A Bioinspired Metal-Organic Framework for Trace CO2

A Bioinspired Metal-Organic Framework for Trace CO2...
0 downloads 0 Views 898KB Size
Subscriber access provided by Kaohsiung Medical University

Communication 2

A Bioinspired Metal-Organic Framework for Trace CO Capture Caitlin E Bien, Kai K Chen, Szu-Chia Chien, Benjamin R. Reiner, Li-Chiang Lin, Casey R. Wade, and W. S. Winston Ho J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b06109 • Publication Date (Web): 26 Sep 2018 Downloaded from http://pubs.acs.org on September 26, 2018

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

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

Page 1 of 5 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

A Bioinspired Metal-Organic Framework for Trace CO2 Capture Caitlin E. Bien,† Kai K. Chen,‡ Szu-Chia Chien,# Benjamin R. Reiner,† Li-Chiang Lin,‡ Casey R. Wade,*,† and W. S. Winston Ho‡,# †

Department of Chemistry and Biochemistry, The Ohio State University, Columbus, OH 43210, United States William G. Lowrie Department of Chemical and Biomolecular Engineering, The Ohio State University, Columbus, OH 43210, United States # Department of Materials Science and Engineering, The Ohio State University, Columbus, OH 43210, United States Supporting Information Placeholder ‡

ABSTRACT:

A Zn benzotriazolate MOF, [Zn(ZnO2CCH3)4(bibta)3] (1, bibta2– = 5,5’-bibenzotriazolate), has been subjected to a mild CH3CO2–/HCO3– ligand exchange procedure followed by thermal activation to generate nucleophilic Zn– OH groups that resemble the active site of α-carbonic anhydrase. The postsynthetically modified MOF, [Zn(ZnOH)4(bibta)3] (2*), exhibits excellent performance for trace CO2 capture and can be regenerated at mild temperatures. IR spectroscopic data and DFT calculations reveal that inter-cluster hydrogen bonding interactions augment a Zn–OH/Zn–O2COH fixation mechanism.

Direct air capture of CO2 has been considered as a means of halting or reversing the effects of anthropogenic CO2 emissions on climate change.1–3 Moreover, CO2 is acutely toxic at concentrations above 5000 ppm, leading to the need for efficient CO2 removal systems in confined environments such as submarines and spacecraft.4,5 Liquid amine scrubbing and causticization are well-studied technologies for trace CO2 removal.1,2 However, these processes are energy-intensive and may not be economically viable on large scale. As a result, solid adsorbents have been explored as more energy efficient alternatives.6,7 Aminefunctionalized porous oxides, polymers, and metal-organic frameworks (MOFs) have been widely studied for trace CO2 capture.8–15 However, in some cases, thermal regeneration can result in adsorbent degradation via amine oxidation or volatilization.14,16,17 Recently developed hybrid ultramicroporous materials (HUMs) such as SIFSIX-3-Cu, TIFSIX-3-Ni, and NbOFFIVE-1-Ni employ a combination of optimized pore size and strong physisorptive interactions that enable adsorption of CO2 at very low pressures while maintaining mild regeneration temperatures.18–22 Although significant advances have been made toward developing solid adsorbents for trace CO2 capture applications, further work is needed to improve capacity, stability, selectivity, and regeneration requirements. Compared to amine-functionalized adsorbents, porous materials containing nucleophilic hydroxyl binding sites have been less studied for CO2 capture applications.23–25 Zhang and co-workers recently demonstrated that MOFs containing CoIII–OH or MnIII– OH groups exhibit good CO2 uptake under simulated flue gas conditions, but only modest capacities were observed at atmospheric CO2 concentrations.26 These materials were prepared by postsynthetic oxidation of CoII/MnII-based MOFs, limiting the approach to metals with accessible MII/III redox couples. Inspired by the CO2 fixation mechanism of α-carbonic anhydrases (α-CA), we considered that the use of more strongly nucleophilic Zn–OH

groups analogous to the enzyme active site might provide materials with improved low-pressure CO2 uptake. Although Chen and co-workers have reported the use of Zn-based MOFs to mediate the conversion of CO2 to HCO3– in aqueous solution, they only observed substoichiometric turnovers, and the generation of Zn– OH groups for gas-solid phase adsorption has not been investigated.27 Herein we describe a mild and convenient postsynthetic modification protocol for generating biomimetic Zn–OH species within the benzotriazolate MOF CFA-1.28 The resulting material displays excellent CO2 adsorption at ultralow concentrations resulting from a CO2/HCO3– chemisorption mechanism and cooperative inter-cluster hydrogen bonding interactions reminiscent of secondary coordination sphere interactions in α-CA.

Figure 1. a) Synthesis of 1 and structures of the framework and Zn(ZnOAc)4 SBUs. b) Synthesis of 2 and 2* and mechanism of reversible CO2 fixation.

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

Zn(ZnOAc)4(bibta)3 (1, OAc = CH3COO–) has been synthesized as a light brown microcrystalline powder by modification of the reported procedure (Figure 1a).28 The powder X-ray diffraction (PXRD) pattern of 1 closely matches the pattern simulated from the reported single crystal structure of CFA-1, confirming that the modified synthesis conditions have no impact on the framework structure (Figure S1). 1 contains Kuratowski-type secondary building units (SBUs) with an octahedrally coordinated, central zinc site (ZnC) and crystallographically distinct, peripheral zinc sites (ZnA and ZnE) that bear a strong resemblance to the α-carbonic anhydrase active site.29 Furthermore, the ZnA–OAc and ZnE–OAc groups are exposed to the internal surface of the framework channels and are present in high density (~3.16 mmol g-1).

THF molecules per formula unit) and loss of CO2 from decomposition of the putative Zn–O2COH groups both occur during the initial mass loss. Samples of 2 were activated by heating at 100 °C under high vacuum to generate 2*. N2 adsorption isotherms measured at 77 K for 2* and a similarly activated sample of 1 provided calculated Brunauer-Emmett-Teller (BET) surface areas of 2522 and 2075 m2 g-1, respectively (Figure S4). The larger surface area and slight increase in average pore diameter of 2* is consistent with the exchange of OAc– for smaller OH– ligands (Figure S5). CO2, N2, and O2 adsorption isotherms measured at 300 K reveal that 2* exhibits remarkably steep CO2 uptake at low pressures but negligible N2 and O2 adsorption (Figure 2a). On the other hand, 1 shows almost no CO2 uptake in the low-pressure region, supporting the generation of Zn–OH chemisorption sites in 2*. At 0.4 mbar, which is roughly equivalent to the concentration of CO2 in Earth’s atmosphere (~400 ppm), 2* exhibits a CO2 capacity of 2.20 mmol g-1 (Figure 2b). This uptake exceeds that of the CoIII– OH functionalized MAF-27ox (~1.1 mmol g-1) and NbOFFIVE-1Ni (1.3 mmol/g) but is lower than that observed for the aminefunctionalized MOFs Mg2(dobdc)-N2H4 (3.89 mmol g-1) and Mg2(dobpdc)-en (2.83 mmol g-1).10,13,20,26 CO2 adsorption isotherms were measured over the 300-353 K range (Figure S8) and the isosteric heat of adsorption (Qst) was calculated as a function of the amount of CO2 adsorbed (n) using the Clausius-Clapeyron equation (Figure 2c, see Supporting Information for details). The plot shows an unusual increase in Qst up to ~2.0 mmol g-1 followed by a sharp decrease to a value consistent with pore filling and weak physisorptive processes (Qst = 24 kJ mol-1). The large uncertainties in the calculated Qst values suggest that the steep low-pressure CO2 uptake causes equilibration and/or pressure measurement issues which may be responsible for the unusual behavior at low coverage. Nevertheless, the observed maximum of 71 kJ mol-1 reflects a strong chemisorption mechanism and is consistent with enthalpies of reaction determined for solution-state CO2 fixation by homogeneous transition metal hydroxide complexes (49-61 kJ mol-1).32–35

Figure 2. a) CO2, N2 and O2 adsorption isotherms of 1 and 2* up to 1 bar at 300 K. b) Low pressure region of the CO2 adsorption isotherm of 2*. c) Isosteric heat of adsorption (Qst) for 2* as a function of the amount of CO2 adsorbed (n). The shaded area represents the uncertainty calculated from the linear fit of the isosteres (see Supporting Information). Attempts to carry out direct Cl–/OH– exchange in MOFs containing Kuratowski-type Zn SBUs have been reported to result in framework decomposition.30,31 Considering that CO2 capture should be reversible via Zn–OH/Zn–O2COH interconversion, we decided to examine OAc–/HCO3– exchange in 1 as an indirect route to generate Zn–OH groups. Gratifyingly, the ligandexchanged MOF 2 was conveniently prepared by soaking 1 in an aqueous solution of NaHCO3 (0.1 M, pH 8.0-8.3) followed by washing with H2O and THF (Figure 1b). PXRD analysis of the product indicates that 2 retains crystallinity and does not undergo any major structural changes (Figure S1). Only residual amounts of HOAc (< 0.1:1 HOAc:H2bibta) are observed in the 1H NMR spectra of acid-digested samples of 2, confirming the successful exchange of the OAc– ligands. (Figure S2). Thermogravimetric analysis (TGA) shows a steep 40 % mass loss up to 100 °C followed by a plateau until the onset of framework decomposition around 400 °C (Figure S3). The apparent stability of 2 up to 400 °C indicates that volatilization of guest solvent molecules (~ 8

Figure 3. Adsorption-desorption cycling of 2* under simulated air conditions (395 ppm CO2) showing changes in sample mass (red) and temperature swing (blue). Thermal swing adsorption (TSA) measurements were carried out to interrogate the stability of 2* to adsorption-desorption cycling and determine its capacity for CO2 uptake from simulated air. At 300 K, the MOF exhibits CO2 uptakes of 9 and 14 wt % upon exposure to 5 % CO2 in N2 and pure CO2, respectively (Figures S10 and S11). Upon cycling with a simulated air mixture (395 ppm CO2, 21 % O2, N2 balance), 2* shows 5.8 wt % uptake (Figure 3). Only small losses in adsorption capacity were ob-

ACS Paragon Plus Environment

Page 2 of 5

Page 3 of 5 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 served after 5-10 adsorption-desorption cycles. Interestingly, the CO2 uptakes measured for 2* in the TSA experiments are lower than predicted from the single component isotherms (9.7 % at 0.4 mbar, 13.2 % at 50 mbar, and 24.6 % at 1 bar). The origin of this disparity is unclear, but similar behavior has been observed for other MOF CO2 sorbents.12,13 The performance of 2* under simulated air conditions is slightly better than the 4.6 wt % uptake observed for Mg2(dobdc)-mmen, but modest compared to Mg2(dobpdc)-en (11.8 wt %).9,10 However, the temperature required to regenerate 2* (100 °C) is considerably lower than for these amine-grafted materials (150 °C). Column breakthrough curves were also measured for 2* at 298 K using dry and humidified gas mixtures containing 3890 ppm CO2 in N2 (Figures S13 and S14). Under anhydrous conditions, the CO2 uptake of the MOF adsorbent in the column was in the range of 2.7-3.2 mmol g1 over three runs, close to the expected value based on the single component isotherm (2.45 mmol g-1). However, with a watersaturated gas stream, CO2 breakthrough occurred almost immediately, indicating that 2* does not perform well under humid conditions. Nevertheless, the material is stable to ambient conditions and maintains its CO2 uptake capacity in single component isotherm measurements after exposure to air for extended periods of time (Figure S15).

Figure 4. Evolution of the IR spectrum of 2-CO2 upon heating under a N2 atmosphere. Diffuse-reflectance IR spectroscopy (DRIFTS) experiments were carried out to gain further insight into the CO2 adsorption mechanism of 2*. At room temperature, the DRIFTS spectrum of 2-CO2 features bands attributable to Zn–O2COH groups at 1340, 1501, 1585, and 2657 cm-1 (Figure 4). These assignments were confirmed by comparison with the IR spectrum of the isotopologue 2-13CO2 (Figure S16). Upon heating under a flow of dry N2, these bands gradually disappear concomitant with the emergence of a new band at 3692 cm-1 corresponding to the O–H stretch of the regenerated Zn–OH groups. The bands appearing at 1340 and 1585 cm-1 have been assigned as symmetric and asymmetric C–O stretches of the Zn–O2COH groups.36,37 Based on observed trends in homogeneous transition metal bicarbonate complexes, the appearance of the asymmetric C–O stretch at 1585 cm-1 suggests that the bicarbonate ligands in 2-CO2 should adopt a bidentate coordination mode.38,39 Intriguingly, the low energy O–H stretch observed at 2657 cm-1 for 2-CO2 indicates that the Zn–O2COH groups are engaged in hydrogen-bonding interactions. This fea-

ture has further prompted assignment of the band at 1501 cm-1 as an O–H–O bend.36,37 These spectral features are consistent with those observed for hydrogen-bonded dimers of transition metal bicarbonate complexes.40–42

Figure 5. a) Inter-cluster hydrogen bonding interactions resulting from sequential CO2 chemisorption at the ZnE–OH sites. b) Optimized structures of 2-A, 2-E1-proximal, and 2-E2. The structure of CFA-1 shows that the distance between ZnE centers of neighboring SBUs (~7.7 Å) is sufficiently short to accommodate inter-cluster hydrogen bonding between coordinated ligands (Figure 5a). Thus we considered that these interactions might play a significant role in CO2 capture by 2* in the same way that secondary coordination sphere interactions regulate the activity of α-CA.43 Periodic DFT calculations were used to explore the nature and strength of the putative hydrogen bonding interactions. All optimized structures and calculated CO2 binding energies can be found in the Supporting Information. Three different hydrogen bonding orientations were investigated for structures containing one ZnE–O2COH group per pair of adjacent ZnE sites (Figure S17). In the lowest energy structure, 2-E1-proximal, the ZnE–O2COH group engages in a strong hydrogen bonding interaction with the nearby ZnE–OH sites, yielding a CO2 binding energy of -71.4 kJ mol-1. The optimized structure containing two adjacent ZnE–O2COH groups (2-E2) shows the formation of strongly hydrogen-bonded bicarbonate dimers with a binding energy of -68 kJ per mol of CO2. In both cases, the formation of inter-cluster hydrogen bonding interactions results in a significant increase in CO2 binding energy versus the isolated ZnA sites (37.5 kJ mol-1, 2-A). The heat of adsorption plot in Figure 2c shows a steep drop-off around 2 mmol g-1, which is in between the expected uptake of one CO2 per pair of ZnE sites (~1.36 mmol g-1) and binding at both adjacent sites (2.73 mmol g-1). This ambiguity coupled with the similar CO2 binding energies calculated for 2-E1-proximal and 2-E2 make it difficult to presently establish which of these hydrogen bonding motifs is operative in 2-CO2. However, it is clear that cooperative hydrogen bonding interactions play an important role in CO2 adsorption by 2*, and further investigations are underway to more fully elucidate their nature and implications. In summary, a mild postsynthetic ligand exchange procedure followed by thermal activation has been used to generate nucleophilic Zn–OH sites in CFA-1. A CO2/HCO3– chemisorption process aided by inter-cluster hydrogen bonding interactions enables CO2 uptake at partial pressures relevant for trace CO2 capture from air. Although cooperative binding mechanisms have been implicated in gas adsorption by MOFs, they remain rare.44–47 The present findings illustrate the potential of using bioinspired strate-

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

gies to design cooperative binding motifs and develop next generation materials for gas adsorption applications.

(18)

(19)

ASSOCIATED CONTENT

(20)

Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Experimental details, analytical data, ATR-IR spectra, gas sorption isotherms and data, TGA and thermal swing adsorption data, breakthrough experiment data, computational details and results

(21)

(22)

(23)

AUTHOR INFORMATION Corresponding Author

(24)

*E-mail: [email protected]

(25)

Notes The authors declare no competing financial interests.

(26)

ACKNOWLEDGMENTS C.R.W acknowledges The Ohio State University for start-up funding. We thank Prof. Nick Brunelli and Aamena Parulkar for assistance with DRIFTS experiments. The authors also thank the Ohio Supercomputer Center for computational resources.48

(2) (3)

(4)

(5)

(6) (7)

(8) (9)

(10)

(11) (12) (13)

(14) (15) (16)

(17)

(28)

(29) (30)

REFERENCES (1)

(27)

Goeppert, A.; Czaun, M.; Surya Prakash, G. K.; Olah, G. A. Energy Environ. Sci. 2012, 5, 7833–7853. Sanz-Pérez, E. S.; Murdock, C. R.; Didas, S. A.; Jones, C. W. Chem. Rev. 2016, 116, 11840–11876. Madden, D. G.; Scott, H. S.; Kumar, A.; Chen, K.-J.; Sanii, R.; Bajpai, A.; Lusi, M.; Curtin, T.; Perry, J. J.; Zaworotko, M. J. Philos. Trans. R. Soc., A 2017, 375, 20160025. Law, J.; Watkins, S.; Alexander, D. In-Flight Carbon Dioxide Exposures and Related Symptoms: Association, Susceptibility, and Operational Implications; Technical Publication TP–2010– 216126; NASA: Houston, 2010. James, J. T.; Macatangay, A. Carbon Dioxide−Our Common “Enemy”; Technical Report No. JSC-CN-18669; NASA: Houston, 2009. Xiang, S.; He, Y.; Zhang, Z.; Wu, H.; Zhou, W.; Krishna, R.; Chen, B. Nat. Commun. 2012, 3, 954. Ye, Y.; Xiong, S.; Wu, X.; Zhang, L.; Li, Z.; Wang, L.; Ma, X.; Chen, Q.-H.; Zhang, Z.; Xiang, S. Inorg. Chem. 2016, 55, 292– 299. Demessence, A.; D’Alessandro, D. M.; Foo, M. L.; Long, J. R. J. Am. Chem. Soc. 2009, 131, 8784–8786. McDonald, T. M.; Lee, W. R.; Mason, J. A.; Wiers, B. M.; Hong, C. S.; Long, J. R. J. Am. Chem. Soc. 2012, 134, 7056– 7065. Lee, W. R.; Hwang, S. Y.; Ryu, D. W.; Lim, K. S.; Han, S. S.; Moon, D.; Choi, J.; Hong, C. S. Energy Environ. Sci. 2014, 7, 744–751. Darunte, L. A.; Oetomo, A. D.; Walton, K. S.; Sholl, D. S.; Jones, C. W. ACS Sustain. Chem. Eng. 2016, 4, 5761–5768. Li, H.; Wang, K.; Feng, D.; Chen, Y.-P.; Verdegaal, W.; Zhou, H.-C. ChemSusChem 2016, 9, 2832–2840. Liao, P.-Q.; Chen, X.-W.; Liu, S.-Y.; Li, X.-Y.; Xu, Y.-T.; Tang, M.; Rui, Z.; Ji, H.; Zhang, J.-P.; Chen, X.-M. Chem. Sci. 2016, 7, 6528–6533. Didas, S. A.; Choi, S.; Chaikittisilp, W.; Jones, C. W. Acc. Chem. Res. 2015, 48, 2680–2687. Lu, W.; Sculley, J. P.; Yuan, D.; Krishna, R.; Zhou, H.-C. J. Phys. Chem. C 2013, 117, 4057–4061. Siegelman, R. L.; McDonald, T. M.; Gonzalez, M. I.; Martell, J. D.; Milner, P. J.; Mason, J. A.; Berger, A. H.; Bhown, A. S.; Long, J. R. J. Am. Chem. Soc. 2017, 139, 10526–10538. Milner, P. J.; Martell, J. D.; Siegelman, R. L.; Gygi, D.; Weston, S. C.; Long, J. R. Chem. Sci. 2018, 9, 160–174.

(31) (32) (33) (34) (35)

(36) (37) (38) (39) (40) (41) (42) (43) (44) (45)

(46)

(47)

(48)

Kumar, A.; Madden, D. G.; Lusi, M.; Chen, K.-J.; Daniels, E. A.; Curtin, T.; Perry, J. J.; Zaworotko, M. J. Angew. Chem., Int. Ed. 2015, 54, 14372–14377. Shekhah, O.; Belmabkhout, Y.; Chen, Z.; Guillerm, V.; Cairns, A.; Adil, K.; Eddaoudi, M. Nat. Commun. 2014, 5, 4228. Bhatt, P. M.; Belmabkhout, Y.; Cadiau, A.; Adil, K.; Shekhah, O.; Shkurenko, A.; Barbour, L. J.; Eddaoudi, M. J. Am. Chem. Soc. 2016, 138, 9301–9307. Kumar, A.; Hua, C.; Madden, D. G.; O’Nolan, D.; Chen, K.-J.; Keane, L.-A. J.; Perry, J. J.; Zaworotko, M. J. Chem. Commun. 2017, 53, 5946–5949. Chen, K.-J.; Yang, Q.-Y.; Sen, S.; Madden, D. G.; Kumar, A.; Pham, T.; Forrest, K. A.; Hosono, N.; Space, B.; Kitagawa, S.; et al. Angew. Chem., Int. Ed. 2018, 57, 3332–3336. Gassensmith, J. J.; Furukawa, H.; Smaldone, R. A.; Forgan, R. S.; Botros, Y. Y.; Yaghi, O. M.; Stoddart, J. F. J. Am. Chem. Soc. 2011, 133, 15312–15315. Wu, D.; Gassensmith, J. J.; Gouvêa, D.; Ushakov, S.; Stoddart, J. F.; Navrotsky, A. J. Am. Chem. Soc. 2013, 135, 6790–6793. Hartlieb, K. J.; Peters, A. W.; Wang, T. C.; Deria, P.; Farha, O. K.; Hupp, J. T.; Stoddart, J. F. Chem. Commun. 2017, 53, 7561– 7564. Liao, P.-Q.; Chen, H.; Zhou, D.-D.; Liu, S.-Y.; He, C.-T.; Rui, Z.; Ji, H.; Zhang, J.-P.; Chen, X.-M. Energy Environ. Sci. 2015, 8, 1011–1016. Jin, C.; Zhang, S.; Zhang, Z.; Chen, Y. Inorg. Chem. 2018, 57, 2169–2174. Schmieder, P.; Denysenko, D.; Grzywa, M.; Baumgärtner, B.; Senkovska, I.; Kaskel, S.; Sastre, G.; van Wüllen, L.; Volkmer, D. Dalton Trans. 2013, 42, 10786–10797. Lindskog, S. Pharmacol. Ther. 1997, 74, 1–20. Denysenko, D.; Grzywa, M.; Jelic, J.; Reuter, K.; Volkmer, D. Angew. Chem., Int. Ed. 2014, 53, 5832–5836. Denysenko, D.; Jelic, J.; Reuter, K.; Volkmer, D. Chem. Eur. J. 2015, 21, 8188–8199. Lohr, T. L.; Piers, W. E.; Parvez, M. Dalton Trans. 2013, 42, 14742. Martínez-Prieto, L. M.; Real, C.; Ávila, E.; Álvarez, E.; Palma, P.; Cámpora, J. Eur. J. Inorg. Chem. 2013, 2013, 5555–5566. Sattler, W.; Parkin, G. Polyhedron 2012, 32, 41–48. Huang, D.; Makhlynets, O. V.; Tan, L. L.; Lee, S. C.; RybakAkimova, E. V.; Holm, R. H. Proc. Natl. Acad. Sci. 2011, 108, 1222–1227. Nakamoto, K.; Sarma, Y. A.; Ogoshi, H. J. Chem. Phys. 1965, 43, 1177–1181. Yoshida, T.; Thorn, D. L.; Okano, T.; Ibers, J. A.; Otsuka, S. J. Am. Chem. Soc. 1979, 101, 4212–4221. Looney, A.; Han, R.; McNeill, K.; Parkin, G. J. Am. Chem. Soc. 1993, 115, 4690–4697. Sattler, W.; Parkin, G. Chem. Sci. 2012, 3, 2015–2019. Tsuda, T.; Chujo, Y.; Saegusa, T. J. Am. Chem. Soc. 1980, 102, 431–433. Palmer, D. A.; Van Eldik, R. Chem. Rev. 1983, 83, 651–731. Ganguly, S.; Mague, J. T.; Roundhill, D. M. Inorg. Chem. 1992, 31, 3831–3835. Zastrow, M. L.; Pecoraro, V. L. Biochemistry 2014, 53, 957– 978. Vaidhyanathan, R.; Iremonger, S. S.; Shimizu, G. K. H.; Boyd, P. G.; Alavi, S.; Woo, T. K. Science 2010, 330, 650–653. Lee, C.-H.; Huang, H.-Y.; Liu, Y.-H.; Luo, T.-T.; Lee, G.-H.; Peng, S.-M.; Jiang, J.-C.; Chao, I.; Lu, K.-L. Inorg. Chem. 2013, 52, 3962–3968. McDonald, T. M.; Mason, J. A.; Kong, X.; Bloch, E. D.; Gygi, D.; Dani, A.; Crocellà, V.; Giordanino, F.; Odoh, S. O.; Drisdell, W. S.; Vlaisavljevich, B.; Dzubak, A. L.; Poloni, R.; Schnell, S. K,; Planas, N.; Lee, K.; Pascal, T.; Wan, L. F.; Prendergast, D.; Neaton, J. B.; Smit, B.; Kortright, J. B.; Gagliardi, L.; Bordiga, S.; Reimer, J. A.; Long J. R. Nature 2015, 519, 303–308. Reed, D. A.; Keitz, B. K.; Oktawiec, J.; Mason, J. A.; Runčevski, T.; Xiao, D. J.; Darago, L. E.; Crocellà, V.; Bordiga, S.; Long, J. R. Nature 2017, 550, 96–100. Ohio Supercomputer Center. http://osc.edu/ark:/19495/f5s1ph73 (accessed Sep 14, 2018).

ACS Paragon Plus Environment

Page 4 of 5

Page 5 of 5 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

TOC Graphic

ACS Paragon Plus Environment