Use of a “Shoe-Last” Solid-State Template in the Mechanochemical

Aug 1, 2018 - Use of a “Shoe-Last” Solid-State Template in the Mechanochemical ... Georgetown University , Washington , D.C. 20057 , United States...
0 downloads 0 Views 724KB Size
Subscriber access provided by NAGOYA UNIV

Communication

Use of a 'Shoe-Last' Solid-State Template in the Mechanochemical Synthesis of High-Porosity RHO-Zinc Imidazolate Ivana Brekalo, Christopher M. Kane, Amanda N. Ley, Joseph R. Ramirez, Tomislav Friscic, and K. Travis Holman J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b05471 • Publication Date (Web): 01 Aug 2018 Downloaded from http://pubs.acs.org on August 1, 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 7 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

Use of a ‘Shoe-Last’ Solid-State Template in the Mechanochemical Synthesis of High-Porosity RHO-Zinc Imidazolate Ivana Brekalo,1 Christopher M. Kane,1 Amanda N. Ley,1 Joseph R. Ramirez,1 Tomislav Friščić*,2, K. Travis Holman*,1 1

Department of Chemistry, Georgetown University, Washington, D.C. 20057, United States

2

Department of Chemistry, McGill University, Montreal. QC, H3A 0B8, Canada, United States

Supporting Information Placeholder ABSTRACT: We report the first use of a non-ionic solid

(NIS) as a template in mechanosynthesis of a metalorganic framework. Via eight intermolecular C-H···O hydrogen bonds, the macrocyclic MeMeCH2 template predictably functions as a ‘shoe-last’ for the assembly of double-eight rings in the liquid-assisted reaction of ZnO and imidazole (ImH). The resulting new form of ZnIm2 (namely xMeMeCH2@RHO-Zn16Im32) is available in multi-gram amounts, highly porous, and thermally stable.

Zeolitic metal organic frameworks (ZMOFs1,2) and zeolitic/tetrahedral/boron imidazolate frameworks (ZIFs,3,4 TIFs,5 BIFs6) form a class of MOFs that are expanded metal-organic analogs of industrially important zeolites. The imidazolate frameworks are constructed of tetrahedral metal nodes (Mn+) and angular ditopic imidazolate (RIm-) ligands, providing the same angles/connectivity as zeolites, and therefore the same potential for topological diversity, while being functionalizable. Though the role of templates in the topological control of zeolites is relatively well understood, structural design principles controlling the topologies of ZIFs are not. ZIF/ZMOF topologies are generally affected by the metal ion radius, reaction conditions, solvents/cosolvents (as putative templates7,8,9,10), counterions (for charged ZMOFs), and steric demands of the ligand(s).11,12 For example, 4,5-disubstituted imidazoles have been observed to often give the RHO13 topology11 and a series of extra-large pore diameter ZIFs have recently been achieved by a mixed ligand/steric index strategy.14 Topological diversity also poses challenges for ZIF synthesis. ZIFs made from smaller imidazolate ligands tend to be highly polymorphic. For example, the “simplest” ZIF system, ZnIm2 is one of the most polymorphic known compounds, reported so far in 14 different topological variants: zni,15 coi,7 neb,16 nog,8 cag,8 crb/BCT,8 a ten-membered ring form,9 zec,8 DFT,8 AFI,10 CAN,10 SOD,17 GIS,4 and MER4,18 (in order of increasing porosity, ε). Unfortunately, this topological diversity also makes the reproducible synthesis of phasepure forms of ZnIm2 challenging, especially for the potentially more important high porosity forms, which are

less thermodynamically stable.19 For example, GISZnIm24 and, until recently,18 MER-ZnIm2 (one of the highest theoretical porosity ZIFs, ε = 0.61) had only been isolated as individual single crystals.4 Mechanochemistry has recently emerged as an excellent means of exploring the compositional and structural/topological landscape of MOFs.20,21,22 It exhibits many advantages over traditional solution syntheses (e.g., speed, high yield, greener production, low energy input)23 and often gives phase-pure and/or previously unknown materials. In the context of ZIFs, it was shown that direct reaction of metal oxides (ZnO, CoO) with various imidazoles can produce porous ZIFs in high yields, under a variety of conditions (neat grinding (NG), liquid-assisted grinding (LAG), ion- and liquidassisted grinding (ILAG), accelerated aging).24,25,26,27,28 For example, reaction of ZnO with 2-methylimidazole (MeImH) quantitatively gives commercially viable ZIF8 (SOD-Zn(MeIm)2),24 and real-time monitoring29 of the reaction revealed a topologically novel intermediate.25 Importantly, mechanochemical ZIF syntheses are consistent with Ostwald’s rule of stages, first yielding less thermodynamically stable, high porosity phases (with appropriate pore-filling agents).19,24

Figure 1. Illustration of the MeMeCH2 cavitand as a ‘shoelast’ template for the d8r motif (the ‘shoe’) and possible ZnIm2 topological outcomes, listed in order of their d8r density (Table S8).

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

Despite these advantages, mechanochemistry has not yet solved the generally important problem of topological control of highly polymorphic MOFs. Indeed, the mechanosynthesis of ZnIm2 has yielded only lowporosity topologies or mixtures, the outcome being highly dependent on subtle differences in reaction conditions (liquid/salt additives, milling times),24 suggesting that the reaction may be easily influenced by templates. In this work we use for the first time a non-ionic solid (NIS) as a template in the mechanosynthesis of a MOF. This approach provides access to decagram quantities of the previously unknown RHO-ZnIm2 phase, which has the lowest tetrahedral density of any ZnIm2 phase and exhibits high porosity and thermal stability. Since the imidazolate C-H bonds of ZIFs are effective hydrogen-bond donors, we hypothesized that a molecule with hydrogen-bond acceptor sites arranged in complement to a desired framework topological motif may function as a sort of ‘shoe-last’ for the assembly of said motif (the ‘shoe’). We recently discovered that one of Cram’s readily accessible macrocyclic cavitands, MeMeCH2 (Figure 1), functions as an effective template for the double-8-ring (d8r) motif in solvothermal ZnIm2 syntheses.18 The reaction of Zn(NO3)2·6H2O with imidazole in the presence of MeMeCH2 invariably gave MeMeCH2@MER-Zn16Im32, wherein the d8rs are assembled around the 'shoe-last' via eight C-H···O hydrogen bonds (Figure 1). Unfortunately, the solvothermal synthesis was low-yielding, required a large excess of the template (>70-fold) and gave only the densest of the d8r-containing topologies (MER, PAU, RHO, SBE, TSC, in order of ε). To improve upon the solvothermal synthesis and/or access one of the more porous d8rcontaining phases, we sought to systematically explore the templation effect of the MeMeCH2 ‘shoe-last’ in ZnIm2 mechanosynthesis. Notably, though liquids and salts have been used as templates/catalysts in mechanochemistry,24 NISs have not yet been explored. First, nanoparticulate ZnO (nano-ZnO)28 was briefly milled with imidazole in the presence of MeMeCH2 (“NG Procedure”, SI-6.1), yielding only the known30 dense phase, Zn4(ImH)Im8 (Figure 2d). To stabilize more porous phases via pore-filling, the milled mixture of nano-ZnO, ImH, and MeMeCH2 was aged in different liquid vapors (“Aging Procedure”, SI-6.2). Aging reactions in vapors of N,N-dimethylformamide (DMF) and N,N-diethylformamide (DEF) revealed signs of slow formation of the previously unknown d8rcontaining RHO-ZnIm2 phase and were chosen for further optimization (SI-6.3-6.4). Ultimately, Aging a mixture of nano-ZnO:HIm:MeMeCH2 in a 1:2:0.5 ratio in vapors of DEF for over 3 months provided nearquantitative conversion (by thermogravimetric analysis (TGA) of washed material, SI-4.1) of the nano-ZnO to an RHO-ZnIm2 phase, with traces of Zn4(HIm)Im8 (Figure 2e). As Zn4(ImH)Im8 is the product obtained in the absence of solvent vapor, it seems clear that its pres-

ence is due to competition between the dry reaction and that promoted by DEF. Controls showed that Aging a nano-ZnO and HIm mixture in the absence of MeMeCH2 gives known dense phases (SI-5.1), and no RHO product. It follows that both the MeMeCH2 d8r ‘shoe-last’ and an appropriate liquid additive (as both pore-filling agent and mechanochemical activator) are necessary to produce the RHO-ZnIm2 phase. Seeking to decrease the reaction time and eliminate the Zn4(HIm)Im8 impurity, the liquid additive (DMF or DEF ≈ 4 molar equivalents compared to Zn) was added directly to the mechanically activated reaction mixture and left to age as a thick paste. This “Paste Procedure” (SI-4.2) resulted in a pure RHO-ZnIm2 phase (after washing, Figure 2f) within approximately 2 months. Analysis revealed the material to have a composition of 0.51MeMeCH2@RHO-Zn16Im32. It exhibited relatively high porosity (BETN2,77K = 1190 m2/g, Vpore = 0.529 cm3/g) and high thermal stability.

Figure 2. PXRD patterns of a)-b) zni-ZnIm2 (IMIDZB06), and Zn4(Im)Im8 (KUMXEW) (calc.); c) product of the LAG Procedure, without MeMeCH2, d) product of the NG Procedure, e)-h) products of the Aging (90 days), Paste (75 days), Salt Paste (19 days), and LAG (14 days) Procedures, respectively. Patterns f-h) are pure RHO-ZnIm2 phases; indexing lines for the Pm-3m, a = 28.907(2) Å unit cell of 0.90MeMeCH2@RHO-Zn16Im32 are shown. Dashed lines denote peak positions of ZnO.

To accelerate the reaction, several protic salts (30 mol%:Zn) were explored as additives during milling (SI6.5). Adding NaHSO4·H2O afforded full conversion to an RHO-topology product after 20 days. (“Salt-Paste Procedure”, SI-4.3, Figure 2g). Though PXRD analysis of the Salt-Paste product showed no bulk NaHSO4·H2O, TGA clearly showed steps consistent with salt decomposition, implying that it remains contained in the pores and isn’t washed out (Figure S27). This is reinforced by sorption analysis of the Salt-Paste product, which exhib-

ACS Paragon Plus Environment

Page 2 of 7

Page 3 of 7 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

its relatively low porosity (BETN2,77K = 565 m2/g, Vpore = 0.292 cm3/g). Finally, a no-salt LAG Procedure (SI-4.4) was explored by adding DEF directly into the milling vessel with the other reagents (ZnO:HIm:MeMeCH2:DEF=1:2:0.5:4). Milling the mixture for 2 min (30 Hz), and leaving it in a capped vial allowed essentially complete conversion (96% by TGA) to highly porous 0.90MeMeCH2@RHOZn16Im32 after only 14 days (BETN2,77K = 1650 m2/g, Vpore = 0.703cm3/g, after washing; Figures 2h, 3). In fact, conversion was nearly complete (87%) after only two days. The PXRD pattern of 0.90MeMeCH2@RHOZn16Im32 reliably indexes to a Pm-3m unit cell, a = 28.907(2) Å (SI 8.1.), confirming the RHO topology by comparison to other RHO ZIFs. Similar to the other methods (SI-5.2-5.4), applying the LAG Procedure without MeMeCH2 gave no RHO topology phases (Figure 2c), confirming the templation effect. Furthermore, conducting the final LAG procedure with 2methylimidazole instead of imidazole provided only the known SOD-ZnMeIm2 product, showing that the protic 2-position of imidazole is crucial to enabling d8r templation. (See SI 6.8.)

attractive. Though it should be possible to increase the pore volume further by removal of the template (ε = 0.54 with template), at present we have been unable to remove it by washing (CHCl3, AcMe, EtOH, PhMe) or heating. Table 1. Surface areas of RHO-topology ZIFs. Name

Composition

CoNIm (RHO) CdIF-9 CdIF-4 RHO-Zn(Im)2 SALEM-1 MAF-6 ZTIF-2 Cu/Zn-ZIF-108 ZIF-25 Na+ rho-ZMOF ZIF-71 ZIF-71(Zn,Mn) ZIF-96 ZIF-93 C9@SIM1 RHO ZIF ZIF-97 C6@SIM1 ZIF-11 ZIF-12

Co(2-NO2Im)2 Cd(2-NO2Im)2 Cd(2-EtIm)2 xMeMeCH2@Zn16(Im)32 Cd(2-MeIm)2 Zn(2-EtIm)2 Zn(2-PrIm)(MTz) Cu0.97Zn0.03(2-NO2Im)2 Zn(4,5-Me2Im)2 In(diCarbIm)2Na Zn(4,5-Cl2Im)2 Zn0.88Mn0.12(4,5-Cl2Im)2 Zn(4-NH2-5-CNIm)2 Zn(fIm)2 Zn2(fIm)1.86(Im-a)0.14 Zn(2-NO2Im)(Pur)c Zn(Im-b)2 Zn2(fIm)1.14(Im-c)0.86 Zn(BIm)2 Co(BIm)2

BET, m2/g 1858 1668 1658 1650 1385 1343 1287 1182 1110 1067b 1050 966 960 765 687 587 564 365 d e

Ref. 31 32, 33 32, 33 this work 33 34 35 36 11 37 11, 38 39 11 11, 40 40 41 11 40 4 4

BIm=benzimidazolate; diCarbIm=4,5-dicarboxyimidazolate; fIm=4-formyl-5-methylimidazolate; Im-a=4-methyl-5((hexylimino)methyl)imidazolate; Im-b=4-methyl-5hydroxymethylimidazolate; Im-c=5-((hexylimino)methyl)-4methylimidazolate; MTz=5-methyltetrazolate, Pur=purinate. a estimated (SI 7). bLangmuir. cContains exogenous zincnitroimidazolates. dnot porous to N2 (77 K), enot reported.

Figure 3. a) Dinitrogen sorption/desorption isotherm (77 K) of activated 0.90MeMeCH2@RHO-Zn16Im32 obtained by the LAG Procedure. Inset shows the indexed PXRD pattern of the post-sorption material. b) 13 g of xMeMeCH2@RHO-Zn16Im32.

Though the RHO topology has been observed for a number of M(RIm)2 compositions (Table 1), the highest surface area RHO materials are either made with highly toxic metals (Cd2+) or require impractically expensive ligands, such as 2-nitroimidazole (2-NO2ImH, ~$300/g). As the RHO ZIF topology has among the highest theoretical porosities (ε = 0.68), an economical route to highly porous, non-toxic RHO-ZnIm2 material is important. As use of the inexpensive MeMeCH2 shoe-last template in repeated mechanosyntheses gave access to >13g of xMeMeCH2@RHO-Zn16Im32 (vessel size is the only apparent limit to scale-up), and the use of nonnanoparticulate ZnO (~$110/kg) in the LAG Procedure is possible, (88% conversion after 12 days, BETN2,77K = 1083 m2/g), this mechanochemical approach is clearly

Serendipitously, exactly one single crystal of MeMeCH2@RHO-Zn16Im32·solvent was found in one of the many solution preparations of [email protected] X-ray analysis (100 K) revealed a cubic Pm3m space group with a unit cell of 28.672(2) Å (Figure 4), matching the bulk xMeMeCH2@RHO-Zn16Im32 material obtained by mechanosynthesis. The ZnIm2 framework adopts the RHO zeolitic topology, consisting of alternating lta cages and d8rs such that every imidazolate ligand is part of a d8r. The d8rs are constructed of imidazolates that build the 8-rings (‘8r imidazolates’) and those that constitute the ‘struts’ of the d8rs. Almost identical to the previously reported structure of MeMeCH2@MER-Zn16Im32, every cavitand is nestled within a d8r, bound by eight C-H...O hydrogen bonds (C(H)...O = 3.24(1) Å) originating from the 2positions of the strut imidazolates. Because of the symmetry and steric requirements of the MeMeCH2, each d8r must locally adopt 4mm symmetry: all eight imidazolates of one 8r, but only every other imidazolate of the other 8r, are oriented approximately perpendicular to the 8r plane (Figure 4c). The mismatch in symmetry between the 4mm cavitand@d8rs and the required 4/mmm symmetry of d8rs in an isotropic RHO structure

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

requires the cavitands to be disordered over two positions (green/black, Figure 4). This is notably unlike the tetragonal I4mm MeMeCH2@MER-Zn16Im32 material,18 in which the cavitand induces bulk polarity in the crystals. Also unlike the MER material, only half of the d8rs of MeMeCH2@RHO-Zn(Im)2 are filled with cavitand, for steric reasons (see SI-8.3). As the density of d8rs in RHO vs. MER MeMeCH2@Zn16(Im)32 is doubled, however, the two materials have identical composition. Why then does the solvothermal synthesis reliably provide MeMeCH2@MER-Zn(Im)2, and mechanosynthesis always gives MeMeCH2@RHO-Zn(Im)2? We hypothesize that, since the near solventless environment of mechanosynthesis provides the highest possible concentration of the d8r template, it yields the framework with the highest density of d8rs (RHO vs. MER, Table S8). It is possible that, in the early stages of ZIF formation, the MeMeCH2 cavitands can template, yet disassociate from, some of the d8r motifs, thus promoting the structure with the highest d8r density (RHO), even if not all d8rs are occupied in the final product.

templates to guide the assembly of various structural motifs (‘shoes’) and/or provide topological control in polymorphic or topologically diverse MOF systems or molecular assemblies. ASSOCIATED CONTENT The Supporting Information is available free of charge on the ACS Publications website at DOI: XXXXXX. Details of synthetic procedures with control experiments and full characterization (NMR, TGA, PXRD, and sorption analyses). Derivation and equations for TGA conversion analyses. Properties of all reported RHO ZIFs. Crystallographic data (CIFs) for MeMeCH2@RHOZn16Im32·xsolvent, MeMeCH2·2DEF, and MeMeCH2·2DMF (also deposited with the CCDC, submission numbers 1844379-1844381, respectively).

AUTHOR INFORMATION Corresponding Author

[email protected] [email protected] Funding Sources No competing financial interests have been declared.

ACKNOWLEDGMENT This work was supported by the NSF (DMR-1610882). I. B. acknowledges the financial support from Georgetown University (Kunin Fellowship, GSAS Dissertation research Grant), the ICDD (Ludo Frevel Scholarship), and NSERC (CREATE grant). We would like to thank Dr. Krunoslav Užarevic and Dr. Ivan Halasz of InSolido Technologies, as well as Form-Tech Scientific for providing us with milling jars.

REFERENCES

Figure 4. The crystal structure of MeMeCH2@RHOZn16(Im)32·solvent, illustrating top (a) and side-on (b-c) views of the MeMeCH2@d8r ‘last@shoe’ motif. b) Cutaway showing both orientations of disordered cavitand/imidazolates. c) One of the two disordered orientations. d-e) Wireframe showing the RHO connectivity (d) and ball-and-stick representation (e) of the RHO-ZnIm2 framework, occupied by disordered cavitands (spacefill).

In conclusion, we have exploited a designer d8r NIS template and the efficiency of mechanosynthesis to obtain decagram quantities of a new, highly porous form of ZnIm2 (RHO) in near-quantitative yield. This work opens the door for the design/employ of other 'shoe-last'

(1) Eddaoudi, M.; Sava, D. F.; Eubank, J. F.; Adil, K.; Guillerm, V. Zeolite-like metal-organic frameworks (ZMOFs): design, synthesis, and properties. Chem. Soc. Rev. 2015, 44, 228. (2) Tan, Y. X.; Wang, F.; Zhang, J. Design and synthesis of multifunctional metal-organic zeolites. Chem. Soc. Rev. 2018, 47, 2130. (3) Tian, Y.-Q.; Cai, C.-X.; Ji, Y.; You, X.-Z.; Peng, S.-M.; Lee, G.-H. [Co5(im)10⋅2 MB]∞: A Metal-Organic Open-Framework with Zeolite-Like Topology. Angew. Chem. Int. Ed. 2002, 41, 1384. (4) Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R.;UribeRomo, F. J.; Chae, H. K.; O’Keeffe, M.; Yaghi, O. M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 10186. (5) Wu, T.; Bu, X.; Zhang, J.; Feng, P. New Zeolitic Imidazolate Frameworks: From Unprecedented Assembly of Cubic Clusters to Ordered Cooperative Organization of Complementary Ligands. Chem. Mater. 2008, 20, 7377. (6) Zhang, J.; Wu, T.; Zhou, C.; Chen, S.; Feng, P.; Bu, X. Zeolitic Boron Imidazolate Frameworks. Angew. Chem., Int. Ed. Engl. 2009, 48, 2542. (7) Tian, Y. Q.; Cai, C. X.; Ren, X. M.; Duan, C. Y.; Xu, Y.; Gao, S.; You, X. Z. The Silica‐Like Extended Polymorphism of Cobalt(II) Imidazolate Three‐Dimensional Frameworks: X‐ray Single‐Crystal Structures and Magnetic Properties. Chem. - Eur. J. 2003, 9, 5673. (8) Tian, Y. Q.; Zhao, Y. M.; Chen, Z. X.; Zhang, G. N.; Weng, L. H.; Zhao, D. Y. Design and Generation of Extended Zeolitic Metal–

ACS Paragon Plus Environment

Page 4 of 7

Page 5 of 7 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 Organic Frameworks (ZMOFs): Synthesis and Crystal Structures of Zinc(II) Imidazolate Polymers with Zeolitic Topologies. Chem. - Eur. J. 2007, 13, 4146. (9) Shi, Q.; Kang, X.; Shi, F.; Dong, J. Zn10(Im)20·4DBF: an unprecedented 10-nodal zeolitic topology with a 10-MR channel and 10 crystallographically independent Zn atoms. Chem. Commun. 2015, 51, 1131. (10) Shi, Q.; Xu, W. J.; Huang, R. K.; Zhang, W. X.; Li, Y.; Wang, P.; Shi, F. N.; Li, L.; Li, J.; Dong, J. Zeolite CAN and AFI-Type Zeolitic Imidazolate Frameworks with Large 12-Membered Ring Pore Openings Synthesized Using Bulky Amides as Structure-Directing Agents. J. Am. Chem. Soc. 2016, 138, 16232. (11) Morris, W.; Leung, B.; Furukawa, H.; Yaghi, O. K.; He, N.; Hayashi, H.; Houndonougbo, Y.; Asta, M.; Laird, B. B.; Yaghi, O. M. A Combined Experimental-Computational Investigation of Carbon Dioxide Capture in a Series of Isoreticular Zeolitic Imidazolate Frameworks. J. Am. Chem. Soc. 2010, 132, 11006. (12) Banerjee, R.; Furukawa, H.; Britt, D.; Knobler, C.; O'Keeffe, M.; Yaghi, O. M. Control of Pore Size and Functionality in Isoreticular Zeolitic Imidazolate Frameworks and Their Carbon Dioxide Selective Capture Properties. J. Am. Chem. Soc. 2009, 131, 3875. (13) For topological notation, it is common to use a three-letter code, where capitalized letters indicate a true zeolitic topology found in the International Zeolite Association (IZA) Database (http://www.iza-structure.org/databases/) and lower case bolded letters indicate alternative topologies, many of which are recorded in the Reticular Chemistry Structure Resource (http://rcsr.anu.edu.au/), O'Keeffe, M.; Peskov, M. A.; Ramsden, S. J.; Yaghi, O. M. The Reticular Chemistry Structure Resource (RCSR) Database of, and Symbols for, Crystal Nets. Acc. Chem. Res. 2008, 41, 1782. (14) Yang, J.; Zhang, Y. B.; Liu, Q.; Trickett, C. A.; GutierrezPuebla, 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. (15) Lehnert, R.; Seel, F. Z. Z. Darstellung und Kristallstruktur des Mangan(II)‐ und Zink(II)‐Derivates des Imidazols. Anorg. Allg. Chem. 1980, 464, 187. (16) Lanchas, M.; Vallejo-Sánchez, D.; Beobide, G.; Castillo, O.; Aguayo, A. T.; Luque, A.; Román, P. A direct reaction approach for the synthesis of zeolitic imidazolate frameworks: template and temperature mediated control on network topology and crystal size. Chem. Commun. 2012, 48, 9930. (17) Karagiaridi, O.; Lalonde, M. B.; Bury, W.; Sarjeant, A. A.; Farha, O. K.; Hupp, J. T. Opening ZIF-8: A Catalytically Active Zeolitic Imidazolate Framework of Sodalite Topology with Unsubstituted Linkers. J. Am. Chem. Soc. 2012, 134, 18790. (18) Ramirez, J. R.; Yang, H.; Kane, C. M.;Ley, A. N.; Yang, H.; Holman, T. K. Reproducible Synthesis and High Porosity of merZn(Im)2 (ZIF-10): Exploitation of an Apparent Double-Eight Ring Template. J. Am. Chem. Soc. 2016, 138, 12017. (19) Akimbekov, Z.; Katsenis, A. D.; Nagabhushana, G. P.; Ayoub, G.; Arhangelskis, M.; Morris, A. J.; Friscic, T.; Navrotsky, A. Experimental and Theoretical Evaluation of the Stability of True MOF Polymorphs Explains Their Mechanochemical Interconversions. J. Am. Chem. Soc. 2017, 139, 7952. (20) Pichon, A.; Lazuen-Garay, A.; James, S. L. Solvent-free synthesis of a microporous metal–organic framework. CrystEngComm 2006, 8, 211. (21) Julien, P. A.; Užarević, K.; Katsenis, A. D.; Kimber, S. A. J.; Wang, T.; Farha, O. K.; Zhang, Y.; Casaban, J.; Germann, L.; Etter, M.; Dinnebier, R. E.; James, S. L.; Halasz, I.; Friščić ,T. In situ monitoring and mechanism of the mechanochemical formation of a microporous MOF-74 framework J. Am. Chem. Soc. 2016, 138, 29292932. (22) Mottillo, C.; Friščić, T. Advances in Solid-State Transformations of Coordination Bonds: From the Ball Mill to the Aging Chamber. Molecules 2017, 22. (23) James, S. L.; Adams, C. J.; Bolm, C.; Braga, D.; Collier, P.; Friščić, T.; Grepioni, F.; Harris, K. D.; Hyett, G.; Jones, W.; Krebs, A.; Mack, J.; Maini, L.; Orpen, A. G.; Parkin, I. P.; Shearouse, W. C.;

Steed, J. W.; Waddell, D. C. Mechanochemistry: opportunities for new and cleaner synthesis. Chem. Soc. Rev. 2012, 41, 413. (24) Beldon, P. J.; Fabian, L.; Stein, R. S.; Thirumurugan, A.; Cheetham, A. K.; Friščić, T. Rapid Room-Temperature Synthesis of Zeolitic Imidazolate Frameworks by Using Mechanochemistry. Angew. Chem., Int. Ed. Engl. 2010, 49, 9640. (25) Katsenis, A. D.; Puskaric, A.; Strukil, V.; Mottillo, C.; Julien, P. A.; Uzarević, K.; Pham, M. H.; Do, T. O.; Kimber, S. A.; Lazic, P.; Magdysyuk, O.; Dinnebier, R. E.; Halasz, I.; Friščić, T. In situ X-ray diffraction monitoring of a mechanochemical reaction reveals a unique topology metal-organic framework. Nat. Commun. 2015, 6, 6662. (26) Mottillo, C.; Lu, Y.; Pham, M.-H.; Cliffe, M. J.; Dob, T.-O.; Friščić, T. Mineral neogenesis as an inspiration for mild, solvent-free synthesis of bulk microporous metal–organic frameworks from metal (Zn, Co) oxides. Green Chem. 2013, 15, 2121. (27) Cliffe, M. J.; Mottillo, C.; Stein, R. S.; Bučar, D.-K.; Friščić, T. Accelerated aging: a low energy, solvent-free alternative to solvothermal and mechanochemical synthesis of metal–organic materials. Chem. Sci. 2012, 3, 2495. (28) Tanaka, S.; Kida, K.; Nagaoka, T.; Ota, T.; Miyake, Y. Mechanochemical dry conversion of zinc oxide to zeolitic imidazolate framework. Chem. Commun. 2013, 49, 7884. (29) Friščić, T.; Halasz, I.; Beldon, P. J.; Belenguer, A. M.; Adams, F.; Kimber, S. A.; Honkimaki, V.; Dinnebier, R. E. Real-time and in situ monitoring of mechanochemical milling reactions. Nat. Chem. 2013, 5, 66. (30) Martins, G. A.; Byrne, P. J.; Allan, P.; Teat, S. J.; Slawin, A. M.; Li, Y.; Morris, R. E. The use of ionic liquids in the synthesis of zinc imidazolate frameworks. Dalton Trans. 2010, 39, 1758. (31) Biswal, B. P.; Panda, T.; Banerjee, R. Solution mediated phase transformation (RHO to SOD) in porous Co-imidazolate based zeolitic frameworks with high water stability. Chem. Commun. 2012, 48, 11868. (32) Tian, Y. Q.; Yao, S. Y.; Gu, D.; Cui, K. H.; Guo, D. W.; Zhang, G.; Chen, Z. X.; Zhao, D. Y. Cadmium Imidazolate Frameworks with Polymorphism, High Thermal Stability, and a Large Surface Area. Chem.-Eur. J. 2010, 16, 1137. (33) Karagiaridi, O.; Bury, W.; Sarjeant, A. A.; Stern, C. L.; Farha, O. K.; Hupp, J. T. Synthesis and characterization of isostructural cadmium zeolitic imidazolate frameworks via solvent-assisted linker exchange. Chem. Sci. 2012, 3, 3256. (34) a) Huang, X. C.; Lin, Y. Y.; Zhang, J. P.; Chen, X. M. LigandDirected Strategy for Zeolite-Type Metal-Organic Frameworks: Zinc(II) Imidazolates With Unusual Zeolitic Topologies. Angew. Chem. Int. Ed. 2006, 45, 1557. b) BET surface area: He, C. T.; Jiang, L.; Ye, Z. M.; Krishna, R.; Zhong, Z. S.; Liao, P. Q.; Xu, J.; Ouyang, G.; Zhang, J. P.; Chen, X. M. Exceptional Hydrophobicity of a LargePore Metal-Organic Zeolite. J. Am. Chem. Soc. 2015, 137, 7217. (35) Wang, F.; Fu, H.-R.; Kang, Y.; Zhang, J., A new approach towards zeolitic tetrazolate-imidazolate frameworks (ZTIFs) with uncoordinated N-heteroatom sites for high CO2 uptake. Chem. Commun. 2014, 50, 12065. (36) Ban, Y.; Li, Y.; Peng, Y.; Jin, H.; Jiao, W.; Liu, X.; Yang, W. Metal-Substituted Zeolitic Imidazolate Framework ZIF-108: GasSorption and Membrane-Separation Properties. Chem. Eur. J. 2014, 20, 11402. (37) Liu, Y.; Kravtsov, V.; Larsen, R.; Eddaoudi, M. Molecular building blocks approach to the assembly of zeolite-like metalorganic frameworks (ZMOFs) with extra-large cavities. Chem. Commun. 2006, 1488. (38) Khay, I.; Chaplais, G.; Nouali, H.; Ortiz, G.; Marichal, C.; Patarin, J. Assessment of the energetic performances of various ZIFs with SOD or RHO topology using high pressure water intrusionextrusion experiments. Dalton Trans. 2016, 45, 4392. (39) Fei, H.; Cahill, J. F.; Prather, K. A.; Cohen, S. M. Tandem Postsynthetic Metal Ion and Ligand Exchange in Zeolitic Imidazolate Frameworks. Inorg. Chem. 2013, 52, 4011. (40) Cacho-Bailo, F.; Etxeberría-Benavides, M.; Karvan, O.; Téllez, C.; Coronas, J. Sequential amine functionalization inducing structural transition in an aldehyde-containing zeolitic imidazolate frame-

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

work: application to gas separation membranes. CrystEngComm 2017, 19, 1545. (41) Kahr, J.; Mowat, J. P.; Slawin, A. M.; Morris, R. E.; FairenJimenez, D.; Wright, P. A. Synthetic control of framework zinc puri-

nate crystallisation and properties of a large pore, decorated, mixedlinker RHO-type ZIF. Chem. Commun. 2012, 48, 6690.

ACS Paragon Plus Environment

Page 6 of 7

Page 7 of 7 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

ACS Paragon Plus Environment

7