Next Article in Journal
Diels–Alder Cycloaddition to the Bay Region of Perylene and Its Derivatives as an Attractive Strategy for PAH Core Expansion: Theoretical and Practical Aspects
Next Article in Special Issue
Transformation of 3-(Furan-2-yl)-1,3-di(het)arylpropan-1-ones to Prop-2-en-1-ones via Oxidative Furan Dearomatization/2-Ene-1,4,7-triones Cyclization
Previous Article in Journal
An Overview on Graphene-Metal Oxide Semiconductor Nanocomposite: A Promising Platform for Visible Light Photocatalytic Activity for the Treatment of Various Pollutants in Aqueous Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of New Efficient Hoveyda–Grubbs Catalysts Comprising an N→Ru Coordinate Bond in a Six-Membered Ring for the Synthesis of Natural Product-Like Cyclopenta[b]furo[2,3-c]pyrroles

by
Alexandra S. Antonova
1,
Marina A. Vinokurova
1,
Pavel A. Kumandin
1,
Natalia L. Merkulova
1,
Anna A. Sinelshchikova
2,
Mikhail S. Grigoriev
2,
Roman A. Novikov
3,
Vladimir V. Kouznetsov
4,
Kirill B. Polyanskii
1,* and
Fedor I. Zubkov
1,*
1
Organic Chemistry Department, Faculty of Science, Peoples’ Friendship University of Russia (RUDN University), Miklukho-Maklaya St., 6, 117198 Moscow, Russia
2
A. N. Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, Leninsky pr. 31, bld. 4, 119071 Moscow, Russia
3
V. A. Engelhardt Institute of Molecular Biology, Russian Academy of Sciences, Vavilov Street, 32, 119991 Moscow, Russia
4
Laboratorio de Química Orgánica y Biomolecular, CMN, Universidad Industrial de Santander, Parque Tecnológico Guatiguara, Km 2 vía refugio, Piedecuesta A.A. 681011, Colombia
*
Authors to whom correspondence should be addressed.
Molecules 2020, 25(22), 5379; https://doi.org/10.3390/molecules25225379
Submission received: 3 November 2020 / Revised: 13 November 2020 / Accepted: 14 November 2020 / Published: 17 November 2020
(This article belongs to the Special Issue New Insights into Furans Transformations)

Abstract

:
The ring rearrangement metathesis (RRM) of a trans-cis diastereomer mixture of methyl 3-allyl-3a,6-epoxyisoindole-7-carboxylates derived from cheap, accessible and renewable furan-based precursors in the presence of a new class of Hoveyda–Grubbs-type catalysts, comprising an N→Ru coordinate bond in a six-membered ring, results in the difficult-to-obtain natural product-like cyclopenta[b]furo[2,3-c]pyrroles. In this process, only one diastereomer with a trans-arrangement of the 3-allyl fragment relative to the 3a,6-epoxy bridge enters into the rearrangement, while the cis-isomers polymerize almost completely under the same conditions. The tested catalysts are active in the temperature range from 60 to 120 °C at a concentration of 0.5 mol % and provide better yields of the target tricycles compared to the most popular commercially available second-generation Hoveyda–Grubbs catalyst. The diastereoselectivity of the intramolecular Diels–Alder reaction furan (IMDAF) reaction between starting 1-(furan-2-yl)but-3-en-1-amines and maleic anhydride, leading to 3a,6-epoxyisoindole-7-carboxylates, was studied as well.

Graphical Abstract

1. Introduction

The discovery of metathesis reactions has allowed the noticeable expansion of possibilities for the transformation of different unsaturated substrates, including those aimed at obtaining complex naturally occurring compounds and pharmacologically meaningful molecules, which is reflected in a number of recent reviews [1,2,3,4,5,6,7,8]. In particular, the ring-rearrangement metathesis (RRM) of bi- and polycyclic alkenes has made it possible to obtain systems that are practically inaccessible by means of other synthetic pathways [9,10,11,12,13,14,15,16,17,18]. Most often, for the rearrangement of bridged alkenes, the commercially available Grubbs (G) and Hoveyda–Grubbs (HG) ruthenium catalysts are used, which permits the reliable production of scaffolds possessing a wide range of potential biological activity due to the characteristics of their carbon skeletons [19,20,21,22,23]. Representative examples of the above-mentioned metathesis reactions clearly demonstrate a variety of possible transformation routes that give the structural diversity of the available final products (Scheme 1). As can be seen, the direction of the reaction may depend not only on the structure of the substrate, but also on the reaction conditions selected and the chosen catalyst.
Moreover, among all the variety of metathesis products known to date, there is a sole successful example of the HG-catalyzed RRM of 3a,6-epoxyisondole 1 which has produced the previously unknown heterocyclic system of cyclopenta[b]furo[2,3-c]pyrrole 2, but in modest yield (31%) [24] (Scheme 2).
Hence, further investigations of the influence of catalysts and the structure of unsaturated substrates on the selectivity of metathesis reactions remain an ongoing trend in modern organic synthesis. The selection of main starting materials for the suitable unsaturated substrates like 1 is also an important task. In this context, it should be noted that the starting material for 1 is furfural, which is a renewable chemical and one of the most available agricultural byproducts from biomass [25,26]. In regard to design of new catalysts for RRM, our group recently prepared a new class of effective ruthenium catalysts comprising an N→Ru coordinate bond in a six-membered ring (Cat.1 and Cat.2) (Figure 1) and found that they revealed a high level of activity (up to 10−2 mol%) in standard models for the olefin metathesis (styrene, allylbenzene, diethyl diallylmalonate, diallyltosylamide, norbornen/hex-1-ene, etc.) [27,28].
According to the statements above and with the knowledge that there are no reports on the efficient synthetic methods for diversely polyfunctionalized cyclopenta[b]furo[2,3-c]pyrroles using easily available 3-allyl-3a,6-epoxyisoindoles derived from 1-(furan-2-yl)-N-arylmethanimines through the intramolecular Diels–Alder furan (IMDAF) reaction, our research was focused on: (i) Establishing the optimal conditions for the diastereoselective preparation of 3-allyl-3a,6-epoxyisoindoles, according to the variables: solvent, acid catalysts, reaction time and temperature; (ii) With the determined conditions in hand, preparing diverse 3a,6-epoxyisondol-7-carboxylic acids and their methyl esters; (iii) Corroborating the spatial structure of isomeric acids and their esters; (iv) Studying the catalytic activity of ruthenium catalysts 1 and 2 in the RRM of 3a,6-epoxyisoindole-7-carboxylates in comparison with HG-II catalyst; and (v) With the established conditions in hand, preparing the desired cyclopenta[b]furo[2,3-c]pyrroles. All this is in order to develop new, short and efficient synthesis of natural product-like scaffolds from renewable furan-based precursors.

2. Results and Discussion

The principal synthetic scheme for the required 3-allyl-3a,6-epoxyisoindole-7-carboxylic acids 5 preparation includes two key steps and is based on readily available starting materials. The condensation of aryl amines (or benzyl amines) and furfural gives the corresponding Schiff bases 3, which, being treated with Grignard reagents (allyl magnesium bromide or methallyl magnesium chloride), give rise to homoallylamines 4 in multigram quantities (Scheme 3) [29,30,31,32].
The subsequent reaction of furfurylamines 4 with maleic anhydride easily provides the corresponding isoindolocarboxylic acids 5, which were isolated in good overall yields. The IMDAF reaction is highly stereoselective [33,34]; in the course of the tandem N-acylation/[4+2] cycloaddition sequence, only exo-adducts form as a mixture of two diastereomers (trans-5A and cis-5B) based on the orientation of the 3-allyl (or 3-methallyl) substituent relative to the 3a,6-epoxy-bridge (Figure 2).
As we showed earlier [29,30,31,32], the synthesized 3-allyl-3a,6-epoxyisoindole-7-carboxylic acids 5 are excellent precursors for heterocyclization into the corresponding isoindoloquinolines and isoindolobenzazepines. In this process, the isomeric composition of adducts 5A/5B had no strong influence on the yield and stereochemistry of the target heterocycles. In contrast, as was demonstrated in the first experiments in this study, only derivatives of the trans-isomer 5A are capable of further transformation under RRM conditions. Thus, in the second stage of this investigation, we tried to resolve the problem of the diastereoselectivity of the IMDAF reaction.
Considering the fact that reversible Diels–Alder reactions are most influenced by the reaction temperature and the polarity of the solvent [35,36,37,38], we compared the isomeric composition of the most diverse acids 5a,b,g,i,m formed under the conditions indicated in Table 1. The application of Lewis acids (10–25 mol % of AlEt3, BF3·OEt2, TiCl4) as catalysts in MeCN, PhH, PhMe, or CH2Cl2 turned out to be ineffective, probably due to their fast binding by a free carboxyl group of the products and due to the difficulties associated with the purification of the resulting reaction mixtures. Without catalyst addition, the pure target products precipitated from all tested solvents after the period of reaction time indicated in Table 1, wherein acetonitrile showed the worst results both at room and at elevated temperatures, apparently due to the partial dissolution of the target acids. As a result, it was found that according to 1H-NMR analysis, the cis-isomer B, inactive in the metathesis reaction, mainly forms at r.t in CH2Cl2 (average predominance 92%), while the desired trans-isomer A is mainly produced at −16 °C in the same solvent (average predominance 68%).
Decreasing the temperature to below −30 °C sharply slowed down the rate of interaction between maleic anhydride and secondary amines 4; however, at the same time, analysis of the data in Table 1 shows that the yields of the isomer 5A/5B mixtures obtained in CH2Cl2 are 1.5–2 times lower compared to the mixtures isolated from benzene or toluene. Besides, the last two solvents provide a slight predominance of the isomer 5A at both r.t and heating. It is also important to note that the reaction time was the shortest in boiling toluene (~1 h); nevertheless, for a reasonable comparison of the results in Table 1, all reactions were carried out in boiling solvents for 3 h. Thus, for further experiments, we synthesized mixtures of the trans-5A/cis-5B isomers either in dichloromethane at −16 °C or in PhMe at 3 h reflux, as indicated in Table 2.
The third part of the work is related to the study of the catalytic activity of ruthenium catalysts 1 and 2 (Figure 1) under the RRM protocol. According to the previous paper [27], the N,N-dimethyl catalyst 2 is the most thermally stable, while the morpholine derivative catalyst 1 is the most active (as it has the longest coordination N→Ru bond ~2.32 Å).
With this in mind, only two of these complexes were selected for the present study. The direct introduction of acids 5 into the RRM was obviously impossible due to their poor solubility in common organic solvents and due to the presence of a free carboxylic group in their structure, which is destructive to Hoveyda–Grubbs catalysts. Thus, methyl esters 6 of corresponding acids 5 were prepared according to the classical procedure (Scheme 4, Table 2).
As the entries a, h, k, n in Table 2 reveal, the trans/cis isomeric ratio of acids 5 does not change noticeably during the esterification process, i.e., retro-DA reaction or epimerization do not proceed under selected conditions. Esters 6 obtained in this way are crystalline, easily isolated substances that are readily soluble in dichloromethane.
Elucidation of the spatial structure of isomeric acids similar to 5A/5B and esters similar to 6A/6B has been previously done in the literature more than once [29,30,31,32,37,39,40,41,42], while as a rule, 2D NOE NMR data have been used for evidence of their structures. Here we present X-ray structural information for a pair of isomeric methyl esters 6eA/6eB, which confirms the preceding conclusions concerning the location of the substituent in the third position of the heterocyclic core (Figure 3).
Single crystals of diastereoisomers 6e were grown from an ether solution, and single-crystal XRD analysis revealed that both isomers comprised 3a,6-epoxyisondolo-7-carboxylates with one molecule in an asymmetric unit. Trans-6eA isomer crystallizes in the triclinic crystal system P-1 space group, while the cis-6eB isomer crystallizes in the orthorhombic crystal system P212121 space group. The configuration of the tricyclic system is identical for both isomers (see Tables S2–S6 for bond lengths and angles, and Figure S1 for the overlay of the two molecules). The substituents in the third position of the heterocyclic core are located in trans or cis positions that influence the distance between atoms, which are key for an NOE analysis (Table 3). The phenyl rings of 3a,6-epoxyisondolo-7-carboxylates are twisted at different angles with respect to the heterocycle due to the different orientations of the bulky substituent in the third position. In the case of cis-isomer 6eB, the angle between the C6 plane of phenyl atoms and the C(1)N(2)C(3) plane is equal to 44°, while for trans-isomer 6eA, this angle is 125°. The orientation of the methyl esters group differs only slightly for both isomers (Figure 3).
The indicated X-ray parameters (for 6eA and 6eB) are in good agreement with the data from 2D NOESY NMR experiments carried out for trans- and cis-isomers of compounds 5e,k,m,p and 6e,k,m,p (A/B) (Figure 4).
All these compounds have a similar configuration of the tricyclic skeleton and side groups for different substituents and gave the same cross-peaks in the NOESY spectra. Key NOE interactions are indicated in Figure 4. There are strong NOE interactions between H(4)–H(7a) and H(5)–H(7) for both diastereomers, indicating their identical endo-configurations for protons H(7) and H(7a). The difference lies in the side allyl/methallyl group, which is present in trans- or cis-form. There are strong resulting cross-peaks in the NOESY spectra between H(4)–H2C for trans-diastereomer A and H(4)–H(3) and H(3)–H(7a) for cis-diastereomer B (green arrows in Figure 4), which indicates their spatial proximity (<3 Å) and fully corresponds to the XRD data (Table 3). Cross-peaks between these protons but for other diastereomers (B and A, respectively)—between H(4)–H(3) and H(3)–H(7a) for trans-diastereomer A and H(4)–H2C for cis-diastereomer B—are absent in the NOESY spectra almost completely (crossed-out red arrows in Figure 4). The side allyl/methallyl group is conformationally quite mobile in solution, and NOE data for it cannot be appropriately compared with XRD distances.
1H and 13C-NMR spectra for series 5/6 of trans- and cis-diastereomers A/B have very few characteristic signals by which they could be easily distinguished. However, they have several characteristic patterns of correlation peaks in 2D HSQC spectra (Figure 5 and see the ESI with assignments data), for example, for CH(4) and CH(5) alkene protons, ortho-CH/meta-CH for para-substituted aromatics at N(2), and aliphatic CH2-groups in allyl/methallyl fragment. In the whole, the absolute values of chemical shifts are non-characteristic, since there is a fairly large scatter of chemical shifts depending on the substituents. However, for a number of protons and carbons, the sequence of the signals and difference of their chemical shifts are characteristic and the same for all compounds. The H(3), H(4), H(7a), CH2 protons, and the CH(3), C(3a), CH2 carbons have such features. Data on them are given in Tables S12 and S13 (see the ESI).
Therefore, the combination of the single-crystal XRD and NMR data allowed us to make an unambiguous assignment of isomers of all compounds 5 and 6 to the A or B series.
Considering that we were unable to achieve a noticeable predominance of the desired acids 5A by varying the IMDAF reaction conditions (see Table 1), we paid close attention to finding preparative methods for the separation of isomeric esters 6A/6B. We failed to find a suitable chromatographic system for the separation of isomers 6A/6B on silica gel due to the closeness of their retention factors. Mixtures of diastereomers 6aA/6aB (79/21), 6eA/6eB (16/84), 6fA/6fB (29/71) and 6pA/6pB (57/43) were used as model compounds, and mixtures of EtOAc/hexane, MeOH/THF and CHCl3/MeOH were applied as eluents. However, our attempts to separate the same esters by fractional crystallization were more fruitful. For this purpose, fine powders of the stereoisomer mixtures were stirred for a half-hour at 45–50 °C in a methanol solution; then, the least soluble isomer 6B was filtered off and the mother liquor was concentrated. After this treatment, the purity of individual isomers 6A and 6B exceeded 92% (diastereomeric excess (de) > 84%) in total yields of more than 94%. Repeating this procedure one more time resulted in virtually pure isomers. According to the described protocol, the diastereomers of 3-allyl substituted esters 6aA/6aB, 6fA/6fB and 6eA/6eB were separated and tested as metathesis objects (Scheme 5), described in the third part of this manuscript. Separation of 3-methallyl substituted esters 6pA/6pB by this way was less efficient, resulting in the formation of a 65/35 mixture of isomers after one “recrystallization” from MeOH.
The first experiments showed that solutions of the individual cis-isomers 6aB, 6eB and 6fB in CH2Cl2 at both room and elevated temperatures underwent polymerization with catalysts 1, 2 or HG-II (0.5–5 mol%). We were unable to isolate individual substances from metathesis product mixtures obtained under either an argon or an ethylene atmosphere (2 bar), except in two cases 6eB (R1 = H, R2 = 4-ClC6H4) and 6fB (R1 = H, R2 = 4-BrC6H). In these tests under an argon atmosphere, divinyl furo[2,3-c]pyrroles 8, probably the products of a series of the sequenced intermolecular cross metathesis reactions, were localized in a very low yield (Scheme 6). The formation of this product only occurs if the reaction is carried out for 15 min or less; upon longer heating, products 8 are completely transformed into a polymer.
In these two cases, it turned out to be possible to isolate products using column chromatography. The spatial structure of one of them (8b) was entirely confirmed by X-ray analysis (Figure 6). Compounds 8a and 8b also can be obtained by reaction of the esters 6e or 6f with ethylene (2 bar, 100 °C, 6 h) in dichloromethane in presence of catalyst 2 in 5–8% yields.
However, it was found that the pure trans-isomers 6aA, 6eA and 6pA are able to form tricyclic systems of cyclopenta[b]furo[2,3-c]pyrrole 7 in the RRM reaction (the last column of Table 2 and Table 4). Moreover, in contrast to the communications [9,10,11,17,19,21,22], in our case, the ring-rearrangement metathesis does not require the supporting application of ethylene that facilitates practical synthesis.
To achieve the maximum yield of the target tricycles 7, the metathesis reaction was optimized in terms of solvent, temperature, duration and amount of catalyst. Due to the poor solubility of esters 6eA (3-allyl) and 6pA (3-methallyl) in benzene and o-dichlorobenzene, we chose not to use aromatic solvents; similarly, polar solvents such as THF or MeCN proved to be ineffective in view of their capacity to interact with the Hoveyda–Grubbs type catalysts at elevated temperatures. The best of the tested mediums turned out to be halogenomethanes (CH2Cl2 and CHCl3), and hence they were used as solvents in all further experiments.
None of the selected catalysts exhibited activity at temperatures below 50 °C, and no rearrangement products 7e,p were fixed. Being the most active, the morpholine containing Cat. 1 started to work at the temperature of boiling chloroform (~60 °C), causing the rearrangement of the test 3-allyl isoindolone 6e. The more sterically demanding methallylic substrate 6pA are not recyclized under these conditions. An increase in temperature up to 100 °C under microwave irradiation led to a rapid decomposition of Cat. 1 (at 100 °C, the emerald-green color of the chloroform solution of Cat. 1 turned brown almost immediately), and accordingly to low yields of the product 7p. Obviously, higher temperatures and a more thermally stable catalyst should be used for a successful cyclization of methallyl derivatives 6ip.
N,N-Dimethylamino derivative (Cat. 2), with a shorter bond length between nitrogen and ruthenium (the N→Ru bond is 2.24 Å [27]), is stable at least up to 140 °C [27], and therefore turned out to be effective for the preparation of cyclopenta[b]furo[2,3-c]pyrrole 7p at temperatures above 100 °C in microwave-assisted experiments (entries 17–26, Table 4). It should be noted that the amount of catalysts 1 and 2 required for the complete conversion of the initial esters 6eA, 6pA did not exceed 0.5 mol % in both cases, while the reaction time was much shorter than in the papers cited above [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23].
After all attempts, the following conditions were found to be optimal for the recyclizations of the allyl derivative (6eA): boiling chloroform, 0.5 mol % of Cat. 1, 30 min. For synthesis of the methyl substituted 6pA, the best results were achieved under microwave activation in CH2Cl2 at 120 °C, 10 min, Cat. 2 (0.5 mol %).
It was interesting to compare the potential of the commercially available HG-II catalyst with catalysts Cat.1 and Cat.2 (see entries 4, 5, 24 and 25 in light gray in Table 4). Analysis of the isolated yields of products 7e,p allows us to conclude that the most popular commercial catalyst (HG-II) is less efficient.
The optimized protocols discussed above were planned to be extended to the recyclization of the remaining esters 6 (Table 2). However, the obvious disadvantage of the method is the need to separate mixtures of diastereomers 6A/6B. We tried to get around this obstacle, taking into account that cis-isomers 6aB6pB can be considered as undesirable ballast in the target tricycles 7 synthesis, and that the proposed pathway is based on readily available starting materials and reliable synthetic procedures, which allows the synthesis to be scaled up to kilogram quantities. Thereby, the developed approaches were further applied to the metathesis of mixtures of isomers 6A/6B (Scheme 5). To our satisfaction, the admixtures of cis-isomers 6B had no significant impact on the reaction progress; the polymerization products formed from them could be easily separated during purification by flash or column chromatography to provide the pure cyclopenta[b]furo[2,3-c]pyrroles 7 in good yields (based on the trans-isomer 6A). Thus, the majority of the RRM reactions presented in Table 2 were carried out using mixtures of isomers 6A/6B as starting materials.
It worth mentioning here that by revising the ethylene-supported transformations of the similar substrates [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23], the question of the sequence of metathesis stages arose. In theory, two paths exist for the rearrangement of 6 to 7 in an ethylene atmosphere: ROM/RCM, or vice versa, RCM/ROM successions [43]. We suppose that the plausible mechanism of the metathesis of trans-isomers 6A does not differ from the generally accepted one, as it is represented in Scheme 7.
We did not set ourselves the goal of investigating in detail the reaction mechanism; however, from general considerations, it can be assumed that the ring-closing metathesis (RCM) stage has to precede the opening of the 7-oxabicycloheptene cycle (ROM). This statement is supported by the observed behavior of trans-3-allyl derivatives 6aA6hA, which regroup more quickly and under more mild conditions compared to their 3-methallyl analogs 6iA6pA. This fact is probably related to the more rapid cycloaddition of the double bond of the allylic fragment to the Ru=CH-R center of the catalyst with the formation of structure I1 (Scheme 7). This statement is in good accordance with recently published data [43,44].
During the metathesis of cis-isomers 6B, which are not capable of RCM, the polymer similar to D is apparently formed through the cross-metathesis intermediate C (Scheme 8). Although the exo-cyclic allylic double bond is more reactive, it is also possible to obtain the short-lived triene 8a which is generated in the course of the catalytic cleavage of the endo-cyclic double bond in the starting molecules 6B (ruthenium complexes E and F).

3. Materials and Methods

3.1. General Remarks

All reagents were purchased from commercial suppliers (Acros Organics, Morris Plains, NJ, USA and Merck KGaA, Darmstadt, Germany) and used without further purification. Metathesis reactions required solvents (CH2Cl2 and CHCl3) pre-dried over anhydrous P2O5 and an inert atmosphere (dry Ar). Thin layer chromatography was carried out on aluminum backed silica pre-coated plates «Sorbfil» and «Alugram». The plates were visualized using a water solution of KMnO4 or UV (254 nm). After extraction, organic layers were dried over anhydrous MgSO4. IR spectra were obtained in KBr pellets or in thin films using an Infralum FT-801 IR-Fourier. NMR spectra were run in deuterated solvents (>99.5 atom % D) on Jeol JNM-ECA 600 (600.1 MHz for 1H and 150.9 MHz for 13C) spectrometer for 2–5% solutions in CDCl3 or DMSO-d6 at 22–23 °C using residual solvent signals (7.26/77.0 ppm for 1H/13C in CDCl3 and 2.50/39.5 ppm for 1H/13C in DMSO-d6) or TMS as an internal standard. CFCl3 was used as an internal standard for 19F-NMR spectra.

3.2. Experimental Procedures

The initial homoallylamines 4ap were synthesized according to the previously described procedures [45,46]. Carboxylic acids 5ap were synthesized using the known methods [29,30,31,32]. The detailed description of the preparation for all new compounds obtained in this work is given in the ESI section.

4. Conclusions

Our systematic studies on furan chemistry and metathesis reactions resulted in the development of general, effective method for the synthesis of natural product-like cyclopenta[b]furo[2,3-c]pyrroles from easily available, renewable furan-based precursors. The target tricyclic molecules were formed in high yields from esters of 3-allyl or 3-methallyl substituted 3a,6-epoxyisondolo-7-carboxylic acids in the presence of a new type of Hoveyda–Grubbs catalysts. The influence of the length of the N→Ru coordinate bond in the ruthenium complexes on the rate of the reaction was examined, and selective catalysts for the RRM of the allyl and methallyl derivatives were discovered. It was also proven that the recyclization is strictly stereoselective; only trans-isomer from two possible diastereoisomers of the initial methyl 3a,6-epoxyisondolo-7-carboxylates can be involved in the metathesis reaction.

Supplementary Materials

The following are available online, Figure S1: Structure of 6eA (296 K) C19H18ClNO4 (CCDC number 2023634), Figure S2: Structure of 6eB (296 K) C19H18ClNO4 (CCDC number 2023635), Figure S3: The overlay of molecules of trans-6eA (blue) and cis-6eB (red) methyl 3-allyl-2-(4-chlorophenyl)-1-oxo-1,2,3,6,7,7a-hexahydro-3a,6-epoxyisoindole-7-carboxylates according to the single crystal XRD, Figure S4: Structure of 8b (CCDC number is 2025199), Table S1: Crystal data and structure refinement for 6eA and 6eB, Table S2: Bond Lengths for 6eA, Table S3: Bond Angles for 6eA, Table S4: Hydrogen Bonds for 6eA, Table S5: Bond Lengths for 6eB, Table S6: Bond Angles for 6eB, Table S7: Hydrogen Bonds for 6eB, Table S8: Crystal data and structure refinement for 8b, Table S9: Bond Lengths for 8b, Table S10: Bond Angles for 8b, Table S11: Hydrogen Bonds for 8b, Table S12: Selected 1H-NMR chemical shifts for trans- and cis-isomers, Table S13: Selected 13C-NMR chemical shifts for trans- and cis-isomers.

Author Contributions

A.S.A., M.A.V., P.A.K., and N.L.M. performed the synthesis and purification of the obtained compounds; A.A.S. and M.S.G. performed the X-ray diffraction experiments; K.B.P. conducted the NMR data analysis; R.A.N. performed the NMR experiments; V.V.K. organized the drafts’ preparation and participated in the editing; F.I.Z. conceived, designed the experiments and supervised all processes of this investigation. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Russian Science Foundation (RSF) for funding this work through a general research project under grant number (project No. 18-13-00456).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ogba, O.M.; Warner, N.C.; O’Leary, D.J.; Grubbs, R.H. Recent advances in ruthenium-based olefin metathesis. Chem. Soc. Rev. 2018, 47, 4510–4544. [Google Scholar] [CrossRef] [Green Version]
  2. Kotha, S.; Dipak, M.K. Strategies and tactics in olefin metathesis. Tetrahedron 2012, 68, 397–421. [Google Scholar] [CrossRef]
  3. Hughes, D.; Wheeler, P.; Ene, D. Olefin Metathesis in Drug Discovery and Development- Examples from Recent Patent Literature. Org. Process Res. Dev. 2017, 21, 1938–1962. [Google Scholar] [CrossRef]
  4. Higman, C.S.; Lummiss, J.A.; Fogg, D. Olefin Metathesis at the Dawn of Implementation in Pharmaceutical and Specialty-Chemicals Manufacturing. Angew. Chem. Int. Ed. 2016, 55, 3552–3565. [Google Scholar] [CrossRef]
  5. Yu, M.; Lou, S.; Gonzalez-Bobes, F. Closing Metathesis in Pharmaceutical Development: Fundamentals, Applications, and Future Directions. Org. Process Res. Dev. 2018, 22, 918–946. [Google Scholar] [CrossRef]
  6. Müller, D.S.; Baslé, O.; Mauduit, M.A. Tutorial review of stereoretentive olefin metathesis based on ruthenium dithiolate catalysts. Beilstein J. Org. Chem. 2018, 14, 2999–3010. [Google Scholar] [CrossRef]
  7. Montgomery, T.P.; Johns, A.M.; Grubbs, R.H. Recent Advancements in Stereoselective Olefin Metathesis Using Ruthenium Catalysts. Catalysts 2017, 7, 87. [Google Scholar] [CrossRef] [Green Version]
  8. Nolan, S.P.; Clavier, H. Chemoselective olefin metathesis transformations mediated by ruthenium complexes. Chem. Soc. Rev. 2010, 39, 3305–3316. [Google Scholar] [CrossRef]
  9. Lam, J.K.; Schmidt, Y.; Vanderwal, C.D. Complex polycyclic scaffolds by metathesis rearrangement of Himbert arene/allene cycloadducts. Org. Lett. 2012, 14, 5566–5569. [Google Scholar] [CrossRef]
  10. Kotha, S.; Meshram, M.; Khedkar, P.; Banerjee, S.; Deodhar, D. Recent applications of ring-rearrangement metathesis in organic synthesis. Beilstein J. Org. Chem. 2015, 11, 1833–1864. [Google Scholar] [CrossRef] [Green Version]
  11. Kotha, S.; Pulletikurti, S.A. Metathetic Approach to [5/5/6] Aza-Tricyclic Core of Dendrobine, Kopsanone, and Lycopalhine A Type of Alkaloids. Synthesis 2019, 51, 3981–3988. [Google Scholar] [CrossRef]
  12. Kotha, S.; Meshram, M.; Dommaraju, Y. Design and Synthesis of Polycycles, Heterocycles, and Macrocycles via Strategic Utilization of Ring-Closing Metathesis. Chem. Rec. 2018, 18, 1613–1632. [Google Scholar] [CrossRef] [PubMed]
  13. Kiss, L.; Kardos, M.; Vass, C.; Fülöp, F. Application of metathesis reactions in the synthesis and transformations of functionalized beta-amino acid derivatives. Synthesis 2018, 50, 3571–3588. [Google Scholar] [CrossRef] [Green Version]
  14. Acharyya, R.K.; Rej, R.K.; Nanda, S. Exploration of Ring Rearrangement Metathesis Reaction: A General and Flexible Approach for the Rapid Construction [5,n]-Fused Bicyclic Systems en Route to Linear Triquinanes. J. Org. Chem. 2018, 83, 2087–2103. [Google Scholar] [CrossRef] [PubMed]
  15. Clavier, H.; Broggi, J.; Nolan, S.P. Ring-Rearrangement Metathesis (RRM) Mediated by Ruthenium-Indenylidene Complexes. Eur. J. Org. Chem. 2010, 2010, 937–943. [Google Scholar] [CrossRef]
  16. Datta, R.; Ghosh, S. Domino ring-opening–ring-closing enyne metathesis vs. enyne metathesis of norbornene derivatives with alkynyl side chains. Construction of condensed polycarbocycles. Beilstein J. Org. Chem. 2018, 14, 2708–2714. [Google Scholar] [CrossRef] [Green Version]
  17. Kotha, S.; Ravikumar, O. Design and synthesis of oxa-bowls via Diels–Alder reaction and ring-rearrangement metathesis as key steps. Tetrahedron Lett. 2014, 55, 5781–5784. [Google Scholar] [CrossRef]
  18. Kotha, S.; Ravikumar, O. Ring-Rearrangement-Metathesis Approach to Polycycles: Substrate-Controlled Stereochemical Outcome During Grignard Addition. Eur. J. Org. Chem. 2016, 3900–3906. [Google Scholar] [CrossRef]
  19. Funel, J.A.; Prunet, J. Domino metathesis reactions for the synthesis of fused tricyclic frameworks. Synlett 2005, 235–238. [Google Scholar] [CrossRef]
  20. Kotha, S.; Ravikumar, O. Diversity−Oriented Approach to Carbocycles and Heterocycles through Ring− Rearrangement Metathesis, Fischer Indole Cyclization, and Diels—Alder Reaction as Key Steps. Eur. J. Org. Chem. 2014, 2014, 5582–5590. [Google Scholar] [CrossRef]
  21. Kotha, S.; Pulletikurti, S. Synthesis of propellanes containing a bicyclo[2.2.2]octene unit via the Diels–Alder reaction and ring-closing metathesis as key steps. RCS Adv. 2018, 8, 14906–14915. [Google Scholar] [CrossRef]
  22. Kotha, S.; Aswar, V.R. Target specific tactics in olefin metathesis: Synthetic approach to cis-syn-cis-triquinanes and propellanes. Org. Lett. 2016, 18, 1808–1811. [Google Scholar] [CrossRef] [PubMed]
  23. Cheng-Sánchez, I.; Sarabia, F. Recent Advances in Total Synthesis via Metathesis Reactions. Synthesis 2018, 50, 3749–3786. [Google Scholar]
  24. Firth, J.D.; Craven, P.G.E.; Lilburn, M.; Pahl, A.; Marsden, S.P.; Nelson, A. A biosynthesis-inspired approach to over twenty diverse natural product-like scaffolds. Chem. Commun. 2016, 52, 9837–9840. [Google Scholar] [CrossRef] [Green Version]
  25. Mariscal, R.; Maireles-Torres, P.; Ojeda, M.; Sádaba, I.; Granados, M.L. Furfural: A renewable and versatile platform molecule for the synthesis of chemicals and fuels. Energy Environ. Sci. 2016, 9, 1144–1189. [Google Scholar] [CrossRef]
  26. Bozell, J.J.; Petersen, G.R. Technology development for the production of biobased products from biorefinery carbohydrates—The US Department of Energy’s “Top 10” revisited. Green Chem. 2010, 12, 539–554. [Google Scholar] [CrossRef]
  27. 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]
  28. Polyanskii, K.B.; Alekseeva, K.A.; Kumandin, P.A.; Atioğlu, Z.; Akkurt, M.; Toze, F.A. Crystal structure of [1,3-bis(2,4,6-trimethylphenyl)imidazolidin-2-ylidene]dichlorido{2-[1-(dimethylamino)ethyl]benzylidene} ruthenium including an unknown solvate. Acta Cryst. E 2019, 75, 342–345. [Google Scholar] [CrossRef]
  29. Zubkov, F.I.; Boltukhina, E.V.; Turchin, K.F.; Borisov, R.S.; Varlamov, A.V. New synthetic approach to substituted isoindolo[2,1-a]quinoline carboxylic acids via intramolecular Diels–Alder reaction of 4-(N-furyl-2)-4-arylaminobutenes-1 with maleic anhydride. Tetrahedron 2005, 61, 4099–4113. [Google Scholar] [CrossRef]
  30. Boltukhina, E.V.; Zubkov, F.I.; Nikitina, E.V.; Varlamov, A.V. Novel approach to isoindolo[2,1-a]quinolines: Synthesis of 1-and 3-halo-substituted 11-oxo-5,6,6a,11-tetrahydroisoindolo [2,1-a]quinoline-10-carboxylic acids. Synthesis 2005, 2005, 1859–1875. [Google Scholar] [CrossRef]
  31. Zubkov, F.I.; Boltukhina, E.V.; Nikitina, E.V.; Varlamov, A.V. Study of regioselectivity of intramolecular cyclization of N-(m-R-phenyl)- and N-(α-naphthyl)-2-allyl(methallyl)-6-carboxy-4-oxo-3-aza-10-oxatricyclo[5.2.1.01,5]dec-8-enes. Russ. Chem. Bull. 2004, 53, 2816–2829. [Google Scholar] [CrossRef]
  32. Varlamov, A.V.; Boltukhina, E.V.; Zubkov, F.I.; Nikitina, E.V.; Turchin, K.F. Intramolecular [4+2] cycloaddition of furfurylsubstituted homoallylamines to allylhalides, acryloyl chloride and maleic anhydride. J. Heterocyclic Chem. 2006, 43, 1479–1495. [Google Scholar] [CrossRef]
  33. Zubkov, F.I.; Nikitina, E.V.; Varlamov, A.V. Thermal and catalytic intramolecular [4+2]-cycloaddition in 2-alkenylfurans. Russ. Chem. Rev. 2005, 74, 639–669. [Google Scholar] [CrossRef]
  34. Padwa, A.; Flick, A.C. Intramolecular Diels–Alder Cycloaddition of Furans (IMDAF) for Natural Product Synthesis. Adv. Heterocycl. Chem. 2013, 110, 1–41. [Google Scholar]
  35. Kiselev, V.D.; Kornilov, D.A.; Sedov, I.A.; Konovalov, A.I. Solvent Influence on the Diels-Alder Reaction Rates of 9-(Hydroxymethyl)anthracene and 9,10-Bis(hydroxymethyl)anthracene with Two Maleimides. Int. J. Chem. Kinet. 2017, 49, 61–68. [Google Scholar] [CrossRef]
  36. Widstrom, A.L.; Lear, B.J. Structural and solvent control over activation parameters for a pair of retro Diels-Alder reactions. Sci. Rep. 2019, 9, 18267. [Google Scholar] [CrossRef]
  37. Gandini, A.; Coelho, D.; Silvestre, A.J. Reversible click chemistry at the service of macromolecular materials. Part 1: Kinetics of the Diels–Alder reaction applied to furan–maleimide model compounds and linear polymerizations. Eur. Polym. J. 2008, 44, 4029–4036. [Google Scholar] [CrossRef]
  38. Kotha, S.; Banerjee, S. Recent developments in the retro-Diels-Alder reaction. RSC Adv. 2013, 3, 7642–7666. [Google Scholar] [CrossRef]
  39. Pedrosa, R.; Andrés, C.; Nieto, J. A short diastereoselective synthesis of enantiopure highly substituted tetrahydroepoxyisoindolines. J. Org. Chem. 2000, 65, 831–839. [Google Scholar] [CrossRef]
  40. Zubkov, F.I.; Boltukhina, E.V.; Turchin, K.F.; Varlamov, A.V. An efficient approach to isoindolo [2,1-b][2] benzazepines via intramolecular [4+2] cycloaddition of maleic anhydride to 4-α-furyl-4-N-benzylaminobut-1-enes. Tetrahedron 2004, 60, 8455–8463. [Google Scholar] [CrossRef]
  41. Pedrosa, R.; Sayalero, S.; Vicente, M.; Casado, B. Chiral Template Mediated Diastereoselective Intramolecular Diels−Alder Reaction Using Furan as a Diene. Toward the Synthesis of Enantiopure Trisubstituted Tetrahydroepoxyisoindolones. J. Org. Chem. 2005, 70, 7273–7278. [Google Scholar] [CrossRef] [PubMed]
  42. Suzuki, M.; Okada, T.; Taguchi, T.; Hanzawa, Y.; Iitaka, Y. Intramolecular Diels-Alder reactions of furan derivatives: Steric and electronic effects of trifluoromethyl groups. J. Fluorine Chem. 1992, 57, 239–243. [Google Scholar] [CrossRef]
  43. Holub, N.; Blechert, S. Ring-Rearrangement Metathesis. Chem. Asian J. 2007, 2, 1064–1082. [Google Scholar] [CrossRef] [PubMed]
  44. Harrity, J.P.; Visser, M.S.; Gleason, J.D.; Hoveyda, A.H. Ru-Catalyzed Rearrangement of Styrenyl Ethers. Enantioselective Synthesis of Chromenes through Zr-and Ru-Catalyzed Processes. J. Am. Chem. Soc. 1997, 119, 1488–1489. [Google Scholar] [CrossRef]
  45. Kouznetsov, V.; Öcal, N.; Turgut, Z.; Zubkov, F.; Kaban, S.; Varlamov, A. Allylation and Heterocycloaddition reactions of Aldimines: Furan- and Quinolinecarboxaldhydes. Monatsh. Chem. 1998, 129, 671–677. [Google Scholar] [CrossRef]
  46. Urbina, J.M.; Cortés, J.C.; Palma, A.; López, S.N.; Zacchino, S.A.; Enriz, R.D.; Ribas, C.; Kouznetsov, V.V. Inhibitors of the fungal cell wall. Synthesis of 4-aryl-4-N-arylamine-1-butenes and related compounds with inhibitory activities on β(1–3) glucan and chitin synthases. Bioorg. Med. Chem. 2000, 8, 691–698. [Google Scholar] [CrossRef]
Scheme 1. Selected examples of the ring-rearrangement metathesis (RRM) of tri- and tetracyclic bridged systems closest to the topic of this study.
Scheme 1. Selected examples of the ring-rearrangement metathesis (RRM) of tri- and tetracyclic bridged systems closest to the topic of this study.
Molecules 25 05379 sch001
Scheme 2. Synthesis of cyclopenta[b]furo[2,3-c]pyrrole core 2 via ring-rearrangement metathesis (HG-II—the second-generation Hoveyda–Grubbs catalyst, TBME—methyl tert-butyl ether) [24].
Scheme 2. Synthesis of cyclopenta[b]furo[2,3-c]pyrrole core 2 via ring-rearrangement metathesis (HG-II—the second-generation Hoveyda–Grubbs catalyst, TBME—methyl tert-butyl ether) [24].
Molecules 25 05379 sch002
Figure 1. Hoveyda–Grubbs-type catalysts [27] used for metathesis reactions in the present work.
Figure 1. Hoveyda–Grubbs-type catalysts [27] used for metathesis reactions in the present work.
Molecules 25 05379 g001
Scheme 3. Synthesis of the starting 3a,6-epoxyisoindole-7-carboxylic acids 5 [29,30,31,32].
Scheme 3. Synthesis of the starting 3a,6-epoxyisoindole-7-carboxylic acids 5 [29,30,31,32].
Molecules 25 05379 sch003
Figure 2. Structures of the cis and trans diastereoisomers of 3a,6-epoxyisoindole-7-carboxylic acids 5.
Figure 2. Structures of the cis and trans diastereoisomers of 3a,6-epoxyisoindole-7-carboxylic acids 5.
Molecules 25 05379 g002
Scheme 4. Esterification of 3a,6-epoxyisondol-7-carboxylic acids 5 with methanol affording respective 3a,6-epoxyisoindole-7-carboxylates (6).
Scheme 4. Esterification of 3a,6-epoxyisondol-7-carboxylic acids 5 with methanol affording respective 3a,6-epoxyisoindole-7-carboxylates (6).
Molecules 25 05379 sch004
Figure 3. X-ray crystal structures of trans-6eA (at the top) and cis-6eB (at the bottom) methyl 3-allyl-2-(4-chlorophenyl)-1-oxo-1,2,3,6,7,7a-hexahydro-3a,6-epoxyisoindole-7-carboxylates.
Figure 3. X-ray crystal structures of trans-6eA (at the top) and cis-6eB (at the bottom) methyl 3-allyl-2-(4-chlorophenyl)-1-oxo-1,2,3,6,7,7a-hexahydro-3a,6-epoxyisoindole-7-carboxylates.
Molecules 25 05379 g003
Figure 4. (A.) Key observed NOE interactions in 2D NOESY NMR spectra for both representative trans- and cis-diastereomers 5e,k,m,p and 6e,k,m,p (A/B). (B.) Representative part of 2D NOESY spectra for mixture of trans- (violet) and cis- (red) diastereomers of the allyl derivative 6e with assignments of signals and NOE.
Figure 4. (A.) Key observed NOE interactions in 2D NOESY NMR spectra for both representative trans- and cis-diastereomers 5e,k,m,p and 6e,k,m,p (A/B). (B.) Representative part of 2D NOESY spectra for mixture of trans- (violet) and cis- (red) diastereomers of the allyl derivative 6e with assignments of signals and NOE.
Molecules 25 05379 g004
Figure 5. Typical characteristic pattern of CH(4) and CH(5) correlation peaks in 2D HSQC spectra (mixture of trans and cis diastereomers of 6k as an example) used for assignments of diastereomers.
Figure 5. Typical characteristic pattern of CH(4) and CH(5) correlation peaks in 2D HSQC spectra (mixture of trans and cis diastereomers of 6k as an example) used for assignments of diastereomers.
Molecules 25 05379 g005
Scheme 5. Synthesis of cyclopenta[b]furo[2,3-c]pyrroles 7.
Scheme 5. Synthesis of cyclopenta[b]furo[2,3-c]pyrroles 7.
Molecules 25 05379 sch005
Scheme 6. Results of metathesis of cis-isomers 6B.
Scheme 6. Results of metathesis of cis-isomers 6B.
Molecules 25 05379 sch006
Figure 6. X-ray crystal structure of methyl (2RS,3SR,3aRS,6SR,6aSR)-6-allyl-5-(4-bromophenyl)-4-oxo-2,6a-divinylhexahydro-2H-furo[2,3-c]pyrrole-3-carboxylate (8b).
Figure 6. X-ray crystal structure of methyl (2RS,3SR,3aRS,6SR,6aSR)-6-allyl-5-(4-bromophenyl)-4-oxo-2,6a-divinylhexahydro-2H-furo[2,3-c]pyrrole-3-carboxylate (8b).
Molecules 25 05379 g006
Scheme 7. Possible reaction sequence for RRM by the example of the conversion of trans-isomer 6aA to 7a in the presence of Cat. 1.
Scheme 7. Possible reaction sequence for RRM by the example of the conversion of trans-isomer 6aA to 7a in the presence of Cat. 1.
Molecules 25 05379 sch007
Scheme 8. Possible behavior of cis-isomer 6aB under metathesis reaction conditions.
Scheme 8. Possible behavior of cis-isomer 6aB under metathesis reaction conditions.
Molecules 25 05379 sch008
Table 1. Influence of solvents and temperature on the yields and ratios of trans-5A/cis-5B isomers of 3-allyl-1-oxo-3a,6-epoxyisoindole-7-carboxylic acids (5).
Table 1. Influence of solvents and temperature on the yields and ratios of trans-5A/cis-5B isomers of 3-allyl-1-oxo-3a,6-epoxyisoindole-7-carboxylic acids (5).
Compd.SubstituentsConditions, Ratios of Trans-5A/cis-5B Isomers a, and Total Yields (%)
R1R2CH2Cl2
−16 °C, 3 d
CH2Cl2
r.t, 3 d
MeCN
r.t, 3 d
MeCN
∆, 3 h
PhH
∆, 3 h
PhMe
r.t, 3 d
PhMe
∆, 3 h
5aHPh76/24
(43)
7/93
(58)
47/53
(45)
44/56
(15)
59/41
(82)
48/52
(85)
59/41
(90)
5bH3-MeC6H463/37
(49)
6/94
(65)
59/44
(41)
52/48
(25)
58/42
(85)
55/45
(90)
57/43
(92)
5gH4-IC6H475/25
(52)
0/100
(67)
24/75
(38)
28/72
(23)
52/48
(86)
78/22
(90)
71/29
(91)
5iMePh68/32
(63)
23/77
(75)
53/47
(62)
50/50
(47)
59/41
(95)
67/33
(97)
67/33
(98)
5mMeBn60/40
(64)
4/96
(77)
55/45
(65)
58/42
(54)
59/41
(96)
59/41
(98)
56/44
(99)
Average values68/32
(54)
8/92
(68)
48/52
(50)
46/54
(33)
57/43
(89)
61/39
(92)
62/38
(94)
a The ratios of the trans-5A/cis-5B isomers were established by 1H-NMR analysis of the isolated solids.
Table 2. Substituents R, isomer ratios, and yields for compounds 57.
Table 2. Substituents R, isomer ratios, and yields for compounds 57.
EntryR1R2Trans-5A/Cis-5B RatioYield of Acid 5, % Trans-6A/Cis-6B RatioYield of Ester 6, %Yield of Tricycle 7, %
aHPh76/2490 b79/218568 c
95/5 e53 e83 c
2/98 e12 e0 c
bH3-MeC6H457/4395 a69/316357 c
cHBn90/1057 b85/158565 c
dH3-ClC6H476/2448 b70/308057 c
eH4-ClC6H480/2091 a70/307273 c
98/2 e46 e89 c
8/92 e18 e0 c
fH4-BrC6H480/2092 a72/287063 c
99/1 e46 e87 c
5/95 e12 e0 c
gH4-IC6H471/2952 b59/416348 c
hH3-Cl,4-FC6H357/4390 a66/348654 c
iMePh68/3263 b55/455350 d
jMe3-MeC6H460/4067 b45/557338 d
kMe4-MeC6H451/4993 a50/509844 d
lMe4-iPrC6H471/2988 a87/138262 d
mMeBn56/4464 a42/585739 d
nMe2-ClC6H471/2963 b75/259559 d
oMe3-ClC6H468/3269 b83/177861 d
pMe4-BrC6H476/2464 b57/435652 d
65/35 e45 e61 d
Conditions: a PhMe, ∆, 3 h; b Cat.1, CH2Cl2, −16 °C, 3 d; c Cat.2, CHCl3, Δ, 30 min, under argon; d CH2Cl2, MW, 120 °C, 10 min, under argon; e In light grey—ratios and yields are presented after separation of the above indicated isomer 6A/6B mixture (in dark gray) with methanol.
Table 3. Selected interatomic distances (Å) in esters 6eA and 6eB according to single-crystal XRD.
Table 3. Selected interatomic distances (Å) in esters 6eA and 6eB according to single-crystal XRD.
Distances (Å)Trans-6eACis-6eB
H(3)H(7a)3.781(1)2.915(1)
H(3)H(4)3.493(1)2.871(1)
C(10) aH(4)4.521(2)3.772(5)
a C(10) is the methylene carbon atom of the allyl fragment.
Table 4. Optimization of the RRM protocol using the test esters 6eA and 6pA.
Table 4. Optimization of the RRM protocol using the test esters 6eA and 6pA.
EntryStarting EsterCatalyst (mol %)ConditionsTarget TricycleYield, % a
16eACat. 1 (0.5)CHCl3, Δ, 3 h, Ar7e85
26eACat. 1 (0.5)CHCl3, Δ, 1.5 h, Ar7e80
36eACat. 1 (0.5)CHCl3, Δ, 1.0 h, Ar7e82
46eACat. 1 (0.5)CHCl3, Δ, 0.5 h, Ar7e80 b
56eAHG-II (0.5)CHCl3, Δ, 0.5 h, Ar7e74
66eACat. 1 (0.5)CHCl3, Δ, 0.25 h, air7e35
76eACat. 1 (0.5)CHCl3, Δ, 0.25 h, Ar7e83 b
86eACat. 1 (0.5)CH2Cl2, Δ, 3 h, Ar7e58
96eACat. 1 (0.5)CH2Cl2, Δ, 1.5 h, Ar7e52
106eACat. 1 (0.5)CH2Cl2, Δ, 0.5 h, Ar7e37
116eACat. 1 (1.0)CHCl3, Δ, 0.25 h, Ar7e85
126eACat. 1 (1.0)CHCl3, Δ, 0.5 h, Ar7e85
136eACat. 1 (0.1)CHCl3, Δ, 0.25 h, Ar7e32
146eACat. 1 (0.1)CHCl3, Δ, 1.0 h, Ar7e51
156eACat. 1 (0.1)CHCl3, Δ, 1.5 h, Ar7e54
166eACat. 1 (0.1)CHCl3, Δ, 3 h, Ar7e55
176pACat. 2 (0.5)CHCl3, Δ, 30 min, Ar, MW, 120 °C7p55
186pACat. 2 (0.5)CHCl3, Δ, 20 min, Ar, MW, 120 °C7p52
196pACat. 2 (0.5)CHCl3, Δ, 10 min, Ar, MW, 120 °C7p45
206pACat. 2 (0.5)CH2Cl2, Δ, 30 min, Ar, MW, 120 °C7p85
216pACat. 2 (0.5)CH2Cl2, Δ, 20 min, Ar, MW, 120 °C7p84
226pACat. 2 (0.5)CH2Cl2, Δ, 15 min, Ar, MW, 120 °C7p80
236pACat. 2 (0.5)CH2Cl2, Δ, 10 min, Ar, MW, 100 °C7p0
246pACat. 2 (0.5)CH2Cl2, Δ, 10 min, Ar, MW, 120 °C7p82 b
256pAHG-II (0.5)CH2Cl2, Δ, 10 min, Ar, MW, 120 °C7p42
266pACat. 2 (1.0)CH2Cl2, Δ, 10 min, Ar, MW, 120 °C7p83
256pACat. 2 (1.0)CH2Cl2, Δ, 15 min, Ar, MW, 120 °C7p85
266pACat. 2 (0.1)CH2Cl2, Δ, 10 min, Ar, MW, 120 °C7p43
276pACat. 2 (0.1)CH2Cl2, Δ, 15 min, Ar, MW, 120 °C7p47
a Isolated yields of products 7 after column chromatography. b In gray—the best conditions and yields (in bold) are highlighted.
Sample Availability: Samples of the compounds 57 are available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Antonova, A.S.; Vinokurova, M.A.; Kumandin, P.A.; Merkulova, N.L.; Sinelshchikova, A.A.; Grigoriev, M.S.; Novikov, R.A.; Kouznetsov, V.V.; Polyanskii, K.B.; Zubkov, F.I. Application of New Efficient Hoveyda–Grubbs Catalysts Comprising an N→Ru Coordinate Bond in a Six-Membered Ring for the Synthesis of Natural Product-Like Cyclopenta[b]furo[2,3-c]pyrroles. Molecules 2020, 25, 5379. https://doi.org/10.3390/molecules25225379

AMA Style

Antonova AS, Vinokurova MA, Kumandin PA, Merkulova NL, Sinelshchikova AA, Grigoriev MS, Novikov RA, Kouznetsov VV, Polyanskii KB, Zubkov FI. Application of New Efficient Hoveyda–Grubbs Catalysts Comprising an N→Ru Coordinate Bond in a Six-Membered Ring for the Synthesis of Natural Product-Like Cyclopenta[b]furo[2,3-c]pyrroles. Molecules. 2020; 25(22):5379. https://doi.org/10.3390/molecules25225379

Chicago/Turabian Style

Antonova, Alexandra S., Marina A. Vinokurova, Pavel A. Kumandin, Natalia L. Merkulova, Anna A. Sinelshchikova, Mikhail S. Grigoriev, Roman A. Novikov, Vladimir V. Kouznetsov, Kirill B. Polyanskii, and Fedor I. Zubkov. 2020. "Application of New Efficient Hoveyda–Grubbs Catalysts Comprising an N→Ru Coordinate Bond in a Six-Membered Ring for the Synthesis of Natural Product-Like Cyclopenta[b]furo[2,3-c]pyrroles" Molecules 25, no. 22: 5379. https://doi.org/10.3390/molecules25225379

Article Metrics

Back to TopTop