Unconventional Method for Fabricating Valence Tautomeric Materials

Apr 15, 2019 - Due to the structural advantages displayed by Metal–Organic Frameworks (MOFs), integrating Valence Tautomerism (VT) systems within MO...
0 downloads 0 Views 670KB Size
Subscriber access provided by UNIV OF LOUISIANA

Communication

Unconventional Method for Fabricating Valence Tautomeric Materials: Integrating Redox Center within a Metal-Organic Framework Bao Li, Yu-meng Zhao, Angelo Kirchon, Jian-dong Pang, Xin-yu Yang, Gui-lin Zhuang, and Hong-Cai Zhou J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b02375 • Publication Date (Web): 15 Apr 2019 Downloaded from http://pubs.acs.org on April 15, 2019

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

Unconventional Method for Fabricating Valence Tautomeric Materials: Integrating Redox Center within a Metal-Organic Framework Bao Li,*,1,2 Yu-Meng Zhao,1 Angelo Kirchon,2 Jian-Dong Pang,2 Xin-Yu Yang,2 Gui-Lin Zhuang*,3 and Hong-Cai Zhou*,2,4 1 Key

laboratory of Material Chemistry for Energy Conversion and Storage, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan, Hubei 430074, People’s Republic of China. 2 Department of Chemistry, Texas A&M University, College Station, Texas 77843-3255, United States.3 Institute of Industrial Catalysis, College of Chemical Engineering, Zhejiang University of Technology, Zhejiang, 310023, People’s Republic of China.4 Department of Materials Science and Engineering, Texas A&M University, College Station, Texas 77842, United States.

Supporting Information Placeholder ABSTRACT: Due to the structural advantages displayed by Metal-Organic Frameworks (MOFs), integrating Valence Tautomerism (VT) systems within MOFs could be an effective strategy in order to break through the constraints of the traditional ones. Herein, we report the first successful integration of a VT system into a MOF termed VT-MOF-1. The structural characteristics of VT-MOF-1, such as di-nuclear cobalt-catechol clusters and solvent-accessible pores, are both innovative and novel, potentially yielding new vitality within VT field. In addition, VT-MOF-1 exhibits specific behaviors responsive to temperature and different solvent molecules as n-butanol, tert-butanol and isopropanol. The entropy values and configurations of the solvent molecules might be responsible for the tunable sensing behaviors.

One major challenge for novel multi-functional sensing or memory materials in today’s world is keeping up with the urgent demands of our rapidly evolving technology.1-3 However, some technologies, for example Valence Tautomerism (VT) , have experienced much slower advancement due to a lack of simplified synthetic strategies. 7-10 VT refers to a redox property which involves a spin transition of a metal center as well as an electron transfer between a metal center and an electro-active ligand, which has been shown to act as controllable and programable molecular sensors or switches. 11-14 Since the first mono-nuclear VT system was reported, the related field has experienced a major boom in attention.15-17 However, due to the limitation of the traditional construction strategy, most of the VT models are limited to isolated mono- and di-nuclear systems, along with a limited number of reports about 1D and 2D ones. 18-23 Inspired by the interesting examples of Spin-crossover MetalOrganic Frameworks (SCO-MOFs), 3D VT-MOFs have been of the long goal in order to prompt the development of VT systems.24-29 Not only MOFs exhibit three-dimensional regularity

of components to be infinitely arranged, but also provide a neutral environment for chemistry to be manipulated in. Considering the structural similarities and complementarities of VT systems and MOFs, if VT systems were integrated into a MOF skeleton, the composite material, termed VT-MOF, is expected to not only surmount the structural defects of VT systems by expanding the structural scope, but also exhibit composite properties(Scheme 1).30-33 The integration of two aspects into one material would take the full advantages of each systems: for VT system, the synergistic effect of transition centers would be enhanced due to the regular connection of bridging linkers and elimination of counter-anions by virtues of structural characteristics of MOFs; whilst special sensing functions would be introduced into MOFs materials via the incorporation of VT components.

Scheme 1. The illustration of the strategy for fabricating novel bistable material, termed as VT-MOF. MOF skeleton adapts the typical PtS configuration. In order to fulfill the desire of assembling a VT-MOF, the rational selection of raw components must play a key role in constructing a 3D framework with permanent porosities and ensuring the occurrence of VT phenomenon.34-36 With the above consideration in mind, multiple pyridinate ligands paired with cobalt ions and 3,5-di-tert-butylcatechol ( 3,5-DBcat ) were tentatively selected in order to construct the first VT-MOF,

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

[Co2(OAc)(3,5-DBcat)2(L)] (VT-MOF-1, L = 3,3’,5,5’tetrakis(4-pyridyl)bimesityl), which could be seen as one novel, multi-functional composite due to the special sensing behaviors responsive to temperature and different solvents. The title sample was synthesized using a solvent diffusion method. Variable-temperature single-crystal X-ray diffraction studies of VT-MOF-1 were carried out at three different temperatures ( listed in Table S1-S3 ). At 100 K, VT-MOF-1 crystallizes in tetragonal I-4c2 space group, and its asymmetric unit contains two cobalt ions, one acetate, two 3,5-DBcat units and two half L ligands (Figure 1). Each cobalt ion adopts an octahedral coordination sphere, and is occupied by two cispyridines, four oxygen atoms from two 3,5-DBcat units and one acetate unit. The two cobalt ions are combined together via the syn-syn acetate and two μ2-oxygen atoms from different 3,5-DBcat units to form the di-nuclear cobalt-catechol secondary building unit (SBU). For Co1, the Co-N and -O bond lengths fall in the range of 1.883-1.996 Å; for Co2, the ranges vary from 1.814 to 2.017 Å. All of these values are located in the middle of typical values for Co(II) and Co(III) states.37-38 No obvious difference in bond lengths for the two cobalt ions could be summarized, which might be caused by electronic delocalization phenomenon since the spin density of cobalt ions and the catechol units overlap with each other. For the two catechol units, the representative C-O bond lengths range from 1.327(2) to 1.368(1) Å, along with the related C-C bond lengths of 1.378(2) and 1.385(2) Å, are similar to the typical values for catecholate species.39-41 Therefore, the formation of di-nuclear clusters at low temperatures should be postulated as [CoIICoIII(OAc)(3,5-DBcat)2]. Each di-nuclear node are further inter-connected via tetra-pyridinate ligands to form PtS topology (Figure S1). 1D channel left in the final structure, and the solvent accessible volume in VT-MOF-1 is calculated as 59.2%. The reservation of the permanent porosities have been validated by N2 adsorption isotherms(Figure S2), even the crystallinity could not be reserved under the vacuum condition(Figure S3).

Figure 1. The partial perspective view of the framework of VTMOF-1 (H atoms have been omitted for clarity, green atoms represent the catechol units) Along with increasing temperature, the parameters of cobalt ions get closer to the typical values of cobalt(II) states. For the catechol chelated onto Co(1), C-O and C-C bond lengths indicate the species of semiquinonate radical, while the second retains the catechol di-anion. Therefore, the formula of the di-nuclear node should be [CoIICoII(OAc)(3,5-DBcat)(3,5-DBsq)]( 3,5-DBsq = 3,5-di-tert-butylsemiquinone) at high temperature regions. The origination of two different states should be ascribed to the redox process that occurs as a one electron transfers from catechol to Co(1) in response to the variation of temperature(Figure 2). Furthermore, VT behavior of VT-MOF-1 could be also detected by variable temperature EPR measurements (Figure S8). The response signal around g=2.00 at room temperature has been vanished at low temperature due to the occurrence of VT process.42-43

Figure 2. The illustration of VT process in dinuclear bistable centers along with the definite transition positions( the black and blue octahedrons represent the Co(III) and Co(II) state; Cat and Sq- represent catechol and semiquinone units )

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 Figure 3. Variable temperature MT values vs. T plots of different samples. Inset: EPR spectrum of VT-MOF-1 at 100 K ( blue line ) and 293K ( red line ). The related VT process could be also validated by variabletemperature magnetic properties (Figure 3). At room temperature, the MT value is about 5.76 cm3 K mol−1, smaller than the expected value for the system containing two cobalt(II) ions and one semiquinonate radical anions, which might be caused by the antiferromagnetic interaction in the dinuclear cluster (Table S9).45-46 Along with the temperature decreasing, the MT values remains relatively constant until 275 K and then decreases gradually to 3.11 cm3 K mol−1 around 150 K. Next, the MT values are stable to 50 K, and decrease gradually to 1.54 cm3 K mol−1 at 2 K. The first decrease of the MT values should be attributed to the VT transition. Around 150 K, the experimental values are in agreement with the system of [CoIICoIII(OAc)(3,5-DBcat)2]. The second decrease in the low temperature regions might be assigned to the zero-field splitting and/or magnetic anisotropy of a paramagnetic cobalt(II) ion. The existence of cobalt(II) ion with large anisotropy at low temperature could be validated by the plots of M vs. H/T and the presentation of slow magnetic relaxation (Figure 3, S4-7). Furthermore, the VT transition could be also validated by the variable temperature EPR spectra (Figure 3: inset, S8) The VT process in VT-MOF-1 could be affected not only by the temperature, but also by the different guest molecules since the reservation of porosities in framework ( Figure 3 ). The transition temperature of samples with n-hexane, tert-Butanol and isopropanol obviously shift to higher temperature regions, compared to the minor movement of samples with n-butanol. Theoretical simulations for the different transition behaviors had been carried out according to the van’s Hoff equation ( eq. S2 and S3 ), and the fitted value parameters for different samples are listed in Table S7. The incorporated solvent molecules exhibit the different configurations and entropy values, which might make significant changes in the total entropy of the resulted composite systems. For example, for the composite systems containing nbutanol, tert-butanol and isopropanol, the equilibrium temperature ( T1/2 ) moves to the higher temperature regions along with the decrease trend of entropy values of the solvent molecules ( Table S8 ). Although the different host-guest interaction might be also of the origination of the changes, the matched tendencies between transition regions and entropy values provide the other insight into the VT-MOF systems. The experimental and theoretical results clearly manifest the important role of guest solvents to VT transitions since the reservation of porous structures, and further emphasize the advantages of the combination of MOFs and VT system, which is consistent with our assumptions.

Figure 4. Up :View of the frontier singly-occupied molecular orbitals (SOMOs) of 173 in low-spin state of S = 3/2 and highspin state of S = 2×(3/2)+1/2; Bottom: The total spin density of dinuclear cluster (Co1: left; Co2: right. density cuttoff = 0.030au). To present deep illustration of the VT process in VT-MOF-1, spin-polarized density functional theory calculations were performed(See ESI). Considering the dinuclear VT SBUs and high cost of computation for the whole framework, the theoretical model was simplified to the components of [Co2(OAc)(3,5DBcat)2(pyridine)4], and one electron charge transfer takes place in the di-nuclear cluster. From the view of thermodynamics, the free-energy differences between the two spin states of S = 3/2 and S = 2×(3/2)+1/2 is 4.50 kcal/mol, indicating the easy transition between these spin states. Broken-symmetry calculations for S = 3/2 and S = 2 × (3/2)+1/2 were also identified (Table S4). For S = 3/2, spin densities for Co1 and Co2 atoms were calculated as 0.049 and 2.482 respectively. Meanwhile, the spin values for all oxygen atoms on the catechol units fall in the range of 0.01 to 0.13. The whole spin in the S=3/2 state mostly concentrates on Co2 ion (Figure 4 and S9). In this way, Co1 features a +3 valence state; Co2 ion should exhibit high-spin state. For S= 2×(3/2)+1/2, spin values for Co1 and Co2 atoms became 2.531 and 2.532, indicating the high-spin configuration.The spin densities of oxygen atoms on the two electro-active ligands exhibit the slight difference due to the electron delocalization caused by the inter-connection modes of di-nuclear cobalt. The two bridging μ2-O ( O1 and O3 ) atoms tend to delocalize the spin distribution of Co(II) centers, and a certain degree of spin density were still attained on these oxygen atoms, in consistent with the observation of X-ray diffraction studies. However, the different spin distributions of carbon rings in two electro-active ligands could be observed, clearly indicating

ACS Paragon Plus Environment

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

the different forms. Moreover, the evident difference of dipole moment ( 10.058 D for S=3/2 state, 7.024 D for S= 2×(3/2)+1/2 state ) also validates the existence of electronic transition between Co ion and catechol ligand. Viewed from the frontier singlyoccupied molecular orbitals (Figure S9-S10), together with geometrical configurations, the transitional mechanism in VTMOF-1 belongs to VT process with class II mixed-valences species.47-48 All of these theoretical investigations exhibit the according results with the above experimental data, clearly illustrating the occurrence of a redox process in the dinuclear SBUs. In conclusion, the first sample integrating VT within MOF has been successfully constructed. The combination of two different fields has not only expanded the scope of VT phenomenon including structural characteristics and synthetic methods, but also endowed MOFs with novel sensing ability. The resulted material integrates multiple properties such as charge transfer related to temperature, slow magnetic relaxation, permanent porosity, and exhibits potential as a sensing material to different guest molecules. In terms of the structural characteristics and composite performance, all can be thought of as an innovative work in the field of VT and MOFs, which could provide scientific significance for guiding the preparation of related materials in the future.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: . The detailed experimental methods, crystal data, theoretical calculation methods and results (PDF)

AUTHOR INFORMATION Corresponding Author * E-mail: [email protected] (B.L.). * E-mail: [email protected] (H.-C. Z.). * E-mail: [email protected] (G.-l.Z.).

ORCID Bao Li: 0000-0003-1154-6423 Angelo Kirchon: 0000-0003-1082-9739 Hong-Cai Zhou: 0000-0002-9029-3788

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENTS This work was partially supported by the financial supports of National Science Foundation of China ( 21471062 and 21671172 ), the Center for Gas Separations Relevant to Clean Energy Technologies, an Energy Frontier Research Center (EFRC) funded by the U.S. Department of Energy (DOE), Office of Science, and Office of Basic Energy Sciences (DESC0001015), Office of Fossil Energy, the National Energy Technology

Laboratory (DE-FE0026472), and the Robert A. Welch Foundation through a Welch Endowed Chair to HJZ (A-0030).

REFERENCES (1) Feng, D.; Liu, T.; Su, J.; Bosch, M.; Wei, Z.; Wan, W.; Yuan, D.; Chen, Y.; Wang, X.; Wang, K.; Lian, X.; Gu, Z.; Park, J.; Zou, X.; Zhou, H. Stable metal-organic frameworks containing single-molecule traps for enzyme encapsulation Nat. commun. 2015, 6, 5979. (2) O Keeffe, M.; Yaghi, O. M. Deconstructing the Crystal Structures of Metal–Organic Frameworks and Related Materials into Their Underlying Nets. Chem. Rev. 2011, 112, 675-702. (3) Stock, N.; Biswas, S. Synthesis of Metal-Organic Frameworks (MOFs): Routes to Various MOF Topologies, Morphologies, and Composites. Chem. Rev. 2011, 112, 933-969. (4) Vázquez-Mera, N. A.; Novio, F.; Roscini, C.; Bellacanzone, C.; Guardingo, M.; Hernando, J.; Ruiz-Molina, D. Switchable colloids, thinfilms and interphases based on metal complexes with non-innocent ligands: the case of valence tautomerism and their applications. J. Mater. Chem. C. 2016, 4, 5879-5889. (5) Drath, O.; Boskovic, C. Switchable cobalt coordination polymers: Spin crossover and valence tautomerism. Coord. Chem. Rev. 2018, 375, 256-266. (6) Ok-Sang, J.; Cortlandt, G. P. Photomechanical Polymers. Synthesis and Characterization of a Polymeric Pyrazine-Bridged Cobalt Semiquinonate-Catecholate Complex. J. Am. Chem. Soc. 1994, 116, 22292230. (7) Dei, A. Photomagnetic Effects in Polycyanometallate Compounds: An Intriguing Future Chemically Based Technology? Angew. Chem. Int. Ed. 2005, 44, 1160-1163. (8) Guardingo, M.; González-Monje, P.; Novio, F.; Bellido, E.; Busqué, F.; Molnár, G.; Bousseksou, A.; Ruiz-Molina, D. Synthesis of Nanoscale Coordination Polymers in Femtoliter Reactors on Surfaces. ACS Nano. 2016, 10, 3206-3213. (9) Nador, F.; Novio, F.; Ruiz-Molina, D. Coordination Polymer Particles with ligand-centred pH-responses and spin transition. Chem. Commun. 2014, 50, 14570-14572. (10) Sato, O.; Tao, J.; Zhang, Y. Control of Magnetic Properties through External Stimuli. Angew. Chem. Int. Ed. 2007, 46, 2152-2187. (11) Tezgerevska, T.; Alley, K. G.; Boskovic, C. Valence tautomerism in metal complexes: Stimulated and reversible intramolecular electron transfer between metal centers and organic ligands. Coordin. Chem. Rev. 2014, 268, 23-40. (12) Drath, O.; Gable, R. W.; Moubaraki, B.; Murray, K. S.; Poneti, G.; Sorace, L.; Boskovic, C. Valence Tautomerism in One-Dimensional Coordination Polymers. Inorg. Chem. 2016, 55, 4141-4151. (13) Kanegawa, S.; Shiota, Y.; Kang, S.; Takahashi, K.; Okajima, H.; Sakamoto, A.; Iwata, T.; Kandori, H.; Yoshizawa, K.; Sato, O. Directional Electron Transfer in Crystals of [CrCo] Dinuclear Complexes Achieved by Chirality-Assisted Preparative Method. J. Am. Chem. Soc. 2016, 138, 14170-14173. (14) Drath, O.; Gable, R. W.; Poneti, G.; Sorace, L.; Boskovic, C. One Dimensional Chain and Ribbon Cobalt-Dioxolene Coordination Polymers: A New Valence Tautomeric Compound. Cryst. Growth Des. 2017, 17, 3156-3162. (15) Alley, K. G.; Poneti, G.; Robinson, P. S. D.; Nafady, A.; Moubaraki, B.; Aitken, J. B.; Drew, S. C.; Ritchie, C.; Abrahams, B. F.; Hocking, R. K.; Murray, K. S.; Bond, A. M.; Harris, H. H.; Sorace, L.; Boskovic, C. Redox Activity and Two-Step Valence Tautomerism in a Family of Dinuclear Cobalt Complexes with a Spiroconjugated Bis(dioxolene) Ligand. J. Am. Chem. Soc. 2013, 135, 8304-8323.

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 (16) Witt, A.; Heinemann, F. W.; Sproules, S.; Khusniyarov, M. M. Modulation of Magnetic Properties at Room Temperature: CoordinationInduced Valence Tautomerism in a Cobalt Dioxolene Complex. Chem. Eur. J. 2014, 20, 11149-11162. (17) Imaz, I.; Maspoch, D.; Rodríguez-Blanco, C.; Pérez-Falcón, J. M.; Campo, J.; Ruiz-Molina, D. Valence-Tautomeric Metal-Organic Nanoparticles. Angew. Chem. Int. Ed. 2008, 47, 1857-1860. (18) Rupp, F.; Chevalier, K.; Graf, M.; Schmitz, M.; Kelm, H.; Grün, A.; Zimmer, M.; Gerhards, M.; van Wüllen, C.; Krüger, H.; Diller, R. Spectroscopic, Structural, and Kinetic Investigation of the Ultrafast Spin Crossover in an Unusual Cobalt(II) Semiquinonate Radical Complex. Chem. - Eur. J. 2017, 23, 2119-2132. (19) Lannes, A.; Suffren, Y.; Tommasino, J. B.; Chiriac, R.; Toche, F.; Khrouz, L.; Molton, F.; Duboc, C.; Kieffer, I.; Hazemann, J.; Reber, C.; Hauser, A.; Luneau, D. Room Temperature Magnetic Switchability Assisted by Hysteretic Valence Tautomerism in a Layered TwoDimensional Manganese-Radical Coordination Framework. J. Am. Chem. Soc. 2016, 138, 16493-16501. (20) Affronte, M.; Beni, A.; Dei, A.; Sorace, L. Valence tautomerism interconversion triggers transition to stable charge distribution in solid polymeric cobalt–polyoxolene complexes. Dalton Trans. 2007, 5253. (21) Novio, F.; Campo, J.; Ruiz-Molina, D. Controlling Spin Transition in One-Dimensional Coordination Polymers through Polymorphism. Inorg. Chem. 2014, 53, 8742-8748. (22) Novio, F.; Evangelio, E.; Vazquez-Mera, N.; González-Monje, P.; Bellido, E.; Mendes, S.; Kehagias, N.; Ruiz-Molina, D. Robust spin crossover platforms with synchronized spin switch and polymer phase transition. Sci. Rep. 2013, 3, 1708. (23) Abhervé, A.; Grancha, T.; Ferrando-Soria, J.; Clemente-León, M.; Coronado, E.; Waerenborgh, J. C.; Lloret, F.; Pardo, E. Spin-crossover complex encapsulation within a magnetic metal-organic framework. Chem. Commun. 2016, 52, 7360-7363. (24) Zhao, T.; Boldog, I.; Spasojevic, V.; Rotaru, A.; Garcia, Y.; Janiak, C. Solvent-triggered relaxative spin state switching of [Fe(HB(pz)3)2] in a closed nano-confinement of NH2-MIL-101(Al). J. Mater. Chem. C. 2016, 4, 6588-6601. (25) Guardingo, M.; Busqué, F.; Novio, F.; Ruiz-Molina, D. Design and Synthesis of a Noninnocent Multitopic Catechol and Pyridine Mixed Ligand: Nanoscale Polymers and Valence Tautomerism. Inorg. Chem. 2015, 54, 6776-6781. (26) González-Monje, P.; Novio, F.; Ruiz-Molina, D. Covalent Grafting of Coordination Polymers on Surfaces: The Case of Hybrid Valence Tautomeric Interphases. Chem. Eur. J. 2015, 21, 10094-10099. (27) Poneti, G.; Poggini, L.; Mannini, M.; Cortigiani, B.; Sorace, L.; Otero, E.; Sainctavit, P.; Magnani, A.; Sessoli, R.; Dei, A. Thermal and optical control of electronic states in a single layer of switchable paramagnetic molecules. Chem. Sci. 2015, 6, 2268-2274. (28) Vázquez-Mera, N. A.; Roscini, C.; Hernando, J.; Ruiz-Molina, D. Liquid-Filled Valence Tautomeric Microcapsules: A Solid Material with Solution-Like Behavior. Adv. Funct. Mater. 2015, 25, 4129-4134. (29) Li, B.; Chen, L.; Wei, R.; Tao, J.; Huang, R.; Zheng, L.; Zheng, Z. Thermally Induced and Photoinduced Valence Tautomerism in a TwoDimensional Coordination Polymer. Inorg. Chem. 2011, 50, 424-426. (30) Tulchinsky, Y.; Hendon, C. H.; Lomachenko, K. A.; Borfecchia, E.; Melot, B. C.; Hudson, M. R.; Tarver, J. D.; Korzyński, M. D.; Stubbs, A. W.; Kagan, J. J.; Lamberti, C.; Brown, C. M.; Dincă, M. Reversible Capture and Release of Cl2 and Br2 with a Redox-Active Metal–Organic Framework. J. Am. Chem. Soc. 2017, 139, 5992-5997. (31) Bin-Salamon, S.; Brewer, S.; Franzen, S.; Feldheim, D. L.; Lappi, S.; Shultz, D. A. Supramolecular Control of Valence-Tautomeric Equilibrium on Nanometer-Scale Gold Clusters. J. Am. Chem. Soc. 2005, 127, 5328-5329.

(32) 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. (33) Min, K. S.; Dipasquale, A.; Rheingold, A. L.; Miller, J. S. Roomtemperature spin crossover observed for [(TPyA)FeII(DBQ(2))Fe(II)(TPyA)](2+) [TPyA = Tris(2-pyridylmethyl)amine; DBQ(2-) = 2,5-Di-tert-butyl-3,6-dihydroxy-1,4-benzoquinonate]. Inorg. Chem. 2007, 46, 1048-1050. (34) Lu, W.; Wei, Z.; Yuan, D.; Tian, J.; Fordham, S.; Zhou, H. Rational Design and Synthesis of Porous Polymer Networks: Toward High Surface Area. Chem. Mater. 2014, 26, 4589-4597. (35) Natarajan, R.; Savitha, G.; Dominiak, P.; Wozniak, K.; Moorthy, J. N. Corundum, Diamond, and PtS Metal-Organic Frameworks with a Difference: Self-Assembly of a Unique Pair of 3-ConnectingD2dSymmetric 3,3’,5,5’-Tetrakis(4-pyridyl)bimesityl. Angew. Chem. Int. Ed. 2005, 44, 2115-2119. (36) Seth, S.; Venugopalan, P.; Moorthy, J. N. Porous Coordination Polymers of Diverse Topologies Based on a Twisted Tetrapyridylbiaryl: Application as Nucleophilic Catalysts for Acetylation of Phenols. Chem. Eur. J. 2015, 21, 2241-2249. (37) Liu, W.; Holden Thorp, H. Bond Valences Sum Analysis of MetalLigand Bond Lengths in Metalloenzymes and Model Complexes. 2 . Refined Distances and Other Enzymes. Inorg. Chem. 1993, 32, 4102-4105. (38) Holden Thorp, H. Bond Valences Sum Analysis of Metal-Ligand Bond Lengths in Metalloenzymes and Model Complexes. Inorg. Chem. 1992, 31, 1585-1588. (39) Ribeiro, M. A.; Stasiw, D. E.; Pattison, P.; Raithby, P. R.; Shultz, D. A.; Pinheiro, C. B. Cryst.Growth  Des. 2016, 16, 2385-2393. (40) Witt, A.; Heinemann, F. W.; Khusniyarov, M. M. Bidirectional photoswitching of magnetic properties at room temperature: ligand-driven light-induced valence tautomerism. Chem. Sci. 2015, 6, 4599-4609. (41) Graf, M.; Wolmershäuser, G.; Kelm, H.; Demeschko, S.; Meyer, F.; Krüger, H. Temperature-Induced Spin-Transition in a Low-Spin Cobalt(II) Semiquinonate Complex. Angew. Chem. Int. Ed. 2010, 49, 950-953. (42) Nurdin, L.; Spasyuk, D. M.; Piers, W. E.; Maron, L. Reactions of Neutral Cobalt(II) Complexes of a Dianionic Tetrapodal Pentadentate Ligand: Cobalt(III) Amides from Imido Radicals. Inorg. Chem. 2017, 56, 4157-4168. (43) Clarke, R. M.; Hazin, K.; Thompson, J. R.; Savard, D.; Prosser, K. E.; Storr, T. Electronic Structure Description of a Doubly Oxidized Bimetallic Cobalt Complex with Proradical Ligands. Inorg. Chem. 2015, 55, 762-774. (44) Andrez, J.; Guidal, V.; Scopelliti, R.; Pécaut, J.; Gambarelli, S.; Mazzanti, M. Ligand and Metal Based Multielectron Redox Chemistry of Cobalt Supported by Tetradentate Schiff Bases. J. Am. Chem. Soc. 2017, 139, 8628-8638. (45) Evangelio, E.; Ruiz-Molina, D. Valence tautomerism: More actors than just electroactive ligands and metal ions. C. R. Chimie 2008, 11, 1137-1154. (46) Dapporto, P.; Dei, A.; Poneti, G.; Sorace, L. Complete Direct and Reverse Optically Induced Valence Tautomeric Interconversion in a Cobalt-Dioxolene Complex. Chem. - Eur. J. 2008, 14, 10915-10918. (47) Robin, M. B.; Day, P. Mixed Valence Chemistry-A survey and Classification. Adv. Inorg. Chem. Radiochem. 1967, 10, 247–422. (48) Bendix, J.; Clark, K. M. Delocalization and Valence Tautomerism in Vanadium Tris(iminosemiquinone) Complexes. Angew. Chem. Int. Ed. 2016, 55, 2748 –2752.

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

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