Next Article in Journal
In Situ Removal of Benzene as a Biomass Tar Model Compound Employing Hematite Oxygen Carrier
Next Article in Special Issue
Comparison of Catalytic Properties of the Easily Interconvertible, Water-Soluble [RuHCl(CO)(mtppms-Na)3] and [RuH(H2O)(CO)(mtppms-Na)3][BF4]
Previous Article in Journal
Origin of the Unexpected Enantioselectivity in the Enzymatic Reductions of 5-Membered-Ring Heterocyclic Ketones Catalyzed by Candida parapsilosis Carbonyl Reductases
Previous Article in Special Issue
Pd-Catalyzed Hirao P–C Coupling Reactions with Dihalogenobenzenes without the Usual P-Ligands under MW Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Supported Polymeric Ruthenium Complexes as Olefin Metathesis Catalysts in Synthesis of Heterocyclic Compounds

Biological and Chemical Research Centre, Faculty of Chemistry, University of Warsaw, Żwirki i Wigury 101, 02-089 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1087; https://doi.org/10.3390/catal12101087
Submission received: 30 August 2022 / Revised: 16 September 2022 / Accepted: 19 September 2022 / Published: 21 September 2022
(This article belongs to the Special Issue Catalysis in Heterocyclic and Organometallic Synthesis II)

Abstract

:
New ruthenium olefin metathesis catalysts containing N-heterocyclic carbene (NHC) connected by a linker tether to a benzylidene ligand were studied. Such obtained self-chelated Hoveyda–Grubbs type complexes existed in the form of an organometallic polymer but could still catalyze olefin metathesis after being dissolved in an organic solvent. Although these polymeric catalysts exhibited a slightly lower activity compared to structurally related nonpolymeric catalysts, they were successfully used in a number of ring-closing metathesis reactions leading to a variety of heterocyclic compounds, including biologically and pharmacologically related analogues of cathepsin K inhibitor and sildenafil (Viagra™). In the last case, a good solubility of a polymeric catalyst in toluene allowed the separation of the product from the catalyst via simple filtration.

1. Introduction

Unquestioned is the fact that ruthenium olefin metathesis precatalysts initiation requires first the dissociation of a neutral ligand, which after a single turnover leads to the release of 14e propagating species [1,2]. An illustrious example of this process is the initiation step in Hoveyda–Grubbs catalysts, which concerns the dissociation of the 2-(isopropoxy)benzylidene ligand and the release of 2-(isopropoxy)styrene moiety [3,4]. In 1999, it was shown that the release of this styrene was reversible, meaning that the created 14e species could reuptake the styrene back, leading to the re-creation of the former complex [5]. This process was called the release/reuptake (or “boomerang”) mechanism (Figure 1a) and since then, the effect has been carefully studied [6,7,8,9,10,11]. However, the results of experiments brought no consensus about the existence and importance of the effect. The studies from 2008 depicted an OM reaction in the presence of deuterium-labeled styrene leading to the reisolation of a Hoveyda–Grubbs type precatalyst containing the labeled benzylidene ligand with decent yield [7]. In 2014, Fogg carried out complementary experiments using a 13C-labeled Hoveyda–Grubbs catalyst and the results provided evidence of the existence of the boomerang effect [10]. On the other hand, the investigation of this effect in low-loading reaction conditions revealed the existence of the boomerang mechanism questionable. The utilization of a fluorophore-tagged Hoveyda–Grubbs complex by Plenio did not show evidence of the release–reuptake mechanism; however, it should be noticed that the complex he used belonged to the fast-activating EWG-substituted Hoveyda–Grubbs class of complexes [8]. In 2020, we described an application of the ammonium-group-tagged Hoveyda–Grubbs type catalyst immobilized on metal–organic frameworks (MOFs) [11] The structure contained ammonium tags on both the NHC and benzylidene parts. To show the possibility of the boomerang effect, the MOF support was doped with an ammonium-tagged styrene derivative. The results showed an ambiguous behavior of this system in terms of a better stability at the cost of a decrease of reactivity in comparison to a nondoped MOF support.
In 2008, Sang-gi described a polymeric ruthenium structure that contained a covalent linker between NHC ligand and alkoxybenzylidene moiety based on an additional ether function in the aromatic part (Figure 1b) [12]. The authors claimed that during olefin metathesis with complex C, they obtained the polymeric, self-supported structure D. Considering the previous reports, we came up with an idea to investigate the boomerang mechanism based on the scaffold of a precatalyst containing the covalent bond between ligand L and alkoxybenzylidene (Figure 1c). Based on Sang-gi’s results, we expected that this kind of structure upon reuptake process could lead to the polymeric complex (intermolecular complex G); however, the creation of a macrocyclic ruthenium complex (intermolecular complex E) was also possible, as described by Golder and co-workers [13,14].
Inspired by these results, we proposed the structure containing an aliphatic linker between the benzyl moiety on the NHC part and the benzylidene ligand via a ruthenium-chelating oxygen atom (Figure 2). The choice of this kind of structure endeavored to reconcile the simplicity of the ligand synthesis and its activity [15,16,17,18,19]. We assumed that the length of six to eight carbons of linker could have been sufficient to furnish the macrocyclic structures; however, we did not reject the possibility of the creation of polymeric species.

2. Results

The synthesis started from O–alkylation of 2–(prop–1–enyl)phenol (1) with corresponding terminal bromoalcohols (C6 and C8) providing the desired products 2a and 2b with moderate yields. The transformation of hydroxyl to tosyloxy function followed by another O–alkylation with 2-salicylaldehyde led to the formation of ethers 4a and 4b with good yields. The next steps provided an imidazoline core and utilized the hydroamination reaction with N-mesitylethylenediamine, furnishing the diamine products with good yields, followed by condensation with triethylorthoformate to provide the final ligand precursors 6a and 6b as imidazolinium tetrafluoroborates (Scheme 1). Having both ligand precursors in hand, we obtained the Grubbs type second-generation complexes (Scheme 2). The deprotonation of the corresponding imidazolinium salts with tBuOK followed by the addition of the Grubbs first-generation precatalyst delivered the desired complexes Ru-2C6 and Ru-2C8 with acceptable yields (for experimental details, see Supplementary Materials).
To create the proposed structure of ruthenium complexes, we decided to utilize common methods such as the addition of CuCl or submission under OM conditions; nevertheless, none of this led to the obtention of macrocyclic nor polymeric ruthenium species. Thus, we decided to use HCl, which we supposed could easily protonate the phosphine ligand and force the exchange of the benzylidene moieties. As we suspected, this approach converted the second generation precatalyst to produce the self-supported polymeric species Ru-3C6 and Ru-3C8 without any traces of macrocyclic ruthenium structures.
The polymeric Ru compounds exhibited a good solubility in many organic solvents such as toluene, chlorinated solvents, and ethyl acetate; however, it could easily precipitate from n-hexane or n-pentane. The nuclear magnetic resonance analysis of the obtained complexes showed a shifting of the benzylidene signal from 18.90 ppm for Ru-2C6 to a broad set of signals at 16.21 ppm for Ru-3C6, which corresponds to O-chelated Hoveyda–Grubbs type second-generation complexes. The fact that this signal was broad and was composed of several singlet signals could be interpreted as undefined polymeric species.
We compared the activity of the obtained polymeric ruthenium complexes with that of the closest structural analog, Ru-4. For further research, we chose the complex Ru-3C6. We performed the RCM reaction of diethyl diallylmalonate (7a) using 1 mol% of pre-catalyst (Figure 3) and calculated the turnover frequency (TOF) parameter to quantify and compare the activity of the complexes (Table 1).
The polymeric species exhibited a very low activity compared to the Hoveyda–Grubbs type complex Ru-4, which performed almost a quantitative conversion within 2 h at 40 °C and in 15 min at 60 °C. The TOFs calculated for 15 min of reactions at 40 and 60 °C were an order of magnitude lower for the polymeric Ru-3C6 compared to Ru-4. The elevation of the temperature to 80 °C significantly increased the activity of Ru-3C6 as these conditions allowed to obtain a quantitative conversion in 2 h, which was comparable with that of Ru-4 at 40 °C. Moreover, the TOF measured for Ru-3C6 at 80 °C was comparable to the one measured for Ru-4 at 40 °C (2.47 · 10−1 s−1 vs. 1.03 · 10−1 s−1, respectively) and very similar to Ru-4 at 80 °C (1.03 · 10−1 s−1. This phenomenon can be explained by a very low rate of depolymerization process of Ru-3C6 to metathetically active species at lower temperatures (for a process scheme: see Figure 1c, reaction G to F). It places the polymeric ruthenium species in a family of high-temperature activating ruthenium precatalysts.
As the properties of the ruthenium complex allowed to precipitate it from the reaction mixture, we decided to investigate the possibility of its recycling. To do so, we performed a standard RCM reaction of 7a at 80 °C for 2 h followed by a precipitation of polymeric species with n-pentane. The first run provided almost full conversion, but the amount of recycled catalyst showed a significant loss of polymeric material as only 38% of the initial amount was recovered (Table 2, entry 1). In the next run, the conversion slightly decreased with a moderate improvement on Ru recycling (Table 2, entry 2). The third run showed a significant decrease in the activity (only 57% of the conversion) of the polymeric complex, therefore we did not perform further recycling runs (Table 2, entry 3). The results indicated that these conditions (including the polymeric nature of the complex) did not promote or support the boomerang effect. The first indication was that the recovery of Ru species was not quantitative, thus it was not possible to assert whether metathetically active monomeric species turned back into polymeric form (and also how much polymeric Ru formed a monomeric complex). Second, it seemed that the catalyst quickly decomposed since a low activity was noticed already in third run; however, it was still comparable with the recycling results obtained by Sang-gi [7,20].
Despite the unsuccessful recovery of the polymeric Ru complex, we evaluated the synthetic utility of this catalyst in the ring-closing metathesis providing heterocyclic compounds (Scheme 3). We performed the reaction with a number of aromatic amides and sulfonamides to produce 2,5-dihydropyrrolyl-containing products. N–tosyl dihydropyrrole (8b) was obtained with an excellent yield and short reaction time (2 h). Moreover, benzoyl-substituted pyrroles containing functional groups such as fluorine (8c), bromine (8d), and dimethylamino substituent (8e) could be easily produced in moderate to good yields. More importantly, the polymeric complex could be utilized in the synthesis of biologically relevant heterocyclic structures. The seven-membered ring cathepsin K inhibitor analogue 8f [22] was obtained with a very good yield and the sildenafil analogue 8g was also synthesized in acceptable yield [23,24] (a higher catalyst loading was used—5 mol%). Moreover, since product 8g was not soluble in toluene (in contrary to polymeric ruthenium complex Ru-3C6), we exploited this condition to purify the compound via simple filtration, to obtain an analytically pure product without the utilization of column chromatography.

3. Conclusions

This paper provided a design and synthesis of a new ligand containing a linker tethering the NHC precursor part with the benzylidene moiety. Applying the ligand on first-generation Grubbs type catalyst successfully provided a second-generation Grubbs type complex, which could be further transformed to a polymeric self-supported second-generation Hoveyda–Grubbs type. The polymeric species exhibited a lower activity compared to structurally similar nonpolymeric precatalysts; however, it could be successfully used in metathetical transformations such as ring-closing metathesis, providing a variety of heterocyclic compounds, also biologically and pharmacologically related, in which case the solubility properties of polymer allowed to ease the purification of products. Since these examples did not confirm a “boomerang” effect, further studies on the impact of a linker architecture and properties are under way.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12101087/s1, Table S1. Gas chromatography data. Table S2. Setup of recycling reactions with reagent amounts. Table S3. Gas chromatography data [25,26,27,28,29,30,31,32,33].

Author Contributions

Conceptualization, A.K. and K.G.; funding acquisition, K.G.; investigation, A.A.R.; supervision, A.K. and K.G.; visualization, A.A.R.; writing—original draft, A.A.R. and K.G.; writing—review & editing, A.A.R., A.K., and K.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Centre, Poland grant number DEC-2019/34/A/ST4/00372.

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 cofinanced by the European Union from the European Regional Development Fund under the Operational Programme Innovative Economy, 2007–2013. Authors want to thank Ewelina Wielgus from the Molecular and Macromolecular Research Centre of Polish Academy of Science (Łódź) for the MALDI-TOF analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grela, K. Olefin Metathesis: Theory and Practice; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2014. [Google Scholar]
  2. Grubbs, R.H.W.; Wenzel, A.G.; O’Leary, D.J.; Khosravi, E. Handbook of Metathesis; Wiley-VCH: Weinheim, Germany, 2015. [Google Scholar]
  3. Kingsbury, J.S.; Hoveyda, A.H. Regarding the Mechanism of Olefin Metathesis with Sol−Gel-Supported Ru-Based Complexes Bearing a Bidentate Carbene Ligand. Spectroscopic Evidence for Return of the Propagating Ru Carbene. J. Am. Chem. Soc. 2005, 127, 4510–4517. [Google Scholar] [CrossRef] [PubMed]
  4. Ashworth, I.W.; Hillier, I.H.; Nelson, D.J.; Percy, J.M.; Vincent, M.A. What is the initiation step of the Grubbs-Hoveyda olefin metathesis catalyst? Chem. Commun. 2011, 47, 5428–5430. [Google Scholar] [CrossRef] [PubMed]
  5. 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]
  6. Ahmed, M.; Barrett, A.G.M.; Braddock, D.C.; Cramp, S.M.; Procopiou, P.A. A recyclable ‘boomerang’ polymer-supported ruthenium catalyst for olefin metathesis. Tetrahedron Lett. 1999, 40, 8657–8662. [Google Scholar] [CrossRef]
  7. Bieniek, M.; Michrowska, A.; Usanov, D.L.; Grela, K. In an Attempt to Provide a User’s Guide to the Galaxy of Benzylidene, Alkoxybenzylidene, and Indenylidene Ruthenium Olefin Metathesis Catalysts. Chem. Eur. J. 2008, 14, 806–818. [Google Scholar] [CrossRef] [PubMed]
  8. Vorfalt, T.; Wannowius, K.J.; Thiel, V.; Plenio, H. How Important Is the Release–Return Mechanism in Olefin Metathesis? Chem. Eur. J. 2010, 16, 12312–12315. [Google Scholar] [CrossRef] [PubMed]
  9. Núñez-Zarur, F.; Solans-Monfort, X.; Pleixats, R.; Rodríguez-Santiago, L.; Sodupe, M. DFT Study on the Recovery of Hoveyda–Grubbs-Type Catalyst Precursors in Enyne and Diene Ring-Closing Metathesis. Chem. Eur. J. 2013, 19, 14553–14565. [Google Scholar] [CrossRef]
  10. Bates, J.M.; Lummiss, J.A.M.; Bailey, G.A.; Fogg, D.E. Operation of the Boomerang Mechanism in Olefin Metathesis Reactions Promoted by the Second-Generation Hoveyda Catalyst. ACS Catal. 2014, 4, 2387–2394. [Google Scholar] [CrossRef]
  11. 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]
  12. Chen, S.-W.; Kim, J.H.; Shin, H.; Lee, S.-g. A new type of self-supported, polymeric Ru-carbene complex for homogeneous catalysis and heterogeneous recovery: Synthesis and catalytic activities for ring-closing metathesis. Org. Biomol. Chem. 2008, 6, 2676–2678. [Google Scholar] [CrossRef]
  13. Wang, T.-W.; Huang, P.-R.; Chow, J.L.; Kaminsky, W.; Golder, M.R. A Cyclic Ruthenium Benzylidene Initiator Platform Enhances Reactivity for Ring-Expansion Metathesis Polymerization. J. Am. Chem. Soc. 2021, 143, 7314–7319. [Google Scholar] [CrossRef] [PubMed]
  14. Morrison, C.M.; Golder, M.R. Ring-Expansion Metathesis Polymerization Initiator Design for the Synthesis of Cyclic Polymers. Synlett 2022, 33, 699–704. [Google Scholar] [CrossRef]
  15. Ablialimov, O.; Kędziorek, M.; Torborg, C.; Malińska, M.; Woźniak, K.; Grela, K. New Ruthenium(II) Indenylidene Complexes Bearing Unsymmetrical N-Heterocyclic Carbenes. Organometallics 2012, 31, 7316–7319. [Google Scholar] [CrossRef]
  16. Ablialimov, O.; Kędziorek, M.; Malińska, M.; Woźniak, K.; Grela, K. Synthesis, Structure, and Catalytic Activity of New Ruthenium(II) Indenylidene Complexes Bearing Unsymmetrical N-Heterocyclic Carbenes. Organometallics 2014, 33, 2160–2171. [Google Scholar] [CrossRef]
  17. Smoleń, M.; Kośnik, W.; Loska, R.; Gajda, R.; Malińska, M.; Woźniak, K.; Grela, K. Synthesis and catalytic activity of ruthenium indenylidene complexes bearing unsymmetrical NHC containing a heteroaromatic moiety. RSC Adv. 2016, 6, 77013–77019. [Google Scholar] [CrossRef]
  18. Małecki, P.; Gajda, K.; Ablialimov, O.; Malińska, M.; Gajda, R.; Woźniak, K.; Kajetanowicz, A.; Grela, K. Hoveyda–Grubbs-Type Precatalysts with Unsymmetrical N-Heterocyclic Carbenes as Effective Catalysts in Olefin Metathesis. Organometallics 2017, 36, 2153–2166. [Google Scholar] [CrossRef]
  19. 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]
  20. Chen, S.-W.; Kim, J.H.; Song, C.E.; Lee, S.-g. Self-Supported Oligomeric Grubbs/Hoveyda-Type Ru−Carbene Complexes for Ring-Closing Metathesis. Org. Lett. 2007, 9, 3845–3848. [Google Scholar] [CrossRef]
  21. Mmongoyo, J.A.; Mgani, Q.A.; Mdachi, S.J.M.; Pogorzelec, P.J.; Cole-Hamilton, D.J. Synthesis of a kairomone and other chemicals from cardanol, a renewable resource. Eur. J. Lipid Sci. Technol. 2012, 114, 1183–1192. [Google Scholar] [CrossRef]
  22. Yamashita, D.S.; Marquis, R.W.; Xie, R.; Nidamarthy, S.D.; Oh, H.-J.; Jeong, J.U.; Erhard, K.F.; Ward, K.W.; Roethke, T.J.; Smith, B.R.; et al. Structure Activity Relationships of 5-, 6-, and 7-Methyl-Substituted Azepan-3-one Cathepsin K Inhibitors. J. Med. Chem. 2006, 49, 1597–1612. [Google Scholar] [CrossRef]
  23. Reichenberger, F.; Voswinckel, R.; Steveling, E.; Enke, B.; Kreckel, A.; Olschewski, H.; Grimminger, F.; Seeger, W.; Ghofrani, H.A. Sildenafil treatment for portopulmonary hypertension. Eur. Respir. J. 2006, 28, 563–567. [Google Scholar] [CrossRef] [PubMed]
  24. Goldstein, I.; Burnett, A.L.; Rosen, R.C.; Park, P.W.; Stecher, V.J. The Serendipitous Story of Sildenafil: An Unexpected Oral Therapy for Erectile Dysfunction. Sex. Med. Rev. 2019, 7, 115–128. [Google Scholar] [CrossRef] [PubMed]
  25. BouzBouz, S.; Boulard, L.; Cossy, J. Ruthenium-Catalyzed Cross-Metathesis between Diallylsilanes and Electron-Deficient Olefins. Org. Lett. 2007, 9, 3765–3768. [Google Scholar] [CrossRef] [PubMed]
  26. So, C.M.; Kume, S.; Hayashi, T. Rhodium-Catalyzed Asymmetric Hydroarylation of 3-Pyrrolines Giving 3-Arylpyrrolidines: Protonation as a Key Step. J. Am. Chem. Soc. 2013, 135, 10990–10993. [Google Scholar] [CrossRef]
  27. César, V.; Zhang, Y.; Kośnik, W.; Zieliński, A.; Rajkiewicz, A.A.; Ruamps, M.; Bastin, S.; Lugan, N.; Lavigne, G.; Grela, K. Ruthenium Catalysts Supported by Amino-Substituted N-Heterocyclic Carbene Ligands for Olefin Metathesis of Challenging Substrates. Chem. Eur. J. 2017, 23, 1950–1955. [Google Scholar] [CrossRef]
  28. 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]
  29. Rajkiewicz, A.A.; Skowerski, K.; Trzaskowski, B.; Kajetanowicz, A.; Grela, K. 2-Methyltetrahydrofuran as a Solvent of Choice for Spontaneous Metathesis/Isomerization Sequence. ACS Omega 2019, 4, 1831–1837. [Google Scholar] [CrossRef]
  30. Yamashita, D.S.; Marquis, R.W.; Xie, R.; Nidamarthy, S.D.; Oh, H.-J.; Jeong, J.U.; Erhard, K.F.; Ward, K.W.; Roethke, T.J.; Smith, B.R.; et al. Structure Activity Relationships of 5-, 6-, and 7-Methyl-Substituted Azepan-3-one Cathepsin K Inhibitors. J. Med. Chem. 2006, 49, 1597–1612. [Google Scholar] [CrossRef]
  31. Monsigny, L.; Piątkowski, J.; Trzybiński, D.; Woźniak, K.; Nienałtowski, T.; Kajetanowicz, A.; Grela, K. Activated Hoveyda-Grubbs Olefin Metathesis Catalysts Derived from a Large Scale Produced Pharmaceutical Intermediate–Sildenafil Aldehyde. Adv. Synth. Catal. 2021, 363, 4590–4604. [Google Scholar] [CrossRef]
  32. 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]
  33. Chołuj, A.; Krzesiński, P.; Ruszczyńska, A.; Bulska, E.; Kajetanowicz, A.; Grela, K. Noncovalent Immobilization of Cationic Ruthenium Complex in a Metal–Organic Framework by Ion Exchange Leading to a Heterogeneous Olefin Metathesis Catalyst for Use in Green Solvents. Organometallics 2019, 38, 3397–3405. [Google Scholar] [CrossRef]
Figure 1. Release/reuptake in olefin metathesis catalysts. (a) The schematic representation of the boomerang mechanism. (b) Polymeric self-supported ruthenium complex described by Sang-gi. (c) Idea of the boomerang effect in Ru complexes containing a covalent linker between NHC and benzylidene moieties.
Figure 1. Release/reuptake in olefin metathesis catalysts. (a) The schematic representation of the boomerang mechanism. (b) Polymeric self-supported ruthenium complex described by Sang-gi. (c) Idea of the boomerang effect in Ru complexes containing a covalent linker between NHC and benzylidene moieties.
Catalysts 12 01087 g001
Figure 2. Proposed structures of macrocyclic ruthenium complexes.
Figure 2. Proposed structures of macrocyclic ruthenium complexes.
Catalysts 12 01087 g002
Scheme 1. Synthetic pathways for the NHC ligand precursors. Conditions: (a) Br-(CH2)6-OH or Br(CH2)8-OH, K2CO3, DMF, 80 °C, 16 h; (b) TsCl, Et3N, CH2Cl2, 0 °C to RT, 16 h; (c) salicylaldehyde, K2CO3, DMF, 60 °C, 16 h; (d) N-mesitylethylenediamine, Na2SO4, HCOOH (cat.), MeOH, RT, 24 h then NaBH4, MeOH, RT, 24 h; (e) NH4BF4, TEOF, 105 °C, 18 h. RT—room temperature, TOEF—triethylorthoformate.
Scheme 1. Synthetic pathways for the NHC ligand precursors. Conditions: (a) Br-(CH2)6-OH or Br(CH2)8-OH, K2CO3, DMF, 80 °C, 16 h; (b) TsCl, Et3N, CH2Cl2, 0 °C to RT, 16 h; (c) salicylaldehyde, K2CO3, DMF, 60 °C, 16 h; (d) N-mesitylethylenediamine, Na2SO4, HCOOH (cat.), MeOH, RT, 24 h then NaBH4, MeOH, RT, 24 h; (e) NH4BF4, TEOF, 105 °C, 18 h. RT—room temperature, TOEF—triethylorthoformate.
Catalysts 12 01087 sch001
Scheme 2. Synthesis of 2nd-generation Grubbs type precatalysts and polymeric species Ru-3C6 and Ru-3C8. Gru-I—benzylidene-bis(tricyclohexylphosphine)dichlororuthenium, RT—room temperature.
Scheme 2. Synthesis of 2nd-generation Grubbs type precatalysts and polymeric species Ru-3C6 and Ru-3C8. Gru-I—benzylidene-bis(tricyclohexylphosphine)dichlororuthenium, RT—room temperature.
Catalysts 12 01087 sch002
Figure 3. Reaction course time plot of RCM of diethyl diallylmalonate (7a). Black marks: Ru-4, red marks: Ru-3C6. Circles: reaction performed at 40 °C; triangles: reaction performed at 60 °C; squares: reaction performed at 80 °C. The inset shows the first 30 min period in detail. Lines are for visual aid only. Conversion was calculated by GC analysis with internal standard (1,2,4,5-tetramethylbenzene).
Figure 3. Reaction course time plot of RCM of diethyl diallylmalonate (7a). Black marks: Ru-4, red marks: Ru-3C6. Circles: reaction performed at 40 °C; triangles: reaction performed at 60 °C; squares: reaction performed at 80 °C. The inset shows the first 30 min period in detail. Lines are for visual aid only. Conversion was calculated by GC analysis with internal standard (1,2,4,5-tetramethylbenzene).
Catalysts 12 01087 g003
Scheme 3. Scope and limitation studies on synthesis of heterocyclic compounds via ring-closing metathesis reaction. Standard condition: 2 mol% of Ru-3C6, toluene (c = 0.2 M), 80 °C, 18 h. The yields correspond to the isolated product. a 1 mol% of Ru-3C6, 2 h of reaction time. b 1 mol% of Ru-3C6. c 5 mol% of Ru-3C6.
Scheme 3. Scope and limitation studies on synthesis of heterocyclic compounds via ring-closing metathesis reaction. Standard condition: 2 mol% of Ru-3C6, toluene (c = 0.2 M), 80 °C, 18 h. The yields correspond to the isolated product. a 1 mol% of Ru-3C6, 2 h of reaction time. b 1 mol% of Ru-3C6. c 5 mol% of Ru-3C6.
Catalysts 12 01087 sch003
Table 1. Turnover frequency parameter calculated on the basis of reaction course time plot (Figure 3). For reactions performed at 40 and 60 °C, TOF is given after 15 min; for reactions performed at 80 °C, TOF is given after 5 min.
Table 1. Turnover frequency parameter calculated on the basis of reaction course time plot (Figure 3). For reactions performed at 40 and 60 °C, TOF is given after 15 min; for reactions performed at 80 °C, TOF is given after 5 min.
[Ru]TOF (T = 40 °C, t = 15 min) (s−1)TOF (T = 60 °C, t = 15 min) (s−1)TOF (T = 80 °C, t = 5 min) (s−1)
Ru-3C62.46 · 10−25.74 · 10−22.47 · 10−1
Ru-41.03 · 10−11.10 · 10−13.31 · 10−1
Table 2. Catalyst recycling in ring-closing metathesis of diethyl diallylmalonate (7a) using Ru-3C6 complex. Reactions were carried out using 1.0 mol% of Ru-3C6 at 80 °C in toluene (c = 0.2 M). a Conversion was calculated by GC analysis with internal standard (1,2,4,5-tetramethylbenzene). b Yield of Ru recycling was calculated as a percentage ratio of precatalyst mass used compared to next run.
Table 2. Catalyst recycling in ring-closing metathesis of diethyl diallylmalonate (7a) using Ru-3C6 complex. Reactions were carried out using 1.0 mol% of Ru-3C6 at 80 °C in toluene (c = 0.2 M). a Conversion was calculated by GC analysis with internal standard (1,2,4,5-tetramethylbenzene). b Yield of Ru recycling was calculated as a percentage ratio of precatalyst mass used compared to next run.
EntryScale (mmoL)Mass of Ru Used in Reaction [21]Conversion (%) aYield of Ru Recycling (%) b
15.0032.7>9938
21.9112.59566
31.278.357-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rajkiewicz, A.A.; Kajetanowicz, A.; Grela, K. Self-Supported Polymeric Ruthenium Complexes as Olefin Metathesis Catalysts in Synthesis of Heterocyclic Compounds. Catalysts 2022, 12, 1087. https://doi.org/10.3390/catal12101087

AMA Style

Rajkiewicz AA, Kajetanowicz A, Grela K. Self-Supported Polymeric Ruthenium Complexes as Olefin Metathesis Catalysts in Synthesis of Heterocyclic Compounds. Catalysts. 2022; 12(10):1087. https://doi.org/10.3390/catal12101087

Chicago/Turabian Style

Rajkiewicz, Adam A., Anna Kajetanowicz, and Karol Grela. 2022. "Self-Supported Polymeric Ruthenium Complexes as Olefin Metathesis Catalysts in Synthesis of Heterocyclic Compounds" Catalysts 12, no. 10: 1087. https://doi.org/10.3390/catal12101087

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