Next Article in Journal
Fatty Acid Hydratases: Versatile Catalysts to Access Hydroxy Fatty Acids in Efficient Syntheses of Industrial Interest
Next Article in Special Issue
Epoxide Syntheses and Ring-Opening Reactions in Drug Development
Previous Article in Journal
Photochemical Study of a New Bimolecular Photoinitiating System for Vat Photopolymerization 3D Printing Techniques under Visible Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Manganese Catalysts with Tetradentate N-donor Pyridine-Appended Bipiperidine Ligands for Olefin Epoxidation Reactions: Ligand Electronic Effect and Mechanism

1
Department of Chemistry, Tufts University, Medford, OR 02148, USA
2
College of Chemical Engineering, Nanjing Forestry University, Nanjing 210037, China
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(3), 285; https://doi.org/10.3390/catal10030285
Submission received: 31 January 2020 / Revised: 12 February 2020 / Accepted: 18 February 2020 / Published: 2 March 2020
(This article belongs to the Special Issue Catalytic Asymmetric Epoxidation: Recent Progress)

Abstract

:
In this work, we described an electron-rich manganese mesoPYBP catalyst, Mn-SR-mesoPYBP(ClO4)2, by introducing electron-donating substituents on the mesoPYBP ligand. We optimized the catalytic performance in olefin epoxidation with H2O2 in the presence of acetic acid. The electron paramagnetic resonance (EPR) and cyclic voltammetry (CV) studies supported that an electronic effect could stabilize the high-valent intermediates in the catalytic cycles of the catalyst, which largely improved the catalytic performance and the reactivity of olefin epoxidation.

Graphical Abstract

1. Introduction

Epoxides are important building blocks in various chemical reactions for pharmaceuticals and fine chemicals. They transform readily to a wide variety of functional groups including ketones, ethers and alcohols by reduction, rearrangement or ring-opening reactions [1,2,3,4,5]. The increasing demand for epoxides has raised research interest in developing efficient, low-cost and environmentally-benign synthetic methods [6,7,8]. Transition metals coordinated to tetradentate N-donor (4N) aminopyridine ligands, such as BPMEN (N,N′-dimethyl-N,N′-bis(2-pyridylmethyl)-1,2-diaminoethane) [9,10], BPMCN (N,N′-bis(2-pyridylmethyl)-N,N′-dimethyl-trans-1,2-diaminocyclohexane) [11,12,13,14] and BPBP (N,N′-bis(2-picolyl)-2,2′-bipyrrolidine) [14,15,16,17,18,19], are now proven, efficient catalytic systems for the epoxidation of unfunctionalized olefins by using H2O2 as a green and atom-economical oxygen source. Carboxylic acid additives, such as acetic acid, promote the efficiencies of these epoxidations [4,15,18]. The first-row transition metals that complex with these ligands are abundant, cheap and environmentally friendly [17,20]. In a recent study, these metal complexes, however, were found to suffer from suboptimal selectivity, as the mixtures of cis-diols and epoxide products that result from these reactions are not easy to separate [1,21]. Addressing this challenge requires new olefin epoxidation catalysts.
In response to this challenge, our research group recently designed a family of bioinspired tetradentate N-donor pyridine-appended bipiperidine ligands (PYBP—1,1′-bis[(pyridin-2-yl)methyl]-2,2′-bipiperidine) [1,22,23]. Those newly developed PYBP ligands were derived from bipiperidine, a six-member-ring platform, separated into racemic and mesomeric diastereomers, and then complexed with manganese (Mn) or iron (Fe) to catalyze olefin epoxidation (Scheme 1). We have previously reported that iron racPYBP catalyzed olefin epoxidation with faster rates and a higher regioselectivity than the catalysts prepared from the aforementioned ligands [23]. We attribute this improved performance to the rigidity in the ligand backbone frameworks, preventing metal decomplexation and the decomposition of the catalysts [24,25,26]. We have also investigated the impact of the isomer MII-racPYBP and MII-mesoPYBP complexes (MII = Mn and Fe) on catalysis [1,12,23]. Catalysts based on the racPYBP ligand showed an excellent epoxidation reactivity at low loadings (0.5 mol%), while metal complexes based on the mesoPYBP ligand had either a slower reactivity (with Fe), or were inactive (with Mn). In this case, improving the reactivity of the mesomeric diastereomer complexes makes the PYBP catalytic system a more efficient and cheaper candidate for olefin epoxidation.
Herein, we describe a new complex, Mn-SR-mesoPYBP(ClO4)2 (Complex 2), which introduces electron-donating substituents on the pyridine rings of the PYBP ligand and complexes with manganese (Scheme 1). Several investigations in the field have discussed the ligand electronic effects on manganese-catalyzed alkene epoxidation, and we chose methoxy substituents on the PYBP ligands, because they are known to stabilize high oxidation state metallic intermediates in the catalytic cycle of BPBP catalysts [16,17,21,27,28]. Compared with the inactive Mn-mesoPYBP(ClO4)2 (Complex 1), Complex 2 yields faster olefin epoxidations, and exhibits a general relationship between the electronics of the ligands and the catalytic activity. The electron paramagnetic resonance (EPR) and cyclic voltammetry (CV) studies support our conclusion that the electronic effects can stabilize the high-valent intermediates in these catalytic cycles.

2. Results and Discussion

Scheme 2 shows our approach for preparing manganese Complex 2 from meso-bipiperidine, which incorporates electron donating methyl and methoxy groups into the axial pyridine ligands. We prepared fully reduced bipiperidines (BP), as reported in the literature [1,6,23]. The GC-MS (Gas Chromatography-Mass Spectroscopy) analysis of the mixture resulting from the reduction of 2,2′-bipyridine revealed that 70% of the mixture was mesomeric stereoisomers. The addition of HBr and heating to 40 °C produced the mesoBP ammonium salt, as a white precipitate. The alkylation of bipyridine with commercially available 4-methoxy-3,5-dimethylpicolyl chloride yielded the SR-mesoPYBP ligand. Complexation with manganese salt under argon protection and isolation by filtration gave the target Complex 2. ESI-MS (Electrospray Ionization-Mass Spectroscopy), NMR, elemental analysis and single crystal X-ray crystallography (see in Supplementary Information) were used to characterize the identity and purity of the SR-mesoPYBP ligand and its manganese complex.
The catalytic performance of Complex 1 and 2 with cyclooctene in H2O2 and acetic acid was investigated and optimized. The epoxidation yields were quantified by GC-MS, with chlorobenzene as an internal standard. As shown in Figure 1, Complex 2 exhibited some catalytic activity—up to a 100% conversion and 76% yield within 30 min—while Complex 1 was catalytically inactive.
The investigation of the acid loading revealed that the addition of a large excess of acid improved the olefin epoxidation yield with Complex 2 up to a point. As shown in Figure 2, the epoxidation yield increased dramatically from 15% to 70% by adding 2000 equiv. acetic acid. However, a high acid loading may also challenge the stability of Complex 2, and too much acid may lead to a complex decomposition and reduced catalytic activity. It is also worth noting that unlike Complex 1, Complex 2 is catalytically active in olefin epoxidation, even in the absence of acetic acid.
The limiting reagent test was designed to find the limiting reagent in this epoxidation reaction. The results supported that the consumption of the active Complex 2 limited the catalytic yield of olefin epoxidation. This is mainly caused by the leakage of manganese ions from the complexes during the reaction, for which ESI-MS confirmed the demetallation and the formation of free ligands (see Figure S5). As shown in Figure 3, the epoxidation reactions were initially performed for 30 min under a standard condition, and adding a fresh catalyst (with oxidant) could resume the catalytic activity of Complex 2. However, adding substrates, oxidants or substrates and oxidants cannot restore the catalytic activity. Moreover, adding fresh manganese salt to the reaction after 30 min did not assist the catalysis. Similar experiments have been conducted with Mn-mesoPYBP(ClO4)2 and Mn(ClO4)2, but both showed no catalytic reactivity.
The cyclic voltammetry test demonstrated that Complex 2 is more stable than Complex 1 at higher oxidation states. Complexes 1 or 2 (1 mM) with 0.1 M electrolytes (tetra-n-butyl ammonium hexafluorophosphate) were prepared in 5 mL of acetonitrile. Two drops of 0.1 mM Ferrocene were then added as a reference (red stars in Figure 4.). Complex 1 showed one reversible process (E1/2) at 0.56 V and Complex 2 showed a similar reversible process at 0.44 V. These two redox potentials indicate the redox couples of MnIII/II [20,29]. Earlier studies have suggested that the high electron density (basicity) and better steric accessibility of the donor atoms to the metal can lower the redox potentials and stabilize the higher oxidation states [1,20]. The lower redox potential of Complex 2 indicates that the higher electron density of the ligands (electron-rich PYBP ligands) contributes to the stability of the high oxidation states in the catalytic cycle
The EPR study was performed with 1 mM of Complexes 1 or 2, with 20 mM H2O2 and 20 mM acetic acid after 30 mins. All of the samples were prepared in acetonitrile, and then frozen in liquid nitrogen at 120 K. As seen in Figure 5, Complex 2 showed a six-line spectrum of MnII free ions superimposed over a 16-line spectrum of the mixed-valence binuclear MnIII/MnIV intermediates [26,30]. However, Complex 1 only showed the MnII ion spectrum, and could not be oxidized to a higher oxidation state, which indicates the catalytic inactivity of Complex 1. It is also worth noting that the EPR data of Complex 2 started to form a broad peak at g = 4.54, which represents an MnIV intermediate [31]. This evidence indirectly supported that the electronic effect assisted in maintaining the manganese in the center of Complex 2 and successfully stabilizing the high-valent intermediates. Therefore, Complex 2 was catalytically active, while Complex 1 decomposed.
An acid-assisted mechanism is proposed in Scheme 3, similar to the related work with BPBP and other ligand systems [4,15,18,19,32,33]. The active MnII catalysts can react with H2O2 in the presence of acetic acid, and form MIIIOOH(HCOOR) intermediates through an acid-assisted pathway. Talsi and co-workers speculated that high-valent MV = O was the key intermediate, which allowed oxygen transformation to happen in analogous metal catalyst systems [34]. However, there was no direct evidence to prove the existence of an MV = O intermediate in the PYBP catalytic cycle. Instead, mixed-valence binuclear MnIII/MnIV intermediates were present after the heterolytic O-O bond cleavage based on the EPR data. Therefore, further mechanistic studies are needed to confirm the existence of MnV = O in the PYBP system.

3. Materials and Methods

3.1. General

All of the chemicals and solvents were purchased from Sigma Aldrich (St. Louis, MO, USA ) or Fisher Scientific (Waltham, MA, USA) and were used without additional purification, unless otherwise noted. The ESI-MS data were monitored using ThermoFinnigan Electrospray Mass Spectrometry (ESI-MS) and a LTQ-LCQ ion trap mass spectrometer in the positive ion detection mode before and after the workup (ThermoFinnigan, San Jose, CA, USA). Pure acetonitrile was used as the solvent. 1H NMR and 13C NMR were performed on a Bruker Avance III NMR spectrometer at 500 MHz (Bruker, Billerica, MA, USA). CD3Cl was used as the solvent. The elemental analysis was conducted at the Microanalysis Laboratory School of Chemical Sciences, University of Illinois, Urbana-Champaign. The GC-MS experiments were carried out using a Shimadzu GC-17A gas chromatograph (Rtx-xLB column) with a GC-MS-QP 5050A mass detector (Shimadzu, Japan). The cyclic voltammetry experiments were performed on an EG&G PAR 273 potentiostat using a three-electrode cell with a glassy carbon working electrode and platinum wire counter- and reference-electrodes. The EPR spectra were acquired on a Bruker EMX EPR spectrometer at 120 K (Bruker, Billerica, MA, USA).

3.2. Synthesis of SR-MesoPYBP Ligand

Meso BP·2HBr (3.19 g, 9.66 mmol) and SR-picolyl chloride hydrochloride (4.29 g, 19.32 mmol, 2 equiv) were separately dissolved in deionized water. An NaOH solution (12 mL, 77.28 mmol, 8 equiv) was prepared by mixing a concentrated NaOH aqueous solution (4.1 mL, 17 M) and deionized water (7.9 mL). A diluted NaOH solution (6 mL, 4 equiv) was added into the meso BP·2HBr solution and stirred for 1 h. SR-picolyl chloride hydrochloride and the rest of the diluted NaOH solution (6 mL, 4 equiv) were then added dropwisely. CH2Cl2 (30 mL) was added to make the bi-phase solution and was vigorously stirred for 2–4 days. The progress of the reaction can be verified by ESI-MS. The additional concentrated NaOH (aq) (1 mL, 17 M) was treated with the bi-phase solution. The aqueous layer was extracted four times with 30 mL CH2Cl2. The combined CH2Cl2 layers were dried over MgSO4. Yellow semi-oil/semi-solid products were achieved by rotary-evaporation under reduced pressure. White powder was obtained by recrystallizing from MeOH. The best yield was around 2.58 g, 57%. ESI-MS m/z 467.45 (experimental) 467.33 (calculated); 1H NMR (CDCl3, 500 Hz): δ 8.07 (s, 2H); 4.05 (d, J = 12.5 Hz, 2H); 3.67 (s, 6H); 3.64 (d, J = 10.4 Hz, 2H); 2.75 (s, 2H); 2.52–2.58 (m, 2H); 2.31 (d, J = 13.9 Hz, 2H); 2.20 (s, 6H); 2.15 (s, 6H); 1.64 (t, J = 12.5 Hz, 2H); 1.44–1.51 (m, 4H); 0.98–1.14 (m, 2H); 0.95 (d, J = 11.0 Hz, 2H); 0.73 (s, 2H). 13C NMR (CDCl3, 500 MHz): δ 164.00; 158.72; 148.00; 126.43; 124.71; 59.69; 58.98; 54.51; 48.42; 21.05; 20.72; 19.59; 13.21; 10.92.

3.3. Synthesis of Manganese SR-mesoPYBP Complex

The reaction was performed under an Ar atmosphere. SR-mesoPYBP ligand (0.40 g, 0.86 mmol) was dissolved in CH2Cl2. Mn(ClO4)2·xH2O (x was not necessarily to be determined since excess Mn(ClO4)2 was added) (0.65 g, 2.58 mmol) was dissolved in minimal acetonitrile (no more than 1 mL). The solution was combined and stirred for 2 days. The progress of the reaction was tested by ESI-MS. Any precipitant was removed by filtration (excess manganese salts). The white powder was filtered by the dropwise addition of the solution into a large amount of Et2O (>1:50). The best yield was around 0.41 g, 66%. ESI-MS (Mn-SR-mesoPYBP(ClO4)+) m/z 620.36 (experimental) 620.22 (calculated). Elemental analysis (average of two runs): measured C%, 46.53 (calculated 46.68); H%, 5.82 (5.88); N%, 7.51 (7.78); Mn%, 7.23 (7.62).

3.4. General Catalysis Procedure

The olefin stock solution was prepared by dissolving cyclooctene (1.3224 g, 12 mmol) in 3 mL of acetonitrile. Chlorobenzene (0.5403 g, 4.8 mmol) was added into the solution as an internal standard for GC testing. Complex 1 or 2 (0.5 mL, 1 mM) was introduced into the olefin stock solution (25 μL, 100 μmol cyclooctene and 40 μmol chlorobenzene) with appropriate amount of additives (0–5000 equiv HOAc) at room temperature. A 5-μL solution was injected into 1 mL of ether as the initial sample. An appropriate amount of H2O2 was then injected into the solution in five portions in order to start the catalysis reaction. A 5-μL solution was injected into 1 mL of ether as the final sample. These samples were quantitatively analyzed by GC-MS. All of the samples were run at least three times, and the arithmetic means were used to report the data.

4. Conclusions

In conclusion, introducing electron-donating substituents on the mesoPYBP ligand improved the catalytic activity of the catalyst in olefin epoxidation with H2O2 in the presence of acetic acid. We successfully synthesized Complex 2 and optimized its catalytic performance by manipulating the acid loading. Mechanistic studies supported that an electronic effect can stabilize the high-valent intermediates in the PYBP catalytic cycle by maintaining manganese in the complex center. In addition to the metal racPYBP catalysts, the development of active electron-rich metal mesoPYBP catalysts makes the PYBP catalytic system a better candidate for olefin epoxidation. In the future, we want to introduce metal PYBP catalysts to a larger substrate scope, and hopefully offer another high-efficient, low-cost and environmental-friendly way to synthesize epoxides.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/3/285/s1. Figure S1: ESI-MS of SR-mesoPYBP ligand, Figure S2: 1H NMR of SR-mesoPYBP ligand, Figure S3: 13C NMR of SR-mesoPYBP ligand, Figure S4: ESI-MS of complex 2, Figure S5: ESI-MS of Entry 1 solution solution after 60 min reaction time, Figure S6: EPR spectrum for complex 1 at 120K at 6 min, 30 min, 1 day or more than 1 day, Figure S7: EPR spectrum for complex 2 at 120K at 6 min, 30 min, 1 day or more than 1 day, Figure S8: OPTRP diagram of SR-mesoPYBP ligand, Table S1: Complex 1 or 2 catalyzed olefin epoxidation reaction, Table S2: Limiting reagent test.

Author Contributions

Conceptualization, F.Z. and E.V.R.-A.; data curation, F.Z. and A.J.Z.; funding acquisition, E.V.R-A.; investigation, F.Z., G.Y. and A.J.Z.; methodology, F.Z. and G.Y.; project administration, E.V.R-A.; supervision, X.Z.; writing (original draft), F.Z.; writing (review and editing), X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by U.S. Department of Energy, DE-FG02-06ER15799 and National Science Foundation Grant, CHE 1412909.

Acknowledgments

This work was supported by the U.S. Department of Energy (grant DE-FG02-06ER15799) and the National Science Foundation (grant CHE 1412909). The ESI-MS spectrometer and NMR at Tufts University were supported by the National Science Foundation (grants CHEM-MRI 0320783 and CHE-MRI 0821508). We appreciate Elena Mihilova for the fundamental research of this project. We also would like to thank Samuel W. Thomas, Sergiy Kryatov and Jilin Han for their comments, which greatly improved the manuscript, and Terry Hass for the instrumental supplement.

Conflicts of Interest

The founding sponsors had no role in the design of the study; in the collection, analyses or interpretation of the data; in the writing of the manuscript and in the decision to publish the results.

References

  1. Mikhalyova, E.A.; Makhlynets, O.V.; Palluccio, T.D.; Filatov, A.S.; Rybak-Akimova, E.V. A new efficient iron catalyst for olefin epoxidation with hydrogen peroxide. Chem. Commun. 2012, 48, 687. [Google Scholar] [CrossRef] [PubMed]
  2. Jürgens, E.; Wucher, B.; Rominger, F.; Törnroos, K.W.; Kunz, D. Selective rearrangement of terminal epoxides into methylketones catalysed by a nucleophilic rhodium–NHC–pincer complex. Chem. Commun. 2015, 51, 1897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Krishnan, K.K.; Thomas, A.M.; Sindhu, K.S.; Anilkumar, G. Recent advances and perspectives in the manganese-catalysed epoxidation reactions. Tetrahedron 2016, 72, 1. [Google Scholar] [CrossRef]
  4. Ottenbacher, R.V.; Samsonenko, D.G.; Talsi, E.P.; Bryliakov, K.P. Highly Enantioselective Bioinspired Epoxidation of Electron-Deficient Olefins with H2O2 on Aminopyridine Mn Catalysts. ACS Catal. 2014, 4, 1599. [Google Scholar] [CrossRef]
  5. Ottenbacher, R.V.; Talsi, E.P.; Bryliakov, K.P. Bioinspired Mn-aminopyridine catalyzed epoxidations of olefins with various oxidants: Enantioselectivity and mechanism. Catal. Today 2016, 278, 30. [Google Scholar] [CrossRef]
  6. Lane, B.S.; Burgess, K. Metal-catalyzed epoxidations of alkenes with hydrogen peroxide. Chem. Rev. 2003, 103, 2457. [Google Scholar] [CrossRef]
  7. Ottenbacher, R.V.; Talsi, E.P.; Bryliakov, K.P. Direct selective oxidative functionalization of C–H bonds with H2O2: Mn-Aminopyridine complexes challenge the dominance of non-heme Fe catalysts. Molecules 2016, 21, 1454. [Google Scholar] [CrossRef] [Green Version]
  8. Milan, M.; Bietti, M.; Costas, M. Aliphatic C–H Bond Oxidation with Hydrogen Peroxide Catalyzed by Manganese Complexes: Directing Selectivity through Torsional Effects. Org. Lett. 2018, 20, 2720. [Google Scholar] [CrossRef]
  9. White, M.C.; Doyle, A.G.; Jacobsen, E.N. A Synthetically Useful, Self-Assembling MMO Mimic System for Catalytic Alkene Epoxidation with Aqueous H2O2. J. Am. Chem. Soc. 2001, 123, 7194. [Google Scholar] [CrossRef]
  10. Costas, M.; Tipton, A.K.; Chen, K.; Jo, D.H.; Que, J. Modeling Rieske Dioxygenases:  The First Example of Iron-Catalyzed Asymmetric cis-Dihydroxylation of Olefins. J. Am. Chem. Soc. 2001, 123, 6722. [Google Scholar] [CrossRef]
  11. Murphy, A.; Dubois, G.; Stack, T.D.P. Efficient epoxidation of electron-deficient olefins with a cationic manganese complex. J. Am. Chem. Soc. 2003, 125, 5250. [Google Scholar] [CrossRef] [PubMed]
  12. Makhlynets, O.V.; Oloo, W.N.; Moroz, Y.S.; Belaya, I.G.; Palluccio, T.D.; Filatov, A.S.; Müller, P.; Cranswick, M.A.; Que, L.; Rybak-Akimova, E.V. H 2 O 2 activation with biomimetic non-haem iron complexes and AcOH: Connecting the g = 2.7 EPR signal with a visible chromophore. Chem. Commun. 2014, 50, 645. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, X.J.; Yang, X.N.; Lee, Y.M.; Nam, W.; Sun, L.C. A nonheme manganese (IV)–Oxo species generated in photocatalytic reaction using water as an oxygen source. Chem. Commun. 2015, 51, 4013. [Google Scholar] [CrossRef] [PubMed]
  14. Ottenbacher, R.V.; Bryliakov, K.P.; Talsi, E.P. Non-Heme Manganese Complexes Catalyzed Asymmetric Epoxidation of Olefins by Peracetic Acid and Hydrogen Peroxide. Adv. Synth. Catal. 2011, 353, 885. [Google Scholar] [CrossRef]
  15. Lyakin, O.Y.; Ottenbacher, R.V.; Bryliakov, K.P.; Talsi, E.P. Dramatic effect of carboxylic acid on the electronic structure of the active species in Fe (PDP)-catalyzed asymmetric epoxidation. ACS Catal. 2012, 2, 1196. [Google Scholar] [CrossRef]
  16. Cussó, O.; Garcia-Bosch, I.; Font, D.; Ribas, X.; Lloret-Fillol, J.; Costas, M. Highly Stereoselective Epoxidation with H2O2 Catalyzed by Electron-Rich Aminopyridine Manganese Catalysts. Org. Lett. 2013, 15, 6158. [Google Scholar] [CrossRef]
  17. Cussó, O.; Garcia-Bosch, I.; Ribas, X.; Lloret-Fillol, J.; Costas, M. Asymmetric Epoxidation with H2O2 by Manipulating the Electronic Properties of Non-heme Iron Catalysts. J. Am. Chem. Soc. 2013, 135, 14871. [Google Scholar] [CrossRef]
  18. Zima, A.M.; Lyakin, O.Y.; Ottenbacher, R.V.; Bryliakov, K.P.; Talsi, E.P. Iron-Catalyzed Enantioselective Epoxidations with Various Oxidants: Evidence for Different Active Species and Epoxidation Mechanisms. ACS Catal. 2017, 7, 60–69. [Google Scholar] [CrossRef]
  19. Cussó, O.; Serrano-Plana, J.; Costas, M. Evidence of a Sole Oxygen Atom Transfer Agent in Asymmetric Epoxidations with Fe-pdp Catalysts. ACS Catal. 2017, 7, 5046. [Google Scholar] [CrossRef]
  20. Hölzl, S.M.; Altmann, P.J.; Kück, J.W.; Kühn, F.E. Speciation in iron epoxidation catalysis: A perspective on the discovery and role of non-heme iron (III)-hydroperoxo species in iron-catalyzed oxidation reactions. Coord. Chem. Rev. 2017, 352, 517. [Google Scholar] [CrossRef]
  21. Yazerski, V.A.; Spannring, P.; Gatineau, D.; Woerde, C.H.M.; Wieclawska, S.M.; Lutz, M.; Kleijn, H.; Klein Gebbink, R.J.M. Making Fe (BPBP)-catalyzed C–H and C [double bond, length as m-dash] C oxidations more affordable. Org. Biomol. Chem. 2014, 12, 2062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yang, G.; Rybak-Akimova, E.V.; Campana, C. {1, 1′-Bis [(pyridin-2-yl) methyl]-2, 2′-bipiperidyl}(perchlorato) copper (II) perchlorate. Acta Crystallogr. Sect. E Crystallogr. Commun. 2017, 73, 1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Yang, G. Bioinspired mononuclear transition-metal aminopyridine complexes for olefin epoxidation and molecular oxygen activation: Synthesis, characterization, catalysis and mechanistic studies. Ph.D. Thesis, Tufts University, Medford, MA, USA, 2017. [Google Scholar]
  24. Costas, M.; Que, L., Jr. Ligand topology tuning of iron-catalyzed hydrocarbon oxidations. Angew. Chem. Int. Ed. 2002, 41, 2179. [Google Scholar] [CrossRef]
  25. Saravanan, N.; Sankaralingam, M.; Palaniandavar, M. Manganese(ii) complexes of tetradentate 4N ligands with diazepane backbones for catalytic olefin epoxidation: Effect of nucleophilicity of peroxo complexes on reactivity. RSC Adv. 2014, 4, 12000. [Google Scholar] [CrossRef]
  26. Hubin, T.J.; McCormick, J.M.; Collinson, S.R.; Buchalova, M.; Perkins, C.M.; Alcock, N.W.; Kahol, P.K.; Raghunathan, A.; Busch, D.H. New iron (II) and manganese (II) complexes of two ultra-rigid, cross-bridged tetraazamacrocycles for catalysis and biomimicry. J. Am. Chem. Soc. 2000, 122, 2512. [Google Scholar] [CrossRef]
  27. Cavallo, L.; Jacobsen, H. J Electronic effects in (salen)Mn-based epoxidation catalysts. J. Org. Chem. 2003, 68, 6202. [Google Scholar] [CrossRef]
  28. Goldsmith, C.R.; Jiang, W. Alkene epoxidation catalyzed by manganese complexes with electronically modified 1,10-phenanthroline ligands. Inorg. Chim. Acta. 2012, 384, 340. [Google Scholar] [CrossRef]
  29. Odom, B.; Hanneke, D.; D’urso, B.; Gabrielse, G. New measurement of the electron magnetic moment using a one-electron quantum cyclotron. Phys. Rev. Lett. 2006, 97, 6. [Google Scholar] [CrossRef] [Green Version]
  30. Ottenbacher, R.V.; Bryliakov, K.P.; Talsi, E.P. Nonheme manganese-catalyzed asymmetric oxidation. A Lewis acid activation versus oxygen rebound mechanism: Evidence for the “Third Oxidant”. Inorg. Chem. 2010, 49, 8620. [Google Scholar] [CrossRef]
  31. Larrow, J.F.; Jacobsen, E.N. (R, R)-N, N′-BIS (3, 5-DI-TERT-BUTYLSALICYLIDENE)-1, 2-CYCLOHEXANEDIAMINO MANGANESE (III) CHLORIDE, A HIGHLY ENANTIOSELECTIVE EPOXIDATION CATALYST:(MANG ANESE, CHLORO 2, 2′-1, 2-CYCLOHEXANEDIYLBIS (NITRILOMETHYLIDYNE)-BIS4, 6-BIS (1, 1-DIMETHYLETHYL) PHENALATO (2-)-N, N′, O, O′-, SP-5-1 3-(1 R-TRANS--). Org. Synth. 1998, 75, 1. [Google Scholar]
  32. Mas-Balleste, R.; Que, L. Iron-catalyzed olefin epoxidation in the presence of acetic acid: Insights into the nature of the metal-based oxidant. J. Am. Chem. Soc. 2007, 129, 15964. [Google Scholar] [CrossRef] [PubMed]
  33. Du, J.; Miao, C.; Xia, C.; Lee, Y.M.; Nam, W.; Sun, W. Mechanistic Insights into the Enantioselective Epoxidation of Olefins by Bioinspired Manganese Complexes: Role of Carboxylic Acid and Nature of Active Oxidant. ACS Catal. 2018, 8, 4528. [Google Scholar] [CrossRef]
  34. Ottenbacher, R.V.; Talsi, E.P.; Bryliakov, K.P. Chiral Manganese Aminopyridine Complexes: The Versatile Catalysts of Chemo- and Stereoselective Oxidations with H2O2. Chem. Rec. 2018, 18, 78. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. PYBP (1,1′-bis[(pyridin-2-yl)methyl]-2,2′-bipiperidine)-catalyzed olefin epoxidation reaction.
Scheme 1. PYBP (1,1′-bis[(pyridin-2-yl)methyl]-2,2′-bipiperidine)-catalyzed olefin epoxidation reaction.
Catalysts 10 00285 sch001
Scheme 2. Synthesis of Complex 2.
Scheme 2. Synthesis of Complex 2.
Catalysts 10 00285 sch002
Figure 1. Catalytic performance of olefin epoxidation: (a) no catalyst, (b) manganese mesoPYBP and (c) SR-mesoPYBP in GC-MS. Reaction conditions: catalyst (0.5 μmol), cyclooctene (100 μmol), chlorobenzene internal standard (40 μmol), H2O2 (30% aqueous solution), HOAc (glacial) in 0.5 mL of acetonitrile, at room temperature. H2O2 was added in five equal portions in the first 4 min. Standard: 5.5 min, cyclooctene: 6 min, cyclooctene oxide: 9 min.
Figure 1. Catalytic performance of olefin epoxidation: (a) no catalyst, (b) manganese mesoPYBP and (c) SR-mesoPYBP in GC-MS. Reaction conditions: catalyst (0.5 μmol), cyclooctene (100 μmol), chlorobenzene internal standard (40 μmol), H2O2 (30% aqueous solution), HOAc (glacial) in 0.5 mL of acetonitrile, at room temperature. H2O2 was added in five equal portions in the first 4 min. Standard: 5.5 min, cyclooctene: 6 min, cyclooctene oxide: 9 min.
Catalysts 10 00285 g001
Figure 2. Investigation on acid loading for Mn-SR-mesoPYBP(ClO4)2-catalyzed epoxidation.
Figure 2. Investigation on acid loading for Mn-SR-mesoPYBP(ClO4)2-catalyzed epoxidation.
Catalysts 10 00285 g002
Figure 3. Limiting reagent test by GC for Complex 2 catalyzed epoxidation.
Figure 3. Limiting reagent test by GC for Complex 2 catalyzed epoxidation.
Catalysts 10 00285 g003
Figure 4. Cyclic voltammetry of Complexes 1 (red line) and 2 (blue line; Ferrocene was referenced by the star).
Figure 4. Cyclic voltammetry of Complexes 1 (red line) and 2 (blue line; Ferrocene was referenced by the star).
Catalysts 10 00285 g004
Figure 5. Electron paramagnetic resonance (EPR) study of manganese Complexes 1 (red line) and 2 (blue line) with 20 mM of H2O2 in the presence of 20 mM of AcOH after 30 min.
Figure 5. Electron paramagnetic resonance (EPR) study of manganese Complexes 1 (red line) and 2 (blue line) with 20 mM of H2O2 in the presence of 20 mM of AcOH after 30 min.
Catalysts 10 00285 g005
Scheme 3. Proposed mechanisms for the PYBP catalytic system.
Scheme 3. Proposed mechanisms for the PYBP catalytic system.
Catalysts 10 00285 sch003

Share and Cite

MDPI and ACS Style

Zhu, F.; Yang, G.; Zoll, A.J.; Rybak-Akimova, E.V.; Zhu, X. Manganese Catalysts with Tetradentate N-donor Pyridine-Appended Bipiperidine Ligands for Olefin Epoxidation Reactions: Ligand Electronic Effect and Mechanism. Catalysts 2020, 10, 285. https://doi.org/10.3390/catal10030285

AMA Style

Zhu F, Yang G, Zoll AJ, Rybak-Akimova EV, Zhu X. Manganese Catalysts with Tetradentate N-donor Pyridine-Appended Bipiperidine Ligands for Olefin Epoxidation Reactions: Ligand Electronic Effect and Mechanism. Catalysts. 2020; 10(3):285. https://doi.org/10.3390/catal10030285

Chicago/Turabian Style

Zhu, Fengfan, Guang Yang, Adam J. Zoll, Elena V. Rybak-Akimova, and Xinbao Zhu. 2020. "Manganese Catalysts with Tetradentate N-donor Pyridine-Appended Bipiperidine Ligands for Olefin Epoxidation Reactions: Ligand Electronic Effect and Mechanism" Catalysts 10, no. 3: 285. https://doi.org/10.3390/catal10030285

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