O (76.2 mg), Cu(NO3)2∙3H2O (135.6 mg) and H2PyC (34.8 mg) were completely dissolved in 10 mL of DMF by ultrasonic in a 20 mL Pyrex vial. The mixture was heated at 100 °C for 15 hours to yield green crystals after cooling down to room temperature. For the synthesis of MOF919-Sc, AlCl3∙6H2O (43.1 mg), Cu(NO3)2∙3H2O (207.1 mg), H2PyC (39.0 mg) and trifluoroacetic acid (50 μL) were

ACS Paragon Plus Environment

Page 3 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

completely dissolved in 10 mL of DMF by ultrasonic in a 20 mL Pyrex vial. The mixture was heated at 100 °C for 10 hours to yield blue crystals after cooling down to room temperature. The activation procedures for all MOF-919-Fe, MOF-919-Sc and MOF-919-Al were identical. The as-synthesized MOF-919 crystals were immersed in DMF for 3 days, during which the solvent exchange was performed for five times using fresh DMF. This was followed by solvent exchange with ethanol for 3 days, 3 times per day. The solvent exchanged MOF-919 samples were dried in vacuum at room temperature for 2 hours and then at 150 °C for 12 hours to achieve activated MOF-919. Structural characterizations of MOFs. The powder X-ray diffraction (PXRD) was performed using both lab-based and synchrotron X-ray source. Rigaku Smartlab 9 kW X-ray diffractometer with copper target Kα1 (λ = 1.54056 Å) was used to collected the PXRD data of all these MOFs in both BraggBrentano (BB) and transmission mode. The PXRD data for Rietveld refinement were collected at beamline BL14B1 in transmission mode with λ = 1.2398 Å at Shanghai Synchrotron Radiation Facility (SSRF). Two-dimensional small angle X-ray scattering (2D-SAXS) were measured on a Rigaku 3.5 m NANOPIX system. The samples in all the X-ray diffraction/scattering measurements were directly exposed to air due to their chemical stability. Transmission electron microscope (TEM) images and electron diffraction patterns were collected on JEOL JEM-2100 Plus with LaB6 electron source. Scanning electron microscope (SEM) images were obtained by a FEI Verios 460 at 0.5 keV at various magnifications. Elemental analysis of the activated MOF samples was performed by an inductively coupled plasmaoptical emission spectrum instrument (Leeman Labs Prodigy7, USA). Thermogravimetric analysis (TGA) of MOFs were performed on a TA Instruments Q-500 series thermal gravimetric analyzer with sample held in aluminum oxide pans in a continuous nitrogen flow atmosphere. N2 adsorptions experiment of activated MOFs was measured on a Quantachrome Autosorb-1 automatic volumetric instrument. X-ray adsorption spectra (XAS) of MOFs were collected at beamline BL08U Shanghai Radiation Facility (SSRF). X-ray photoelectron spectroscopy (XPS) of MOFs on a Thermo Fisher Scientific ESCALAB250Xi X-ray Photoelectron Spectrometer. Infrared (IR) spectra of MOFs were recorded on a Therno Nicolet NEXUS IR spectroscopy. Detailed characterization conditions are provided in the SI. RESULTS AND DISCUSSION Systematical analysis on the connectivity of vertices for the construction of molecular base cages. The connectivity of vertices in polyhedra for both supramolecules and MOFs is defined by the number of edges (x) joining at the vertices within the polyhedra, and marked as x-c. This is different from the connectivity of vertices in the 3D nets in the topological analysis of MOFs, where all edges linking to the vertices are counted.10d,21 For example, in the case of MOF-5, the zinc oxide SBUs are linked to six edges in 3D nets, but only three edges within a single cubic polyhedron, thus the connectivity of this SBU as vertices in polyhedron is 3-c, instead of 6-c (Figure 1B). In the case of polyhedra constructed by multiple types of vertices, their connectivity is listed as combinations. For example, the ada cage in ZIF-23 contains two types of vertices with different connectivity,16b 2-c and 3-c, thus the connectivity of these polyhedra is represented as [2-c, 3-c] (Figure 1B). The extraction of polyhedra from 3D nets of MOFs makes them comparable to isolated polyhedra in supramolecules. The sizes of polyhedra are plotted against the number of their

compositional vertices, where the size of the edge is normalized to provide fair comparison. Figure 1A presents selected supramolecular cages and 3D extended cages in MOFs for the clear illustration of principles regarding the connectivity of vertices, while a complete plot is provided in Figure S3, SI, revealing the same trend. The actual size of these cages in their crystal structures are also provided and compared in Figure S1 to S2, SI. In the detailed analysis of the cages, there are two typical classes of materials different in connectivity but showing the same trend. One class is supramolecules linked by 4-c vertices, marked in black;4,15 the other one is zeolitic imidazolate frameworks (ZIFs), a type of MOF with zeolitic topologies and imidazolate-Zn/Co coordination bond, where 3-c vertices present, marked in blue.16,18 Clear trends are observed for both connectivity, leading to the discovery of the first principle. The size of polyhedra increases as the number of vertices grows without altering the underlying connectivity (Figure 1A). Specifically, the size of supramolecules increases in strict order from M6L12, M12L24, M24L48, M30L60 and M48L96, where the number of vertices is reflected in the subscript of M.3,25 The size of cage in ZIF also increases in the order from sod, pau, lta, to ucb in ZIF-8, ZIF-10, ZIF-71 and ZIF-412, respectively, as the number of vertices grows from 24 to 144.16,18 The cages in other MOFs with the connectivity of 4-c or 3-c also fit into the corresponding trend. The MOF cage with [3-c, 4-c] connectivity, such as cage IV in FDM-319a and moz in ZIF-100,18 also exhibits relatively large polyhedra size, as the number of vertices is 72 and 264, respectively, but negative curvature is observed at the external surface of the polyhedra. Similar trend is also observed in the MOFs with the connectivity of [2-c, 3-c], where polyhedra size increases from ada, wuh, yys, liu in ZIF-23, MOF-818, and MOF-919 (last two), respectively. There are other predicted polyhedra with [2-c, 3-c] connectivity yet to be synthesized, such as cage 2 and cage 3 in mov, cage 3 and cage 4 in moo-a.20 These polyhedra are extremely large in size and the number of vertices extends from 90 to 168. All examples above comply with the first principle. It is worth noting that, in the classes of materials above, the structure factor directing presents within the organic linkers, while the construction of large cages in ZIFs relies on the steric repulsion between the adjacent linkers in the same ring.16c,18 In contrast, the formation of wuh, yys and liu cages with [2-c, 3-c] connectivity is governed by a different structure directing factor, where neither the coordination angle within the linker nor repulsion between the adjacent linkers exist. These cages are formed strictly by the lowest symmetry principle in reticular chemistry, where the presence of two SBUs with different connectivity dictates the angles between the linkers. Detailed structure design can be found below. Another important trend is observed in between polyhedra with different connectivity. When the number of vertices is relatively the same, polyhedra with less connectivity usually exhibit larger size. This comes as the second principle. Indeed, sod cage in ZIF-816a is bigger than that of M24L48,15a both having 24 vertices in the polyhedra (Figure 1B). The difference is in their connectivity, where less connectivity 3-c presents in sod cage while 4-c for M24L48. Similarly, the polyhedra size increases gradually from M30L60, pau in ZIF-10, to wuh cage in MOF-818, all three have vertices number between 28 and 32 (Figure 1B and Table S1, SI). In general, the connectivity decreases from 4-c, 3-c to [2-c, 3-c] accordingly. Another example is the comparison between yys cage in MOF-919, lta cage in ZIF-71 and M48L96, with the number of vertices in the range of 48 to 50 (Figure 1B). The polyhedra size increases as

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the connectivity reduce from 4-c, 3-c to [2-c, 3-c] accordingly. The growth rate in polyhedra size for each connectivity can be estimated roughly by fitting the dots in Figure S3, SI. In general, reduced connectivity has higher slope, indicating faster growth in polyhedra size. [2-c, 3-c] is the minimum connectivity required for the construction of any polyhedra, also the one with the fastest growth rate for the creation of large molecule-

Page 4 of 16

based cages. When the actual cage size and linker size in the crystal structure are considered, liu cage in this study exhibits the largest ratio of cage to linker size, 15, among all cages reported so far (Figure 4I). The geometric explanation for the second

Figure 2. Structures of MOF-818, including structure of Zr-SBU (A), Cu-SBU (B) used to construct mesoporous cages, supertetrahedron (I) (C), structure of mesoporous wuh cage (D), structure of MOF-818 (E), spn topology (F) and corresponding tiling (G). principle is that less vertices are required to construct a cage when the connectivity is low. Furthermore, it is more likely to form larger angles at the vertices when the connectivity is reduced, thus leading to a larger curvature (Figure S3, SI). Following these two principles, a series of cages with [2-c, 3-c] connectivity is possible to be created including wuh, yys, liu, along with other giant cages in 3D nets of spn, moo, num, oge, rra, nul, fav, mov and moo-a.8g,13,14,20 Here, the design and synthesis of MOF-818 and MOF-919 just provide a humble example. Structure of wuh cage in MOF-818. The construction of MOF-818 is based on spn topology selected from Reticular Chemistry Structure Resource (RCSR) database.20 The 6-c nodes and 3-c nodes in the spn topology are achieved here by Zr-SBU and Cu-SBU, respectively. The Zr-SBU is formed by coordinating carboxylate in H2PyC with Zr6O8, while the CuSBU is formed by coordinating pyrazole in H2PyC with Cu, which was rare in MOFs19,22 (Figure 2A). This allows us to simulate the crystal structure of MOF-818, where the largest cage wuh presents. The formula of MOF-818 is supposed to be [Zr6(µ3-O)4(µ3-OH)4(OH)6(H2O)6][Cu3(µ3-O)(µ-PyC)3(H2O)6]2 based on the charge balance and previous reports.13,19 The wuh

cage is formed by 10 supertetrahedron (I) units sharing vertices (Figure 2F). Each supertetrahedron (I) unit is composed of four Cu-SBUs and four Zr-SBUs (Figure 2C). Within the wuh cage, both Cu-SBUs and Zr-SBUs function as the vertices to give a total number of 28 (Figure 2G). All 12 Zr-SBUs correspond to 3-c vertices in the polyhedra, while 4 Cu-SBUs function as 3-c vertices and 12 Cu-SBUs as 2-c vertices. The size of wuh cage and accurate angle at both types of vertices are given by the crystal structure of MOF-818 (Table S7, SI). The face symbol of wuh cage is [124] where four 12-member rings present (Figure 2). The crystal structure of MOF-818 is resolved by combining electron diffractions (ED) of the single crystal under TEM and refinement using PXRD pattern collected at synchrotron X-ray source. Prior to structure determination, MOF-818 samples were analyzed by SEM. The observed singular octahedra morphology of the crystals confirmed the phase purity of this MOF, with an average particle size of 400 nm (Figure S18, SI). The presence of sharp peaks in both ED and PXRD patterns reveal the high crystalline feature of this MOF (Figure S11, S12 and S21, SI). Detailed analysis of the ED pattern along [001] and [111] directions revealed the following diffractions conditions:

ACS Paragon Plus Environment

Page 5 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

hkl: h+k = 2n, h+l = 2n and k+l = 2n; 0hl: h+l = 4n; hhl: h+l = 2n; 00l: l = 4n, which narrowed down the possibility to two cubic space groups: Fd-3 and Fd-3m.23 Here, we use Fd-3m space group with a higher symmetry for the structure refinement. The Rietveld refinement of synchrotron PXRD data against the simulated pattern of MOF-818 shows excellent fitness with satisfactory residue, Rwp = 5.9 % (Figure 4C). This is further confirmed by the 2D-SAXS patterns of the same sample (Figure

4A and 4B). Rietveld refinement of the SAXS profile covering q = 0.04 to 0.4 Å-1 [q = 4π·sin(θ)/λ] also shows low residue. Based on the crystal structure resolved, the size of the supertertrahedra is 1.8 nm, while the diameter and aperture of wuh cage are 3.8 nm and 3.1 nm, respectively (3.7 nm and 3.0 nm, respectively considering the van der Waals radii of the corresponding

Figure 3. Structures of MOF-919-Sc, including structure of Sc-SBU and Cu-SBU used to construct mesoporous cages (A), supertetrahedron (II) (B), tiling (C), structure (D), topology (E) and tile (F) of mesoporous yys cage and structure (G), topology (H) and tile (I) of mesoporous liu cage. atoms) (Figure 2). The size is measured by the distance between centers of the corresponding atoms in the structure. Structure of yys and liu cages in MOF-919. The construction of MOF-919 is based on moo topology chosen from Reticular Chemistry Structure Resource (RCSR) database.10d,20 The 6-c nodes and 3-c nodes in the moo topology are achieved here by Fe/Sc/Al-SBU and Cu-SBU, respectively. The Fe/Sc/Al-SBU is formed by coordinating

carboxylate in H2PyC with Fe3O/Sc3O/Al3O, while the Cu SBU is identical to that in MOF-818 (Figure 3A). This allows us to simulate the crystal structure of MOF-919, where the large cages yys and liu polyhedra present. The formula of MOF-919 is supposed to be [M3(µ3-O)(OH)3][Cu3(µ3-O)(µ-PyC)3(H2O)6]2 (M = Sc, Al and Fe) based on the charge balance and previous reports.13,14 Each yys cage is formed by linking 20 supertertrahedra (Figure 3E), where the supertertrahedra is

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

constructed by 4 Fe/Sc/Al-SBUs and 4 Cu-SBUs (Figure 3B). Each liu cage is formed by linking 28 supertertrahedra (Figure 3H). Within the yys and liu cage, both Cu-SBUs and Fe/Sc/AlSBUs function as the vertices to give a total number of 50 and 70, respectively. All 30 Fe/Sc/Al-SBUs correspond to 2-c vertices in the polyhedra and 20 Cu-SBUs function as 3-c vertices in the yys cage, while 42 Fe/Sc/Al-SBUs and 28 Cu-SBU in liu cage act as 2-c and 3-c vertices, respectively (Figure 3F

Page 6 of 16

and 3I). The size of yys and liu cage and accurate angle at both types of vertices are given by the crystal structure of MOF-919 (Table S8, SI). The face symbol of yys cage is [1012], where twelve 10-member rings present. The face symbol of liu cage is [1012.124], where twelve 10-member rings and four 12member rings present (Figure 3). The observed singular octahedra

Figure 4. (A) 2D SAXS image of MOF-818 (λ = 1.54056 Å). Rietveld refinement of experimental SAXS data (B) and synchrotron data of MOF-818 (D), where blue circles are experimental data, red line is calculated data, black line is the difference and purple bars are Bragg positions (λ = 1.54056 Å for SAXS data and 1.2398 Å for synchrotron data). (C) N2 adsorption isotherms at 77 K for MOF-818 (insert, pore distribution based on N2 adsorption isotherms). (E) 2D SAXS image of MOF-919-Sc (λ = 1.54056 Å). Rietveld refinement of experimental SAXS data (F) and synchrotron data of MOF-919-Sc (H), where blue circles are experimental data, red line is calculated data, black line is the difference and purple bars are Bragg positions (λ = 1.54056 Å for SAXS data and 1.2398 Å for

ACS Paragon Plus Environment

Page 7 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

synchrotron data). (G) N2 adsorption isotherms at 77 K for MOF-919-Sc (insert, pore distribution based on N2 adsorption isotherms). (I) Comparison of the ratio between cage size and linker size of large cages in classical supramolecules and mesoporous MOFs. Cages created by both di-topic organic linker (marked in red) and other types (marked in blue) are listed here. If there are more than one types of cages in the structure, only the largest cage is used here for comparison. The largest dimension of the linker is used to represent the linker size to provide fair comparison. When multiple types of linkers were used, an average linker size is used where the largest dimension of each linker is weight by the linker ratios. Details are described in Table S2, SI. morphology of the crystals con-firmed the phase purity of MOF-919, with an average particle size of 4 μm (Figure S19, SI). The sharp peaks in the PXRD patterns obtained before and after activation indicate the high crystallinity of MOF-919 (Figure 4G, S8, S9 and S12 to S14, SI). MOF-919 can be formed by linking Cu-SBUs with Fe-SBUs, Al-SBU and Sc-SBUs, to generate two iso-reticular structures, MOF-919-Fe, MOF-919-Al and MOF-919-Sc, respectively. The PXRD pattern of MOF-919-Sc collected at synchrotron X-ray source exhibit interpretable diffractions beyond q = 1.76 Å-1 (Figure 4G and S15, SI). Rietveld refinement of the PXRD profile leads to the resolution of MOF-919-Sc, where the intensity and position of most diffraction peaks match well with those calculated on the basis of simulated structure with the space group of Fd-3m. Relatively high residue, however, was generated (Rwp = 23.2 %). One reason is the huge number of atoms (116000) in the large cubic unit cell [a = 116.6990(0) Å], and the large pore volume, where the influence by solvent molecules is not negligible. Another reason is the extremely low q values of the first few peaks (111, 220, 311, etc), making the geometry and optics of the diffractometer critical for the collection of low q peaks with accurate intensities. The optics of synchrotron beamline designed for wide-angle X-ray diffraction is no longer suitable to give the accurate peak intensity at low q. This is reflected in the obvious deviation of the low q peak intensity in comparison to that of the high angle peak in the Rietveld refinement of the synchrotron PXRD data here. In light of this, we collected the 2D-SAXS patterns of the same sample to access the accurate intensities of the low q peaks. Rietveld refinement of the SAXS pattern showed less deviation from the simulated pattern, especially at the low q region, and smaller residue was achieved, Rwp = 19.4 %, even with limited number of independent diffractions (Figure 4E, 4F and S17, SI). The structure of MOF-919-Sc is further validated by the presence of two pore-filling steps corresponding to yys and liu cage at high P/P0, details of which are discussed in the following section. Based on the crystal structure resolved, the size of the supertertrahedra (II) is 1.8 nm. The diameter and aperture of yys cage are 4.9 nm and 2.0 nm, respectively (4.8 nm and 1.9 nm, respectively considering the van der Waals radii of the corresponding atoms) (Figure 3). The larger liu cage has two kinds of apertures: one, 2.0 nm, is the same as the aperture in yys cage at the 10-member ring, while the size of other kind at the 12-member ring replaced by pore opening is 2.5 nm (2.4 nm considering the van der Waals radii of the corresponding atoms). The diameter of liu cage is 6.0 nm (5.9 nm considering the van der Waals radii of the corresponding atoms) (Figure 3G and 3H). The size is measured by the distance between centers of the corresponding atoms in the structure. Both structures of MOF-818 and MOF-919 are [3-c, 6-c] nets. The different topologies are resulted from the different geometries of their SBUs. The Zr-SBUs in MOF-818 has a symmetry of 3m (D3d) with a staggered pattern (Figure 2A), while the symmetry is 6m2 (D3h) for Fe/Sc/Al-SBUs in MOF919 and it is an eclipsed pattern (Figure 3A). Combining these SBUs with the triangular SBU leads inevitably to spn and moo respectively as the simplest (minimal transitivity) possible nets.8g,13,14,20

Characterizations of porosity and stability of MOFs. Prior to the gas adsorption analysis, the solvent molecules in the pores were removed by degassing. The complete removal of solvent residues was confirmed by thermal gravimetric analysis (TGA), where no weight loss was observed before 280 °C for both MOF-818 and MOF-919 (Figure S21-S23, SI). The perma

Figure 5. (A) Scheme of insulin in MOF-818. (B) Fluorescence intensity decrease of insulin in MOFs versus inclusion time. (C) PXRD of MOF-818 after inclusion of insulin and immersed in HCl and NaOH aqueous solution for 10 days. Where pH = 2, 3 and 4 solution are HCl aqueous solution, pH = 7 is pure water and pH = 10, 11 and 12 are NaOH aqueous solution, respectively. (D) Comparison in N2 adsorption of MOF-818 before and after the inclusion of insulin, marked in red and blue, respectively. (E) PXRD of MOF-919-Sc after inclusion of insulin and immersed in HCl and NaOH aqueous solution for 2 hours. Where pH = 4, 5 and 6 solution are HCl aqueous solution, pH = 7 is pure water and pH = 11 and 12 are NaOH aqueous solution, respectively. nent porosity of these MOFs in guest-free form was confirmed by N2 adsorption measurements at 77 K. The N2 adsorption isotherm of MOF-818 is type IV, which is typical for mesoporous materials, and the starting point of the second step is observed at P/P0 = 0.27. The pore distribution analysis based on N2 adsorption isotherm using quenched solid state functional theory (QSDFT) showed there are two kinds of pores with width 3.7 nm in MOF-818, which is in good agreement

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

with its crystal structure (Figure 4D). The N2 adsorption isotherm of MOF-919-Sc contains two steps with starting pressures of P/P0 = 0.29 and 0.40, which consistent with the sequential filling of yys and liu cages in MOF-919-Sc, respectively. QSDFT analysis of the N2 adsorption isotherm gives pore distribution with the sizes of 2.0, 3.9 and 6.0 nm, which is in good consistence with the crystal structure (Figure 4H). There are no hysteresis loops observed in the N2 adsorptions of MOF-818 and MOF-919. This is because that the increasing steps of the isotherms are below the limit of hysteresis of P/P0 = 0.42 for N2 at 77 K.24 The calculated Brunauer−Emmett−Teller (BET) surface areas of MOF-818 and MOF-919-Sc are 2050 and 2740 m2/g, respectively. Both MOF-818 and MOF-919 exhibits excellent chemical stabilities in air and humidity, reflected in the unaltered PXRD patterns and gas adsorption uptakes after exposure to air for a week (Figure 5C and 5E). MOF-818 is also stable in aqueous solution with various pH range, as confirmed by PXRD studies. After immersed into water, HCl aqueous solution with pH = 2 and NaOH aqueous solution with pH = 12 for 10 days, the PXRD of MOF-818 still matches well with the simulated one, indicating its high chemical stability (Figure 5C and S27, SI). The large porosity and chemical stability of these MOFs make them ideally suited for the inclusion of large molecules in water. A common drug, vitamin B12 (VB12), with the size of 2.7 nm, and a natural protein, insulin, with the size of 3.4 nm, are used in the inclusion test (Figure 5C, S38 and S41, SI). Both molecules exhibit sizes smaller than the opening of wuh cage, and they are successfully introduced into this cage by immersion of MOFs in the aqueous solution of the corresponding guest molecules. The uptake of VB12 in MOF-818 and insulin in MOF-818 and MOF-919-Sc are measured by ultraviolet-visible (UV-Vis) and fluorescent spectrophotometry, respectively (Figure S38 and S41, SI). The decrease in the fingerprint peak at 550 nm in the UV-vis spectra of the VB12 solution after immersion of MOFs confirmed the inclusion of VB12 (Fig. S38, SI). Standard curves of various concentrations of VB12 were used to prove accurate uptake of VB12 and the calculated uptake is 4.4 nmol/mg. Similar experiments were performed for the inclusion of insulin in MOF-818 and MOF-919-Sc, which showed a continuous decrease in the concentration of insulin. The calculated uptake of insulin was 2.8 nmol/mg and 3.79 nmol/mg for MOF-818 and MOF-919-Sc, respectively based on the standard curve. (Figure 5B). The inclusion of insulin was also reflected in the systematical variation of peak intensity in the PXRD pattern of MOF-818, while adhesion on crystal surface will not alter the intensity of the peaks (Figure S43, SI). N2 adsorption isotherm was performed on the MOF samples after inclusion. The uptake at relatively low P/P0 (< 0.02) remains nearly unaltered, while the uptake after pore filling decrease dramatically, confirming the inclusion of insulin into the mesoporous wuh cage (Figure 5D).

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: ××× ×××. Information on synthesis, characterization, structural determination of MOFs and biomolecules inclusion details (PDF) Crystal data for MOF-818 (CIF)

Page 8 of 16

Crystal data for MOF-919-Sc (CIF)

AUTHOR INFORMATION Corresponding Author *[email protected]

Author Contributions ▲Q.

L. and Y. S. contributed equally.

Note The authors declare no competing financial interest.

ACKNOWLEDGMENT The synchrotron XRD were performed using beamline BL14B1 and XAS were collected at beamline BL08U in Shanghai Synchrotron Radiation Facility (SSRF). We thank Dr. Kazuki Omoto, Dr. Yayoi Taniguchi, Dr. Kazuki Ito (Rigaku) for their invaluable help and discussion on SAXS. We also thank Zhuo Jiang, Yushan Wu, Jing Cao, Wei Yan, Deng Ding, Binglin Bie and Ziyi Chen from Wuhan University for their assistance. Financial support was provided by National Key Basic Research Program of China (2014CB239203), Natural Science Foundation of China (21471118, 91545205, 91622103) and General and Special Financial Grant from China Postdoctoral Science Foundation (2017M622500 and 2018T110791). Crystallographic data for the reported crystal structures have been deposited at the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk with the codes 1873015 (MOF-818) and 1873016 (MOF-919-Sc).

REFERENCES 1. Furukawa, H., Cordova, K. E., O’Keeffe, M., Yaghi, O. M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. 2. Xuan, W., Zhu, C., Liu, Y., Cui, Y. Mesoporous metal–organic framework materials. Chem. Soc. Rev. 2012, 41, 1677. 3. Cook, T. R., Stang, P. J. Recent developments in the preparation and chemistry of metallacycles and metallacages via coordination. Chem. Rev. 2015, 115, 7001. 4. Fujita, D.; Ueda, Y.; Sato, S.; Mizuno, N.; Kumasaka, T.; Fujita, M. Selfassembly of tetravalent Goldberg polyhedra from 144 small components. Nature 2016, 540, 563. 5. (a) Li, H., Eddaoudi, M., Groy, T. L., Yaghi, O. M. Establishing microporosity in open metal− organic frameworks: gas sorption isotherms for Zn (BDC)(BDC= 1, 4-benzenedicarboxylate). J. Am. Chem. Soc. 1998, 120, 8571. (b) Cho, H. S., Deng, H., Miyasaka, K., Dong, Z., Cho, M., Neimark, A. V., Kang, J. K., Yaghi, O. M., Terasaki, O. Extra adsorption and adsorbate superlattice formation in metal-organic frameworks. Nature 2015, 527, 503. 6. (a) Sato, H., Kosaka, W., Matsuda, R., Hori, A., Hijikata, Y., Belosludov, R. V., Sakaki, S., Takata, M., Kitagawa, S. Self-accelerating CO sorption in a soft nanoporous crystal. Science 2014, 343, 167. (b) Cadiau, A., Adil, K., Bhatt, P. M., Belmabkhout, Y., Eddaoudi, M. A metal-organic framework–based splitter for separating propylene from propane. Science 2016, 353, 137. (c) Cui, X., Chen, K., Xing, H., Yang, Q., Krishna, R., Bao, Z., Wu, H., Zhou, W., Dong, X., Han, Y., Li, B., Ren, Q., Zaworotko, M. J., Chen, B. Pore chemistry and size control in hybrid porous materials for acetylene capture from ethylene. Science 2016, 353, 141. (d) Liao, P. Q., Huang, N. Y., Zhang, W. X., Zhang, J. P., Chen, X. M. Controlling guest conformation for efficient purification of butadiene. Science 2017, 356, 1193.

ACS Paragon Plus Environment

Page 9 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

7. (a) Li, G., Kobayashi, H., Taylor, J. M., Ikeda, R., Kubota, Y., Kato, K., Takata, M., Yamamoto, T., Toh, S., Matsumura, S., Kitagawa, H. Hydrogen storage in Pd nanocrystals covered with a metal–organic framework. Nature Mater. 2014, 13, 802. (b) Ji, P., Song, Y., Drake, T., Veroneau, S. S., Lin, Z., Pan, X., Lin, W. Titanium (III)-oxo clusters in a metal–organic framework support single-site Co (II)-hydride catalysts for arene hydrogenation. J. Am. Chem. Soc. 2017, 140, 433. (c) Diercks, C. S., Liu, Y., Cordova, K. E., Yaghi, O. M. The role of reticular chemistry in the design of CO2 reduction catalysts. Nature Mater. 2018, 17, 301. (d) Zhang, X.; Huang, Z.; Ferrandon, M.; Yang, D.; Robison, L.; Li, P.; C Wang, T. C.; Delferro, M.; Farha, O. K. Catalytic chemoselective functionalization of methane in a metal-organic framework. Nature Catal. 2018, 1, 356. 8. (a) Horcajada, P., Serre, C., Vallet‐Regí, M., Sebban, M., Taulelle, F., Férey, G. Metal-organic frameworks as efficient materials for drug delivery. Angew. Chem. 2006, 118, 6120. (b) Dong, Z., Sun, Y., Chu, J., Zhang, X., Deng, H. Multivariate Metal-organic frameworks for dialingin the binding and programming the release of drug molecules. J. Am. Chem. Soc. 2017, 139, 14209. (c) Lykourinou, V.; Chen, Y.; Wang, X.-S.; Meng, L.; Hoang, T.; Ming, L.-J.; Musselman, R. L.; Ma, S. Immobilization of MP-11 into a mesoporous metal–organic framework, MP-11@ mesoMOF: a new platform for enzymatic catalysis. J. Am. Chem. Soc. 2011, 133, 10382. (d) Deng, H.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.; Kazumori, H.; O'Keeffe, M.; Terasaki, O.; Stoddart, J. F.; Yaghi, O. M. Large-pore apertures in a series of metalorganic frameworks. Science 2012, 336, 1018. (e) Lu, K., He, C., Lin, W. A chlorin-based nanoscale metal-organic framework for photodynamic therapy of colon cancers. J. Am. Chem. Soc. 2015, 137, 7600. (f) Lian, X., Fang, Y., Joseph, E., Wang, Q., Li, J., Banerjee, S., Lollar, C., Wang, X., Zhou, H. C. Enzyme–MOF (metal-organic framework) composites. Chem. Soc. Rev. 2017, 46, 3386. (g) Li, P.; Vermeulen, N. A.; Malliakas, C. D.; GómezGualdrón, D. A.; Howarth, A. J.; Mehdi, B. L.; Dohnalkova, A.; Browning, N. D.; O’Keeffe, M.; Farha, O. K. Bottom-up construction of a superstructure in a porous uranium-organic crystal. Science 2017, 356, 624. (h) Chen, T. T., Yi, J. T., Zhao, Y. Y., Chu, X. Biomineralized metalorganic framework nanoparticles enable intracellular delivery and endo-lysosomal release of native active proteins. J. Am. Chem. Soc. 2018, 140, 9912. 9. (a) Peng, S., Bie, B., Sun, Y., Liu, M., Cong, H., Zhou, W., Xia, Y., Tang, H., Deng, H., Zhou, X. Metal-organic frameworks for precise inclusion of single-stranded DNA and transfection in immune cells. Nature Commun. 2018, 9, 1293. (b) Riccò, R., Liang, W., Li, S., Gassensmith, J. J., Caruso, F., Doonan, C., Falcaro, P. Metal–organic frameworks for cell and virus biology: a perspective. ACS Nano 2018, 12, 13. 10. (a) Baburin, I. A., Blatov, V. A., Carlucci, L., Ciani, G., Proserpio, D. M. Interpenetrating metal-organic and inorganic 3D networks: a computer-aided systematic investigation. Part II [1]. Analysis of the Inorganic Crystal Structure Database (ICSD). J. Solid State Chem. 2005, 178, 2452. (b) O’Keeffe, M. Tetrahedral frameworks TX2 with T–X–T angle= 180°: Rationalization of the structures of MOF-500 and of MIL100 and MIL-101. Mat. Res. Bull. 2006, 41, 911. (c) 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, 675. (d) Li, M., Li, D., O’Keeffe, M., Yaghi, O. M. Topological analysis of metal–organic frameworks with polytopic linkers and/or multiple building units and the minimal transitivity principle. Chem. Rev. 2014, 114, 1343. 11. (a) Jie, K., Zhou, Y., Li, E., Li, Z., Zhao, R., Huang, F. Reversible iodine capture by nonporous pillar [6] arene crystals. J. Am. Chem. Soc. 2017, 139, 15320. (b) Cai, L. X., Li, S. C., Yan, D. N., Zhou, L. P., Guo, F., Sun, Q. F. Water-Soluble Redox-Active Cage Hosting Polyoxometalates for Selective Desulfurization Catalysis. J. Am. Chem. Soc. 2018, 140, 4869. (c) Wang, Y., Fang, H., Tranca, I., Qu, H., Wang, X., Markvoort, A. J., Tian, Z., Cao, X. Elucidation of the origin of chiral amplification in discrete molecular polyhedra. Nature Commun. 2018, 9, 488. 12. (a) Li, H., Eddaoudi, M., O'Keeffe, M., Yaghi, O. M. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature 1999, 402, 276. (b) Eddaoudi, M., Kim, J., Rosi, N., Vodak, D., Wachter, J., O'keeffe, M., Yaghi, O. M. Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage. Science 2002, 295, 469. 13. (a) Furukawa, H., Gándara, F., Zhang, Y. B., Jiang, J., Queen, W. L., Hudson, M. R., Yaghi, O. M. Water adsorption in porous metal-organic frameworks and related materials. J. Am. Chem. Soc. 2014, 136, 4369.

(b) Feng, D., Wang, K., Su, J., Liu, T. F., Park, J., Wei, Z., Bosch, M., Yakovenko, A., Zou, X., Zhou, H. C. A highly stable zeotype mesoporous zirconium metal-organic framework with ultralarge pores. Angew. Chem. Int. Ed. 2015, 54, 149. 14. (a) Férey, G., Serre, C., Mellot-Draznieks, C., Millange, F., Surblé, S., Dutour, J., Margiolaki, I. A hybrid solid with giant pores prepared by a combination of targeted chemistry, simulation, and powder diffraction. Angew. Chem. 2004, 116, 6456. (b) Park, Y. K.; Choi, S. B.; Kim, H.; Kim, K.; Won, B-H.; Choi, K.; Choi, J-S.; Ahn, W-S.; Won, N.; Kim, S.; Dong, H. J.; Choi, S-H.; Kim, G-H.; Cha, S-S.; Jhon, Y. H.; Yang, J. K.; Kim. J. Crystal structure and guest uptake of a mesoporous metal-organic framework containing cages of 3.9 and 4.7 nm in diameter. Angew. Chem. In. Ed. 2007, 119, 8378. (c) Feng, D., Liu, T. F., Su, J., Bosch, M., Wei, Z., Wan, W., Yuan, D., Chen, Y., Wang, X., Wang, K., Lian, X., Gu, Z., Park, J., Zou, X., Zhou, H. C., Stable metal-organic frameworks containing singlemolecule traps for enzyme encapsulation. Nature Commun. 2015, 6, 5979. (d) Lian, X., Chen, Y. P., Liu, T. F., Zhou, H. C. Coupling two enzymes into a tandem nanoreactor utilizing a hierarchically structured MOF. Chem. Sci. 2016, 7, 6969. 15. (a) Sun, Q. F., Iwasa, J., Ogawa, D., Ishido, Y., Sato, S., Ozeki, T., Sei, Y., Yamaguchi, K., Fujita, M. Self-assembled M24L48 polyhedra and their sharp structural switch upon subtle ligand variation. Science 2010, 1188605. (b) Harris, K.; Fujita, D.; Fujita, M. Giant hollow MnL2n spherical complexes: structure, functionalisation and applications. Chem Commun. 2013, 49, 6703. (c) Fujita, D.; Ueda, Y.; Sato, S.; Yokoyama, H.; Mizuno, N.; Kumasaka, T.; Fujita, M. Self-assembly of M30L60 icosidodecahedron. Chem 2016, 1, 91. 16. (a) Park, K. S., Ni, Z., Côté, A. P., Choi, J. Y., Huang, R., Uribe-Romo, F. J., Chae, H. K., O'keeffe, M., Yaghi, O. M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. 2006, 103, 10186. (b) Banerjee, R., Phan, A., Wang, B., Knobler, C., Furukawa, H., O'keeffe, M., Yaghi, O. M. High-throughput synthesis of zeolitic imidazolate frameworks and application to CO2 capture. Science 2008, 319, 939. (c) Yang, J.; Zhang, Y. B.; Liu, Q.; Trickett, C. A.; Gutiérrez-Puebla, E.; Monge, M. A.; Cong, H.; Aldossary, A.; Deng, H.; Yaghi, O. M. Principles of designing extra-large pore openings and cages in zeolitic imidazolate frameworks. J. Am. Chem. Soc. 2017, 139, 6448. 17. Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surblé, S.; Margiolaki, I. A chromium terephthalate-based solid with unusually large pore volumes and surface area. Science 2005, 309, 2040. 18. Wang, B.; Côté, A. P.; Furukawa, H.; O’Keeffe, M.; Yaghi, O. M. Colossal cages in zeolitic imidazolate frameworks as selective carbon dioxide reservoirs. Nature 2008, 453, 207. 19. (a) Tu, B.; Pang, Q.; Ning, E.; Yan, W.; Qi, Y.; Wu, D.; Li, Q. Heterogeneity within a mesoporous metal-organic framework with three distinct metal-containing building units. J. Am. Chem. Soc. 2015, 137, 13456. (b) Tu, B., Pang, Q., Xu, H., Li, X., Wang, Y., Ma, Z., Weng, L., Li, Q. Reversible redox activity in multicomponent metal-organic frameworks constructed from trinuclear copper pyrazolate building blocks. J. Am. Chem. Soc. 2017, 139, 7998. 20. Howarth, A. J., Li, P., Farha, O. K., O’Keeffe, M. Bottom-up design and generation of complex structures: a new twist in reticular chemistry. Cryst. Growth Des. 2018, 18, 449. 21. Tranchemontagne, D. J., Ni, Z., O'Keeffe, M., Yaghi, O. M. Reticular chemistry of metal-organic polyhedra. Angew. Chem. Int. Ed. 2008, 47, 5136. 22. (a) Hulsbergen, F. B.; ten Hoedt, R. W.; Verschoor, G. C.; Reedijk, J. Synthesis, magnetic properties, and X-ray structure of catena-µ3nitrato-O, O′, O″-[µ3-hydroxo-1-nitrato-1,2;1,3;2,3-tris(µ-pyrazolato-N, N′)-2,3-bis(pyrazole-N2)tricopper(II) monohydrate]. An unusual chain of trinuclear copper clusters. J. Chem. Soc. Dalton Trans. 1983, 539. (b) He, J., Yin, Y. G., Wu, T., Li, D., Huang, X. C. Design and solvothermal synthesis of luminescent copper (I)-pyrazolate coordination oligomer and polymer frameworks. Chem. Commun. 2006, 2845. (c) Zhang, J. P.; Kitagawa, S. Supramolecular isomerism, framework flexibility, unsaturated metal center, and porous property of Ag(I)/Cu(I) 3, 3', 5, 5'-tetrametyl-4, 4'-bipyrazolate. J. Am. Chem. Soc. 2008, 130, 907. 23. International Tables for Crystallography. Vol. A. Springer, 2005, Chapter 3.1, pp 53. 24. (a) Rouquerol, F., Sing, K. Absorption by Powders and Porous Solids: Principles; Methodology and Applications. Elsevier Science & Technology, 1999; pp 204-205. (b) Li, W., Liu, J., Zhao, D. Mesoporous

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

materials for energy conversion and storage devices. Nature Rev. Mater. 2016, 1, 16023.

TOC

ACS Paragon Plus Environment

Page 10 of 16

Page 11 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 1

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2

ACS Paragon Plus Environment

Page 12 of 16

Page 13 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Figure 3

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Revised Figure 4 in main text

ACS Paragon Plus Environment

Page 14 of 16

Page 15 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Revised Figure 5 in main text

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC figure

ACS Paragon Plus Environment

Page 16 of 16