Next Article in Journal
Two Novel, Flavin-Dependent Halogenases from the Bacterial Consortia of Botryococcus braunii Catalyze Mono- and Dibromination
Next Article in Special Issue
Catalytic Combustion of Toluene over Highly Dispersed Cu-CeOx Derived from Cu-Ce-MOF by EDTA Grafting Method
Previous Article in Journal
Magnetically Agitated Nanoparticle-Based Batch Reactors for Biocatalysis with Immobilized Aspartate Ammonia-Lyase
Previous Article in Special Issue
Catalytic Performance of CPM-200-In/Mg in the Cycloaddition of CO2 and Epoxides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Molecular Cage Promoted Aerobic Oxidation or Photo-Induced Rearrangement of Spiroepoxy Naphthalenone

1
College of Chemistry and Material Science, Fujian Normal University, Fuzhou 350007, China
2
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(4), 484; https://doi.org/10.3390/catal11040484
Submission received: 20 March 2021 / Revised: 6 April 2021 / Accepted: 7 April 2021 / Published: 9 April 2021

Abstract

:
Herein, we report a Pd4L2-type molecular cage (1) and catalyzed reactions of spiroepoxy naphthalenone (2) in water, where selective formation of 2-(hydroxymethyl)naphthalene-1,4-dione (3) via aerobic oxidation, or 1-hydroxy-2-naphthaldehyde (4) via photo-induced rearrangement under N2 have been accomplished. Encapsulation of four molecules of guest 2 within cage 1, i.e., (2)41, has been confirmed by NMR, and a final host-guest complex of 31 has also been determined by single crystal X-Ray diffraction study. While the photo-induced ring-opening isomerization from 2 to 4 are known, appearance of charge-transfer absorption on the host-guest complex of (2)41 allows low-power blue LEDs irradiation to promote this process.

Graphical Abstract

1. Introduction

Artificial supramolecular architectures constructed via self-assembly with metal ions and organic ligands have attached tremendous attentions due to aesthetic structural diversity [1,2,3,4,5,6,7,8,9,10,11,12] and bio-mimic catalysis [13,14]. Among them, coordination molecular cages have been extensively studied as artificial enzyme-mimics to promote various reactions of the contained molecules within the confined space [15,16,17,18,19]. Cage catalyzed reactions can be classified into three pathways [20]: (1) cage promoted reactions directly; (2) modified cage to improve catalysis performance; (3) cooperation with other catalysts. For the cage promoted reactions, it is generally accepted that enhanced effective concentration and the pre-organization effect imposed to the guest molecules by the cage cavity contribute to the accelerated reaction rate and the improved product selectivity. To date, cage promoted cycloaddition [21,22], rearrangement [23,24], photocatalysis reactions [25,26,27], oxidation or reduction [28,29,30,31,32], nucleophilic addition [33,34], elimination [35] and acid or base-catalyzed addition [36,37,38] have been reported, and a part of works exhibit rate enhancement, amazing selectivity and wide substrate scope. Moreover, highly-charged molecular cages can perform catalysis in water, like enzymes do in nature. Recently, we have developed a redox-active coordination-assembled cage 1 (Scheme 1) based on bridged pyridinium ligands, which has been employed as a water-soluble bio-mimic catalyst [39]. The large hydrophobic cavity of cage 1 allows a variety of guests to be encapsulated, and the big apertures on the cage favor the product turn-over during catalysis [40].
Spiroepoxy-cyclohexdienones, one type of vinyl epoxides [41,42], are important building blocks for organic synthesis due to their highly reactive nature toward self-coupling Diels-Alder reactions [43]. We anticipate that new reaction pathways may be discovered if spiroepoxy-cyclohexdienones are entrapped within molecular cages. Herein, we report that cage 1 promoted aerobic oxidation and photo-induced isomerization reactions in water from spiroepoxy naphthalenone 2. In the presence of cage 1, both naphthoquinone 3 and aldehyde 4 can be selectively obtained by regulating the reaction conditions. It is worth to note that as a typical spiroepoxy-cyclohexdienones compound, reactivity of 2 has seldom been studied, with only one type of acid mediated rearrangement reaction being reported previously [44].

2. Results and Discussion

Cage 1 and substrate 2 were synthesized according to the previous reports [40,44]. We started from the investigation of their host-guest chemistry. After excess amount of guest 2 was added to cage 1 in D2O, the formation of a 1:4 host-guest complex was indicated by 1H NMR spectra. Compared to the signals of free 2 (Figure 1a) and empty cage 1 (Figure 1b), the host–guest complex shows significant changes. Obvious up-field shifts for the aromatic signals on 2 from 8.1–5.9 ppm to 6.4–4.5 ppm, and the methylene signals from 3.4–3.1 ppm to 2.0–2.2 ppm were observed in 1H NMR spectrum (Figure 1c), which indicates the efficient guest encapsulation within the hydrophobic cavity of cage 1. Moreover, integral ratio (Figure S3) confirms that four molecules of 2 could be encapsulated by 1. Diffusion-ordered 1H NMR spectroscopy (1H DOSY, Figure S4) also confirmed the formation of a single host-guest species with a diameter of 1.70 nm estimated from the Stokes–Einstein equation. The optimized host-guest complex model indicates that four molecules of 2 completely occupied the space of cavity within the cage 1 (Figure 2a).
Interestingly, after heating at 50 °C for 2 h, the color of the solution for the initial (2)41 complex changed gradually from yellow to deep red, and the dramatic change in 1H NMR spectrum (Figure 1d) suggests the formation of a new host-guest complex. Characteristic signals for the encapsulated 2 (methylene signals) disappeared with the evolution of a new set of guest signals. After extraction by CDCl3, the final product was determined to be 2-(hydroxymethyl)naphthalene-1,4-dione (3) by NMR and GC-MS, with an NMR yield of over 99% (Figure 1e and Figures S5–S7). A series of control experiments were carried out to confirm the indispensable role of cage 1 during this reaction. First of all, no such conversion was detected without cage 1, under N2, or with ligand L only (entry 2–4). With the Pd salt (TMEDA)Pd(NO3)2 as catalyst, product 3 could only be obtained in a 39% yield (entry 5). Considering previous reported work [39,45,46,47,48,49,50], we propose that redox-active nature of cage 1 plays a key role in this aerobic Wacker oxidation process.
To further confirm that efficient hydrophobic encapsulation is crucial for this conversion, a tetraphenylboron anion (Ph4B-, sodium salt) was chosen as a competition inhibitor which is known to strongly bind to cage 1 [40]. Although partial guest 2 can also enter the cavity of cage 1 in the presence of an inhibitor (Figure S8), a lower yield of 36% was observed under the same conditions (entry 6, Figure S9). More importantly, when excess amount of guest 2 (100 equiv.) was added to the solution of cage 1 as a suspension, product 3 was formed in 95% yield after heating for 4 h, leading to a TON of 95 (entry 7).
DFT calculation indicated that the C2-O2 bond between the naphthalene ring and the epoxy group was longer than the O2-C5 bond, suggesting that the epoxy group tended to undergo ring-opening reaction (Figure 3a). In fact, a previous report [51] revealed that substrate 2 can undergo a photo-induced ring-opening rearrangement to form 1-hydroxy-2-naphthaldehyde 4 by purple LEDs irradiation. UV-vis spectra (Figure 3b) measurements suggest that a new shoulder peak tailing up to visible region appeared for the (2)41 host-guest complex, assignable to host-guest charge transform (CT) absorption. Indeed, under blue LEDs irradiation, cage 1 promoted isomerization reaction of 2 to 4 has been observed in 90% yield after 8 h, which is a significant improvement comparing to that in the absence of cage 1 under the same conditions (Table 1, entry 8 and 9). Considering that conversion from 2 to 3 is a competition pathway under air in the presence of cage 1, this photo-induced isomerization has to be carried out under N2 atmosphere. In this case, cage 1 not only plays as a phase transfer catalyst but also facilitates the visible light absorption by the formation of host-guest CT complex.
To our delight, dark red crystals were obtained by slow evaporation of the final reaction solution at room temperature over one week. The crystals were of sufficient quality and X-ray crystallography finally established the structure of the new host–guest complex, which contained product 3 and cage 1 (Appendix A). X-ray structure revealed that there was only one molecule of 3 sitting inside in the inner cavity of the cage, along with two molecules of 3 binding to the external panels of the cage (Figure 2b). The external binding of the products explains product replacement of start material observed during the above catalysis.

3. Conclusions

To conclude, we have revealed a condition-controlled supramolecular cage catalyzed reaction for a spiroepoxy-naphthalenone guest molecule, where both quinone product 3 and aldehyde product 4 can be obtained selectively in good yields. Such coordination-cage-promoted selective transformations from one starting material toward different products is still rare in supramolecular catalysis. This approach promised a potential application for cage catalyzed oxidation and photo-isomerization reactions. Further development of catalytic reactions by using cage 1 as an artificial enzyme-mimic is still underway.

4. Materials and Methods

4.1. General

Unless otherwise stated, all chemicals and solvents were purchased from commercial companies (AdamasReagent Ltd., Shanghai, China; J&K Scientific Ltd., Beijing, China, and Sigma-Aldrich LLC., Darmstadt, Germany. etc) and used without further purification. 1D and 2D-NMR were measured on a Bruker Biospin Avance III (400 MHz) spectrometer or JEOL JNM-ECZ600R/S1 (600 MHz) spectrometer. 1H-NMR chemical shifts were determined with tetramethylsilane (TMS) or respect to residual signals of the deuterated solvents used (δ = 4.79 for D2O in 1H NMR). The photoreactors used in this research were bought from Wuhan Geao Instruments Science and Technology Co., Ltd. (Wuhan, China) (Purple LEDs, light intensity = 37.4 mw/cm2, λmax = 390 nm; Blue LEDs, λmax = 450 nm; 1 W for every light bulb; every Schlenk tube was irradiated by 6 light bulbs from the side). Gas chromatography mass spectrometry (GC-MS) analyses were performed on a Shimadzu GCMS-QP2010SE instrument. ESI-TOF-MS were recorded on Impact II UHR-TOF from Bruker. Data analysis was conducted with the Bruker Data Analysis software (Version 4.3, Bruker Daltonik GmbH, Bremen, Germany) and simulations were performed with the Bruker Isotope Pattern software. UV-vis adsorption spectra were recorded on UV-2700 UV-visible spectrophotometer from SHIMADZU Corporation.

4.2. Preparation and Catalysis Procedure

The general catalysis procedure and all the characterization data of the products as well as the procedures to prepare are listed in the Supplementary Materials online. The known compounds were recorded in the previous report.

4.3. Computational Methods

Molecular modeling of host-guest complex (2)41 was optimized by force-field calculation carried on Material Studio software by Quasi-Newton Methods which was based on a previously reported cage 1 X-ray structure [39]. Structure optimization of guest 2 which was carried at B3LYP/6-31G level by gaussian software [52].

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11040484/s1. Figures S1–S21: synthesis procedure, catalysis procedure, NMR spectra, ESI-TOF-MS, GC-MS and Single-Crystal information.

Author Contributions

Conceptualization, Q.S. and P.C.; writing—original draft preparation, P.C.; writing—review and editing, Q.S.; P.C. performed most of the experiments and analyzed the data, L.C., D.Y., L.Z. helped with synthesis/characterization and joined the discussion. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 21825107, 21801241), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDB20000000).

Data Availability Statement

All data generated or analyzed during this study are included in this article.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Crystallographic data for the structures of host-guest complex of product 3 and cage 1 have been deposited to the Cambridge Crystallographic Data Centre as supplementary No.: CCDC-2065411.

References

  1. Fujita, D.; Ueda, Y.; Sato, S.; Yokoyama, H.; Mizuno, N.; Kumasaka, T.; Fujita, M. Self-Assembly of M 30 L 60 Icosidodecahedron. Chem 2016, 1, 91–101. [Google Scholar] [CrossRef]
  2. Yang, D.; Greenfield, J.L.; Ronson, T.K.; Von Krbek, L.K.S.; Yu, L.; Nitschke, J.R. LaIII and ZnII Cooperatively Template a Metal–Organic Capsule. J. Am. Chem. Soc. 2020, 142, 19856–19861. [Google Scholar] [CrossRef]
  3. Cheng, P.; Cai, L.; Li, S.; Hu, S.; Yan, D.; Zhou, L.; Sun, Q. Guest-Reaction Driven Cage to Conjoined Twin-Cage Mitosis-Like Host Transformation. Angew. Chem. Int. Ed. 2020, 59, 23569–23573. [Google Scholar] [CrossRef]
  4. Zhang, Z.; Li, Y.; Song, B.; Zhang, Y.; Jiang, X.; Wang, M.; Trumbleson, R.; Liu, C.; Wang, P.; Hao, X.-Q.; et al. Intra- and intermolecular self-assembly of a 20-nm-wide supramolecular hexagonal grid. Nat. Chem. 2020, 12, 468–474. [Google Scholar] [CrossRef]
  5. Cai, L.-X.; Yan, D.-N.; Cheng, P.-M.; Xuan, J.-J.; Li, S.-C.; Zhou, L.-P.; Tian, C.-B.; Sun, Q.-F. Controlled Self-Assembly and Multistimuli-Responsive Interconversions of Three Conjoined Twin-Cages. J. Am. Chem. Soc. 2021, 143, 2016–2024. [Google Scholar] [CrossRef]
  6. Liu, D.; Li, K.; Chen, M.; Zhang, T.; Li, Z.; Yin, J.-F.; He, L.; Wang, J.; Yin, P.; Chan, Y.-T.; et al. Russian-Doll-Like Molecular Cubes. J. Am. Chem. Soc. 2021, 143, 2537–2544. [Google Scholar] [CrossRef] [PubMed]
  7. Fang, Y.; Xiao, Z.; Li, J.; Lollar, C.; Liu, L.; Lian, X.; Yuan, S.; Banerjee, S.; Zhang, P.; Zhou, H.-C. Formation of a Highly Reactive Cobalt Nanocluster Crystal within a Highly Negatively Charged Porous Coordination Cage. Angew. Chem. Int. Ed. 2018, 57, 5283–5287. [Google Scholar] [CrossRef]
  8. Zhang, X.; Dong, X.; Lu, W.; Luo, D.; Zhu, X.-W.; Li, X.; Zhou, X.-P.; Li, D. Fine-Tuning Apertures of Metal–Organic Cages: Encapsulation of Carbon Dioxide in Solution and Solid State. J. Am. Chem. Soc. 2019, 141, 11621–11627. [Google Scholar] [CrossRef]
  9. Wu, G.-Y.; Shi, X.; Phan, H.; Qu, H.; Hu, Y.-X.; Yin, G.-Q.; Zhao, X.-L.; Li, X.; Xu, L.; Yu, Q.; et al. Efficient self-assembly of heterometallic triangular necklace with strong antibacterial activity. Nat. Commun. 2020, 11, 1–11. [Google Scholar] [CrossRef] [PubMed]
  10. Gao, W.-X.; Feng, H.-J.; Guo, B.-B.; Lu, Y.; Jin, G.-X. Coordination-Directed Construction of Molecular Links. Chem. Rev. 2020, 120, 6288–6325. [Google Scholar] [CrossRef] [PubMed]
  11. Lu, Y.; Liu, D.; Lin, Y.-J.; Li, Z.-H.; Jin, G.-X. Self-assembly of metalla[3]catenanes, borromean rings and ring-in-ring complex using a simple π-donor unit. Natl. Sci. Rev. 2020, 7, 1548–1556. [Google Scholar] [CrossRef]
  12. Huang, S.-L.; Lin, Y.-J.; Li, Z.-H.; Jin, G.-X. Self-Assembly of Molecular Borromean Rings from Bimetallic Coordination Rectangles. Angew. Chem. Int. Ed. 2014, 53, 11218–11222. [Google Scholar] [CrossRef] [PubMed]
  13. Brown, C.J.; Toste, F.D.; Bergman, R.G.; Raymond, K.N. Supramolecular Catalysis in Metal–Ligand Cluster Hosts. Chem. Rev. 2015, 115, 3012–3035. [Google Scholar] [CrossRef]
  14. Yoshizawa, M.; Klosterman, J.K.; Fujita, M. Functional Molecular Flasks: New Properties and Reactions within Discrete, Self-Assembled Hosts. Angew. Chem. Int. Ed. 2009, 48, 3418–3438. [Google Scholar] [CrossRef] [PubMed]
  15. Tan, C.; Chu, D.; Tang, X.; Liu, Y.; Xuan, W.; Cui, Y. Supramolecular Coordination Cages for Asymmetric Catalysis. Chem. A Eur. J. 2019, 25, 662–672. [Google Scholar] [CrossRef] [PubMed]
  16. Pan, M.; Wu, K.; Zhang, J.-H.; Su, C.-Y. Chiral metal–organic cages/containers (MOCs): From structural and stereochemical design to applications. Coord. Chem. Rev. 2019, 378, 333–349. [Google Scholar] [CrossRef]
  17. Zhao, L.; Jing, X.; Li, X.; Guo, X.; Zeng, L.; He, C.; Duan, C. Catalytic properties of chemical transformation within the confined pockets of Werner-type capsules. Coord. Chem. Rev. 2019, 378, 151–187. [Google Scholar] [CrossRef]
  18. Morimoto, M.; Bierschenk, S.M.; Xia, K.T.; Bergman, R.G.; Raymond, K.N.; Toste, F.D. Advances in supramolecular host-mediated reactivity. Nat. Catal. 2020, 3, 969–984. [Google Scholar] [CrossRef]
  19. Percástegui, E.G.; Ronson, T.K.; Nitschke, J.R. Design and Applications of Water-Soluble Coordination Cages. Chem. Rev. 2020, 120, 13480–13544. [Google Scholar] [CrossRef]
  20. Fang, Y.; Powell, J.A.; Li, E.; Wang, Q.; Perry, Z.; Kirchon, A.; Yang, X.; Xiao, Z.; Zhu, C.; Zhang, L.; et al. Catalytic reactions within the cavity of coordination cages. Chem. Soc. Rev. 2019, 48, 4707–4730. [Google Scholar] [CrossRef]
  21. Merget, S.; Catti, L.; Piccini, G.; Tiefenbacher, K. Requirements for Terpene Cyclizations inside the Supramolecular Resorcinarene Capsule: Bound Water and Its Protonation Determine the Catalytic Activity. J. Am. Chem. Soc. 2020, 142, 4400–4410. [Google Scholar] [CrossRef] [PubMed]
  22. Hastings, C.J.; Pluth, M.D.; Bergman, R.G.; Raymond, K.N. Enzymelike Catalysis of the Nazarov Cyclization by Supramolecular Encapsulation. J. Am. Chem. Soc. 2010, 132, 6938–6940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Hastings, C.J.; Fiedler, D.; Bergman, R.G.; Raymond, K.N. Aza Cope Rearrangement of Propargyl Enammonium Cations Catalyzed by a Self-Assembled “Nanozyme”. J. Am. Chem. Soc. 2008, 130, 10977–10983. [Google Scholar] [CrossRef] [Green Version]
  24. Hong, C.M.; Morimoto, M.; Kapustin, E.A.; Alzakhem, N.; Bergman, R.G.; Raymond, K.N.; Toste, F.D. Deconvoluting the Role of Charge in a Supramolecular Catalyst. J. Am. Chem. Soc. 2018, 140, 6591–6595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zhao, L.; Cai, J.; Li, Y.; Wei, J.; Duan, C. A host–guest approach to combining enzymatic and artificial catalysis for catalyzing biomimetic monooxygenation. Nat. Commun. 2020, 11, 1–11. [Google Scholar] [CrossRef]
  26. Guo, J.; Fan, Y.; Lu, Y.; Zheng, S.; Su, C. Visible-Light Photocatalysis of Asymmetric [2+2] Cycloaddition in Cage-Confined Nanospace Merging Chirality with Triplet-State Photosensitization. Angew. Chem. Int. Ed. 2020, 59, 8661–8669. [Google Scholar] [CrossRef]
  27. Wang, J.-S.; Wu, K.; Yin, C.; Li, K.; Huang, Y.; Ruan, J.; Feng, X.; Hu, P.; Su, C.-Y. Cage-confined photocatalysis for wide-scope unusually selective [2 + 2] cycloaddition through visible-light triplet sensitization. Nat. Commun. 2020, 11, 1–9. [Google Scholar] [CrossRef]
  28. Zaffaroni, R.; Orth, N.; Ivanović-Burmazović, I.; Reek, J.N.H. Hydrogenase Mimics in M 12 L 24 Nanospheres to Control Overpotential and Activity in Proton-Reduction Catalysis. Angew. Chem. Int. Ed. 2020, 59, 18485–18489. [Google Scholar] [CrossRef]
  29. Lai, Y.-L.; Wang, X.-Z.; Dai, R.-R.; Huang, Y.-L.; Zhou, X.-C.; Zhou, X.-P.; Li, D. Self-assembly of mixed-valence and heterometallic metallocycles: Efficient catalysts for the oxidation of alcohols to aldehydes in ambient air. Dalton Trans. 2020, 49, 7304–7308. [Google Scholar] [CrossRef] [PubMed]
  30. Ding, C.-W.; Luo, W.; Zhou, J.-Y.; Ma, X.-J.; Chen, G.-H.; Zhou, X.-P.; Li, D. Hydroxo Iron(III) Sites in a Metal–Organic Framework: Proton-Coupled Electron Transfer and Catalytic Oxidation of Alcohol with Molecular Oxygen. ACS Appl. Mater. Interfaces 2019, 11, 45621–45628. [Google Scholar] [CrossRef] [PubMed]
  31. Fang, Y.; Li, J.; Togo, T.; Jin, F.; Xiao, Z.; Liu, L.; Drake, H.; Lian, X.; Zhou, H.-C. Ultra-Small Face-Centered-Cubic Ru Nanoparticles Confined within a Porous Coordination Cage for Dehydrogenation. Chem 2018, 4, 555–563. [Google Scholar] [CrossRef] [Green Version]
  32. Jiang, W.-L.; Shen, J.-C.; Peng, Z.; Wu, G.-Y.; Yin, G.-Q.; Shi, X.; Yang, H.-B. Controllable synthesis of ultrasmall Pd nanocatalysts templated by supramolecular coordination cages for highly efficient reductive dehalogenation. J. Mater. Chem. A 2020, 8, 12097–12105. [Google Scholar] [CrossRef]
  33. Wang, J.; Young, T.A.; Duarte, F.; Lusby, P.J. Synergistic Noncovalent Catalysis Facilitates Base-Free Michael Addition. J. Am. Chem. Soc. 2020, 142, 17743–17750. [Google Scholar] [CrossRef]
  34. Zhu, F.-F.; Chen, L.-J.; Chen, S.; Wu, G.-Y.; Jiang, W.-L.; Shen, J.-C.; Qin, Y.; Xu, L.; Yang, H.-B. Confinement Self-Assembly of Metal-Organic Cages within Mesoporous Carbon for One-Pot Sequential Reactions. Chem 2020, 6, 2395–2406. [Google Scholar] [CrossRef]
  35. Cullen, W.; Misuraca, M.C.; Hunter, M.C.M.C.A.; Williams, N.H.; Ward, W.C.N.H.W.M.D. Highly efficient catalysis of the Kemp elimination in the cavity of a cubic coordination cage. Nat. Chem. 2016, 8, 231–236. [Google Scholar] [CrossRef] [PubMed]
  36. Ngai, C.; Sanchez-Marsetti, C.M.; Harman, W.H.; Hooley, R.J. Supramolecular Catalysis of the oxa-Pictet–Spengler Reaction with an Endohedrally Functionalized Self-Assembled Cage Complex. Angew. Chem. Int. Ed. 2020, 59, 23505–23509. [Google Scholar] [CrossRef] [PubMed]
  37. Takezawa, H.; Shitozawa, K.; Fujita, M. Enhanced reactivity of twisted amides inside a molecular cage. Nat. Chem. 2020, 12, 574–578. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, J.; Châtelet, B.; Dufaud, V.; Hérault, D.; Michaud-Chevallier, S.; Robert, V.; Dutasta, J.-P.; Martinez, A. Endohedral Functionalized Cage as a Tool to Create Frustrated Lewis Pairs. Angew. Chem. Int. Ed. 2018, 57, 14212–14215. [Google Scholar] [CrossRef]
  39. Cai, L.-X.; Li, S.-C.; Yan, D.-N.; Zhou, L.-P.; Guo, F.; Sun, Q.-F. Water-Soluble Redox-Active Cage Hosting Polyoxometalates for Selective Desulfurization Catalysis. J. Am. Chem. Soc. 2018, 140, 4869–4876. [Google Scholar] [CrossRef]
  40. Li, S.-C.; Cai, L.-X.; Zhou, L.-P.; Guo, F.; Sun, Q.-F. Supramolecular synthesis of coumarin derivatives catalyzed by a coordination-assembled cage in aqueous solution. Sci. China Ser. B Chem. 2019, 62, 713–718. [Google Scholar] [CrossRef]
  41. Gesson, J.-P.; Mondon, M.; Pokrovska, N. Synthesis of Fused Aromatic [1,3]Dioxoles from 2-Hydroxymethylphenols. Synlett 1997, 12, 1395–1396. [Google Scholar] [CrossRef]
  42. He, J.; Ling, J.; Chiu, P. Vinyl Epoxides in Organic Synthesis. Chem. Rev. 2014, 114, 8037–8128. [Google Scholar] [CrossRef]
  43. Xin, M.; Bugg, T.D.H. Biomimetic Formation of 2-Tropolones by Dioxygenase-Catalysed Ring Expansion of Substituted 2,4-Cyclohexadienones. ChemBioChem 2009, 11, 272–276. [Google Scholar] [CrossRef]
  44. Tius, M.A.; Reddy, N.K. Stereoselective Synthesis of Disubstituted Naphthalene-1,2-oxides. Synth. Commun. 1994, 24, 859–869. [Google Scholar] [CrossRef]
  45. Yoshizawa, M.; Sato, N.; Fujita, M. Selective Enclathration of Linear Alkanols by a Self-assembled Coordination Cage. Application to the Catalytic Wacker Oxidation of ω-Alkenols. Chem. Lett. 2005, 34, 1392–1393. [Google Scholar] [CrossRef]
  46. Mahmudov, K.T.; Gurbanov, A.V.; Guseinov, F.I.; da Silva, M.F.C.G. Noncovalent interactions in metal complex catalysis. Coord. Chem. Rev. 2019, 387, 32–46. [Google Scholar] [CrossRef]
  47. Ma, Z.; Mahmudov, K.T.; Aliyeva, V.A.; Gurbanov, A.V.; Pombeiro, A.J. TEMPO in metal complex catalysis. Coord. Chem. Rev. 2020, 423, 213482. [Google Scholar] [CrossRef]
  48. Ma, Z.; Mahmudov, K.T.; Aliyeva, V.A.; Gurbanov, A.V.; da Silva, M.F.C.G.; Pombeiro, A.J. Peroxides in metal complex catalysis. Coord. Chem. Rev. 2021, 437, 213859. [Google Scholar] [CrossRef]
  49. Higuchi, M.; Yamaguchi, S.; Hirao, T. Construction of Palladium-Polypyrrole Catalytic System in the Wacker Oxidation. Synlett 1996, 1996, 1213–1214. [Google Scholar] [CrossRef]
  50. Mitsudome, T.; Umetani, T.; Nosaka, N.; Mori, K.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Convenient and Efficient Pd-Catalyzed Regioselective Oxyfunctionalization of Terminal Olefins by Using Molecular Oxygen as Sole Reoxidant. Angew. Chem. Int. Ed. 2006, 45, 481–485. [Google Scholar] [CrossRef] [PubMed]
  51. Becker, H.-D.; Bremholt, T.; Adler, E. Oxidative formation and photochemical isomerization of spiro-epoxy-2,4-cyclohexadienones. Tetrahedron Lett. 1972, 13, 4205–4208. [Google Scholar] [CrossRef]
  52. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision E.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
Scheme 1. Cage 1 promoted selective transformation from 2 to 3 or 4.
Scheme 1. Cage 1 promoted selective transformation from 2 to 3 or 4.
Catalysts 11 00484 sch001
Figure 1. 1H NMR (400 MHz, 298 K) spectra of (a) 2 in CDCl3; (b) cage 1 in D2O; (c) host-guest complex of (2)41 in D2O and (d) (2)41 in D2O after heating at 50 °C for 2 h; (e) extracted 3 in CDCl3; (f) (2)41 in D2O after blue LEDs irradiation under N2 atmosphere at r.t. for 8 h; (g) extracted 4 in CDCl3; (: 2, :3, : cage 1, : 4).
Figure 1. 1H NMR (400 MHz, 298 K) spectra of (a) 2 in CDCl3; (b) cage 1 in D2O; (c) host-guest complex of (2)41 in D2O and (d) (2)41 in D2O after heating at 50 °C for 2 h; (e) extracted 3 in CDCl3; (f) (2)41 in D2O after blue LEDs irradiation under N2 atmosphere at r.t. for 8 h; (g) extracted 4 in CDCl3; (: 2, :3, : cage 1, : 4).
Catalysts 11 00484 g001
Figure 2. (a) Optimized structure of (2)41 based on molecular mechanical simulation; (b) X-ray structure of 31 host-guest complex. (cage 1 and guests 2 or 3 are displayed with stick and sphere models, respectively. C: wathet blue, N: blue, O: red, Pd: brown. Counterions are omitted for clarity).
Figure 2. (a) Optimized structure of (2)41 based on molecular mechanical simulation; (b) X-ray structure of 31 host-guest complex. (cage 1 and guests 2 or 3 are displayed with stick and sphere models, respectively. C: wathet blue, N: blue, O: red, Pd: brown. Counterions are omitted for clarity).
Catalysts 11 00484 g002
Figure 3. (a) DFT calculated structure of 2 with selected bond lengths; (b) UV-vis absorption spectra for cage 1, guest 2 and the (2)41 host-guest complex (1 × 10−5 M), with their pictures shown.
Figure 3. (a) DFT calculated structure of 2 with selected bond lengths; (b) UV-vis absorption spectra for cage 1, guest 2 and the (2)41 host-guest complex (1 × 10−5 M), with their pictures shown.
Catalysts 11 00484 g003
Table 1. Cage 1 promoted reaction of guest 2 under different conditions *.
Table 1. Cage 1 promoted reaction of guest 2 under different conditions *.
Catalysts 11 00484 i001
EntryCatalystConditionsLEDsTime (h)SolventYield a
34
11Air50 °C-2 hD2O>99%-
2-Air50 °C-2 hH2O--
31N250 °C-2 hD2O--
4 bLAir50 °C-2 hH2O--
5 cPdAir50 °C-2 hH2O39%-
6Ph4B⊂1Air50 °C-2 hD2O36%-
7 d1Air50 °C-4 hD2O95%-
81N2r.t.blue8 hD2O-90%
9-Airr.t.blue8 hH2O-32%
* Unless otherwise stated, all experiments were carried out with 0.01 mmol of 2 and 25% mol of cage 1 in water. The power of LEDs is ca. 6 W. a: Yields were determined from 1H NMR spectra using 1,3,5-trimethoxybenzene as the inner standard. b: 10% mol of ligand was used as catalyst. c: 10% mol of Pd salt [(TMEDA)Pd(NO3)2] was used as catalyst. d: 1% mol of cage 1 was used and reaction was carried under suspension condition.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cheng, P.; Cai, L.; Yan, D.; Zhou, L.; Sun, Q. Molecular Cage Promoted Aerobic Oxidation or Photo-Induced Rearrangement of Spiroepoxy Naphthalenone. Catalysts 2021, 11, 484. https://doi.org/10.3390/catal11040484

AMA Style

Cheng P, Cai L, Yan D, Zhou L, Sun Q. Molecular Cage Promoted Aerobic Oxidation or Photo-Induced Rearrangement of Spiroepoxy Naphthalenone. Catalysts. 2021; 11(4):484. https://doi.org/10.3390/catal11040484

Chicago/Turabian Style

Cheng, Peiming, Lixuan Cai, Danni Yan, Lipeng Zhou, and Qingfu Sun. 2021. "Molecular Cage Promoted Aerobic Oxidation or Photo-Induced Rearrangement of Spiroepoxy Naphthalenone" Catalysts 11, no. 4: 484. https://doi.org/10.3390/catal11040484

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop