Quinone Reaction in Metal–Organic

Aug 28, 2015 - Zirconium Metal–Organic Frameworks Assembled from Pd and Pt PNP Pincer Complexes: Synthesis, Postsynthetic Modification, and Lewis ...
0 downloads 0 Views 622KB Size
Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Article

Reversible Tuning Hydroquinone/Quinone Reaction in MetalOrganic Framework: Immobilized Molecular Switches in Solid State Bo Gui, Xiangshi Meng, Yi Chen, Jianwu Tian, Guoliang Liu, ChenCheng Shen, Matthias Zeller, Daqiang Yuan, and Cheng Wang Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.5b02648 • Publication Date (Web): 28 Aug 2015 Downloaded from http://pubs.acs.org on August 29, 2015

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 free 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 accessible to all readers and 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.

Chemistry of Materials 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

Chemistry of Materials

Reversible Tuning Hydroquinone/Quinone Reaction in MetalOrganic Framework: Immobilized Molecular Switches in Solid State Bo Gui,a Xiangshi Meng,a Yi Chen,a Jianwu Tian,a Guoliang Liu,b Chencheng Shen,a Matthias Zeller,c Daqiang Yuan*b and Cheng Wang*a a

Key Laboratory of Biomedical Polymers (Ministry of Education), College of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, China b

State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China c

Department of Chemistry, Youngstown State University, One University Plaza, Youngstown, Ohio 44555, USA

ABSTRACT: Transferring the solution-state chemistry of organic-based molecular switches (OMS) into the solid state usually faces several fatal problems, such as spatial confinement or inefficient conversion. As a result, their switching behavior usually cannot be maintained. Herein, we report a redox-switchable metal-organic framework (MOF) that can undergo a reversible single-crystal-to-single-crystal (SCSC) transformation through a hydroquinone/quinone redox reaction. The redox-triggered transformation is quantitatively reversible, while maintaining the crystallinity of the MOF scaffold. In addition, the transformation occurs gradually in the MOF backbone and from the outsurface of MOF to the inside. This study represents a general strategy to enable efficient conversion of the functionality of an OMS from solution into solid state, by incorporation of OMS into the framework of MOF. Furthermore, the material exhibits interesting changes in spectroscopic properties through reversible SCSC transformation, thus may be a starting point for the use of such materials in memory storage or redox-based electronic devices.

INTRODUCTION Molecular switches are essential and possibly even indispensable for the design of nanoscale electronic circuits and component.1-3 Especially interesting are organic-based molecular switches (OMS),4-8 owing to their structural flexibility and a large number of synthetic methods based on centuries of knowledge in the field of organic chemistry. The possibility of targeted design and optimization relates OMS to molecular homogeneous or surface anchored catalysts,9,10 which − in contrast to most heterogeneous surface based catalysts − can be designed and developed with a specific task in mind. For further application, the OMS usually have to be linked to the surface of a device or incorporated into a functional solid-state material.11,12 However, again resembling similar problems in catalysis, transitioning of solution-based functionality of OMS into solid-state devices is problematic. The switches’ solution-state chemistry does not directly translate into their solid-state behavior.13 Several fatal problems, such as spatial confinement or inefficient conversion, often prevent their switching behavior to be maintained in the solid state. An attractive way to circumvent many if these problems would be to incorporate these OMS into the scaffold of a crystalline metal-organic framework (MOF),14-18 as

one of its organic linkers.19 There are several reasons that make this approach particularly attractive. First, similar to the OMS themselves, MOFs can be designed and their properties can be tailored to a given task.20,21 Second, the frameworks of MOF are often flexible enough to tolerate even substantial changes to the bridging ligands, and as such, many MOFs can maintain their crystallinity even upon drastic changes to some of their constituents.22,23 Third, the porous nature of MOFs allows free access for the reactants needed to activate the OMS and, if designed properly, enough space will be available within the MOF for conformational change of the OMS.24,25 Assuming successful implementation, this strategy could not only enable efficient switching of an OMS in the solid state, but would also allow the switching behaviour through reversible single-crystal-to-single-crystal (SCSC) transformation in the MOF, which would deepen our understanding of structure-function relationships at the molecular level and provide important information for development of technologically useful materials (e.g., memory devices26). Although some MOFs can undergo reversible SCSC transformation,15,22,23 the incorporation of an OMS into MOFs as a linker and then performing reversible SCSC transformation upon triggering the OMS have proven to be a substantial challenge.27-29 Only very recently have a

ACS Paragon Plus Environment

Chemistry of Materials

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

Page 2 of 7

few such systems been reported.30-39 For example, Zhou et al. reported a photoswitchable MOF by using azobenzene units as a dangling ligand, which exhibits a remarkable stimulus-responsive adsorption effect.30 It should be noted here that, most of the very few reported examples are of photo-responsive MOFs.30-35 As the construction of such systems is very interesting, more stimuli-responsive MOFs with other triggered mechanism are highly desirable. Moreover, how the switching behavior happen in the MOF state is still not clear. Herein, we would like to report the synthesis and characterization of a redoxswitchable MOF, namely UiO-68-OH, which can undergo a reversible SCSC transformation initiated by a hydroquinone/quinone redox reaction (Scheme 1). Our results demonstrate that the redox-tiggered transformation is quantitatively reversible, while maintaining the crystallinity of the MOF scaffold. By monitoring the switching process of this unique system, we demonstrated that the transformation occurred gradually in the MOF backbone and from the outsurface of MOF to the inside. In addition, this SCSC transformation is accompanied by drastic changes in spectroscopic properties, and thus may find application in memory storage, or even in the design of nanoscale electronic devices.

freshly exchanged three times per day. The crystal samples were kept in anhydrous DCM prior to oxidation. In addition, samples not used for subsequent modification were evacuated in oil pump vacuum at room temperature to yield activated samples. Elemental analyses of the activated sample found [C (49.94%), H (2.80%), N (0.35%)]; This would roughly correspond to a formula of Zr6O4(OH)4(C20H12O6)6(C3H7NO) (CH2Cl2)2 (mw 3012), giving a calculated profile as [C (49.84%), H (2.91%), N (0.47%)].

Scheme 1. Synthesis of UiO-68-OH and the redoxtriggered transformation of hydroquinone/quinone reaction in solution and solid state.

Further reduction of UiO-68-C=O. A crystal sample of freshly prepared UiO-68-C=O (~ 80 mg) was exchanged with anhydrous DMF 3 times and then immersed in 10 mL DMF solution containing ascorbic acid (20 mg mL−1), which was repeatedly agitated by pumping with a pipette for 15min. After that, the crystals were rinsed with anhydrous DMF (5 mL × 3) and then placed in anhydrous DCM (5 mL × 3) for 3 days to exchange and remove DMF. During this period, DCM was freshly exchanged three times per day. The crystal samples were kept in anhydrous DCM for further oxidation. In addition, samples not undergoing subsequent oxidation were evacuated by an oil pump at room temperature. The 1H NMR spectrum and high resolution mass spectrum of digested activated UiO-68-OH(R) is identical to that of compound 1.

Synthesis of UiO-68-C=O. Fresh UiO-68-OH (~ 80 mg) was immersed in a 10 mL DCM solution of iodobenzene diacetate (25 mg mL−1) and was repeatedly agitated by pumping with a pipette during a period of 5 min. After that, the solution was exchanged with anhydrous DCM (5 mL × 3) and kept in anhydrous DCM for further reduction. In addition, samples not used for subsequent modification were evacuated in oil pump vacuum at room temperature to yield activated samples. Elemental analyses on the activated sample found [C (48.76%), H (2.75%), N (0.30%)]; This would roughly correspond to a formula of Zr6O4(OH)4(C20H10O6)6(C3H7NO)(CH2Cl2)3 (mw 3085), giving a calculated profile as [C (49.06%), H (2.52%), N (0.45%)].

RESULTS AND DISCUSSION

EXPERIMENTAL SECTION Synthesis of UiO-68-OH. ZrCl4 (11.2 mg, 0.048 mmol), compound 1 (24.5 mg, 0.070 mmol) and benzoic acid (222.6 mg, 1.82 mmol) were dissolved in DMF (3 mL) and then filtered into a Pyrex glass tube. The mixture was degassed and sealed under nitrogen atmosphere, and the tube was placed in a preheated 120 ºC oven for 13 h. Single crystals with octahedral shape suitable for single crystal diffraction were harvested with a yield of ~ 9 mg. The assynthesized UiO-68-OH was rinsed with DMF overnight to remove any unreacted starting materials and trapped benzoic acid. After that, the sample was allowed to immerse in anhydrous DCM for 3 days to replace and remove DMF. During this period, anhydrous DCM was

To test the possibility of reversible SCSC transformation in an MOF through a redox reaction, we chose to investigate the hydroquinone/quinone system as the prototypical example, as it is usually involved in various biological electron transport systems.40-41 As such, we designed and synthesized a new tailored-made ligand, 2′,5 ′-dihydroxyterphenyl-4,4″-dicarboxylic acid (1).42 We first investigated the redox switching behavior of compound 1 in the solution-state (Scheme 1 and table S1). After reacting with oxidant, compound 1 is quantitatively oxidized to 2, as evidenced by 1H NMR spectroscopy (Figure S1) and high resolution mass spectrometry (Figure S24). Furthermore, compound 2 can be quantitatively reduced back to 1 (see table S2 for details), after addition of reductant. Therefore, compound 1 can undergo a reversible transformation in the solution state through oxidation/reduction cycles.

ACS Paragon Plus Environment

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

Chemistry of Materials UiO-68-OH was obtained as colorless octahedron-shaped crystals in reasonable yield. Single crystal X-ray diffraction studies performed at the Beijing Synchrotron Facility revealed a UiO-type structure for UiO-68-OH with an fcu topology (Figure 1a, and see section S7 for details), which crystallizes in the Fm3m space group and contains threedimensional percolated pore structure with the maximum pore diameter of 17 Å. The powder X-ray diffraction (PXRD) pattern of UiO-68-OH was collected, which closely matched with that simulated from the singlecrystal structure (Figure 2). Furthermore, the 1H NMR spectrum of digested UiO-68-OH (Figure 3a) is the same as that of compound 1 (Figure S1), indicating that the ligands remained intact in the MOF. However, starting from compound 2 (Scheme 1), the hydrothermal reaction performed under the same condition does not form UiO-68C=O.45 This can be possibly attributed to the fact that DMF under hydrothermal conditions can reduce quinone to hydroquinone.46

Figure 1. (a) Perspective view ([110] direction) of the reversible transformation between the crystal structures of UiO-68-OH and UiO-68-C=O. Oxidation condition: iodobenzene diacetate, CH2Cl2; Reduction condition: ascorbic acid, DMF. (b) Crystal images of UiO-68-OH after oxidation with different time. (c) Crystal images of UiO68-C=O after reduction with different time. The pictures were taken with crystals covered by a thin layer of DMF. The scale bar is 50 µm. We then designed and synthesized an MOF system which can favor a reversible SCSC transformation through this hydroquinone/quinone redox reaction. We chose a Zr-MOF as the substrate, due to its robustness compared with common Zn/Cu-centered MOFs.43-44 After treatment of a reaction mixture (sealed in a Pyrex glass tube under nitrogen atmosphere) containing ZrCl4, compound 1, benzoic acid and DMF in a preheated 120 ºC oven for 13 h,

Fresh UiO-68-OH(R)

Fresh UiO-68-C=O

Fresh UiO-68-OH

x8

10

Simulated UiO-68-OH

20

30

2θ (degree)

40

50

Figure 2. PXRD patterns of simulated from the single crystal structure (black line), fresh UiO-68-OH (red line), fresh UiO-68-C=O (blue line) and fresh UiO-68-OH(R) (purple line). The inset shows a magnification on the high 2θ region. The patterns were taken with crystals covered by a thin layer of DMF.

1

Figure 3. H NMR spectra of crystals digested with DMSOd6/HF: (a) UiO-68-OH; (b) oxidation for 5 s; (c) oxidation for 20 s; (d) UiO-68-C=O (oxidation for 5 min.); (e) reduction for 40 s; (f) reduction for 80 s; (g) UiO-68-OH(R) (reduction for 15 min.). The proton for OH could not be detected, due to the existence of HF.

The reversible SCSC transformation in UiO-68-OH through hydroquinone/quinone redox reaction was investigated. Unlike in solution, the chemically reversible oxidation/reduction of hydroquinone units in UiO-68-OH is quite a challenge,47-50 especially if the single crystalline nature of the samples is to be maintained. The choice of oxidant, reductant and even of the solvent, are all important for a successful SCSC transformation (see table S3 and S4 for details). After many unsuccessful or only partially successful attempts, we finally succeeded in reversible tuning SCSC transformation in UiO-68-OH through hydroquinone/quinone redox reaction, using iodobenzene diacetate (PIDA) as the oxidant and ascorbic acid (VC) as the reductant. Addition of an excess of PIDA in dichloromethane (DCM) to the colorless octahedron-shaped crystals (UiO68-OH) induced a change from originally colorless to yellow (Figure 1b) in 10 seconds, without change of shape.

ACS Paragon Plus Environment

Chemistry of Materials

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

After this quick deepening of color, crystals did visibly not change substantially any more. However, as monitored by 1 H NMR spectroscopy of digested crystals after varying oxidation periods (Figure 3a-d), the transformation of the hydroquinone units into quinone units in UiO-68-OH is a gradual process. As such, we believe the ligands on the surface area of UiO-68-OH are oxidized first, obscuring any change of color deeper within the crystals. After 5 min., the 1H NMR spectrum of the digested yellow crystals is identical to that of compound 2, indicating quantitative oxidation of the hydroquinone units in UiO-68-OH. The solid-state 13C NMR (SSNMR) spectrum (Figure 4d) of UiO-68-C=O is distinct and differs from that of UiO-68OH (Figure 4a), which also proves that the transformation has occurred. In addition, the Fourier transform infrared (FT-IR) spectrum (Figure S12) of UiO-68-C=O showed a peak at 1655 cm−1, which should be assigned to the vibrational frequency of C=O stretch and also confirm the transformation. As shown in Figure 2, the position of the PXRD peaks of UiO-68-C=O is similar to that of UiO68-OH, pointing towards retention of the framework topology. Therefore, by using post-synthetic modification approach, UiO-68-C=O was quantitatively obtained from UiO-68-OH via in situ oxidation.

Page 4 of 7

Reduction of UiO-68-C=O back to UiO-68-OH [named UiO-68-OH(R)] was then performed. After addition of an excess of VC in DMF, the yellow octahedron-shaped crystals gradually changed back to colorless, while again retaining their shape (Figure 1c). Interestingly, core-shell liked MOFs (with a yellow center and colorless on the outside, Figure 1c) were observed during this process. The formation of core-shell MOFs can be explained by the slow diffusion of VC into the crystals of UiO-68-C=O, which results the reduction of the outsurfaces of MOF first and then the inside. After further proved by SSNMR spectrum of the UiO-68-C=O crystals with different reduction time (Figure 4e and 4f), we can conclude again the reduction of UiO-68-C=O back to UiO-68-OH definitely occurred within the MOF backbone. Based on the 1 H NMR spectrum of digested crystals (Figure 3g), the quinone units in UiO-68-OH were quantitatively reduced back to hydroquinone within 15 min. Furthermore, the PXRD experiments (Figure 2 and Figure S18) demonstrate that the crystallinity was again retained during this reduction process. Thus, UiO-68-OH can be quantitatively oxidized to UiO-68-C=O, and UiO-68-C=O can be quantitatively reduced back to UiO-68-OH. Such SCSC transformations can be repeated for at least three cycles, with quantitative transformation (Figure S13) and retention of crystallinity (Figure S14).

13

Figure 4. Solid-state C NMR spectra of (a) UiO-68-OH; (b) oxidation for 5 s; (c) oxidation for 20 s; (d) UiO-68-C=O (oxidation for 5 min.); (e) reduction for 40 s; (f) reduction for 80 s.

In order to prove the oxidation transformation occurred within the MOF backbone, we investigated the SSNMR spectra of the UiO-68-OH crystals after different oxidation periods (Figure 4b and 4c). For example, spectrum taken after 5 seconds shows resonances at both 117 and 186 ppm, characteristic of the unique carbon atoms of compound 1 and 2. These spectra indicate the simultaneous presence of compounds 1 and 2 in the MOF backbone, confirming the oxidation transformation happened within the MOF backbone. Furthermore, the PXRD patterns of crystals after different oxidation time match that of UiO-68-OH and UiO-68-C=O (Figure S18), indicating the retention of the framework topology not only before and after full transformation, but also throughout the oxidation process.

Figure 5. Bright-field and PL images (excited with mercury lamp) of fresh UiO-68-OH (a and d), fresh UiO-68-C=O (b and e) and fresh UiO-68-OH(R) (c and f). The pictures were taken with crystals covered by a thin layer of DMF. The scale bar is 100 µm. Solid-state luminescence materials that exhibit a reversible stimulus-responsive change of luminescence color have attracted great attention in recent years for their manifold ap50,51 The reversible SCSC transformation in UiO-68plications. OH is also accompanied by a drastic change in its fluorescence spectrum. As shown in Figure 5, UiO-68-OH shows a blue emission, whereas UiO-68-C=O is almost nonfluorescent. After further reduction of UiO-68-C=O, UiO-68-OH(R) crystals have a blue emission again. This fluorescent switching behavior can be repeated for at least three cycles. We believe this phenomenon to be ascribed to the ligands, as compound 1 and 2 have totally different absorption (Figure S3) and fluorescence (Figure S4) properties matching those of UiO-68-OH and UiO-68-C=O.

ACS Paragon Plus Environment

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

Chemistry of Materials

CONCLUSSION We have reported a novel redox-switchable MOF, which can undergo a reversible SCSC transformation through a hydroquinone/quinone redox reaction. Our investigations demonstrate that − with the right choice of oxidant and reductant − the hydroquinone/quinone redox reaction can not only happen unhindered in the solution state, but also quantitatively take place in the MOF, with retention of crystallinity. We also monitored the switching process, which clearly demonstrates that the redox-triggered transformation occurred gradually in the MOF backbone and from the outsurface to the inside. Since MOFs can be designed and their properties can be tailored to a given task, we believe this study represents a general strategy to enable quantitative conversion of OMS in the solid state, by incorporation of the OMS into the robust framework of an MOF. In addition, drastic spectroscopic properties changes accompany this SCSC transformation process. We are in the process of fabricating UiO-68-OH based thin-film,52,53 with the aim of their implementation in information storage or other redox-based electronic devices.

ASSOCIATED CONTENT Supporting Information. Details of the synthesis, MOF characterization, switching study and crystallographic data. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author Email: [email protected], [email protected]

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This work was supported by National Natural Science Foundation of China (21203140, 21271172), the Research Fund for the Doctoral Program of Higher Education of China (20130141110008), the Outstanding Youth Foundation of Hubei Province (2015CFA045) and the Beijing National Laboratory for Molecular Sciences. We thank Dr. Junfeng Xiang (Institute of Chemistry, Chinese Academy of Science) for SSNMR measurement, Professor Zhiling Zhang and Miss Li Zhang (Wuhan University) for crystal images recording and all the staff members of the 3W1A beamline of the Beijing Synchrotron Radiation Facility (BSRF) for crystallographic data collection.

REFERENCES (1) Sauvage, J. P. Molecular Machines and Motors; SpringerVerlag: Berlin Heidelberg, Germany, 2001. (2) Feringa, B. L.; Brwone, W. R. Molecular Switches; WileyVCH: Weinheim, Germany, 2011. (3) Credi, A.; Silvi, S.; Venturi, M. Molecular Machines and Motors: Recent Advances and Perspectives; Springer International Publishing: Switzerland, 2014. (4) Tian, H.; Yang, S. Recent progresses on diarylethene based photochromic switches. Chem. Soc. Rev. 2004, 33, 85−97.

(5) Cao, D.; Amelia, M.; Klivansky, L. M.; Koshkakaryan, G.; Khan, S. I.; Semeraro, M.; Silvi, S.; Venturi, M.; Credi, A.; Liu, Y. Probing donor-acceptor interactions and co-conformational changes in redox active desymmetrized [2]catenanes. J. Am. Chem. Soc. 2009, 132, 1110−1122. (6) Andreasson, J.; Pischel, U. Smart molecules at workmimicking advanced logic operations. Chem. Soc. Rev. 2010, 39, 174−188. (7) Simão, C.; Mas-Torrent, M.; Crivillers, N.; Lloveras, V.; Artés, J. M.; Gorostiza, P.; Veciana, J.; Rovira, C. A robust molecular platform for non-volatile memory devices with optical and magnetic responses. Nat. Chem. 2011, 3, 359−364. (8) Zhang, J.; Zou, Q.; Tian, H. Photochromic materials: more than meets the eye. Adv. Mater. 2013, 25, 378−399. (9) McMorn, P.; Hutchings, G. J. Heterogeneous enantioselective catalysts: strategies for the immobilisation of homogeneous catalysts. Chem. Soc. Rev. 2004, 33, 108−122. (10) Trindade, A. F.; Gois, P. M. P.; Afonso, C. A. M. Recyclable stereoselective catalysts. Chem. Rev. 2009, 109, 418−514. (11) Browne, W. R.; Feringa, B. L. Light switching of molecules on surfaces. Annu. Rev. Phys. Chem. 2009, 60, 407−428. (12) Fuentes, N.; Martin-Lasanta, A.; Alvarez de Cienfuegos, L.; Ribagorda, M.; Parra, A.; Cuerva, J. M. Organic-based molecular switches for molecular electronics. Nanoscale 2011, 3, 4003−4014. (13) Gahl, C.; Schmidt, R.; Brete, D.; McNellis, E. R.; Freyer, W.; Carley, R.; Reuter, K.; Weinelt, M. Structure and excitonic coupling in self-assembled monolayers of azobenzenefunctionalized alkanethiols. J. Am. Chem. Soc. 2010, 132, 1831−1838. (14) Horike, S.; Shimomura, S.; Kitagawa, S. Soft porous crystals. Nat. Chem. 2009, 1, 695−704. (15) Férey, G.; Serre, C. Large breathing effects in threedimensional porous hybrid matter: facts, analyses, rules and consequences. Chem. Soc. Rev. 2009, 38, 1380−1399. (16) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to metal–organic frameworks. Chem. Rev. 2012, 112, 673−674. (17) Furukawa, H.; Cordova, K. E.; O'Keeffe, M.; Yaghi, O. M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. (18) Zhang, J.-P.; Liao, P.-Q.; Zhou, H.-L.; Lin, R.-B.; Chen, X.M. Single-crystal X-ray diffraction studies on structural transformations of porous coordination polymers. Chem. Soc. Rev. 2014, 43, 5789−5814. (19) Deng, H.; Olson, M. A.; Stoddart, J. F.; Yaghi, O. M. Robust dynamics. Nat. Chem. 2010, 2, 439−443. (20) Xuan, W.; Zhu, C.; Liu, Y.; Cui, Y. Mesoporous metal– organic framework materials. Chem. Soc. Rev. 2012, 41, 1677−1695. (21) Hu, Z.; Deibert, B. J.; Li, J. Luminescent metal-organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815−5840. (22) Kole, G. K.; Vittal, J. J. Solid-state reactivity and structural transformations involving coordination polymers. Chem. Soc. Rev. 2013, 42, 1755−1775. (23) Schneemann, A.; Bon, V.; Schwedler, I.; Senkovska, I.; Kaskel, S.; Fischer, R. A. Flexible metal-organic frameworks. Chem. Soc. Rev. 2014, 43, 6062−6096. (24) Chen, B.; Xiang, S.; Qian, G. Metal−organic frameworks with functional pores for recognition of small molecules. Acc. Chem. Res. 2010, 43, 1115−1124. (25) Deng, H.; Grunder, S.; Cordova, K. E.; Valente, C.; Furukawa, H.; Hmadeh, M.; Gandara, F.; Whalley, A. C.; Liu, Z.; Asahina, S.; Kazumori H.; O'Keeffe, M.; Terasaki, O.; Stoddart, J. F.; Yaghi, O. M. Large-pore apertures in a series of metal-organic frameworks. Science 2012, 336, 1018−1023. (26) Sakata, Y.; Furukawa, S.; Kondo, M.; Hirai, K.; Horike, N.; Takashima, Y.; Uehara, H.; Louvain, N.; Meilikhov, M.; Tsuruoka,

ACS Paragon Plus Environment

Chemistry of Materials

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

T.; Isoda, S.; Kosaka, W.; Sakata, O.; Kitagawa, S. Shape-memory nanopores induced in coordination frameworks by crystal downsizing. Science 2013, 339, 193−196. (27) Modrow, A.; Zargarani, D.; Herges, R.; Stock, N. The first porous MOF with photoswitchable linker molecules. Dalton Trans. 2011, 40, 4217−4222. (28) Coskun, A.; Hmadeh, M.; Barin, G.; Gandara, F.; Li, Q.; Choi, E.; Strutt, N. L.; Cordes, D. B.; Slawin, A. M.; Stoddart, J. F.; Sauvage, J. P.; Yaghi, O. M. Metal-organic frameworks incorporating copper-complexed rotaxanes. Angew. Chem. Int. Ed. 2012, 51, 2160−2163. (29) Patel, D. G.; Walton, I. M.; Cox, J. M.; Gleason, C. J.; Butzer, D. R.; Benedict, J. B. Photoresponsive porous materials: the design and synthesis of photochromic diarylethene-based linkers and a metal-organic framework. Chem. Commun. 2014, 50, 2653−2656. (30) Park, J.; Yuan, D.; Pham, K. T.; Li, J. R.; Yakovenko, A.; Zhou, H. C. Reversible alteration of CO2 adsorption upon photochemical or thermal treatment in a metal-organic framework. J. Am. Chem. Soc. 2012, 134, 99−102. (31) Brown, J. W.; Henderson, B. L.; Kiesz, M. D.; Whalley, A. C.; Morris, W.; Grunder, S.; Deng, H.; Furukawa, H.; Zink, J. I.; Stoddart, J. F.; Yaghi, O. M. Photophysical pore control in an azobenzene-containing metal-organic framework. Chem. Sci. 2013, 4, 2858−2864. (32) Kole, G. K.; Kojima, T.; Kawano, M.; Vittal, J. J. Reversible single-crystal-to-single-crystal photochemical formation and thermal cleavage of a cyclobutane ring. Angew. Chem. Int. Ed. 2014, 53, 2143−2146. (33) Luo, F.; Fan, C. B.; Luo, M. B.; Wu, X. L.; Zhu, Y.; Pu, S. Z.; Xu, W.-Y.; Guo, G.-C. Photoswitching CO2 capture and release in a photochromic diarylethene metal-organic framework. Angew. Chem. Int. Ed. 2014, 53, 9298−9301. (34) Williams, D. E.; Rietman, J. A.; Maier, J. M.; Tan, R.; Greytak, A. B.; Smith, M. D.; Krause, J. A.; Shustova, N. B. Energy transfer on demand: photoswitch-directed behavior of metalporphyrin frameworks. J. Am. Chem. Soc. 2014, 136, 11886−11889. (35) Park, J.; Feng, D.; Yuan, S.; Zhou, H. C. Photochromic metal-organic frameworks: reversible control of singlet oxygen generation. Angew. Chem. Int. Ed. 2015, 54, 430−435. (36) Wade, C. R.; Li, M.; Dincă, M. Facile deposition of multicolored electrochromic metal–organic framework thin films. Angew. Chem. Int. Ed. 2013, 52, 13377−13381. (37) Kung, C. W.; Wang, T. C.; Mondloch, J. E.; Fairen-Jimenez, D.; Gardner, D. M.; Bury, W.; Klingsporn, J. M.; Barnes, J. C.; Van Duyne, R.; Stoddart, J. F.; Wasielewski, M. R.; Farha, O. K.; Hupp, J. T. Metal–organic framework thin films composed of freestanding acicular nanorods exhibiting reversible electrochromism. Chem. Mater. 2013, 25, 5012−5017. (38) 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. (39) Nickerl, G.; Senkovska, I.; Kaskel, S. Tetrazine functionalized zirconium MOF as optical sensor for oxidizing gases. Chem. Commun. 2015, 51, 2280−2282. (40) Turunen, M.; Olsson, J.; Dallner, G. Metabolism and function of coenzyme Q. Biochim. Biophys. Acta 2004, 1660, 171−199. (41) Ma, W.; Long, Y.-T. Quinone/hydroquinonefunctionalized biointerfaces for biological applications from the macro- to nano- scale. Chem. Soc. Rev. 2014, 43, 30−41. (42) Gui, B.; Hu, G.; Zhou, T.; Wang, C. Pore surface engineering in a zirconium metal-organic framework via thiol-ene reaction. J. Solid State Chem. 2015, 223, 79−83.

Page 6 of 7

(43) 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. (44) Yee, K.-K.; Reimer, N.; Liu, J.; Cheng, S.-Y.; Yiu, S.-M.; Weber, J.; Stock, N.; Xu, Z. Effective mercury sorption by thiollaced metal–organic frameworks: in strong acid and the vapor phase. J. Am. Chem. Soc. 2013, 135, 7795−7798. (45) The 1H NMR spectrum of digested samples is totally different from that of compound 2, see Figure S6. (46) Pastoriza-Santos, I.; Liz-Marzán, L. M. Formation and stabilization of silver nanoparticles through reduction by N, Ndimethylformamide. Langmuir 1999, 15, 948−951. (47) Burrows, A. D.; Frost, C. G.; Mahon, M. F.; Richardson, C. Sulfur-tagged metal–organic frameworks and their postsynthetic oxidation. Chem. Commun. 2009, 4218−4220. (48) Wade, C. R.; Corrales-Sanchez, T.; Narayan, T. C.; Dincă, M. Postsynthetic tuning of hydrophilicity in pyrazolate MOFs to modulate water adsorption properties. Energy Environ. Sci., 2013, 6, 2172−2177. (49) Cozzolino, A. F.; Brozek, C. K.; Palmer, R. D.; Yano, J.; Li, M.; Dincă, M. Ligand redox non-innocence in the stoichiometric oxidation of Mn2(2,5-dioxidoterephthalate) (Mn-MOF-74). J. Am. Chem. Soc. 2014, 136, 3334−3337. (50) Sagara, Y.; Kato, T. Mechanically induced luminescence changes in molecular assemblies. Nat. Chem. 2009, 1, 605−610. (51) Cui, Y.; Yue, Y.; Qian, G.; Chen, B. Luminescent functional metal–organic frameworks. Chem. Rev. 2012, 112, 1126−1162. (52) Zacher, D.; Shekhah, O.; Woll, C.; Fischer, R. A. Thin films of metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1418−1429. (53) Fei, H.; Pullen, S.; Wagner, A.; Ott, S.; Cohen, S. M. Functionalization of robust Zr(IV)-based metal-organic framework films via postsynthetic ligand exchange. Chem. Commun. 2015, 51, 66−69.

ACS Paragon Plus Environment

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

Chemistry of Materials

Table of Content

ACS Paragon Plus Environment

7