Next Article in Journal
On-Surface Chemistry on Low-Reactive Surfaces
Previous Article in Journal
An Easy and Reliable Method for the Mitigation of Deuterated Chloroform Decomposition to Stabilise Susceptible NMR Samples
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fast Initiating Furan-Containing Hoveyda-Type Complexes: Synthesis and Applications in Metathesis Reactions

Biological and Chemical Research Centre, Faculty of Chemistry, University of Warsaw, Żwirki i Wigury 101, 02-089 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Chemistry 2022, 4(3), 786-795; https://doi.org/10.3390/chemistry4030056
Submission received: 30 June 2022 / Revised: 2 August 2022 / Accepted: 3 August 2022 / Published: 9 August 2022
(This article belongs to the Topic Catalysis: Homogeneous and Heterogeneous)

Abstract

:
Two new ruthenium complexes with chelating-ether benzylidene ligands bearing a furan moiety were synthesized and characterized, including X-ray crystallography. They initiated fast, also at 0 °C, and were found to be highly active in a variety of ring-closing, ene-yne, and cross-metathesis reactions, including an active pharmaceutical ingredient (API) model, which makes them good candidates for the transformation of complex polyfunctional compounds that require mild reaction conditions.

1. Introduction

In just a few decades, the olefin metathesis evolved from a chemical curiosity discovered accidentally in the 1960s to a useful methodology known to virtually every chemist [1,2]. This was possible due to understanding its mechanism [3], as well as the development of well-defined catalysts based on tungsten, molybdenum, and ruthenium [4,5,6,7]. The latter owe their popularity to their high stability in the presence of moisture and oxygen, tolerance to most known functional groups, mild reaction conditions, and the possibility of fine-tuning their structure to control chemical properties. In this context, N-heterocyclic carbenes (NHCs) [8,9] and their analogues, viz. unsymmetrical N-heterocyclic carbenes (uNHCs) [10] and cyclic (alkyl) (amino) carbenes (CAACs) [11] have received the most attention (Figure 1a). Modifications of the benzylidene ligands in Hoveyda–Grubbs-type complexes also offer wide possibilities to control the catalytic properties. Therefore, the introduction of electron-withdrawing [12,13] or bulky substituents (the latter in the ortho position to the OiPr group) [14] to the aromatic ring of the benzylidene ligand accelerates the initiation rate, while the replacement of the chelating oxygen atom with sulfur [15,16,17], selenium [18,19], or nitrogen [20,21,22] results in latent catalysts activated by light [23,24] or temperature [24]. The structure of the substituent on the chelating oxygen atom also plays an important role. Replacement of the isopropyl substituent with the smaller methyl group (Ru5, Figure 1b) had a significant impact on the activity and stability of the resulting complex [25,26]. The larger isopropyl substituent not only facilitates the dissociation of the oxygen atom from Ru during initiation, but also allows for the more effective protection of the metal center from undesirable side reactions leading to catalyst decomposition. On the other hand, replacement of the iPr group with the phenyl one reduced the steric bulk and, at the same time, decreased the donation of diaryl ether oxygen atoms, leading to stable and rapidly initiating catalysts (Ru6) [27]. Recently, this structural motif has been applied to fast initiating Z-selective catalysts developed by Grubbs [28,29]. Further modifications of the alkyl substituent were independently conducted by Grubbs [30], Diver [31], and Grela [32]. Grubbs et al. studied complexes bearing various small to large substituents at the chelating oxygen (e.g., Ru7) and observed their impact on the strength and length of the Ru-O bond, as well as the catalyst initiation rate [30].
In a similar set of complexes possessing a cyclic fragment (such as Ru8), Grela and co-workers observed an influence of ring size on catalyst activity [32], while Diver noted a significant influence of the axial or equatorial conformation of the differently substituted cyclohexyl ethers (e.g., Ru9) on the initiation rate in ring-closing metathesis reactions [31]. A different approach was presented by Grela et al. [33,34,35], who, by introducing an electron-withdrawing group as a terminal substituent of the ‘leaving’ benzylidene ether group (Ru10, Ru11), boosted the activity of Hoveyda catalysts. At the same time, the authors noted that substituents such as an ester, ketonic, or a malonic group work, as there is an additional coordinating functionality binding to the metal center. In addition, an analogue of Ru11 that contains free carboxylic acid in the ether moiety easily undergoes cyclisation to form a complex Ru13 containing a chelating carboxylate ligand [36], which can be activated in situ by acids and has found some applications in metathesis reactions [37]. Subsequently, the same concept was creatively developed by Skowerski and Olszewski [38], Liu and Wang [39], Matsuto [40], and Al-Awadi [41]. The ethereal substituent in the benzylidene ligand can also serve as a platform to increase the solubility of catalysts in polar media [42] or to allow the immobilization of the resulting complexes [43,44,45]. This short and inevitably fragmentary introduction to a waste collection of olefin metathesis catalysts shows that the ligand engineering within the coordination sphere of the Ru atom is an important field of research, as it can bring about the control of catalyst initiation and productivity and introduce new traits such as solubility in given solvents, immobilization handles, etc., [46,47].
Understanding the influence of the modification of the chelating alkoxy-benzylidene ligand on the structure and catalytic activity of the resulting ruthenium complexes, we decided to synthesize a catalyst containing an oxomethylenefuran group as the ethereal-chelating fragment (the idea is presented as a prototypical structure Ru16 in Figure 2). This design was inspired by a promising catalytic profile exhibited by Ru14 and Ru15 that featured benzyl-ether fragments in the chelating benzylidene ligand [48].

2. Results

We first approached the synthesis of the ligand precursor 4 (Scheme 1). The rationale behind selecting this structure was the known stability of brominated furan 2 and the general reliability of this reaction. In this regard, we performed the bromination of methyl 2-methyl-3-furancarboxylate (1) using NBS in the presence of AIBN and obtained product 2 in 70% yield. We then reacted the resulted bromide with 2-propenylphenol (3) and obtained the desired ligand precursor 4 in 77% yield (Scheme 1). In the alternative approach, propenylbenzene derivative 4 was prepared in a two-step procedure; first, a reaction of 2 with salicylaldehyde was performed, followed by Wittig reaction, giving the desired product with 32% yield (for details, see Supplementary Materials).
With propenylbenzene derivative 4 in hand, we prepared two versions of Hoveyda–Grubbs type complexes containing SIMes and SIPr NHC ligands, respectively. To do so, the reactions between the corresponding indenylidene-type complex, namely Ru2 and its SIPr analogue, and 4 were carried out in DCM at room temperature in the presence of CuCl used as a phosphine scavenger (Scheme 2). In both cases, the desired catalysts were obtained as green crystals in high yields, around 80%.
The new catalysts Ru16 and Ru17 were fully characterized by 1H and 13C NMR spectroscopy, as well as elemental analysis, MS, and IR spectroscopy. The signals of the benzylidene protons in the NMR spectra appeared at 16.54 and 16.34 ppm, which is typical of Hoveyda-type complexes. A single crystal of catalysts Ru16 was grown and also analyzed using XRD (Figure 3). The studied complex crystallizes in P21/c monoclinic space with one molecule in the asymmetric unit. The coordination sphere of the ruthenium atom is slightly distorted from the trigonal bipyramid. The geometrical features of the catalyst were compared with previously reported Hoveyda–Grubbs complex (Ru3) [48]. Most of the bond distances between the metal center and the atom in the first coordination sphere do not differ more than 3σ with the exception of the Ru1-O1 distance that is significantly elongated from 2.256(1) Å for Hoveyda to 2.282(1) Å for Ru16. This bond is even shorter for Ru10′ molecule with methyl ester moiety. Unfortunately, we have not observed any interactions between the oxygen atom, neither in the furan ring nor in the ester group, and the ruthenium center, as the Ru1-O2 distance is 3.352(2) Å, and it is much longer in comparison to 2.536(2) Å for Ru10′.
The molecular overlay presented in Figure 4 revealed differences in the position of the benzylidene and NHC ligand due to the bulky substituent replacement of the isopropoxy ligand. The torsion angle of Ru1-O1-C29-C30 is 69.1(2)° compared to the analog angle of 19.5(4) ° for Ru10′ and −18.2(2)° in Hoveyda–Grubbs catalyst. It can also explain the change in the position of the benzylidene ligand that is pushed back and the Ru1-C22-C23-C24 torsion angle is positive (174.1(1)° for Ru16 and 171.5(3)° for Ru10′) compared to Hoveyda–Grubbs negative value (−173.8(1)°). Additionally, the NHC ligand is twisted in such a way that the methyl groups pointing towards the viewer in Figure 4a are closer to one another by 2 Å comparing the distance between the C21 and C10 atoms equal to 3.908(3) Å for Ru16, 4.292(6) Å for Ru10′ vs. 5.732(2) Å for Hoveyda–Grubbs.
With both complexes in hand, we investigated their activity in model metathesis reactions to check the profile of their applications. The results were compared with two known catalysts, the commercially available Hoveyda–Grubbs complex Ru3 and its analogue with the ester group, Ru10. First, we carried out a model ring-closing metathesis (RCM) reaction of diethyl diallylmalonate (5, Figure 5) in the presence of 1 mol% of the examined complexes at 0 °C. Such a low temperature is rarely used in olefin metathesis reactions, because only the most active catalysts allow satisfactory conversions, but we believed that the system we designed was capable of such a challenging task [33,49].
As expected, the Hoveyda–Grubbs complex Ru3 initiates the slowest and also gives the lower conversion, only 64% after 4 h. All other complexes, viz. Ru10 (18-electron double-chelated complex), new Ru16, and Ru17, behaved in a similar way, each of them initiated relatively fast and reached almost full conversion within 120 min.
Based on this preliminary study, we selected Ru10 as a reference point for further comparison of the activity of newly obtained catalysts.
First, we examined the RCM of a more demanding substrate with a substituted double bond, namely diethyl 2-allyl-2-(2-methylallyl)malonate (7, Table 1, entry 1). When the reaction was performed at room temperature in the presence of 1 mol% of catalyst, in all cases, the conversion was quantitative or almost quantitative; however, Ru10 required three or nine times more time than the furan-containing compounds Ru16 and Ru17. When the catalyst loading was decreased to 0.2 mol% the conversion dropped significantly, but new complexes still allowed for reaching around 60% yield. Ru10 provided the desired product with a 40% yield that only slightly increased to 49% when the catalyst loading was increased to 0.5 mol%. A similar trend was observed in the case of the next RCM reaction, this time a proline derivative 9 (Table 1, entry 2). Moreover, the best result was obtained when the SIPr version of the furan-containing complex was used, meaning that an almost quantitative yield was reached after two hours at room temperature. Ru16 was slightly worse and provided the desired product 10 in 85% yield, while Ru10 reached only 49% of yield, but only when a higher catalyst loading of 0.5 mol% was used. The situation slightly changed in the case of the ene-yne reaction of allyl 1,1-diphenylpropargyl ether (11). Here, all complexes exhibited high activity, Ru10 and Ru17—used in 0.2 mol% loading—reached almost full conversion in 2 h while Ru16 gave a similar result (92%) in only 15 min (Table 1, entry 3). When the loading was raised to 1 mol%, all complexes provided the desired product in 100% yield, but after varying periods of time. In a cross-metathesis reaction between allyl benzene (13) and cis-1,4-diacetoxy-2-butene, the best result was obtained when Ru16 was used as a catalyst while the remaining complexes provided product 14 in a less than 80% yield (Table 1, entry 4). On the other hand, when estrone derivative 15 and methyl acrylate were used as substrates, all catalysts gave similar results, reaching an over 90% yield (Table 1, entry 5).
Encouraged by these results, we turned our attention to compounds with potential biological activity. This time, it was an analogue of Vardenafil, a popular drug utilized in the treatment of erectile dysfunction and pulmonary arterial hypertension, sold inter alia under the trade name Levitra [50]. From a synthetic point of view, the structure of the substrate can cause some problems during a metathesis reaction, as it contains a number of Lewis basic centers that can chelate the propagating ruthenium species, decreasing the activity of the catalyst. After a short optimization, including finding the best solvent, temperature, and reaction time (for details, see Supplementary Materials), we were able to obtain the desired product 18 in 77% yield (Scheme 3). This result is slightly worse than the best one known in the literature [51]; however, in the latter case, 1.5–2 mol% of catalyst bearing an unsymmetrical NHC ligand with thiophene moiety was used to achieve a 91% yield. Nevertheless, when the reaction was repeated in the presence of 2 mol% of Ru17, we were able to achieve the same result, 91%, as reported previously.

3. Conclusions

The straightforward reaction between the easily accessible furan-containing benzylidene ligand precursor 4 and second-generation indenylidene complexes gave two new Hoveyda–Grubbs type catalysts in high yields (≥80%). These new complexes were fully characterized, and their catalytic activity was examined using a diverse set of olefin metathesis reactions. The new complexes were found to be fast initiating and highly efficient. Among others, they exhibited high activity in RCM, ene-yne, and cross-metathesis reactions in low catalyst-loading (0.2–1 mol%), including the transformation of a derivative of the known Vardenafil (Levitra™) API. Interestingly, in a model RCM of diethyl diallylmalonate substrate (5) conducted at 0 °C, the new complexes have shown visibly higher activity than the one exhibited by standard Hoveyda–Grubbs catalyst (Ru3), successfully rivaling with the 18-electron, double-chelated complex Ru10.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry4030056/s1. Detailed experimental procedures and copies of NMR spectra [51,52,53,54,55,56,57,58,59].

Author Contributions

Conceptualization, A.K.; formal analysis, M.N., A.Z. and M.M.; investigation, M.N. and A.Z.; data curation, M.N. and A.K.; writing—original draft preparation, A.K.; writing—review and editing, M.N., M.M. and A.K.; visualization, M.N., M.M. and A.K.; supervision, A.K.; project administration, A.K.; funding acquisition, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research performed within SONATA BIS project and was funded by National Science Centre, Poland, grant number DEC-2021/42/E/ST4/00187.

Data Availability Statement

Data supporting reported results of this study are available in the supplementary material of this article and can be obtained from the corresponding author.

Acknowledgments

The study was carried out at the Biological and Chemical Research Centre, University of Warsaw, established within the project co-financed by European Union from the European Regional Development Fund under the Operational Program Innovative Economy, 2007–2013.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Grela, K. (Ed.) Olefin Metathesis: Theory and Practice; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014. [Google Scholar]
  2. Grubbs, R.H.; Wenzel, A.G.; O’Leary, D.J.; Khosravi, E. (Eds.) Handbook of Metathesis; Wiley-VCH: Weinheim, Germany, 2015. [Google Scholar]
  3. Hérisson, P.J.-L.; Chauvin, Y. Catalyse de Transformation des Oléfines par Les Complexes du Tungstène. II. Télomérisation des Oléfines Cycliques en Présence D’oléfines Acycliques. Makromol. Chem. 1971, 141, 161–176. [Google Scholar] [CrossRef]
  4. Schwab, P.; Grubbs, R.H.; Ziller, J.W. Synthesis and Applications of RuCl2(CHR’)(PR3)2:  The Influence of the Alkylidene Moiety on Metathesis Activity. J. Am. Chem. Soc. 1996, 118, 100–110. [Google Scholar] [CrossRef]
  5. Harrity, J.P.A.; La, D.S.; Cefalo, D.R.; Visser, M.S.; Hoveyda, A.H. Chromenes through Metal-Catalyzed Reactions of Styrenyl Ethers. Mechanism and Utility in Synthesis. J. Am. Chem. Soc. 1998, 120, 2343–2351. [Google Scholar] [CrossRef]
  6. Schrock, R.R.; Murdzek, J.S.; Bazan, G.C.; Robbins, J.; DiMare, M.; O’Regan, M. Synthesis of molybdenum imido alkylidene complexes and some reactions involving acyclic olefins. J. Am. Chem. Soc. 1990, 112, 3875–3886. [Google Scholar] [CrossRef]
  7. Schrock, R.R.; Hoveyda, A.H. Molybdenum and Tungsten Imido Alkylidene Complexes as Efficient Olefin-Metathesis Catalysts. Angew. Chem. Int. Ed. 2003, 42, 4592–4633. [Google Scholar] [CrossRef] [PubMed]
  8. Samojłowicz, C.; Bieniek, M.; Grela, K. Ruthenium-Based Olefin Metathesis Catalysts Bearing N-Heterocyclic Carbene Ligands. Chem. Rev. 2009, 109, 3708–3742. [Google Scholar] [CrossRef]
  9. Vougioukalakis, G.C.; Grubbs, R.H. Ruthenium-Based Heterocyclic Carbene-Coordinated Olefin Metathesis Catalysts. Chem. Rev. 2009, 110, 1746–1787. [Google Scholar] [CrossRef]
  10. Monsigny, L.; Kajetanowicz, A.; Grela, K. Ruthenium Complexes Featuring Unsymmetrical N-Heterocyclic Carbene Ligands–Useful Olefin Metathesis Catalysts for Special Tasks. Chem. Rec. 2021, 21, 3648–3661. [Google Scholar] [CrossRef]
  11. Morvan, J.; Mauduit, M.; Bertrand, G.; Jazzar, R. Cyclic (Alkyl)(amino)carbenes (CAACs) in Ruthenium Olefin Metathesis. ACS Catal. 2021, 11, 1714–1748. [Google Scholar] [CrossRef]
  12. 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]
  13. 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]
  14. Wakamatsu, H.; Blechert, S. A New Highly Efficient Ruthenium Metathesis Catalyst. Angew. Chem. Int. Ed. 2002, 41, 2403–2405. [Google Scholar] [CrossRef]
  15. Ben-Asuly, A.; Tzur, E.; Diesendruck, C.E.; Sigalov, M.; Goldberg, I.; Lemcoff, N.G. A Thermally Switchable Latent Ruthenium Olefin Metathesis Catalyst. Organometallics 2008, 27, 811–813. [Google Scholar] [CrossRef]
  16. Szadkowska, A.; Makal, A.; Woźniak, K.; Kadyrov, R.; Grela, K. Ruthenium Olefin Metathesis Initiators Bearing Chelating Sulfoxide Ligands. Organometallics 2009, 28, 2693–2700. [Google Scholar] [CrossRef]
  17. Tzur, E.; Szadkowska, A.; Ben-Asuly , A.; Makal, A.; Goldberg, I.; Woźniak, K.; Grela , K.; Lemcoff, N.G. Studies on Electronic Effects in O-, N- and S-Chelated Ruthenium Olefin-Metathesis Catalysts. Chem. Eur. J. 2010, 16, 8726–8737. [Google Scholar] [CrossRef]
  18. Monsigny, L.; Cejas Sánchez, J.; Piątkowski, J.; Kajetanowicz, A.; Grela, K. Synthesis and Catalytic Properties of a Very Latent Selenium-Chelated Ruthenium Benzylidene Olefin Metathesis Catalyst. Organometallics 2021, 40, 3608–3616. [Google Scholar] [CrossRef] [PubMed]
  19. Diesendruck, C.E.; Tzur, E.; Ben-Asuly, A.; Goldberg, I.; Straub, B.F.; Lemcoff, N.G. Predicting the Cis−Trans Dichloro Configuration of Group 15−16 Chelated Ruthenium Olefin Metathesis Complexes: A DFT and Experimental Study. Inorg. Chem. 2009, 48, 10819–10825. [Google Scholar] [CrossRef]
  20. Żukowska, K.; Szadkowska, A.; Pazio, A.E.; Woźniak, K.; Grela, K. Thermal Switchability of N-Chelating Hoveyda-type Catalyst Containing a Secondary Amine Ligand. Organometallics 2012, 31, 462–469. [Google Scholar] [CrossRef]
  21. Barbasiewicz, M.; Szadkowska, A.; Bujok, R.; Grela, K. Structure and Activity Peculiarities of Ruthenium Quinoline and Quinoxaline Complexes:  Novel Metathesis Catalysts. Organometallics 2006, 25, 3599–3604. [Google Scholar] [CrossRef]
  22. Polyanskii, K.B.; Alekseeva, K.A.; Raspertov, P.V.; Kumandin, P.A.; Nikitina, E.V.; Gurbanov, A.V.; Zubkov, F.I. Hoveyda–Grubbs catalysts with an N→Ru coordinate bond in a six-membered ring. Synthesis of stable, industrially scalable, highly efficient ruthenium metathesis catalysts and 2-vinylbenzylamine ligands as their precursors. Beilstein J. Org. Chem. 2019, 15, 769–779. [Google Scholar] [CrossRef]
  23. Eivgi, O.; Phatake, R.S.; Nechmad, N.B.; Lemcoff, N.G. Light-Activated Olefin Metathesis: Catalyst Development, Synthesis, and Applications. Acc. Chem. Res. 2020, 53, 2456–2471. [Google Scholar] [CrossRef]
  24. Ivry, E.; Frenklah, A.; Ginzburg, Y.; Levin, E.; Goldberg, I.; Kozuch, S.; Lemcoff, N.G.; Tzur, E. Light- and Thermal-Activated Olefin Metathesis of Hindered Substrates. Organometallics 2018, 37, 176–181. [Google Scholar] [CrossRef]
  25. Kingsbury, J.S.; Harrity, J.P.A.; Bonitatebus, P.J.; Hoveyda, A.H. A Recyclable Ru-Based Metathesis Catalyst. J. Am. Chem. Soc. 1999, 121, 791–799. [Google Scholar] [CrossRef]
  26. Ferré-Filmon, K.; Delaude, L.; Demonceau, A.; Noels, A.F. Stereoselective Synthesis of (E)-Hydroxystilbenoids by Ruthenium-Catalyzed Cross-Metathesis. Eur. J. Org. Chem. 2005, 2005, 3319–3325. [Google Scholar] [CrossRef]
  27. Kos, P.; Savka, R.; Plenio, H. Fast Olefin Metathesis: Synthesis of 2-Aryloxy-Substituted Hoveyda-Type Complexes and Application in Ring-Closing Metathesis. Adv. Synth. Catal. 2013, 355, 439–447. [Google Scholar] [CrossRef]
  28. Xu, Y.; Wong, J.J.; Samkian, A.E.; Ko, J.H.; Chen, S.; Houk, K.N.; Grubbs, R.H. Efficient Z-Selective Olefin-Acrylamide Cross-Metathesis Enabled by Sterically Demanding Cyclometalated Ruthenium Catalysts. J. Am. Chem. Soc. 2020, 142, 20987–20993. [Google Scholar] [CrossRef]
  29. Xu, Y.; Gan, Q.; Samkian, A.E.; Ko, J.H.; Grubbs, R.H. Bulky Cyclometalated Ruthenium Nitrates for Challenging Z-Selective Metathesis: Efficient One-Step Access to α-Oxygenated Z-Olefins from Acrylates and Allyl Alcohols. Angew. Chem. Int. Ed. 2022, 61, e202113089. [Google Scholar] [CrossRef]
  30. Engle, K.M.; Lu, G.; Luo, S.-X.; Henling, L.M.; Takase, M.K.; Liu, P.; Houk, K.N.; Grubbs, R.H. Origins of Initiation Rate Differences in Ruthenium Olefin Metathesis Catalysts Containing Chelating Benzylidenes. J. Am. Chem. Soc. 2015, 137, 5782–5792. [Google Scholar] [CrossRef] [Green Version]
  31. Gregg, Z.R.; Griffiths, J.R.; Diver, S.T. Conformational Control of Initiation Rate in Hoveyda–Grubbs Precatalysts. Organometallics 2018, 37, 1526–1533. [Google Scholar] [CrossRef]
  32. Zieliński, A.; Szczepaniak, G.; Gajda, R.; Woźniak, K.; Trzaskowski, B.; Vidović, D.; Kajetanowicz, A.; Grela, K. Ruthenium Olefin Metathesis Catalysts Systematically Modified in Chelating Benzylidene Ether Fragment: Experiment and Computations. Eur. J. Inorg. Chem. 2018, 2018, 3675–3685. [Google Scholar] [CrossRef]
  33. Bieniek, M.; Bujok, R.; Cabaj, M.; Lugan, N.; Lavigne, G.; Arlt, D.; Grela, K. Advanced Fine-Tuning of Grubbs/Hoveyda Olefin Metathesis Catalysts:  A Further Step toward an Optimum Balance between Antinomic Properties. J. Am. Chem. Soc. 2006, 128, 13652–13653. [Google Scholar] [CrossRef] [PubMed]
  34. Bieniek, M.; Samojłowicz, C.; Sashuk, V.; Bujok, R.; Śledź, P.; Lugan, N.; Lavigne, G.; Arlt, D.; Grela, K. Rational Design and Evaluation of Upgraded Grubbs/Hoveyda Olefin Metathesis Catalysts: Polyfunctional Benzylidene Ethers on the Test Bench. Organometallics 2011, 30, 4144–4158. [Google Scholar] [CrossRef]
  35. Guidone, S.; Blondiaux, E.; Samojłowicz, C.; Gułajski, Ł.; Kędziorek, M.; Malińska, M.; Pazio, A.; Woźniak, K.; Grela, K.; Doppiu, A.; et al. Catalytic and Structural Studies of Hoveyda–Grubbs Type Pre-Catalysts Bearing Modified Ether Ligands. Adv. Synth. Catal. 2012, 354, 2734–2742. [Google Scholar] [CrossRef]
  36. Gawin, R.; Makal, A.; Woźniak, K.; Mauduit, M.; Grela, K. A Dormant Ruthenium Catalyst Bearing a Chelating Carboxylate Ligand: In Situ Activation and Application in Metathesis Reactions. Angew. Chem. Int. Ed. 2007, 46, 7206–7209. [Google Scholar] [CrossRef] [PubMed]
  37. Gawin, R.; Czarnecka, P.; Grela, K. Ruthenium Catalysts Bearing Chelating Carboxylate Ligands: Application to Metathesis Reactions in Water. Tetrahedron 2010, 66, 1051–1056. [Google Scholar] [CrossRef]
  38. Skowerski, K.; Kasprzycki, P.; Bieniek, M.; Olszewski, T.K. Efficient, durable and reusable olefin metathesis catalysts with high affinity to silica gel. Tetrahedron 2013, 69, 7408–7415. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Shao, M.; Zhang, H.; Li, Y.; Liu, D.; Cheng, Y.; Liu, G.; Wang, J. Synthesis and reactivity of oxygen chelated ruthenium carbene metathesis catalysts. J. Organomet. Chem. 2014, 756, 1–9. [Google Scholar] [CrossRef]
  40. Jatmika, C.; Goshima, K.; Wakabayashi, K.; Akiyama, N.; Hirota, S.; Matsuo, T. Second-coordination sphere effects on the reactivities of Hoveyda–Grubbs-type catalysts: A ligand exchange study using phenolic moiety-functionalized ligands. Dalton Trans. 2020, 49, 11618–11627. [Google Scholar] [CrossRef]
  41. Al-Enezi, M.Y.; John, E.; Ibrahim, Y.A.; Al-Awadi, N.A. Highly efficient Ru(II)-alkylidene based Hoveyda–Grubbs catalysts for ring-closing metathesis reactions. RSC Adv. 2021, 11, 37866–37876. [Google Scholar] [CrossRef]
  42. Thurier, C.; Fischmeister, C.; Bruneau, C.; Olivier-Bourbigou, H.; Dixneuf, P.H. Ionic imidazolium containing ruthenium complexes and olefin metathesis in ionic liquids. J. Mol. Catal. A Chem. 2007, 268, 127–133. [Google Scholar] [CrossRef]
  43. Varray, S.; Lazaro, R.; Martinez, J.; Lamaty, F. New Soluble-Polymer Bound Ruthenium Carbene Catalysts: Synthesis, Characterization, and Application to Ring-Closing Metathesis. Organometallics 2003, 22, 2426–2435. [Google Scholar] [CrossRef]
  44. Consorti, C.S.; Aydos, G.L.P.; Ebeling, G.; Dupont, J. On the Immobilization of Ruthenium Metathesis Catalysts in Imidazolium Ionic Liquids. Organometallics 2009, 28, 4527–4533. [Google Scholar] [CrossRef]
  45. Lee, S.; Shin, J.Y.; Lee, S.-G. Imidazolium-Salt-Functionalized Ionic-CNT-Supported Ru Carbene/Palladium Nanoparticles for Recyclable Tandem Metathesis/Hydrogenation Reactions in Ionic Liquids. Chem. Asian J. 2013, 8, 1990–1993. [Google Scholar] [CrossRef] [PubMed]
  46. Michrowska, A.; Grela, K. Quest for the ideal olefin metathesis catalyst. Pure Appl. Chem. 2008, 80, 31–43. [Google Scholar] [CrossRef]
  47. Gladysz, J.A. Recoverable catalysts. Ultimate goals, criteria of evaluation, and the green chemistry interface. Pure Appl. Chem. 2001, 73, 1319–1324. [Google Scholar] [CrossRef] [Green Version]
  48. Barbasiewicz, M.; Bieniek, M.; Michrowska, A.; Szadkowska, A.; Makal, A.; Woźniak, K.; Grela, K. Probing of the Ligand Anatomy: Effects of the Chelating Alkoxy Ligand Modifications on the Structure and Catalytic Activity of Ruthenium Carbene Complexes. Adv. Synth. Catal. 2007, 349, 193–203. [Google Scholar] [CrossRef]
  49. Grela, K.; Harutyunyan, S.; Michrowska, A. A Highly Efficient Ruthenium Catalyst for Metathesis Reactions. Angew. Chem. Int. Ed. 2002, 41, 4038–4040. [Google Scholar] [CrossRef] [Green Version]
  50. Aversa, A.; Pili, M.; Francomano, D.; Bruzziches, R.; Spera, E.; La Pera, G.; Spera, G. Effects of vardenafil administration on intravaginal ejaculatory latency time in men with lifelong premature ejaculation. Int. J. Impot. Res. 2009, 21, 221–227. [Google Scholar] [CrossRef] [Green Version]
  51. Nienałtowski, T.; Szczepanik, P.; Małecki, P.; Czajkowska-Szczykowska, D.; Czarnocki, S.; Pawłowska, J.; Kajetanowicz, A.; Grela, K. Large-Scale Synthesis of a Niche Olefin Metathesis Catalyst Bearing an Unsymmetrical N-Heterocyclic Carbene (NHC) Ligand and its Application in a Green Pharmaceutical Context. Chem. Eur. J. 2020, 26, 15708–15717. [Google Scholar] [CrossRef]
  52. Khatuya, H. On the bromination of methyl 2-methyl-3-furoate. Tetrahedron Lett. 2001, 42, 2643–2644. [Google Scholar] [CrossRef]
  53. Szczepaniak, G.; Urbaniak, K.; Wierzbicka, C.; Kosiński, K.; Skowerski, K.; Grela, K. High-Performance Isocyanide Scavengers for Use in Low-Waste Purification of Olefin Metathesis Products. ChemSusChem 2015, 8, 4139–4148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. APEXII-2008v1.0 Bruker Nonius 2007.
  55. SAINT V7.34A Bruker Nonius 2007.
  56. SADABS-2004/1 Bruker Nonius area detector scaling and absorption correction, 2007.
  57. Sheldrick, G. Phase annealing in SHELX-90: Direct methods for larger structures. Acta Crystallogr. A 1990, A46, 467–473. [Google Scholar] [CrossRef]
  58. Sheldrick, G.M. SHELXL93. Program for the Refinement of Crystal Structures; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  59. Wilson, A.J.C. (Ed.) International Tables for Crystallography; Kluwer: Dordrecht, The Netherlands, 1992. [Google Scholar]
Figure 1. Selected (a) commercially available ruthenium-based general-purpose olefin metathesis catalysts and the structures of NHC, uNHC, and CAAC, and (b) selected ether-modified chelating benzylidene complexes.
Figure 1. Selected (a) commercially available ruthenium-based general-purpose olefin metathesis catalysts and the structures of NHC, uNHC, and CAAC, and (b) selected ether-modified chelating benzylidene complexes.
Chemistry 04 00056 g001
Figure 2. Combined structural characteristics leading to development of a new system. * NHC with Dipp substituent instead of Mes. (For Ru14 and Ru15, see [48]; for Ru10 and Ru11, see [33,34,35]; for Ru12, see [38]).
Figure 2. Combined structural characteristics leading to development of a new system. * NHC with Dipp substituent instead of Mes. (For Ru14 and Ru15, see [48]; for Ru10 and Ru11, see [33,34,35]; for Ru12, see [38]).
Chemistry 04 00056 g002
Scheme 1. Two-step synthesis of ligand precursor 4.
Scheme 1. Two-step synthesis of ligand precursor 4.
Chemistry 04 00056 sch001
Scheme 2. Synthesis of complexes Ru16 and Ru17.
Scheme 2. Synthesis of complexes Ru16 and Ru17.
Chemistry 04 00056 sch002
Figure 3. ORTEP diagram (50% probability ellipsoids) of complexes Ru16. Hydrogen atom omitted for clarity.
Figure 3. ORTEP diagram (50% probability ellipsoids) of complexes Ru16. Hydrogen atom omitted for clarity.
Chemistry 04 00056 g003
Figure 4. Front view (a) and side view (b) of molecular overlay for Hoveyda–Grubbs (violet), Ru10′ (light green) and Ru16 (orange) catalysts. Molecules represented by sticks, and hydrogen atom omitted for clarity.
Figure 4. Front view (a) and side view (b) of molecular overlay for Hoveyda–Grubbs (violet), Ru10′ (light green) and Ru16 (orange) catalysts. Molecules represented by sticks, and hydrogen atom omitted for clarity.
Chemistry 04 00056 g004
Figure 5. Relative conversion rates for a model RCM reaction of 5 using 1 mol% of the catalyst.
Figure 5. Relative conversion rates for a model RCM reaction of 5 using 1 mol% of the catalyst.
Chemistry 04 00056 g005
Scheme 3. Preparation of Vardenafil analogue 18 in RCM reaction catalyzed with Ru17.
Scheme 3. Preparation of Vardenafil analogue 18 in RCM reaction catalyzed with Ru17.
Chemistry 04 00056 sch003
Table 1. Catalytical activity of Ru16 and Ru17 in comparison with Ru10.
Table 1. Catalytical activity of Ru16 and Ru17 in comparison with Ru10.
EntrySubstrateProductCatalyst
(mol%)
Temp (°C)Time (min)Yield (%) a
1 Chemistry 04 00056 i001 Chemistry 04 00056 i002Ru10 (1.0)
Ru16 (1.0)
Ru17 (1.0)
Ru10 (0.2)
Ru16 (0.2)
Ru17 (0.2)
Ru10 (0.5)
23
23
23
23
23
23
23
90
10
30
120
90
90
120
(92)
(97)
(100)
(40)
(59)
(60)
(49)
2 Chemistry 04 00056 i003 Chemistry 04 00056 i004Ru10 (0.5)
Ru16 (0.2)
Ru17 (0.2)
23
23
23
120
120
120
49
85
96
3 Chemistry 04 00056 i005 Chemistry 04 00056 i006Ru10 (0.2)
Ru16 (0.2)
Ru17 (0.2)
Ru10 (1.0)
Ru16 (1.0)
Ru17 (1.0)
23
23
23
23
23
23
120
15
120
90
10
90
(100)
(92)
(98)
(100)
(100)
(100)
4 b Chemistry 04 00056 i007 Chemistry 04 00056 i008Ru10 (1.0)
Ru16 (1.0)
Ru17 (1.0)
30
30
30
60
60
60
70
91
79
5 c Chemistry 04 00056 i009 Chemistry 04 00056 i010Ru10 (1.0)
Ru16 (1.0)
Ru17 (1.0)
23
23
23
120
120
120
93
97
98
Conditions: a Isolated yields after silica gel chromatography. In parentheses are yields determined by GC. b Reaction with two equivalents of cis-1,4-diacetoxy-2-butene. c Reaction with two equivalents of methyl acrylate.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nadirova, M.; Zieliński, A.; Malinska, M.; Kajetanowicz, A. Fast Initiating Furan-Containing Hoveyda-Type Complexes: Synthesis and Applications in Metathesis Reactions. Chemistry 2022, 4, 786-795. https://doi.org/10.3390/chemistry4030056

AMA Style

Nadirova M, Zieliński A, Malinska M, Kajetanowicz A. Fast Initiating Furan-Containing Hoveyda-Type Complexes: Synthesis and Applications in Metathesis Reactions. Chemistry. 2022; 4(3):786-795. https://doi.org/10.3390/chemistry4030056

Chicago/Turabian Style

Nadirova, Maryana, Adam Zieliński, Maura Malinska, and Anna Kajetanowicz. 2022. "Fast Initiating Furan-Containing Hoveyda-Type Complexes: Synthesis and Applications in Metathesis Reactions" Chemistry 4, no. 3: 786-795. https://doi.org/10.3390/chemistry4030056

Article Metrics

Back to TopTop