Next Article in Journal
Octahedral Growth of PtPd Nanocrystals
Next Article in Special Issue
Acidic Stabilization of the Dual-Aromatic Heterocyclic Anions
Previous Article in Journal
Surface Modification of Electrocatalyst for Optimal Adsorption of Reactants in Oxygen Evolution Reaction
Previous Article in Special Issue
Isoquinolone Syntheses by Annulation Protocols
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tandem Olefin Metathesis/α-Ketohydroxylation Revisited

Biological and Chemical Research Centre, Faculty of Chemistry, University of Warsaw, Żwirki i Wigury Street 101, 02-089 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
Submission received: 24 May 2021 / Revised: 7 June 2021 / Accepted: 7 June 2021 / Published: 9 June 2021
(This article belongs to the Special Issue Catalysts for the Synthesis of Heterocyclic Compounds)

Abstract

:
EWG-activated and polar quaternary ammonium salt-tagged ruthenium metathesis catalysts have been applied in a two-step one-pot metathesis-oxidation process leading to functionalized α-hydroxyketones (acyloins). In this assisted tandem process, the metathesis catalyst is used first to promote ring-closing metathesis (RCM) and cross-metathesis (CM) steps, then upon the action of Oxone™ converts into an oxidation catalyst able to transform the newly formed olefinic product into acyloin under mild conditions.

Graphical Abstract

1. Introduction

α-Hydroxyketones (acyloins)—a class of organic compounds that possess a hydroxy group adjacent to a ketone moiety—are not only interesting from a biological point of view (e.g., promoting of cell–cell communication in bacteria [1], exhibiting antibacterial/antifungal properties [2,3] as well as neuroactivity [4], and can be used in the treatment of Alzheimer’s disease [5] and as urease inhibitors [6]) but are also useful building blocks in organic synthesis [7,8]. For example, macrocyclic acyloins are convenient starting materials in the preparation of macrocyclic musk [9,10], and in fact the first industrial synthesis of musk was based on acyloin condensation [11]. In addition to acyloin condensation [7,12,13,14]—an established methodology in α-hydroxyketones synthesis—other transformations like, e.g., oxidation [15,16], epoxide opening [17,18], the Barbier reaction [19], reduction of 1,2-diketones [20,21], asymmetric reductive coupling of alkynes and aldehydes [22], or CO2-promoted regioselective hydration of propargylic alcohols [23], can be used. Most of them provide the desired α-hydroxyketones with good to excellent yields but are limited to specific substrates, which are not always easily available. In light of the above, a two-step protocol composed of olefin metathesis (OM) and oxidation reactions offers an interesting level of retrosynthetic orthogonality (Scheme 1a).
Recently, the Nobel-Prize-winning olefin metathesis reaction catalyzed by well-defined molybdenum- and ruthenium-based catalysts has found widespread application in a diverse range of synthetic protocols [24,25]. Moreover, the Ru-based olefin metathesis catalysts (such as Ru0-Ru2, Figure 1) have been found to efficiently promote numerous non-metathetical transformations [26,27]. Interestingly, olefin metathesis can be conveniently combined with such reactions, including reductions (Scheme 1a (I)) [28,29,30] and oxidations (Scheme 1a (II–III)), conducted in a one-pot step-wise sequence, or in tandem (Scheme 1b) [28,31]. Following this idea, a number of efficient sequential metathesis-dihydroxylation protocols has been developed by Blechert [32], Grubbs [33], and Jarosz [34] (Scheme 1a (II)). At the same time, Plietker reported that various olefins can be oxidized to cis-diols or acyloins with the use of RuCl3. When the reaction is performed in the presence of NaIO4, the in situ formed species “RuO4” afforded diols in high yields, whereas treatment with Oxone™ (potassium peroxymonosulfate) and NaHCO3 provided α-hydroxyketones [35,36,37,38]. Based on this observation, Snapper et al. described a tandem sequence for the preparation of cis-diols and α-hydroxyketones from olefinic precursors, where the Grubbs’ 2nd generation metathesis catalyst Ru0 (used in 5 to 10 mol% ammount) was—after the metathesis step was completed—used to catalyze oxidation of the newly formed C-C double bond utilizing Oxone™ as the stoichiometric oxidant (Scheme 1a (III)) [39].

2. Results

Drawing on the availability of Ru metathesis catalysts bearing EWG-activators [40,41] and N-heterocyclic carbene (NHC) ligands with quaternary ammonium tags [42] (Figure 1), we decided to re-investigate the tandem metathesis/α-ketohydroxylation of dienes and alkenes. Specifically, we hoped that with the help of modern, commercially available activated Ru2 and polar catalysts such as AquaMet (Ru3), FixCat (Ru4), and StickyCat (Ru5), we would be able to reduce the loading of the noble metal catalyst, simplify the purification, and extend the scope of the reaction.
To do so, we examined the efficiency of classical Ru1Ru2 and tagged Ru3Ru5 Ru-catalysts in a model tandem ring-closing metathesis (RCM)/α-ketohydroxylation sequence of diethyl diallylmalonate (S1) (Scheme 2). On the basis of Plietker’s and Snapper’s observations, the oxidation step was performed by Oxone™ as a stoichiometric oxidant in a 6:6:1 v/v/v mixture of MeCN/EtOAc/H2O in the presence of NaHCO3; however, we performed this step at room temperature (RT) and using ten- to five-times lower loading of the Ru catalyst (1 mol%) than it was in the original procedure. As presented in Scheme 2, the best results were found when the tagged catalysts Ru3 and Ru5 were employed as the ruthenium source. It is of note that these results are not worse than those previously reported in the literature with 5 mol% of Ru0 [39]. Interestingly, a heterogeneous system [43], based on Ru5 immobilized on mesoporous silica SBA-15, which we added to the starters in this competition in the hope of achieving higher catalyst activity and easier purification of the reaction mixture, gave the conversion 11 percent-points lower than the one obtained for Ru5 under homogeneous conditions (Scheme 2).
Quaternary ammonium salt tagged-NHC Ru catalysts such as Ru3Ru5 offer an additional practical benefit of easier purification [44,45]. When the RCM/oxidation reaction of S1 was conducted with a polar catalyst (Ru5), we noted that both the reaction mixture and the product P1b isolated by aqueous extraction and column chromatography on silica were visibly less colored (Figure 2) as compared with the same reaction made with the classical Grubbs benzylidene complex Ru0 (previously used in this transformation) [39].
Following the initial trial with diethyl diallylmalonate (S1), the RCM/α-ketohydroxylation sequence was tested with other dienes (Scheme 3). Essentially, we proceeded according to the protocol developed previously (Scheme 2); however, we made a number of small alterations. A diene substrate was treated with 1 mol% of the corresponding Ru catalysts (Conditions A = Ru2; Conditions B = Ru3; C–E = Ru5) in ethyl acetate at RT. The RCM was ≥95% complete within 1 h. Next, MeCN and H2O were added and the mixture was treated with NaHCO3 and solid Oxone™ at RT. After the oxidation step was completed (1 h), the solvents were evaporated and the resultant crude product was purified by column chromatography on silica gel. In Procedure D, an Ru scavenger SN (4.4 mol%) [46,47,48] was added after the metathesis step to fully deactivate the catalyst and see if it facilitated the purification of the reaction mixture. The acyloins were isolated in 26–90% overall yields. In Scheme 3 only the representative conditions individually optimized for each substrate are presented (for full set of optimization data, see SI, Scheme S6 and Table S3). We were pleased to note that 5-, 7-, and 16-membered carbocyclic acyloins (P1b, P2b, Pb3, P4b, P8b/8b’) were obtained in moderate to good yields. Interestingly, two low-molecular-weight 3-pyrroline derivatives (P5b, P6b) were not formed (substrate was not recovered), while a 2,3,6,7-tetrahydro-1H-azepine derivative (P7b) was obtained in a 61% yield. As anticipated [39], in the case of an unsymmetrical diene, a mixture of regioisomers (P8b and P8b’) was formed with 1.5:1 regioselectivity (for other example of macrocyclic acyloin formation, see SI, Scheme S6 and Table S3). In this case, an additive—quinone Qn—was used in the RCM reaction to block the C–C double bond isomerization (shift) that was previously noted during similar RCM macrocyclizations [49] (Conditions E, Scheme 3).
The above established conditions were also used in in situ oxidation of cross-metatheses (CM) products. The results presented in Scheme 4 demonstrate that a variety of functionalized products including acyloins with aliphatic (e.g., P9bP12b) and aromatic substituents (e.g., P13bP15b) and decorated by various functionalities (e.g., P10b, P11b, P12b, P15b) can be obtained in tandem CM/oxidation (for full data please see SI: Scheme S7 and Table S4). As previously in the case of RCM, the transformation involving CM is operationally simple, involving EtOAc as a solvent for CM followed by admixing of two other co-solvents leading to the formation of linear acyloins in 40–55% yields over two steps.
To show compatibility of the worked-out metathesis/α-ketohydroxylation conditions, we decided to convert a more advanced substrate—an estrone-alkene derivative—into substituted acyloin P16b. Despite the complexity of the steroid S16, the reaction proceeded in a surprisingly clean manner and required only a short silica gel pad to provide an analytically pure product P16b in a 52% yield over two steps (Scheme 5).
We next tackled an even more complex substrate S17, a derivative of Sildenafil (Viagra™), a drug used to treat erectile dysfunction and primary pulmonary hypertension [50]. This compound, in addition to numerous functional groups, contains N,N-diallyl-sulfonamide moiety which, when present in S5 and S6 substrates, was found to completely suppress the oxidation step and formation of the desired acyloin. Fortunately, substrate S17 proved to be susceptible to oxidation reaction, and we were able to obtain a new Sildenafil analogue P17b with a satisfactory yield of 47% (over two steps, Scheme 6).

3. Conclusions

In summary, a ruthenium-catalyzed assisted tandem transformation for the preparation of functionalized α-hydroxyketones has been re-investigated. As the catalyst for this transformation, the EWG-activated nitro-Hoveyda–Grubbs complex (Ru2) and the polar ammonium tagged Ru-complexes (Ru3Ru5) were used at room temperature without extra additives, or in the presence of quinone and/or isocyanide metal-scavenger. Importantly, it was possible to reduce the catalyst loading from 5 (in the case of RCM) to 10 times (in the case of CM), still allowing the diversely functionalized acyloins, including a heterocyclic analogue of Sildenafil, to be produced in moderate to good yields. In addition, we found that the tandem processes involving CM of alkenes can be performed in ethyl acetate as opposed to the original conditions [39] where this reaction was made in DCM, which was then evaporated and replaced with EtOAc before the oxidation step. Owing to its ease of use and retrosynthetic simplicity, together with the small optimization steps introduced by us, we believe that the Snapper’s metathesis/α-ketohydroxylation of alkenes and dienes will allow access to functionalized densely oxygenated products in a cost-effective and friendly manner.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11060719/s1. Scheme S1. General scheme of RCM reaction (first step in tandem metathesis/α-ketohydroxylation); Scheme S2. General scheme of macroRCM reaction (first step in tandem metathesis/α-ketohydroxylation); Scheme S3. General scheme of selfCM reaction (first step in tandem metathesis/α-ketohydroxylation); Scheme S4. General scheme of oxidation reaction (second step in tandem metathesis/α-ketohydroxylation); Scheme S5. Tandem metathesis/α-ketohydroxylation of DEDAM (S1); Scheme S6. Tandem metathesis/α-ketohydroxylation of dienes; Scheme S7. Tandem metathesis/α-ketohydroxylation of olefins; Table S1. Influence of time on the oxidation step in tandem metathesis/α-ketohydroxylation of DEDAM (S1) in the presence of 1 mol% of catalyst; Table S2. Influence of time on the oxidation step in tandem metathesis/α-ketohydroxylation of DEDAM (S1) in the presence of 0.1 mol% of catalyst; Table S3. Results of tandem RCM/α-ketohydroxylation reactions; Table S4. Results of tandem selfCM/α-ketohydroxylation reactions.

Author Contributions

Investigation, optimization, M.P., A.Z.; supervision, writing—original draft preparation, review and editing, A.K. and K.G. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful to the “Catalysis for the Twenty-First Century Chemical Industry” project carried out within the TEAM-TECH programme of the Foundation for Polish Science co-financed by the European Union from the European Regional Development under the Operational Programme Smart Growth. The study was carried out at the Biological and Chemical Research Centre, University of Warsaw, established within the project co-financed by the European Union from the European Regional Development Fund under the Operational Programme Innovative Economy, 2007–2013.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tiaden, A.; Spirig, T.; Hilbi, H. Bacterial gene regulation by alpha-hydroxyketone signaling. Trends Microbiol. 2010, 18, 288–297. [Google Scholar] [CrossRef] [PubMed]
  2. Stierle, A.A.; Stierle, D.B.; Decato, D.; Priestley, N.D.; Alverson, J.B.; Hoody, J.; McGrath, K.; Klepacki, D. The Berkeleylactones, Antibiotic Macrolides from Fungal Coculture. J. Nat. Prod. 2017, 80, 1150–1160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Gala, D.; DiBenedetto, D.J.; Clark, J.E.; Murphy, B.L.; Schumacher, D.P.; Steinman, M. Preparations of antifungal Sch 42427/SM 9164: Preparative chromatographic resolution, and total asymmetric synthesis via enzymatic preparation of chiral α-hydroxy arylketones. Tetrahedron Lett. 1996, 37, 611–614. [Google Scholar] [CrossRef]
  4. Lin, Z.; Marett, L.; Hughen, R.W.; Flores, M.; Forteza, I.; Ammon, M.A.; Concepcion, G.P.; Espino, S.; Olivera, B.M.; Rosenberg, G.; et al. Neuroactive diol and acyloin metabolites from cone snail-associated bacteria. Bioorg Med. Chem. Lett. 2013, 23, 4867–4869. [Google Scholar] [CrossRef] [Green Version]
  5. Wallace, O.B.; Smith, D.W.; Deshpande, M.S.; Polson, C.; Felsenstein, K.M. Inhibitors of Aβ production: Solid-phase synthesis and SAR of α-hydroxycarbonyl derivatives. Bioorganic Med. Chem. Lett. 2003, 13, 1203–1206. [Google Scholar] [CrossRef]
  6. Tanaka, T.; Kawase, M.; Tani, S. α-Hydroxyketones as inhibitors of urease. Bioorganic Med. Chem. 2004, 12, 501–505. [Google Scholar] [CrossRef]
  7. Finley, K.T. The Acyloin Condensation as a Cyclization Method. Chem. Rev. 1964, 64, 573–589. [Google Scholar] [CrossRef]
  8. Bloomfield, J.J.; Owsley, D.C.; Nelke, J.M. The Acyloin Condensation. Org. React. 2004, 23, 259–404. [Google Scholar]
  9. Williams, A.S. The Synthesis of Macrocyclic Musks. Synthesis 1999, 1999, 1707–1723. [Google Scholar] [CrossRef]
  10. Sytniczuk, A.; Milewski, M.; Kajetanowicz, A.; Grela, K. Preparation of macrocyclic musks via olefin metathesis: Comparison with classical syntheses and recent advances. Russ. Chem. Rev. 2020, 89, 469–490. [Google Scholar] [CrossRef]
  11. Prelog, V.; Frenkiel, L.; Kobelt, M.; Barman, P. Zur Kenntnis des Kohlenstoffringes. Ein Herstellungsverfahren für vielgliedrige Cyclanone. Helv. Chim. Acta 1947, 30, 1741–1749. [Google Scholar] [CrossRef]
  12. Rühlmann, K. Die Umsetzung von Carbonsäureestern mit Natrium in Gegenwart von Trimethylchlorsilan. Synthesis 1971, 1971, 236–253. [Google Scholar] [CrossRef]
  13. Menon, R.S.; Biju, A.T.; Nair, V. Recent advances in N-heterocyclic carbene (NHC)-catalysed benzoin reactions. Beilstein J. Org. Chem. 2016, 12, 444–461. [Google Scholar] [CrossRef] [Green Version]
  14. Gaggero, N.; Pandini, S. Advances in chemoselective intermolecular cross-benzoin-type condensation reactions. Org. Biomol. Chem. 2017, 15, 6867–6887. [Google Scholar] [CrossRef]
  15. Tamao, K.; Maeda, K. Silafunctional compounds in organic synthesis. 29.1 Oxidation of (alkenyl)alkoxysilanes to α-hydroxy ketones. Tetrahedron Lett. 1986, 27, 65–68. [Google Scholar] [CrossRef]
  16. Chen, C.; Feng, X.; Zhang, G.; Zhao, Q.; Huang, G. An Efficient Method for the Synthesis of α-Hydroxyalkyl Aryl Ketones. Synthesis 2008, 2008, 3205–3208. [Google Scholar]
  17. Gala, D.; DiBenedetto, D.J. A rational approach to chiral α-hydroxy aryl ketones from chiral aryl epoxides via regioselective, stereo retentive oxidative epoxide opening: Its application to the synthesis of antifungal Sch 42427/SM 9164. Tetrahedron Lett. 1994, 35, 8299–8302. [Google Scholar] [CrossRef]
  18. Zhang, J.; Wu, S.; Wu, J.; Li, Z. Enantioselective Cascade Biocatalysis via Epoxide Hydrolysis and Alcohol Oxidation: One-Pot Synthesis of (R)-α-Hydroxy Ketones from Meso- or Racemic Epoxides. ACS Catal. 2015, 5, 51–58. [Google Scholar] [CrossRef]
  19. Nair, V.; Ros, S.; Jayan, C.N. Gallium Mediated Barbier Reactions of 1,2-diones: A Facile Synthesis of α-hydroxy ketones. J. Chem. Res. 2001, 2001, 551–553. [Google Scholar] [CrossRef]
  20. Hayakawa, R.; Sahara, T.; Shimizu, M. Reduction of 1,2-diketones with titanium tetraiodide: A simple approach to α-hydroxy ketones. Tetrahedron Lett. 2000, 41, 7939–7942. [Google Scholar] [CrossRef]
  21. Khan, F.A.; Dash, J.; Sudheer, C. Indium-Mediated Regio- and Diastereoselective Reduction of Norbornyl α-Diketones. Chem.-A Eur. J. 2004, 10, 2507–2519. [Google Scholar] [CrossRef] [PubMed]
  22. Miller, K.M.; Huang, W.-S.; Jamison, T.F. Catalytic Asymmetric Reductive Coupling of Alkynes and Aldehydes:  Enantioselective Synthesis of Allylic Alcohols and α-Hydroxy Ketones. J. Am. Chem. Soc. 2003, 125, 3442–3443. [Google Scholar] [CrossRef] [PubMed]
  23. He, H.; Qi, C.; Hu, X.; Guan, Y.; Jiang, H. Efficient synthesis of tertiary α-hydroxy ketones through CO2-promoted regioselective hydration of propargylic alcohols. Green Chem. 2014, 16, 3729–3733. [Google Scholar] [CrossRef]
  24. Grela, K. (Ed.) Olefin Metathesis: Theory and Practice; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014. [Google Scholar]
  25. Grubbs, R.H.; Wenzel, A.G.; O’Leary, D.J.; Khosravi, E. (Eds.) Handbook of Metathesis; Wiley-VCH: Weinheim, Germany, 2015. [Google Scholar]
  26. Alcaide, B.; Almendros, P. Non-Metathetic Behavior Patterns of Grubbs’ Carbene. Chem.-A Eur. J. 2003, 9, 1258–1262. [Google Scholar] [CrossRef] [PubMed]
  27. Alcaide, B.; Almendros, P.; Luna, A. Grubbs’ Ruthenium-Carbenes Beyond the Metathesis Reaction: Less Conventional Non-Metathetic Utility. Chem. Rev. 2009, 109, 3817–3858. [Google Scholar] [CrossRef] [PubMed]
  28. Zieliński, G.K.; Grela, K. Tandem Catalysis Utilizing Olefin Metathesis Reactions. Chem.-A Eur. J. 2016, 22, 9440–9454. [Google Scholar] [CrossRef] [PubMed]
  29. Zieliński, G.K.; Samojłowicz, C.; Wdowik, T.; Grela, K. In tandem or alone: A remarkably selective transfer hydrogenation of alkenes catalyzed by ruthenium olefin metathesis catalysts. Org. Biomol. Chem. 2015, 13, 2684–2688. [Google Scholar] [CrossRef] [Green Version]
  30. Zieliński, G.K.; Majtczak, J.; Gutowski, M.; Grela, K. A Selective and Functional Group-Tolerant Ruthenium-Catalyzed Olefin Metathesis/Transfer Hydrogenation Tandem Sequence Using Formic Acid as Hydrogen Source. J. Org. Chem. 2018, 83, 2542–2553. [Google Scholar] [CrossRef]
  31. Fogg, D.E.; dos Santos, E.N. Tandem catalysis: A taxonomy and illustrative review. Coord. Chem. Rev. 2004, 248, 2365–2379. [Google Scholar] [CrossRef]
  32. Beligny, S.; Eibauer, S.; Maechling, S.; Blechert, S. Sequential Catalysis: A Metathesis/Dihydroxylation Sequence. Angew. Chem. Int. Ed. 2006, 45, 1900–1903. [Google Scholar] [CrossRef]
  33. Dornan, P.K.; Wickens, Z.K.; Grubbs, R.H. Tandem Z-Selective Cross-Metathesis/Dihydroxylation: Synthesis of anti-1,2-Diols. Angew. Chem. Int. Ed. 2015, 54, 7134–7138. [Google Scholar] [CrossRef] [Green Version]
  34. Malik, M.; Ceborska, M.; Witkowski, G.; Jarosz, S. Ruthenium-catalyzed one-pot ring-closing metathesis/syn-dihydroxylation in the synthesis of bicyclic iminosugars. Tetrahedron Asymmetry 2015, 26, 29–34. [Google Scholar] [CrossRef]
  35. Plietker, B.; Niggemann, M. An Improved Protocol for the RuO4-Catalyzed Dihydroxylation of Olefins. Org. Lett. 2003, 5, 3353–3356. [Google Scholar] [CrossRef]
  36. Plietker, B. RuO4-Catalyzed Ketohydroxylation of Olefins. J. Org. Chem. 2003, 68, 7123–7125. [Google Scholar] [CrossRef]
  37. Plietker, B. The RuO4-Catalyzed Ketohydroxylation. Part 1. Development, Scope, and Limitation. J. Org. Chem. 2004, 69, 8287–8296. [Google Scholar] [CrossRef]
  38. Plietker, B. The RuO4-Catalyzed Ketohydroxylation, Part II: A Regio-, Chemo- and Stereoselectivity Study. Eur. J. Org. Chem. 2005, 2005, 1919–1929. [Google Scholar] [CrossRef]
  39. Scholte, A.A.; An, M.H.; Snapper, M.L. Ruthenium-Catalyzed Tandem Olefin Metathesis−Oxidations. Org. Lett. 2006, 8, 4759–4762. [Google Scholar] [CrossRef]
  40. Kajetanowicz, A.; Grela, K. Nitro and Other Electron Withdrawing Group Activated Ruthenium Catalysts for Olefin Metathesis Reactions. Angew. Chem. Int. Ed. 2021, 60, 13738–13756. [Google Scholar] [CrossRef]
  41. Olszewski, T.K.; Bieniek, M.; Skowerski, K.; Grela, K. A New Tool in the Toolbox: Electron-Withdrawing Group Activated Ruthenium Catalysts for Olefin Metathesis. Synlett 2013, 24, 903–919. [Google Scholar] [CrossRef]
  42. Jana, A.; Grela, K. Forged and fashioned for faithfulness—ruthenium olefin metathesis catalysts bearing ammonium tags. Chem. Commun. 2018, 54, 122–139. [Google Scholar] [CrossRef]
  43. Chołuj, A.; Nogaś, W.; Patrzałek, M.; Krzesiński, P.; Chmielewski, M.J.; Kajetanowicz, A.; Grela, K. Preparation of Ruthenium Olefin Metathesis Catalysts Immobilized on MOF, SBA-15, and 13X for Probing Heterogeneous Boomerang Effect. Catalysts 2020, 10, 438. [Google Scholar] [CrossRef]
  44. Szczepaniak, G.; Kosinski, K.; Grela, K. Towards “cleaner” olefin metathesis: Tailoring the NHC ligand of second generation ruthenium catalysts to afford auxiliary traits. Green Chem. 2014, 16, 4474–4492. [Google Scholar] [CrossRef]
  45. Skowerski, K.; Wierzbicka, C.; Szczepaniak, G.; Gulajski, L.; Bieniek, M.; Grela, K. Easily removable olefin metathesis catalysts. Green Chem. 2012, 14, 3264–3268. [Google Scholar] [CrossRef]
  46. Szczepaniak, G.; Piątkowski, J.; Nogaś, W.; Lorandi, F.; Yerneni, S.S.; Fantin, M.; Ruszczyńska, A.; Enciso, A.E.; Bulska, E.; Grela, K.; et al. An isocyanide ligand for the rapid quenching and efficient removal of copper residues after Cu/TEMPO-catalyzed aerobic alcohol oxidation and atom transfer radical polymerization. Chem. Sci. 2020, 11, 4251–4262. [Google Scholar] [CrossRef] [Green Version]
  47. Szczepaniak, G.; Nogaś, W.; Piątkowski, J.; Ruszczyńska, A.; Bulska, E.; Grela, K. Semiheterogeneous Purification Protocol for the Removal of Ruthenium Impurities from Olefin Metathesis Reaction Products Using an Isocyanide Scavenger. Org. Process Res. Dev. 2019, 23, 836–844. [Google Scholar] [CrossRef]
  48. Szczepaniak, G.; Ruszczyńska, A.; Kosiński, K.; Bulska, E.; Grela, K. Highly efficient and time economical purification of olefin metathesis products from metal residues using an isocyanide scavenger. Green Chem. 2018, 20, 1280–1289. [Google Scholar] [CrossRef]
  49. Sytniczuk, A.; Leszczyńska, A.; Kajetanowicz, A.; Grela, K. Preparation of Musk-Smelling Macrocyclic Lactones from Biomass: Looking for the Optimal Substrate Combination. ChemSusChem 2018, 11, 3157–3166. [Google Scholar] [CrossRef]
  50. Nichols, D.J.; Muirhead, G.J.; Harness, J.A. Pharmacokinetics of sildenafil after single oral doses in healthy male subjects: Absolute bioavailability, food effects and dose proportionality. Br. J. Clin. Pharmacol. 2002, 53, 5S–12S. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. (a) Assisted tandem or sequential catalysis involving olefin metathesis (OM) and reduction or oxidation steps. (b) Classification of one-pot processes.
Scheme 1. (a) Assisted tandem or sequential catalysis involving olefin metathesis (OM) and reduction or oxidation steps. (b) Classification of one-pot processes.
Catalysts 11 00719 sch001
Figure 1. Selected ruthenium olefin metathesis catalysts. Cy = cyclohexyl.
Figure 1. Selected ruthenium olefin metathesis catalysts. Cy = cyclohexyl.
Catalysts 11 00719 g001
Scheme 2. Assisted tandem process involving olefin metathesis and Oxone™ oxidation. Yield determined by GC using durene as an internal standard. Conditions: Metathesis step: Ru-catalyst (1 mol%), 50 °C. 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). a Result in brackets is taken from ref. [39]. Conditions: Ru0 (5 mol %), RT, EtOAc, then NaHCO3, Oxone™, MeCN/H2O (6:1).
Scheme 2. Assisted tandem process involving olefin metathesis and Oxone™ oxidation. Yield determined by GC using durene as an internal standard. Conditions: Metathesis step: Ru-catalyst (1 mol%), 50 °C. 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). a Result in brackets is taken from ref. [39]. Conditions: Ru0 (5 mol %), RT, EtOAc, then NaHCO3, Oxone™, MeCN/H2O (6:1).
Catalysts 11 00719 sch002
Figure 2. Reaction mixture after RCM and oxidation of S1 conducted under identical conditions using (a) Grubbs Ru0 and (b) StickyCat catalyst Ru5. Product P1b isolated by column chromatography (CC) from (c) reaction with Ru0 and (d) with Ru5.
Figure 2. Reaction mixture after RCM and oxidation of S1 conducted under identical conditions using (a) Grubbs Ru0 and (b) StickyCat catalyst Ru5. Product P1b isolated by column chromatography (CC) from (c) reaction with Ru0 and (d) with Ru5.
Catalysts 11 00719 g002
Scheme 3. Scope and limitations in one-pot process involving RCM and oxidation steps. Conversion in RCM step ≥90% in all cases. Conditions: A = Metathesis step: Ru2 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). B = Metathesis step: Ru3 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). C = Metathesis step: Ru5 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). D = Metathesis step: Ru5 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc, then SnachCat™ (SN, 4.4 mol%). Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). E = Metathesis step: Ru5 (2 mol%), 50 °C, 1 h, 10 mM in EtOAc, quinone (Qn, 4 mol%), then concentration to 0.2 M. Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O 6/6/1 (v/v/v).
Scheme 3. Scope and limitations in one-pot process involving RCM and oxidation steps. Conversion in RCM step ≥90% in all cases. Conditions: A = Metathesis step: Ru2 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). B = Metathesis step: Ru3 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). C = Metathesis step: Ru5 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). D = Metathesis step: Ru5 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc, then SnachCat™ (SN, 4.4 mol%). Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v). E = Metathesis step: Ru5 (2 mol%), 50 °C, 1 h, 10 mM in EtOAc, quinone (Qn, 4 mol%), then concentration to 0.2 M. Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O 6/6/1 (v/v/v).
Catalysts 11 00719 sch003
Scheme 4. Scope and limitations in one-pot process involving CM and oxidation. Conditions: A = Metathesis step: Ru2 (1 mol%), Qn (4 mol%), 50 °C, 1 h, 0.5 M in EtOAc. Oxidation step: RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v), NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.). B = Metathesis step: Ru5 (1 mol%), Qn (4 mol%), 50 °C, 1 h, 0.5 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, up to 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v).
Scheme 4. Scope and limitations in one-pot process involving CM and oxidation. Conditions: A = Metathesis step: Ru2 (1 mol%), Qn (4 mol%), 50 °C, 1 h, 0.5 M in EtOAc. Oxidation step: RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v), NaHCO3 (2.5 equiv.), then Oxone™ (5 equiv.). B = Metathesis step: Ru5 (1 mol%), Qn (4 mol%), 50 °C, 1 h, 0.5 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.) then Oxone™ (5 equiv.), RT, up to 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v).
Catalysts 11 00719 sch004
Scheme 5. One-pot CM and oxidation of estrone derivative S16. Conditions A = Metathesis: Ru5 (1 mol%), 50 °C, 1 h, 0.5 M in EtOAc. Oxidation: NaHCO3 (2.5 equiv.), Oxone™ (5 equiv.), RT, 0.5 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v).
Scheme 5. One-pot CM and oxidation of estrone derivative S16. Conditions A = Metathesis: Ru5 (1 mol%), 50 °C, 1 h, 0.5 M in EtOAc. Oxidation: NaHCO3 (2.5 equiv.), Oxone™ (5 equiv.), RT, 0.5 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v).
Catalysts 11 00719 sch005
Scheme 6. One-pot CM and oxidation of an active pharmaceutical ingredient (API) derivative S17. Conditions B = Metathesis: Ru5 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v).
Scheme 6. One-pot CM and oxidation of an active pharmaceutical ingredient (API) derivative S17. Conditions B = Metathesis: Ru5 (1 mol%), 50 °C, 1 h, 0.2 M in EtOAc. Oxidation step: NaHCO3 (2.5 equiv.), Oxone™ (5 equiv.), RT, 1 h, 0.08 M in EtOAc/MeCN/H2O (6/6/1 v/v/v).
Catalysts 11 00719 sch006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patrzałek, M.; Zasada, A.; Kajetanowicz, A.; Grela, K. Tandem Olefin Metathesis/α-Ketohydroxylation Revisited. Catalysts 2021, 11, 719. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11060719

AMA Style

Patrzałek M, Zasada A, Kajetanowicz A, Grela K. Tandem Olefin Metathesis/α-Ketohydroxylation Revisited. Catalysts. 2021; 11(6):719. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11060719

Chicago/Turabian Style

Patrzałek, Michał, Aleksandra Zasada, Anna Kajetanowicz, and Karol Grela. 2021. "Tandem Olefin Metathesis/α-Ketohydroxylation Revisited" Catalysts 11, no. 6: 719. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11060719

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