Linker on the Catalytic Activity of Copper-Catecholate-Decorated

Dec 26, 2017 - ABSTRACT: Two new UiO-68 type of Zr-MOFs featuring redox non-innocent catechol-based linkers of different redox activities have been sy...
1 downloads 8 Views 2MB Size
Research Article www.acsami.org

Cite This: ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Effect of Redox “Non-Innocent” Linker on the Catalytic Activity of Copper-Catecholate-Decorated Metal−Organic Frameworks Xuan Zhang,† Nicolaas A. Vermeulen,† Zhiyuan Huang,‡ Yuexing Cui,† Jian Liu,† Matthew D. Krzyaniak,† Zhanyong Li,† Hyunho Noh,† Michael R. Wasielewski,† Massimiliano Delferro,‡ and Omar K. Farha*,†,§,∥ †

Department of Chemistry, Northwestern University, 2145 Sheridan Road, Evanston, Illinois 60208, United States Chemical Sciences & Engineering Division, Argonne National Laboratory, Argonne, Illinois 60439, United States § Department of Chemical and Biological Engineering, Northwestern University, 2145 Sheridan Road, Evanston, Illinois 60208, United States ∥ Department of Chemistry, Faculty of Science, King Abdulaziz University, Jeddah 21589, Saudi Arabia ‡

S Supporting Information *

ABSTRACT: Two new UiO-68 type of Zr-MOFs featuring redox non-innocent catechol-based linkers of different redox activities have been synthesized through a de novo mixed-linker strategy. Metalation of the MOFs with Cu(II) precursors triggers the reduction of Cu(II) by the phenyl-catechol groups to Cu(I) with the concomitant formation of semiquinone radicals as evidenced by EPR and XPS characterization. The MOF-supported catalysts are selective toward the allylic oxidation of cyclohexene and it is found that the presence of in situ-generated Cu(I) species exhibits enhanced catalytic activity as compared to a similar MOF with Cu(II) metalated naphthalenyl-dihydroxy groups. This work unveils the importance of metal−support redox interactions in the catalytic activity of MOF-supported catalysts which are not easily accessible in traditional metal oxide supports. KEYWORDS: redox-active, metal−organic frameworks, mixed-linker, catechol, copper catalyst, alkene oxidation, UiO-68, non-innocent



INTRODUCTION Metal−organic frameworks (MOFs) have been extensively studied over the past two decades due to their high surface area, porosity, as well as versatile structural and functional tunability.1−7 Redox active MOFs have received particular attention owing to their promising applications in electrocatalysis, magnetism, electrical conductivity, batteries, and molecular switches.8−17 One strategy of imparting redox activity to MOFs is via de novo synthesis by using redoxactive metal nodes and/or organic linkers featuring redoxactive moieties as the building blocks.11,18−27 Additionally, it can be achieved via postsynthetic functionalization by introducing redox-active species onto their redox-innocent inorganic nodes and/or organic linkers.14,15,28,29 In the context of homogeneous catalysis, redox non-innocent organic species have been known to serve as cocatalysts or a redox mediator for metal catalysts to facilitate catalytic reactions.30,31 However, using the redox properties of non-innocent organic ligands to effect catalysis is under-reported in heterogeneous systems. By using a redox “non-innocent” organic ligand with metal anchoring sites, we can access in MOFs the metal−support redox interactions that are not easily attainable in traditional supports such as alumina, silica, and zirconia. In this regard, catechol-based MOFs may serve as a promising platform for the study of metal−support redox © XXXX American Chemical Society

interactions by providing redox-active catechol moieties in the scaffolds of MOFs.32,33 However, previous reports on catecholbased MOFs have only focused on using the catechol groups as metal anchoring sites. For example, studies on catechol-based MOFs and covalent−organic frameworks (COFs) by us and others have demonstrated the use of catechol groups as the metal-anchoring sites for their applications in gas storage/ separation, proton/electron conductivity, and catalysis.32−42 Therefore, investigations into the interplay between the redox non-innocent catechol ligands and redox active metal catalysts in MOFs, especially its effect on catalysis, have yet to be reported. One of the challenges of employing chemically active organic linkers in the direct synthesis of MOFs is the instability of these organic struts under the solvothermal synthetic conditions. To circumvent the thermal decomposition and to alleviate the steric hindrance from the protecting groups of such reactive organic linkers, a doping strategy where the reactive linkers are doped into a MOF matrix with inert linkers has proven to be powerful, either by a de novo mixed-linker synthesis or a postsynthetic solvent-assisted linker-exchange Received: October 10, 2017 Accepted: December 12, 2017

A

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Figure 1. Schematic view of the structure of UiO-68 highlighting the octahedral and tetrahedral cavities.

Scheme 1. Mixed-Linker Syntheses of the UiO-68-PCAT and UiO-68-NCAT MOFs

(SALE) method.43−46 Zr-MOFs such as the UiO (Universitetet i Oslo) series with the general formula of Zr6(OH)4(O)4(dicarboxylate)6 have been widely studied since they are first reported due to the excellent stability of this family of MOFs which makes them great supports for catalysts.47,48 We hypothesized that the redox-activity of the non-innocent catechol linkers in stable UiO-type of zirconiumMOFs will allow us to access the redox interactions between the metal precursors and MOF supports and their effect on catalytic reactions. Herein we report the introduction of redox-active catechol linkers in UiO-68 type of MOFs (Figure 1) through a mixedlinker synthetic method. Subsequent metalation of the phenylcatechol groups with Cu(II) precursors generated Cu(I)

species in situ, which exhibited enhanced catalytic activity toward the allylic oxidation of cyclohexene than the corresponding Cu(II) catalysts.



RESULTS AND DISCUSSION It has been found that when only 2′,3′-di-tertbutyloxycarbonyloxy[1,1′:4′,1″-terphenyl]-4,4″-dicarboxylic acid (L2) or 2,3-bis(methoxymethoxy)-(1,4-naphthalenyl)bis(benzoic acid) (L4) were used in the synthesis of MOFs under solvothermal conditions, either amorphous materials or no solid materials were obtained due to the decomposition of the ligands. Therefore, a mixed-linker synthetic strategy was employed to introduce redox-active catechol groups in the robust backbone of UiO-68 type of Zr-MOFs.48−50 The B

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Figure 2. EPR spectra of the catechol-incorporated MOFs and copper-metalated MOFs (top), as well as the corresponding chemical events (bottom) that resulted in the EPR spectra. a) The semiquinone radical signals of the catechol-incorporated MOFs. b) Cu(II) and semiquinone radical signals of the copper-metalated MOFs. The inset indicates the semiquinone radical signals.

catechol by one electron to produce the semiquinone radical and Cu(I).53 Compared to UiO-68-NCAT-Cu, UiO-68PCAT-Cu exhibits a more pronounced radical signal (g = 2.0032) as shown in the EPR spectra, indicating the oxidation of the catechol groups in the latter by Cu(II) (Figure 2b). This result has been confirmed by the XPS (X-ray photoelectron spectroscopy) spectra of the metalated MOFs. The shoulders at 933.4 and 953.1 eV in the XPS spectrum of UiO-68-PCATCu indicate the presence of Cu(I) species, while only Cu(II) signals were observed in UiO-68-NCAT-Cu (Figures S10 and S11). The total copper loading has been determined from ICPAES to be 1.3 and 1.1 copper per Zr6 node (or ∼73% and ∼52% of the catechol sites) for UiO-68-PCAT-Cu and UiO68-NCAT-Cu, respectively. To provide additional evidence of the presence and to quantify the amount of Cu(I) in UiO-68PCAT-Cu, a click reaction between azidobenzene and acetylenylbenzene, in which Cu(I) can serve as a stoichiometric reducing agent, has been carried out and the triazole product has been observed, indicating that 45% of Cu species in UiO-68-PCAT-Cu is Cu(I) (Figure S13). Cyclohexene oxidation was chosen as a model reaction to test our concept of the effect of metal−support redox interaction on the catalytic activity, as Cu(II) has been previously shown to be active for the oxidation of alkenes. The reaction was carried out by using tert-butylhydroperoxide (TBHP) as the oxidant.54 Two reaction pathways usually occur in cyclohexene oxidation, namely the allylic oxidation and the olefinic epoxidation.18,55,56 It is found that both UiO-68PCAT-Cu and UiO-68-NCAT-Cu are highly selective toward the allylic oxidation and no epoxidation products were observed.57 Among the three allylic oxidation products of 2cyclohexe-1-ol (1), 2-cyclohexe-1-one (2), and tert-butyl-2cyclohexenyl-1-peroxide (3), as confirmed by GC-MS, the Cucontaining catalysts UiO-68-PCAT-Cu and UiO-68-NCAT-Cu are more selective toward 3 than 1 or 2 (Table 1). Cyclohexene turnover numbers (TONs) of 141 ± 7 and 109 ± 4 were achieved with UiO-68-PCAT-Cu and UiO-68NCAT-Cu, respectively, after 6 h at 70 °C (Table 1). The higher initial rate and higher TON of UiO-68-PCAT-Cu compared to UiO-68-NCAT-Cu (Figure S14) are attributed to the presence of Cu(I) species, which has been shown to be more active than Cu(II) in some oxidation reactions in the literature.30,58 When using UiO-68-NCAT-Cu(II) as a catalyst,

protected catechol-bearing linker of L2 is mixed with a structurally analogous noncatechol linker of 2′,3′,5′,6′tetramethyl[1,1′:4′,1″-terphenyl]-4,4″-dicarboxylic acid (L1) during the MOF synthesis process to afford UiO-68-PCAT (Scheme 1). Similarly, the mixed-linkers of L4 and 9,10anthacenyl bis(benzoic acid) (L3) yielded UiO-68-NCAT. The amounts of catechol-containing linkers in the MOFs were determined to be ∼30% and ∼35% with L1/L2 and L3/L4 mixed-linkers respectively by 1H NMR of the digested MOF samples (Figures S3 and S5 of the Supporting Information, SI). The UiO-68 type structure is confirmed by powder X-ray diffraction patterns of UiO-68-PCAT and UiO-68-NCAT, which matches the simulated pattern of PCN-57, a derivative of the UiO-68 type MOF constructed from L1 (Figure S7).51 The octahedron morphology of UiO-68-PCAT and UiO-68NCAT crystals is consistent with the UiO-68 type of MOFs, as evidenced by their SEM images (Figure S9). The presence of excess acetic acid and heat helped to deprotect the catechol groups from the tert-butylcarbonyl (Boc) and methoxymethyl (MOM) protecting groups during the MOF synthesis process, which was confirmed by the absence of the -MOM and -Boc signals in the 1H NMR spectra of the digested MOFs. In comparison, the free ligands digested under the same conditions did not show signs of deprotection (Figures S4 and S6). This in situ deprotection method for the incorporation of catechol functional groups in MOFs is favorable compared to the previously reported postsynthetic deprotection methods as it is a one-pot reaction and does not require additional treatments such as light irradiation and linker exchange.34,39,52 The synthetic conditions employed in this work also engender radical species in the MOFs as confirmed by the EPR spectra of the as-synthesized UiO-68PCAT and UiO-68-NCAT MOFs, which is attributed to the partial aerobic oxidation of the catechol groups after being deprotected (Figure 2a). Additionally, when we attempted to use the less stable -MOM protecting group instead of -Boc in the synthesis of UiO-68-PCAT, we failed to incorporate the PCAT linkers, which was probably due to easy deprotection and loss of protection groups during the MOF synthesis. To study the redox interactions between the catechol groups and anchored metal ions, the MOFs were metalated with Cu(II) (denoted as UiO-68-PCAT-Cu and UiO-68-NCATCu), as it has been reported before to be able to oxidize C

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Mixed-Linker MOFs Synthesis. UiO-68-PCAT. ZrCl4 (93.2 mg, 0.4 mmol), L1 (75.3 mg, 0.2 mmol), L2 (110.2 mg, 0.2 mmol), and acetic acid (688 μL, 1.2 mmol) in 16 mL of deoxygenated DMF were sealed in a microwave reaction tube under nitrogen and heated at 120 °C for 24 h. After cooling to room temperature, the supernatant was decanted, and the solid was washed with DMF three times. Then the DMF was exchanged with acetone three times, and the solid was dried under vacuum to give 93 mg of product (yield 48%). The crystallinity and porosity have been confirmed by PXRD and N2 isotherms. UiO-68-NCAT. ZrCl4 (93.2 mg, 0.4 mmol), L3 (83.7 mg, 0.2 mmol), L4 (97.7 mg, 0.2 mmol), and acetic acid (688 μL, 1.2 mmol) in 16 mL of deoxygenated DMF were sealed in a microwave reaction tube under nitrogen and heated at 120 °C for 24 h. After cooling to room temperature, the supernatant was decanted, and the solid was washed with DMF three times. Then the DMF was exchanged with acetone three times, and the solid was dried under vacuum to give 112 mg of product (yield 57%). The crystallinity and porosity have been confirmed by PXRD and N2 isotherms. Metalation of the MOFs was carried out by adding 40 mg of the MOFs to a degassed solution of Cu precursors (5 mg of Cu(OAc)2 hydrate, or 10 mg of [Cu(CH3CN)4]OTf and 0.84 mL of Et3N) in 10 mL methanol under N2 atmosphere. The mixture was left undisturbed at room temperature for 2 h and then centrifuged and washed with methanol for three times. The Cu metal content was determined by ICP-AES. Cyclohexene Oxidation. The cyclohexene oxidation reaction was carried out by a modification of literature procedure.55 0.2 mL of cychohexene, 1.18 mL of tert-butylhydroperoxide (5.5 M in decane), 50 μL of 1,2-dichlorobenzene (internal standard), and the catalyst (0.45 mol % equivalence Cu) was added and heated with stirring in a sealed microwave tube at 70 °C. 50 μL of aliquots were taken at different time points, diluted with 1 mL of acetone, filtered through an activated alumina-packed glass wool, and subjected to GC-FID analysis. After the reaction, the catalysts were separated from the solution by centrifuge and washed with acetone three times and dried under vacuum for reuse. Fresh catalysts were added in each reuse to compensate the loss of the catalysts during the sampling and washing steps.

Table 1. Catalytic Oxidation of Cyclohexene by Different MOF Catalysts (0.45 mol % Catalyst, 70 °C)

selectivity (%) catalyst

TON (6 h)

conversion (%)

1

2

3

UiO-68-PCAT-Cu(I) UiO-68-PCAT-Cu(I/II) UiO-68-NCAT-Cu(II) UiO-68-PCAT UiO-68-NCAT

152 141 109 19 20

68 63 49 9 9

35 31 26 62 60

19 20 13 24 25

46 49 61 14 15

the allylic oxidation of cyclohexene takes place via a radical route where Cu(III)-hydroxy and Cu(III)-tert-butylperoxy intermediates are possibly present (Figure S17).59,60 In comparison, when Cu(I) serves as the active catalyst, it is likely that the weaker Cu−O bonds in the Cu(II)-hydroxy or Cu(II)-tert-butylperoxy intermediates allow for more ready formation of the tert-butylperoxyl radicals than the Cu(III) intermediates (Figure S17).30,61,62 This postulation is supported by a control reaction where a Cu(I) precursor was used to metalate UiO-68-PCAT (designated as UiO-68PCAT-Cu(I)), which exhibited an even higher turnover of 152 ± 4 (Table 1). Therefore, it demonstrates our concept that non-innocent linkers in UiO-68-PCAT can be employed to modify the catalytic activity of the MOF-supported Cu catalysts by altering the oxidation state of the metal precursor. The catalysts were reused, and no significant metal leaching or loss in the crystallinity and catalytic activity of the catalysts were observed after three cycles (Figures S15 and S16).





The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.7b15326. Details on ligands synthesis, instrumentation, and additional figures (PDF)

CONCLUSIONS Non-innocent catechol groups have been introduced into the backbones of UiO-68-like MOFs and metalation of the phenylcatechol groups with Cu(II) engenders Cu(I) species, which exhibited superior catalytic activity for cyclohexene oxidation. It has been shown that the redox properties of the noninnocent catechol ligands can be employed to alter the catalytic activity of the metalated MOFs by modifying the oxidation state of the active catalyst. We have demonstrated a model system for the study of metal−support redox interactions and their effect on catalytic reactions and this concept can be successfully applied to improve catalytic activities from Cu(II) precursors. This work offers a new designing strategy for MOFsupported catalysts that takes advantage of the redox-noninnocence of organic linkers in the catalytic activities.



ASSOCIATED CONTENT

S Supporting Information *



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (O.K.F.). ORCID

Xuan Zhang: 0000-0001-8214-7265 Zhanyong Li: 0000-0002-3230-5955 Michael R. Wasielewski: 0000-0003-2920-5440 Massimiliano Delferro: 0000-0002-4443-165X Omar K. Farha: 0000-0002-9904-9845 Author Contributions

EXPERIMENTAL SECTION

The manuscript was written through the contributions of all authors. All authors have given approval to the final version of the manuscript.

All chemicals were purchased from commercial sources and were used without further purification unless otherwise stated in the procedures. Tetramethyl[1,1′:4′,1″-terphenyl]-4,4″-dicarboxylic acid (L1), 9,10anthracenyl bis(benzoic acid) (L3) and 1,4-dibromo-2,3-bis(methoxymethoxy)naphthalene are synthesized according to reported procedures.51,63,64 The detailed synthesis and characterization of the two new ligands, 2′,3′-di-tert-butyloxycarbonyloxy[1,1′:4′,1″-terphenyl]-4,4″-dicarboxylic acid (L2) and 2,3-bis(methoxymethoxy)-1,4naphthalenyl bis(benzoic acid) (L4), can be found in the SI.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The synthesis and catalysis work was supported as part of the Inorganometallic Catalyst Design Center, an Energy Frontier D

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces

Via Heterogenous Redox Chemistry. J. Am. Chem. Soc. 2017, 139, 4175−4184. (13) Chen, Q.; Sun, J.; Li, P.; Hod, I.; Moghadam, P. Z.; Kean, Z. S.; Snurr, R. Q.; Hupp, J. T.; Farha, O. K.; Stoddart, J. F. A Redox-Active Bistable Molecular Switch Mounted inside a Metal−Organic Framework. J. Am. Chem. Soc. 2016, 138, 14242−14245. (14) Kung, C.-W.; Mondloch, J. E.; Wang, T. C.; Bury, W.; Hoffeditz, W.; Klahr, B. M.; Klet, R. C.; Pellin, M. J.; Farha, O. K.; Hupp, J. T. Metal−Organic Framework Thin Films as Platforms for Atomic Layer Deposition of Cobalt Ions to Enable Electrocatalytic Water Oxidation. ACS Appl. Mater. Interfaces 2015, 7, 28223−28230. (15) Hod, I.; Sampson, M. D.; Deria, P.; Kubiak, C. P.; Farha, O. K.; Hupp, J. T. Fe-Porphyrin-Based Metal−Organic Framework Films as High-Surface Concentration, Heterogeneous Catalysts for Electrochemical Reduction of CO2. ACS Catal. 2015, 5, 6302−6309. (16) Wang, H.-Y.; Ge, J.-Y.; Hua, C.; Jiao, C.-Q.; Wu, Y.; Leong, C. F.; D’Alessandro, D. M.; Liu, T.; Zuo, J.-L. Photo- and Electronically Switchable Spin-Crossover Iron(II) Metal−Organic Frameworks Based on a Tetrathiafulvalene Ligand. Angew. Chem., Int. Ed. 2017, 56, 5465−5470. (17) Kornienko, N.; Zhao, Y.; Kley, C. S.; Zhu, C.; Kim, D.; Lin, S.; Chang, C. J.; Yaghi, O. M.; Yang, P. Metal−Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2015, 137, 14129−14135. (18) Leus, K.; Vandichel, M.; Liu, Y.-Y.; Muylaert, I.; Musschoot, J.; Pyl, S.; Vrielinck, H.; Callens, F.; Marin, G. B.; Detavernier, C.; Wiper, P. V.; Khimyak, Y. Z.; Waroquier, M.; Van Speybroeck, V.; Van Der Voort, P. The Coordinatively Saturated Vanadium MIL-47 as a Low Leaching Heterogeneous Catalyst in the Oxidation of Cyclohexene. J. Catal. 2012, 285, 196−207. (19) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Metal-Organic Framework Materials as Catalysts. Chem. Soc. Rev. 2009, 38, 1450−1459. (20) Xiao, D. J.; Oktawiec, J.; Milner, P. J.; Long, J. R. Pore Environment Effects on Catalytic Cyclohexane Oxidation in Expanded Fe2(DOBDC) Analogues. J. Am. Chem. Soc. 2016, 138, 14371−14379. (21) Ruano, D.; Díaz-García, M.; Alfayate, A.; Sánchez-Sánchez, M. Nanocrystalline M−MOF-74 as Heterogeneous Catalysts in the Oxidation of Cyclohexene: Correlation of the Activity and Redox Potential. ChemCatChem 2015, 7, 674−681. (22) Tulchinsky, Y.; Hendon, C. H.; Lomachenko, K. A.; Borfecchia, E.; Melot, B. C.; Hudson, M. R.; Tarver, J. D.; Korzynski, M. D.; Stubbs, A. W.; Kagan, J. J.; Lamberti, C.; Brown, C. M.; Dinca, M. Reversible Capture and Release of Cl2 and Br2 with a Redox-Active Metal-Organic Framework. J. Am. Chem. Soc. 2017, 139, 5992−5997. (23) Zhang, X.; Zhang, Z.; Zhao, H.; Mao, J. G.; Dunbar, K. R. A Cadmium TCNQ-Based Semiconductor with Versatile Binding Modes and Non-Integer Redox States. Chem. Commun. 2014, 50, 1429−1431. (24) Narayan, T. C.; Miyakai, T.; Seki, S.; Dinca, M. High Charge Mobility in a Tetrathiafulvalene-Based Microporous Metal-Organic Framework. J. Am. Chem. Soc. 2012, 134, 12932−12935. (25) Farha, O. K.; Malliakas, C. D.; Kanatzidis, M. G.; Hupp, J. T. Control over Catenation in Metal−Organic Frameworks Via Rational Design of the Organic Building Block. J. Am. Chem. Soc. 2010, 132, 950−952. (26) An, W.; Aulakh, D.; Zhang, X.; Verdegaal, W.; Dunbar, K. R.; Wriedt, M. Switching of Adsorption Properties in a Zwitterionic Metal−Organic Framework Triggered by Photogenerated Radical Triplets. Chem. Mater. 2016, 28, 7825−7832. (27) Leong, C. F.; Chan, B.; Faust, T. B.; D’Alessandro, D. M. Controlling Charge Separation in a Novel Donor-Acceptor MetalOrganic Framework Via Redox Modulation. Chem. Sci. 2014, 5, 4724−4728. (28) Hod, I.; Farha, O. K.; Hupp, J. T. Modulating the Rate of Charge Transport in a Metal-Organic Framework Thin Film Using Host:Guest Chemistry. Chem. Commun. 2016, 52, 1705−1708.

Research Center funded by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), under Award DESC0012702 (O.K.F. and M.D.). The EPR work was supported by the Chemical Sciences, Geosciences, and Biosciences Division, Office of Basic Energy Sciences, DOE under grant no. DE-FG02-99ER14999 (M.R.W.). Argonne National Laboratory is supported by the U.S. Department of Energy, Office of Science, and Office of Basic Energy Sciences, under Contract DE-AC02-06CH11357. This work made use of the IMSERC, J. B. Cohen X-ray Diffraction, EPIC, and KECK II facilities of Northwestern University’s NUANCE Center, which has received support from the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF ECCS-1542205); the MRSEC program (NSF DMR1121262) at the Materials Research Center; the International Institute for Nanotechnology (IIN); the Keck Foundation; and the State of Illinois, through the IIN.



REFERENCES

(1) Yaghi, O. M. Reticular ChemistryConstruction, Properties, and Precision Reactions of Frameworks. J. Am. Chem. Soc. 2016, 138, 15507−15509. (2) Wang, C.; Liu, D.; Lin, W. Metal-Organic Frameworks as a Tunable Platform for Designing Functional Molecular Materials. J. Am. Chem. Soc. 2013, 135, 13222−13234. (3) Zhou, H. C.; Long, J. R.; Yaghi, O. M. Introduction to MetalOrganic Frameworks. Chem. Rev. 2012, 112, 673−674. (4) Li, Z.; Peters, A. W.; Bernales, V.; Ortuno, M. A.; Schweitzer, N. M.; DeStefano, M. R.; Gallington, L. C.; Platero-Prats, A. E.; Chapman, K. W.; Cramer, C. J.; Gagliardi, L.; Hupp, J. T.; Farha, O. K. Metal-Organic Framework Supported Cobalt Catalysts for the Oxidative Dehydrogenation of Propane at Low Temperature. ACS Cent. Sci. 2017, 3, 31−38. (5) Islamoglu, T.; Goswami, S.; Li, Z.; Howarth, A. J.; Farha, O. K.; Hupp, J. T. Postsynthetic Tuning of Metal-Organic Frameworks for Targeted Applications. Acc. Chem. Res. 2017, 50, 805−813. (6) Chen, Z.; Weseliński, Ł. J.; Adil, K.; Belmabkhout, Y.; Shkurenko, A.; Jiang, H.; Bhatt, P. M.; Guillerm, V.; Dauzon, E.; Xue, D.-X.; O’Keeffe, M.; Eddaoudi, M. Applying the Power of Reticular Chemistry to Finding the Missing alb-MOF Platform Based on the (6,12)-Coordinated Edge-Transitive Net. J. Am. Chem. Soc. 2017, 139, 3265−3274. (7) Alezi, D.; Spanopoulos, I.; Tsangarakis, C.; Shkurenko, A.; Adil, K.; Belmabkhout, Y.; O'Keeffe, M.; Eddaoudi, M.; Trikalitis, P. N. Reticular Chemistry at Its Best: Directed Assembly of Hexagonal Building Units into the Awaited Metal-Organic Framework with the Intricate Polybenzene Topology, pbz-MOF. J. Am. Chem. Soc. 2016, 138, 12767−12770. (8) Sheberla, D.; Sun, L.; Blood-Forsythe, M. A.; Er, S.; Wade, C. R.; Brozek, C. K.; Aspuru-Guzik, A.; Dinca, M. High Electrical Conductivity in Ni3(2,3,6,7,10,11-Hexaiminotriphenylene)2, a Semiconducting Metal-Organic Graphene Analogue. J. Am. Chem. Soc. 2014, 136, 8859−8862. (9) Sun, L.; Campbell, M. G.; Dinca, M. Electrically Conductive Porous Metal-Organic Frameworks. Angew. Chem., Int. Ed. 2016, 55, 3566−3579. (10) Zhang, X.; Saber, M. R.; Prosvirin, A. P.; Reibenspies, J. H.; Sun, L.; Ballesteros-Rivas, M.; Zhao, H.; Dunbar, K. R. Magnetic Ordering in Tcnq-Based Metal−Organic Frameworks with Host− Guest Interactions. Inorg. Chem. Front. 2015, 2, 904−911. (11) Zhang, Z.; Yoshikawa, H.; Awaga, K. Monitoring the SolidState Electrochemistry of Cu(2,7-Aqdc) (Aqdc = Anthraquinone Dicarboxylate) in a Lithium Battery: Coexistence of Metal and Ligand Redox Activities in a Metal-Organic Framework. J. Am. Chem. Soc. 2014, 136, 16112−16115. (12) DeGayner, J. A.; Jeon, I.-R.; Sun, L.; Dincă, M.; Harris, T. D. 2D Conductive Iron-Quinoid Magnets Ordering up to Tc = 105 K E

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces (29) Wang, T. C.; Hod, I.; Audu, C. O.; Vermeulen, N. A.; Nguyen, S. T.; Farha, O. K.; Hupp, J. T. Rendering High Surface Area, Mesoporous Metal−Organic Frameworks Electronically Conductive. ACS Appl. Mater. Interfaces 2017, 9, 12584−12591. (30) McCann, S. D.; Stahl, S. S. Copper-Catalyzed Aerobic Oxidations of Organic Molecules: Pathways for Two-Electron Oxidation with a Four-Electron Oxidant and a One-Electron RedoxActive Catalyst. Acc. Chem. Res. 2015, 48, 1756−1766. (31) Weers, J. J.; Thomasson, C. E. In ACS Fuels: Washington DC, 1990; Vol. 35 (4), p 1247−1254. (32) Lee, Y.; Kim, S.; Fei, H.; Kang, J. K.; Cohen, S. M. Photocatalytic CO2 Reduction Using Visible Light by MetalMonocatecholato Species in a Metal-Organic Framework. Chem. Commun. 2015, 51, 16549−16552. (33) Totten, R. K.; Weston, M. H.; Park, J. K.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. Catalytic Solvolytic and Hydrolytic Degradation of Toxic Methyl Paraoxon with La(Catecholate)-Functionalized Porous Organic Polymers. ACS Catal. 2013, 3, 1454−1459. (34) Tanabe, K. K.; Allen, C. A.; Cohen, S. M. Photochemical Activation of a Metal−Organic Framework to Reveal Functionality. Angew. Chem., Int. Ed. 2010, 49, 9730−9733. (35) Hmadeh, M.; Lu, Z.; Liu, Z.; Gándara, F.; Furukawa, H.; Wan, S.; Augustyn, V.; Chang, R.; Liao, L.; Zhou, F.; Perre, E.; Ozolins, V.; Suenaga, K.; Duan, X.; Dunn, B.; Yamamto, Y.; Terasaki, O.; Yaghi, O. M. New Porous Crystals of Extended Metal-Catecholates. Chem. Mater. 2012, 24, 3511−3513. (36) Weston, M. H.; Farha, O. K.; Hauser, B. G.; Hupp, J. T.; Nguyen, S. T. Synthesis and Metalation of Catechol-Functionalized Porous Organic Polymers. Chem. Mater. 2012, 24, 1292−1296. (37) Weston, M. H.; Peterson, G. W.; Browe, M. A.; Jones, P.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. Removal of Airborne Toxic Chemicals by Porous Organic Polymers Containing Metal-Catecholates. Chem. Commun. 2013, 49, 2995−2997. (38) Weston, M. H.; Colon, Y. J.; Bae, Y.-S.; Garibay, S. J.; Snurr, R. Q.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. High Propylene/Propane Adsorption Selectivity in a Copper(Catecholate)-Decorated Porous Organic Polymer. J. Mater. Chem. A 2014, 2, 299−302. (39) Fei, H.; Shin, J.; Meng, Y. S.; Adelhardt, M.; Sutter, J.; Meyer, K.; Cohen, S. M. Reusable Oxidation Catalysis Using MetalMonocatecholato Species in a Robust Metal−Organic Framework. J. Am. Chem. Soc. 2014, 136, 4965−4973. (40) Darago, L. E.; Aubrey, M. L.; Yu, C. J.; Gonzalez, M. I.; Long, J. R. Electronic Conductivity, Ferrimagnetic Ordering, and Reductive Insertion Mediated by Organic Mixed-Valence in a Ferric Semiquinoid Metal-Organic Framework. J. Am. Chem. Soc. 2015, 137, 15703−15711. (41) Nguyen, H. G. T.; Weston, M. H.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. A Catalytically Active Vanadyl(Catecholate)-Decorated Metal Organic Framework Via Post-Synthesis Modifications. CrystEngComm 2012, 14, 4115−4118. (42) Nguyen, H. G. T.; Weston, M. H.; Sarjeant, A. A.; Gardner, D. M.; An, Z.; Carmieli, R.; Wasielewski, M. R.; Farha, O. K.; Hupp, J. T.; Nguyen, S. T. Design, Synthesis, Characterization, and Catalytic Properties of a Large-Pore Metal-Organic Framework Possessing Single-Site Vanadyl(Monocatecholate) Moieties. Cryst. Growth Des. 2013, 13, 3528−3534. (43) Yuan, S.; Qin, J.-S.; Zou, L.; Chen, Y.-P.; Wang, X.; Zhang, Q.; Zhou, H.-C. Thermodynamically Guided Synthesis of Mixed-Linker Zr-Mofs with Enhanced Tunability. J. Am. Chem. Soc. 2016, 138, 6636−6642. (44) Manna, K.; Zhang, T.; Greene, F. X.; Lin, W. Bipyridine- and Phenanthroline-Based Metal−Organic Frameworks for Highly Efficient and Tandem Catalytic Organic Transformations Via Directed C−H Activation. J. Am. Chem. Soc. 2015, 137, 2665−2673. (45) Fei, H.; Cohen, S. M. Metalation of a ThiocatecholFunctionalized Zr(IV)-Based Metal−Organic Framework for Selective C−H Functionalization. J. Am. Chem. Soc. 2015, 137, 2191− 2194.

(46) Karagiaridi, O.; Bury, W.; Mondloch, J. E.; Hupp, J. T.; Farha, O. K. Solvent-Assisted Linker Exchange: An Alternative to the De Novo Synthesis of Unattainable Metal−Organic Frameworks. Angew. Chem., Int. Ed. 2014, 53, 4530−4540. (47) Bai, Y.; Dou, Y.; Xie, L.-H.; Rutledge, W.; Li, J.-R.; Zhou, H.-C. Zr-Based Metal-Organic Frameworks: Design, Synthesis, Structure, and Applications. Chem. Soc. Rev. 2016, 45, 2327−2367. (48) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc. 2008, 130, 13850−13851. (49) Chavan, S. M.; Shearer, G. C.; Svelle, S.; Olsbye, U.; Bonino, F.; Ethiraj, J.; Lillerud, K. P.; Bordiga, S. Synthesis and Characterization of Amine-Functionalized Mixed-Ligand Metal−Organic Frameworks of UiO-66 Topology. Inorg. Chem. 2014, 53, 9509−9515. (50) Qin, J.-S.; Yuan, S.; Wang, Q.; Alsalme, A.; Zhou, H.-C. MixedLinker Strategy for the Construction of Multifunctional MetalOrganic Frameworks. J. Mater. Chem. A 2017, 5, 4280−4291. (51) Jiang, H.-L.; Feng, D.; Liu, T.-F.; Li, J.-R.; Zhou, H.-C. Pore Surface Engineering with Controlled Loadings of Functional Groups Via Click Chemistry in Highly Stable Metal−Organic Frameworks. J. Am. Chem. Soc. 2012, 134, 14690−14693. (52) Fei, H.; Pullen, S.; Wagner, A.; Ott, S.; Cohen, S. M. Functionalization of Robust Zr(IV)-Based Metal-Organic Framework Films Via a Postsynthetic Ligand Exchange. Chem. Commun. 2015, 51, 66−69. (53) Thompson, J. S.; Calabrese, J. C. Copper-Catechol Chemistry. Synthesis, Spectroscopy, and Structure of Bis(3,5-Di-Tert-Butyl-OSemiquinato)Copper(II). J. Am. Chem. Soc. 1986, 108, 1903−1907. (54) Allen, S. E.; Walvoord, R. R.; Padilla-Salinas, R.; Kozlowski, M. C. Aerobic Copper-Catalyzed Organic Reactions. Chem. Rev. 2013, 113, 6234−6458. (55) Noh, H.; Cui, Y.; Peters, A. W.; Pahls, D. R.; Ortuno, M. A.; Vermeulen, N. A.; Cramer, C. J.; Gagliardi, L.; Hupp, J. T.; Farha, O. K. An Exceptionally Stable Metal-Organic Framework Supported Molybdenum(VI) Oxide Catalyst for Cyclohexene Epoxidation. J. Am. Chem. Soc. 2016, 138, 14720−14726. (56) Thornburg, N. E.; Thompson, A. B.; Notestein, J. M. Periodic Trends in Highly Dispersed Groups IV and V Supported Metal Oxide Catalysts for Alkene Epoxidation with H2O2. ACS Catal. 2015, 5, 5077−5088. (57) Wang, J.-C.; Hu, Y.-H.; Chen, G.-J.; Dong, Y.-B. Cu(II)/Cu(0) @UiO-66-NH2: Base Metal@MOFs as Heterogeneous Catalysts for Olefin Oxidation and Reduction. Chem. Commun. 2016, 52, 13116− 13119. (58) Hoover, J. M.; Stahl, S. S. Highly Practical Copper(I)/TEMPO Catalyst System for Chemoselective Aerobic Oxidation of Primary Alcohols. J. Am. Chem. Soc. 2011, 133, 16901−16910. (59) Tonigold, M.; Lu, Y.; Bredenkötter, B.; Rieger, B.; Bahnmüller, S.; Hitzbleck, J.; Langstein, G.; Volkmer, D. Heterogeneous Catalytic Oxidation by MFU-1: A Cobalt(II)-Containing Metal−Organic Framework. Angew. Chem., Int. Ed. 2009, 48, 7546−7550. (60) Franklin, C. C.; VanAtta, R. B.; Tai, A. F.; Valentine, J. S. Copper Ion Mediated Epoxidation of Olefins by Iodosylbenzene. J. Am. Chem. Soc. 1984, 106, 814−816. (61) Lu, Z.-R.; Yin, Y.-Q.; Jin, D.-S. Oxidation of Alkene with TButyl Hydroperoxide Catalysed by Copper(II) Complexes of Tris(2Benzimidazolylmethyl)Amine: Model Complexes of Blue Copper Proteins. J. Mol. Catal. 1991, 70, 391−397. (62) Mase, N.; Mizumori, T.; Tatemoto, Y. Aerobic Copper/ TEMPO-Catalyzed Oxidation of Primary Alcohols to Aldehydes Using a Microbubble Strategy to Increase Gas Concentration in Liquid Phase Reactions. Chem. Commun. 2011, 47, 2086−2088. (63) Hauptvogel, I. M.; Biedermann, R.; Klein, N.; Senkovska, I.; Cadiau, A.; Wallacher, D.; Feyerherm, R.; Kaskel, S. Flexible and Hydrophobic Zn-Based Metal−Organic Framework. Inorg. Chem. 2011, 50, 8367−8374. (64) Zhao, Y.-L.; Liu, L.; Zhang, W.; Sue, C.-H.; Li, Q.; Miljanić, O. Š .; Yaghi, O. M.; Stoddart, J. F. Rigid-Strut-Containing Crown Ethers F

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

Research Article

ACS Applied Materials & Interfaces and [2]Catenanes for Incorporation into Metal−Organic Frameworks. Chem. - Eur. J. 2009, 15, 13356−13380.

G

DOI: 10.1021/acsami.7b15326 ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX