Next Article in Journal
Hepatic Alterations in a BTBR T + Itpr3tf/J Mouse Model of Autism and Improvement Using Melatonin via Mitigation Oxidative Stress, Inflammation and Ferroptosis
Previous Article in Journal
Transcription Factors in the Pathogenesis of Lupus Nephritis and Their Targeted Therapy
Previous Article in Special Issue
[3+2]-Cycloaddition of Nitrile Imines to Parabanic Acid Derivatives—An Approach to Novel Spiroimidazolidinediones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Concise and Free-Metal Access to Lactone-Annelated Pyrrolo[2,1-a]isoquinoline Derivatives via a 1,2-Rearrangement Step

by
Arina Y. Obydennik
1,
Alexander A. Titov
1,
Anna V. Listratova
1,
Tatiana N. Borisova
1,
Victor B. Rybakov
2,
Leonid G. Voskressensky
1,* and
Alexey V. Varlamov
1
1
Organic Chemistry Department, Science Faculty, Peoples’ Friendship University of Russia (RUDN University), 6 Miklukho-Maklaya Street, Moscow 117198, Russia
2
Department of Chemistry, Lomonosov Moscow State University, Leninskie Gory, 1-3, Moscow 119991, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(2), 1085; https://doi.org/10.3390/ijms25021085
Submission received: 28 December 2023 / Revised: 10 January 2024 / Accepted: 11 January 2024 / Published: 16 January 2024

Abstract

:
Here, An efficient approach to obtaining previously unknown furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline derivatives from readily available 1-R-1-ethynyl-2-vinylisoquinolines is described. The reaction features a simple procedure, occurs in hexaflouroisopropanol and does not require elevated temperatures. It has been found that the addition of glacial acetic acid significantly increases the yields of the target spirolactone products. Using trifluoroethanol instead of hexaflouroisopropanol results in the formation of pyrido[2,1-a]isoquinolines.

1. Introduction

The γ-lactone moiety is present in many bioactive natural products isolated from various plants and fungal metabolites [1,2,3]. Compounds with lactone and spirolactone fragments are characterized by a broad range of bioactivities and find their application in the field of medicine and agriculture. Thus, trans-dehydrocrotonin exhibits hypolipidemic and hypoglycaemic properties and has anti-cancer activity [4,5,6,7]; tetranorditerpenoids can be used as herbicides [8]; dehydroleucodine has anti-inflammatory and antiulcer activities [9]; and Stemoamide, Stemonamine and Tuberostemospironine, being Stemona alkaloids, possess anti-inflammatory, insecticidal, antitussive activities (Figure 1) [2,10,11]. Lactonic pyrrolizidinone alkaloids—pyrrolizilactone and UCS1025A—demonstrate potent antibacterial and antitumor effects [3,12].
Due to having a wide profile of pharmaceutical activities, spirolactones attract considerable attention from scientists and advance both the development of simple and effective synthetic routes to such structures and the further study of their properties. Recently, numerous methods for the synthesis of spirolactones have been described in the literature [13,14,15,16]. Among a variety of known approaches, those that are based on mild, free-metal and step-economic reactions start from readily available materials and meet the requirements of modern and “advantageous” synthetic chemistry, and so deserve special attention. Domino processes, incorporating the rearrangement and reconstruction of the carbon skeleton and leading to the complexity of a molecule’s structure to be quickly revealed in one step, can be considered as eligible candidates, fitting all the requirements of “advantageous chemistry” [17].

2. Results and Discussion

Herein, we report a study devoted to elucidating the divergent transformations of 1-R-1-ethynyl-2-vinyl-substituted 1,2,3,4-tetrahydroisoquinolines 1ag which occur in protic fluorinated solvents. One of the observed transformations proceeds via a 1,2-rearrangement step in the presence of AcOH/HFIP and opens up access to previously unexplored furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline derivatives 3.
Previously, we have described the chemical behavior of 1-R-1-ethynyl-2-vinyl-substituted 1,2,3,4-tetrahydroisoquinolines in aprotic solvents [18]. It has been shown that the route of the MW-stimulated rearrangements deeply depends on the type of solvent used. The use of toluene favored the formation of pyrrolo[2,1-b][3]benzazepines, while switching to acetonitrile afforded pyrido[2,1-a]isoquinolines in good yields. Encouraged by these unusual results, we decided to examine the influence of protic solvents, particularly fluorinated alcohols—trifluoroethanol and hexafluoroisopropanol (HFIP)—on the disclosed rearrangements. Fluorinated alcohols are characterized by having a low nucleophilicity and high ionizing and solvating power, increased Brønsted acidity in the hydroxyl proton and high polarity, as well as the ability to affect the regio- and chemoselectivity of a reaction and its process rate [19,20,21,22]. In other words, they could open up new directions for these well-known transformations.
The initial 1-R-1-ethynyl-2-vinyl-substituted 1,2,3,4-tetrahydroisoquinolines 1ag were obtained, according to a previously described procedure, from the corresponding 3,4-dihydroisoquinolines and methyl propiolate [18]. We began our study with transformations of tetrahydroisoquinolines 1ag so it would arise in less acidic trifluoroethanol (pKa = 12.4) [19]. To our delight, the conversions did not require elevated temperatures and proceeded smoothly at 20 °C to generate pyrido[2,1-a]isoquinolines, in 55–95% yields, as the sole product (Table 1). To understand what caused the change in the transformation route, the effect of the fluorinated alcohol or simply the presence of a protic solvent, we carried out a reaction of isoquinoline 1a with a non-fluorinated ethanol (pKa = 15.9) [19]. Substrate 1a was transformed into product 2a but the use of ethanol as a solvent slowed down the process three times; in addition, the yield of the target compound decreased to 78%. We have already reported on the synthesis of pyrido[2,1-a]isoquinolines from isoquinolines 1af in acetonitrile in the presence of triphenylphosphine [18]. In that case, the conversions required more severe conditions, which makes it less attractive compared to the present protocol.
Inspired by the results obtained in trifluoroethanol, we decided to explore the intramolecular changes when starting with tetrahydroisoquinolines 1ag in more acidic hexafluoroisopropanol (HFIP) (pKa = 9.3) [19]. Using isoquinoline 1a as a model substrate, we performed a reaction at 20 °C. The transformation proceeded smoothly, but, to our surprise, led to a reaction mixture which consisted of lactonic pyrrolo[2,1-a]isoquinoline 3a (25%) and pyrido[2,1-a]isoquinoline 2a (71%) (Table 2, entry 1). The formation of 3a was completely unexpected. The literature survey has not revealed the analogous structures, and we have succeeded only in finding isomeric ones [23]. It was clear that the acidity of the solvent played a key role. Given our earlier published studies demonstrating that the use of more acidic solvents such as HFIP and AcOH can alter the routes in the transformation of 1-R-ethynyl-decorated tetrahydroisoquinolines in reaction with activated alkynes towards more thermodynamically stable products, we considered that increasing the acidity of the medium with acetic acid would promote the construction of product 3a [24,25]. Indeed, the yield of the desired 3a was improved to 43% by adding 0.5 equiv of glacial acetic acid; however, the formation of compound 2a was still observed (Table 2, entry 2). The best result was achieved with 3.0 equiv of AcOH to produce lactone 3a with a 55% yield. It is noteworthy that a further increase in acetic acid did not have any significant effect on the yield of the target compound 3a (Table 2, entries 3 and 4).
With the optimized conditions in hand, we investigated the scope of the discovered transformation. To estimate the effect of the substituents attached at C-1 during the intramolecular changes to tetrahydroisoquinolines 1bg, experiments with different alkyl and aryl substituents were carried out. Isoquinolines 1bd with isopropyl, benzyl and phenyl groups proved to be good substrates for the transformation, producing lactonic pyrrolo[2,1-a]isoquinolines 3bd in 50–64% yields (Scheme 1). However, the presence of substituents in the phenyl radical at C-1 affected both the composition and the ratio of the reaction mixtures. Thus, isoquinolines 1ef containing electron-donating substituents (-OMe and -F) in the para-position in the phenyl ring provided pyrrolo[2,1-b][3]benzazepines 4 and 5; no traces of lactones were observed (Scheme 2). We have already published a paper describing the construction of the pyrrolo[2,1-b][3]azepines scaffold via [3,3]-sigmatropic rearrangement in vinyl- and ethynyl-substituted di(tetra)hydroisoquinolines [18], but again the present version of the reaction stood out due to its simplicity and mild reaction conditions. para-Nitrophenyl-substituted isoquinoline 1g produced a mixture of products, consisting of pyrido[2,1-a]isoquinoline 2g (47%), 1-ylidene pyrrolo[2,1-a]isoquinoline 6b (19%) and lactone 3g (13%) (Scheme 1).
The structure of 1-ylidene pyrrolo[2,1-a]isoquinoline 6b was assigned on the basis of NOESY, HMQC and HMBC spectra (Figure 2, see Supporting Information, Figures S1–S3). The NOESY spectrum has correlations between H-1 and H-3 in the pyrrole cycle as well as between H-1 and H-5 and H-10 in the isoquinoline moiety. In the HMBC spectrum, there are correlations between H-1 and C-1, C-3, C-10b in the pyrrole cycle; C-5, C-2 in the ester group; and C-6 in the aryl substituent.
Catalytic routes towards lactones where HFIP facilitates the formation of the products [26,27] which are known in the literature. We believe that the transformation commences with the HFIP-assisted polarization of the enamine moiety (Scheme 3). The subsequent formation of the pyrrole ring (A) followed by the migration of a proton from the solvent to the anionic center of the ylidene fragment results in an intermediate (B). The following [1,3]-shift gives cation (C) in which a Wagner–Meerwein rearrangement occurs to furnish the intermediate (D). The final lactonization of the latter leads to the formation of the target products 3.
The ambiguous behavior of isoquinolines in HFIP in the presence of 3.0 equiv of acetic acid returned us to the idea of carrying out these reactions without any additives. At 20 °C in HFIP, isoquinolines 1ag formed multicomponent mixtures, from which the products were isolated using column chromatography. As was expected in the case of the starting compounds 1bd with isopropyl, benzyl and phenyl substituents, the yields of lactones 3 decreased (Scheme 1). But again, the isoquinolines 1eg decorated with para-OMe, para-F and para-NO2 phenyl radicals at C-1 stood out from the general scheme. Now, isoquinolines 1ef having electron-donating groups demonstrated the highest yields of the desired lactone 3. The formation of lactone 3e was accompanied by the formation of product 6a—1-ylidene-substituted pyrrolo[2,1-a]isoquinolines—with a 15% yield (Scheme 1). In the case of isoquinoline 1g with the para-NO2 phenyl radical, we did not find the corresponding lactone 3g; from the reaction mixture we obtained, pyrido[2,1-a]isoquinoline 2g and 1-ylidene-pyrrolo[2,1-a]isoquinoline 6b were isolated with 43% and 31% yields, respectively (Scheme 1).

3. Materials and Methods

3.1. General Information

IR spectra were recorded on an Infralum FT-801 FTIR spectrometer on KBr tablets for crystalline compounds or on a film for amorphous compounds (ISP SB RAS, Novosibirsk, Russia). 1H and 13C NMR spectra were acquired on a 600 MHz NMR spectrometer (JEOL Ltd., Tokyo, Japan) from CDCl3 to acquire compounds with a solvent signal as the internal standard (7.27 ppm for 1H nuclei, 77.2 ppm for 13C nuclei); peak positions were given in parts per million (ppm, δ). Multiplicities are indicated by s (singlet), d (doublet), t (triplet), m (multiplet). Coupling constants, J, are reported in Hertz. HRMS spectra were recorded on an AB SCIEX TripleTOF 5600+ mass-spectrometer (AB Sciex Pte. Ltd., Singapore) using electrospray ionization (ESI). The measurements were conducted in a positive-ion-mode mass range from m/z 100 to 1000. A syringe injection was used for solutions in MeOH (concentration 100 ng/mL, flow rate 100 μL/min). Melting points were determined on SMP-10 apparatus (Bibby Sterilin Ltd., Stone, UK) with open capillary tubes. Sorbfil PTH-AF-A-UF plates (Imid Ltd., Krasnodar, Russia) were used for TLC; visualization was carried out in an iodine chamber or using KMnO4 and H2SO4 solutions. Silica gel (40–60 μm, 60 Å) from Macherey-Nagel GmbH&Co (Loughborough, UK) was used for column chromatography. All reagents (Sigma-Aldrich, St. Louis, MO, USA; Merck, Darmstadt, Germany; J.T. Baker, Phillipsburg, NJ, USA) were used without additional purification. Compounds 1af, 2af and 4 were also prepared earlier according to the described procedures [18].
Deposition Number 2156399 (for 3a) contains the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service (see Supporting Information, Table S1).

3.2. General Procedure for the Synthesis of Compound 1g

Methyl propiolate (3.0 mmol) was added to the solution of corresponding isoquinoline (1.0 mmol) in 7 mL of CH2Cl2. The reaction was carried out at room temperature. The progress of the reaction was monitored by TLC (Sorbfil, EtOAc-hexane, 1:1). The solvent was removed under reduced pressure; in the case of compound 1g, the residue was purified by column chromatography on silica gel (1:5 EtOAc–hexane).
  • Methyl (2E)-3-[6,7-dimethoxy-1-(3-methoxy-3-oxoprop-1-yn-1-yl)-1-(4-nitrophenyl)-3,4-dihydroisoquinolin-2(1H)-yl]prop-2-enoate (1g). Yield 0.397 g (83%), yellow oil. IR spectrum (KBr), υ/cm−1: 2231 (C
  • C), 1717 (C=O), 1519, 1349 (NO2). 1H NMR (600 MHz, CDCl3) δ 8.23–8.21 (m, 2H, H-Ar), 7.68–7.66 (m, 2H, H-Ar), 7.36 (d, J = 13.6 Hz, 1H, -CH =
  • CH-CO2Me), 6.65 (s, 1H, 8-CH), 6.39 (s, 1H, 5-CH), 4.94 (d, J = 13.6 Hz, 1H, -CH =
  • CH-CO2Me), 3.88 (s, 3H, OCH3), 3.82 (s, 3H, OCH3), 3.67 (s, 3H, OCH3), 3.66–3.63 (m, 1H, 3-CH2), 3.62 (s, 3H, OCH3), 3.49–3.46 (m, 1H, 3-CH2), 3.09–3.05 (m, 1H, 4-CH2), 2.95–2.92 (m, 1H, 4-CH2). 13C NMR (150 MHz, CDCl3) δ 169.0, 153.4, 149.3, 149.0, 148.8, 148.5, 148.0, 128.6 (2C), 127.2, 125.6, 124.3 (2C), 111.2, 111.1, 92.8, 84.7, 80.5, 64.3, 56.2, 56.1, 53.2, 51.1, 42.8, 27.9. HRMS (ESI) m/z calc’d for C25H24N2O8 [M+H]⁺ 481.1605, found: 481.1605 (0.0 ppm).

3.3. General Procedure for the Synthesis of Compounds 2ag

Isoquinolines 1ag (0.3 mmol) were dissolved in 2,2,2-trifluoroethanol (7 mL). The reaction was carried at room temperature. The progress of the reaction was monitored by TLC (Sorbfil, EtOAc-hexane, 1:1). The solvent was removed under reduced pressure, the residue was crystallized from Et2O to produce compounds 2a, 2cf; in the case of compounds 2b and 2g, the residue was purified by column chromatography on silica gel (1:5 EtOAc–hexane). Yields of 2af in 2,2,2-trifluoroethanol: 2a (95%), 2b (55%), 2c (56%), 2d (71%), 2e (79%), 2f (80%). The spectral data for compounds 2af are similar to those previously obtained and reported in [18].
  • Dimethyl 11b-(4-nitrophenyl)-9,10-dimethoxy-7,11b-dihydro-6H-pyrido[2,1-a]isoquinoline-2,3-dicarboxylate (2g). Yield 0.098 g (68%), light yellow oil. IR spectrum (KBr), υ/cm−1: 1688 (C=O), 1519, 1347 (NO2). 1H NMR (600 MHz, CDCl3) δ 8.10 (d, J = 8.8 Hz, 2H, H-Ar), 7.81 (s, 1H, 4-CH), 7.54 (s, 1H, 1-CH), 7.27–7.25 (m, 2H, H-Ar), 7.01 (s, 1H, 11-CH), 6.67 (s, 1H, 8-CH), 3.90 (s, 3H, OCH3), 3.76 (s, 6H, 2*OCH3), 3.60–3.56 (m, 1H, 6-CH2), 3.39 (s, 3H, OCH3), 3.34–3.30 (m, 1H, 6-CH2), 3.00–2.96 (m, 1H, 7-CH2), 2.80–2.77 (m, 1H, 7-CH2). 13C NMR (150 MHz, CDCl3) δ 167.0, 164.3, 161.2, 156.0, 149.1, 147.8, 147.6, 147.2, 129.4 (2C), 126.2, 126.1, 123.1 (2C), 112.3, 111.3, 105.7, 104.1, 78.7, 56.1, 55.9, 51.1, 51.0, 42.4, 29.2. HRMS (ESI) m/z calc’d for C25H24N2O8 [M+Na]⁺ 503.1425, found: 503.1421 (−0.8 ppm).

3.4. General Procedure for the Synthesis of Compounds 3ag, 4, 5a,b and 6a,b

(A) Isoquinoline 1 (0.3 mmol) was dissolved in 7 mL HFIP. The reaction was carried out at room temperature. The progress of the reaction was monitored by TLC (Sorbfil, EtOAc-hexane, 1:1). The solvent was removed under reduced pressure, the residues were chromatographed on silica gel (1:3 EtOAc–hexane) to obtained compounds 3ag and 6a,b.
(B) To a solution of isoquinoline 1 (0.3 mmol) in 7 mL HFIP, glacial AcOH (0.9 mmol) was added. The reaction was carried at room temperature. The progress of the reaction was monitored by TLC (Sorbfil, EtOAc-hexane, 1:1). The solvent was removed under reduced pressure; compounds 3ag, 4, 5a,b and 6b were chromatographed on silica gel (1:5 EtOAc–hexane (for 4 and 6a,b); 1:3 EtOAc–hexane (for 3ag, 5a,b)).
  • Methyl 10,11-dimethoxy-3a-methyl-2-oxo-3,3a,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3a). Yield 0.059 g (55%), white solid, mp 210–212 °C. IR spectrum (KBr), υ/cm−1: 1764, 1680 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.19 (s, 1H, 5-CH), 6.69 (s, 1H, H-Ar), 6.60 (s, 1H, H-Ar), 3.90 (s, 3H, OCH3), 3.87 (s, 3H, OCH3), 3.70 (s, 3H, OCH3), 3.68–3.60 (m, 2H, 7-CH2), 3.52 (d, J = 18.2 Hz, 1H, 3-CH2), 2.90 (d, J = 18.2 Hz, 1H, 3-CH2), 2.90–2.85 (m, 1H, 8-CH2), 2.74–2.70 (m, 1H, 8-CH2), 1.03 (s, 3H, CH3). 13C NMR (150 MHz, CDCl3) δ 174.9, 164.6, 150.0, 148.3, 146.5, 129.0, 122.4, 111.5, 109.2, 108.4, 104.9, 56.4, 56.0, 54.1, 50.8, 42.8, 40.5, 29.8, 21.6. HRMS (ESI) m/z calc’d for C₁₉H2₁NO6 [M+H]⁺ 360.1442, found: 360.1451 (2.5 ppm).
  • Methyl 10,11-dimethoxy-2-oxo-3a-(propan-2-yl)-3,3a,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3b). Yield 0.081 g (50%), white solid, mp 237–239 °C. IR spectrum (KBr), υ/cm−1: 1751, 1675 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.35 (s, 1H, 5-CH), 6.67 (s, 1H, H-Ar), 6.62 (s, 1H, H-Ar), 3.90 (s, 3H, OCH3), 3.88 (s, 3H, OCH3), 3.79 (d, J = 17.9 Hz, 1H, 3-CH2), 3.68 (s, 3H, OCH3), 3.66–3.64 (m, 2H, 7-CH2), 2.98–2.93 (m, 1H, 8-CH2), 2.95 (d, J = 17.9 Hz, 1H, 3-CH2), 2.75–2.71 (m, 1H, 8-CH2), 1.85–1.79 (m, 1H, CH(CH3)2), 0.98 (d, J = 6.7 Hz, 3H, CH(CH3)2), 0.43 (d, J = 6.7 Hz, 3H, CH(CH3)2). 13C NMR (150 MHz, CDCl3) δ 174.6, 165.3, 150.1, 148.7, 148.1, 128.9, 122.4, 111.5, 109.7, 105.1, 102.2, 60.8, 56.4, 56.0, 50.7, 42.4, 39.7, 34.3, 29.3, 20.5, 16.4. HRMS (ESI) m/z calc’d for C2₁H25NO6 [M+H]⁺ 388.1755, found: 388.1765 (2.6 ppm).
  • Methyl 3a-benzyl-10,11-dimethoxy-2-oxo-3,3a,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3c). Yield 0.083 g (64%), white solid, mp 218–220 °C. IR spectrum (KBr), υ/cm−1: 1762, 1676 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.07 (t, J = 7.6 Hz, 1H, H-Ph), 7.04 (s, 1H, 5-CH), 6.95 (t, J = 7.6 Hz, 2H, H-Ph), 6.71 (s, 1H, H-Ar), 6.59 (s, 1H, H-Ar), 6.25 (d, J = 7.6 Hz, 2H, H-Ph), 3.95 (s, 3H, OCH3), 3.93 (s, 3H, OCH3), 3.78 (s, 3H, OCH3), 3.65 (d, J = 18.2 Hz, 1H, 3-CH2), 3.46–3.42 (m, 1H, 7-CH2), 3.31 (d, J = 14.1 Hz, 1H, -CH2-Ph), 3.30–3.27 (m, 1H, 7-CH2), 3.07 (d, J = 18.2 Hz, 1H, 3-CH2), 2.65 (d, J = 14.1 Hz, 1H, -CH2-Ph), 2.38–2.34 (m, 1H, 8-CH2), 1.85–1.80 (m, 1H, 8-CH2). 13C NMR (150 MHz, CDCl3) δ 174.3, 165.1, 150.5, 148.6, 147.8, 135.6, 130.3, 130.2 (2C), 127.3 (2C), 126.5, 122.5, 111.5, 109.4, 104.7, 104.1, 58.0, 56.6, 56.3, 51.0, 42.3, 41.2, 39.3, 28.9. HRMS (ESI) m/z calc’d for C25H25NO6 [M+H]⁺ 436.1755, found: 436.1757 (0.5 ppm).
  • Methyl 10,11-dimethoxy-2-oxo-3a-phenyl-3,3a,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3d). Yield 0.077 g (61%), white solid, mp 212–214 °C. IR spectrum (KBr), υ/cm−1: 1759, 1679 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.37 (s, 1H, 5-CH), 7.10–7.08 (m, 2H, H-Ph), 7.06–7.04 (m, 1H, H-Ph), 7.02 (d, J = 7.6 Hz, 2H, H-Ph), 6.53 (s, 1H, H-Ar), 6.14 (s, 1H, H-Ar), 3.85–3.81 (m, 1H, 7-CH2), 3.79 (s, 3H, OCH3), 3.78 (br. D, J = 5.0 Hz, 2H, 3-CH2), 3.75–3.73 (m, 1H, 7-CH2), 3.57 (s, 3H, OCH3), 3.54 (s, 3H, OCH3), 3.04–3.00 (m, 1H, 8-CH2), 2.82–2.79 (m, 1H, 8-CH2). 13C NMR (150 MHz, CDCl3) δ 174.1, 164.1, 149.6, 147.7, 146.9, 138.7, 128.3 (3C), 127.5, 126.4 (2C), 122.7, 110.9, 110.6, 109.4, 106.1, 60.8, 56.0, 55.9, 50.8, 42.4, 38.0, 29.1. HRMS (ESI) m/z calc’d for C24H23NO6 [M+H]⁺ 422.1598, found: 422.1604 (1.4 ppm).
  • Methyl 10,11-dimethoxy-3a-(4-methoxyphenyl)-2-oxo-3,3a,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3e). Yield 0.067 g (50%), light yellow solid, mp 196–198 °C. IR spectrum (KBr), υ/cm−1: 1760, 1675 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.34 (s, 1H, 5-CH), 6.92 (d, J = 8.6 Hz, 2H, H-Ar), 6.62 (d, J = 8.6 Hz, 2H, H-Ar), 6.53 (s, 1H, H-Ar), 6.18 (s, 1H, H-Ar), 3.84–3.80 (m, 1H, 7-CH2), 3.81 (s, 3H, OCH3), 3.75 (br. S, 2H, 3-CH2), 3.73-3.71 (m, 1H, 7-CH2), 3.69 (s, 3H, OCH3), 3.58 (s, 6H, 2*OCH3), 3.03–2.98 (m, 1H, 8-CH2), 2.81–2.78 (m, 1H, 8-CH2). 13C NMR (150 MHz, CDCl3) δ 174.3, 164.2, 158.6, 149.6, 147.7, 146.6, 130.7, 128.3, 127.5 (2C), 122.7, 113.6 (2C), 110.9, 110.6, 109.4, 105.9, 60.4, 56.0, 55.9, 55.2, 50.8, 42.4, 38.2, 29.1. HRMS (ESI) m/z calc’d for C25H25NO7 [M+H]⁺ 452.1704, found: 452.1714 (2.2 ppm).
  • Methyl 3ª-(4-fluorophenyl)-10,11-dimethoxy-2-oxo-3,3ª,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3f). Yield 0.080 g (61%), white solid, mp 206–208 °C. IR spectrum (KBr), υ/cm−1: 1771, 1683 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.35 (s, 1H, 5-CH), 6.99–6.97 (m, 2H, H-Ar), 6.80–6.77 (m, 2H, H-Ar), 6.54 (s, 1H, H-Ar), 6.14 (s, 1H, H-Ar), 3.83–3.80 (m, 1H, 7-CH2), 3.81 (s, 3H, OCH3), 3.76 (d, J = 15.7 Hz, 2H, 3-CH2), 3.75–3.71 (m, 1H, 7-CH2), 3.58 (s, 6H, 2*OCH3), 3.03–2.98 (m, 1H, 8-CH2), 2.82–2.79 (m, 1H, 8-CH2). 13C NMR (150 MHz, CDCl3) δ 173.8, 164.1, 161.8 (d, J = 247.1 Hz, 1C), 149.7, 147.9, 146.9, 134.6, 128.4, 128.1 (d, J = 8.1 Hz, 2C), 122.4, 115.2 (d, J = 21.6 Hz, 2C), 111.0, 110.4, 109.3, 105.8, 60.4, 56.0, 55.9, 50.8, 42.4, 38.2, 29.1. HRMS (ESI) m/z calc’d for C24H22FNO6 [M+H]⁺ 440.1504, found: 440.1500 (−0.9 ppm).
  • Methyl 10,11-dimethoxy-3ª-(4-nitrophenyl)-2-oxo-3,3ª,7,8-tetrahydro-2H-furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinoline-4-carboxylate (3g). Yield 0.018 g (13%), yellow solid, mp 147–149 °C. IR spectrum (KBr), υ/cm−1: 1774, 1682 (C=O), 1519, 1347 (NO2). 1H NMR (600 MHz, CDCl3) δ 7.98 (d, J = 8.8 Hz, 2H, H-Ar), 7.43 (s, 1H, 5-CH), 7.22 (d, J = 8.8 Hz, 2H, H-Ar), 6.58 (s, 1H, H-Ar), 6.12 (s, 1H, H-Ar), 3.90–3.83 (m, 3H, 3-CH2, 7-CH2), 3.81 (s, 3H, OCH3), 3.78 (d, J = 17.4 Hz, 1H, 3-CH2), 3.58 (s, 3H, OCH3), 3.56 (s, 3H, OCH3), 3.09–3.04 (m, 1H, 8-CH2), 2.89–2.85 (m, 1H, 8-CH2). 13C NMR (150 MHz, CDCl3) δ 172.8, 163.7, 150.4, 148.1, 147.2, 147.0, 146.1, 128.8, 127.5 (2C), 123.5 (2C), 121.6, 111.3 (2C), 108.8 (2C), 60.8, 56.1, 56.0, 51.1, 42.5, 38.0, 29.0. HRMS (ESI) m/z calc’d for C24H22N2O8 [M+H]⁺ 467.1449, found: 467.1455 (1.3 ppm).
  • Dimethyl 11-hydroxy-8,9-dimethoxy-11-(4-methoxyphenyl)-6,11-dihydro-5H-pyrrolo[2,1-b][3]benzazepine-1,2-dicarboxylate (5a). Yield 0.079 g (55%), orange oil. IR spectrum (KBr), υ/cm−1: 3521 (OH), 1723, 1709 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.61 (s, 1H, 3-CH), 7.17 (s, 1H, 10-CH), 6.98 (d, J = 8.9 Hz, 2H, H-Ar), 6.77 (d, J = 8.9 Hz, 2H, H-Ar), 6.60 (s, 1H, 7-CH), 4.03–3.99 (m, 1H, 5-CH2), 3.93 (s, 3H, OCH3), 3.90 (s, 3H, OCH3), 3.89 (s, 3H, OCH3), 3.88–3.84 (m, 1H, 5-CH2), 3.80 (s, 3H, OCH3), 3.76 (s, 3H, OCH3), 3.64 (s, 1H, OH), 2.98–2.94 (m, 1H, 6-CH2), 2.86–2.82 (m, 1H, 6-CH2). 13C NMR (150 MHz, CDCl3) δ 169.6, 164.1, 159.6, 148.3, 147.6, 138.5, 137.2, 134.0, 128.4 (2C), 127.4, 127.2, 116.9, 114.0 (2C), 113.4, 113.3, 110.7, 77.7, 56.2, 56.1, 55.4, 52.8, 51.6, 48.2, 33.2. HRMS (ESI) m/z calc’d for C26H27NO8 [M+Na]⁺ 504.1629, found: 504.1641 (2.4 ppm).
  • Dimethyl 11-(4-fluorophenyl)-11-hydroxy-8,9-dimethoxy-6,11-dihydro-5H-pyrrolo[2,1-b][3]benzazepine-1,2-dicarboxylate (5b). Yield 0.062 g (44%), orange oil. IR spectrum (KBr), υ/cm−1: 3449 (OH), 1715 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.59 (s, 1H, 3-CH), 7.17 (s, 1H, 10-CH), 7.06–7.03 (m, 2H, H-Ar), 6.94–6.91 (m, 2H, H-Ar), 6.61 (s, 1H, 7-CH), 4.02–3.98 (m, 1H, 5-CH2), 3.93 (s, 3H, OCH3), 3.90 (s, 3H, OCH3), 3.89 (s, 3H, OCH3), 3.87–3.85 (m, 1H, 5-CH2), 3.80 (s, 3H, OCH3), 3.75 (s, 1H, OH), 2.95–2.91 (m, 1H, 6-CH2), 2.87–2.83 (m, 1H, 6-CH2). 13C NMR (150 MHz, CDCl3) δ 169.6, 164.0, 162.6 (d, J = 248.5 Hz, 1C), 148.5, 147.7, 142.2, 136.9, 133.7, 129.1 (d, J = 8.1 Hz, 2C), 127.5, 127.4, 117.0, 115.6 (d, J = 21.6 Hz, 2C), 113.6, 113.4, 110.6, 77.6, 56.2, 56.1, 52.9, 51.6, 48.4, 33.2. HRMS (ESI) m/z calc’d for C25H24FNO7 [M+Na]⁺ 492.1429, found: 492.1434 (1.0 ppm).
  • Methyl (1E)-8,9-dimethoxy-1-(2-methoxy-2-oxoethylidene)-10b-(4-methoxyphenyl)-1,5,6,10b-tetrahydropyrrolo[2,1-a]isoquinoline-2-carboxylate (6a). Yield 0.021 g (15%), beige solid, mp 227–229 °C. IR spectrum (KBr), υ/cm−1: 1721, 1679 (C=O). 1H NMR (600 MHz, CDCl3) δ 7.31 (s, 1H, 3-CH), 7.21 (d, J = 8.8 Hz, 2H, H-Ar), 6.84 (d, J = 8.8 Hz, 2H, H-Ar), 6.60 (s, 1H, H-Ar), 6.42 (s, 1H, H-Ar), 5.60 (s, 1H, =CH-CO2Me), 3.90 (s, 3H, OCH3), 3.82 (s, 3H, OCH3), 3.79 (s, 3H, OCH3), 3.75 (s, 3H, OCH3), 3.74–3.70 (m, 1H, 5-CH2), 3.69 (s, 3H, OCH3), 3.33–3.30 (m, 1H, 5-CH2), 3.12–3.08 (m, 1H, 6-CH2), 2.69–2.66 (m, 1H, 6-CH2). 13C NMR (150 MHz, CDCl3) δ 169.5, 165.7, 159.5, 148.4, 148.3, 146.8, 136.9, 130.6, 129.8 (2C), 129.3, 126.3, 120.0, 113.7 (2C), 111.4, 109.3, 94.4, 64.7, 56.2, 56.1, 55.4, 52.4, 51.0, 48.0, 28.8. HRMS (ESI) m/z calc’d for C26H27NO7 [M+H]⁺ 466.1860, found: 466.1861 (0.2 ppm).
  • Methyl (1E)-8,9-dimethoxy-1-(2-methoxy-2-oxoethylidene)-10b-(4-nitrophenyl)-1,5,6,10b-tetrahydropyrrolo[2,1-a]isoquinoline-2-carboxylate (6b). Yield 0.045 g (31%), orange oil. IR spectrum (KBr), υ/cm−1: 1733, 1699 (C=O), 1518, 1349 (NO2). 1H NMR (600 MHz, CDCl3) δ 8.19–8.17 (m, 2H, H-Ar), 7.51–7.49 (m, 2H, H-Ar), 7.32 (s, 1H, 3-CH), 6.65 (s, 1H, H-Ar), 6.35 (s, 1H, H-Ar), 5.59 (s, 1H, =CH-CO2Me), 3.91 (s, 3H, OCH3), 3.83 (s, 3H, OCH3), 3.75 (s, 3H, OCH3), 3.71 (s, 3H, OCH3), 3.68–3.64 (m, 1H, 5-CH2), 3.41–3.39 (m, 1H, 5-CH2), 3.16–3.11 (m, 1H, 6-CH2), 2.73–2.70 (m, 1H, 6-CH2). 13C NMR (150 MHz, CDCl3) δ 169.0, 165.3, 150.2, 148.9, 148.7, 147.6, 146.8, 130.7, 129.3 (2C), 129.1, 126.4, 123.8 (2C), 118.8, 111.7, 109.0, 95.9, 64.6, 56.3, 56.1, 52.6, 51.2, 48.2, 28.5. HRMS (ESI) m/z calc’d for C25H24N2O8 [M+H]⁺ 481.1605, found: 481.1605 (0.0 ppm).

4. Conclusions

In summary, we have described a novel procedure for the synthesis of lactonic pyrrolo[2,1-a]isoquinolines and pyrido[2,1-a]isoquinolines through the rearrangements of 1-R-1-ethynyl-2-vinyl-1,2,3,4-tetrahydroisoquinolines in fluorinated alcohols. It has been demonstrated that the rearrangements depend on the acidity of the solvents used. In some cases, the addition of 3 equiv of AcOH increased the yields of the target lactones. The substituent at C-1 in the starting isoquinolines affects the composition and the ratio of the products in the transformation occurring in HFIP both with and without AcOH.

Supplementary Materials

The following supporting information (copies of 1H and 13C NMR spectra of the compounds 13,5,6) can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25021085/s1.

Author Contributions

Conceptualization, A.V.V. and A.A.T.; methodology, A.A.T.; investigation, A.Y.O.; resources, V.B.R.; writing—original draft preparation, A.A.T., A.V.L. and A.Y.O.; writing—review and editing, T.N.B. and A.V.V.; visualization, A.Y.O.; supervision, L.G.V. and A.V.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and in the Supplementary Materials.

Acknowledgments

This paper has been supported by the RUDN University Strategic Academic Leadership Program.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hitotsuyanagi, Y.; Fukaya, H.; Takeda, E.; Matsuda, S.; Saishu, Y.; Zhu, S.; Komatsu, K.; Takeya, K. Structures of stemona-amine B and stemona-lactams M–R. Tetrahedron 2013, 69, 6297–6304. [Google Scholar] [CrossRef]
  2. Xu, Y.; Xiong, L.; Yan, Y.; Sun, D.; Duan, Y.; Li, H.; Chen, L. Alkaloids from Stemona Tuberosa and Their Anti-Inflammatory Activity. Front. Chem. 2022, 10, 847595. [Google Scholar] [CrossRef]
  3. Nogawa, T.; Kawatani, M.; Uramoto, M.; Okano, A.; Aono, H.; Futamura, Y.; Takahashi, S.; Osada, H. Pyrrolizilactone, a new pyrrolizidinone metabolite produced by a fungus. J. Antibiot. 2013, 66, 621–623. [Google Scholar] [CrossRef] [PubMed]
  4. Salatino, A.; Salatino, M.L.F.; Negri, G. Traditional uses, chemistry and pharmacology of Croton species (Euphorbiaceae). J. Braz. Chem. Soc. 2007, 18, 11–33. [Google Scholar] [CrossRef]
  5. Farias, R.; Rao, V.; Viana, G.; Silveira, E.; Maciel, M.; Pino, A. Hypoglycemic Effect of Trans-Dehydrocrotonin, a Nor-Clerodane Diterpene from Croton Cajucara. Planta Med. 1997, 63, 558–560. [Google Scholar] [CrossRef]
  6. Rodriguez, J.A.; Haun, M. Cytotoxicity of Trans-Dehydrocrotonin from Croton Cajucara on V79 Cells and Rat Hepatocytes. Planta Med. 1999, 65, 522–526. [Google Scholar] [CrossRef] [PubMed]
  7. Aleman, J.; del Solar, V.; Martin-Santos, C.; Cubo, L.; Ranninger, C.N. Tandem Cyclization–Michael Reaction by Combination of Metal-and Organocatalysis. J. Org. Chem. 2011, 76, 7287–7293. [Google Scholar] [CrossRef]
  8. Herath, H.B.; Herath, W.H.; Carvalho, P.; Khan, S.I.; Tekwani, B.L.; Duke, S.O.; Tomaso-Peterson, M.; Nanayakkara, N.D. Biologically active tetranorditerpenoids from the fungus Sclerotinia homoeocarpa causal agent of dollar spot in turfgrass. J. Nat. Prod. 2009, 72, 2091–2097. [Google Scholar] [CrossRef]
  9. Ivanescu, B.; Miron, A.; Corciova, A. Sesquiterpene lactones from Artemisia genus: Biological activities and methods of analysis. J. Anal. Methods Chem. 2015, 2015, 247685. [Google Scholar] [CrossRef]
  10. Greger, H. Structural classification and biological activities of Stemona alkaloids. Phytochem. Rev. 2019, 18, 463–493. [Google Scholar] [CrossRef]
  11. Pilli, R.A.; Rosso, G.B.; de Oliveira, M.D.C.F. The chemistry of Stemona alkaloids: An update. Nat. Prod. Rep. 2010, 27, 1908–1937. [Google Scholar] [CrossRef]
  12. Li, L.; Tang, M.C.; Tang, S.; Gao, S.; Soliman, S.; Hang, L.; Xu, W.; Ye, T.; Watanabe, K.; Tang, Y. Genome mining and assembly-line biosynthesis of the UCS1025A pyrrolizidinone family of fungal alkaloids. J. Am. Chem. Soc. 2018, 140, 2067–2071. [Google Scholar] [CrossRef]
  13. Bartoli, A.; Rodier, F.; Commeiras, L.; Parrain, J.L.; Chouraqui, G. Construction of spirolactones with concomitant formation of the fused quaternary centre–application to the synthesis of natural products. Nat. Prod. Rep. 2011, 28, 763–782. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, B.; Zhang, Z.; Li, P.; Miao, T.; Wang, L. Synthesis of Spirolactones via a BF3·Et2O-Promoted Cascade Annulation of α-Keto Acids and 1,3-Enynes. Org. Lett. 2021, 23, 5698–5702. [Google Scholar] [CrossRef] [PubMed]
  15. Sun, X.; Zhou, M.; Zhu, J.P.; Zhang, X.F.; Liu, Z.J.; Li, H.R.; Chen, Y.; Chen, H.-P.; Zhao, J.; Pu, J.-X.; et al. An unexpected photoinduced cyclization to synthesize fully substituted γ-spirolactones via intramolecular hydrogen abstraction with allyl acrylates. Org. Chem. Front. 2022, 9, 2316–2321. [Google Scholar] [CrossRef]
  16. Nair, D.; Basu, P.; Pati, S.; Baseshankar, K.; Sankara, C.S.; Namboothiri, I.N. Synthesis of Spirolactones and Functionalized Benzofurans via Addition of 3-Sulfonylphthalides to 2-Formylaryl Triflates and Conversion to Benzofuroisocoumarins. J. Org. Chem. 2023, 88, 4519–4527. [Google Scholar] [CrossRef] [PubMed]
  17. Delayre, B.; Wang, Q.; Zhu, J. Natural product synthesis enabled by domino processes incorporating a 1,2-rearrangement step. ACS Cent. Sci. 2021, 7, 559–569. [Google Scholar] [CrossRef]
  18. Obydennik, A.Y.; Titov, A.A.; Listratova, A.V.; Borisova, T.N.; Sokolova, I.L.; Rybakov, V.B.; Van der Eycken, E.V.; Voskressensky, L.G.; Varlamov, A.V. Divergent and Nucleophile-Assisted Rearrangement in the Construction of Pyrrolo [2,1-b][3]benzazepine and Pyrido[2,1-a]isoquinoline Scaffolds. Chem.-Eur. J. 2023, e202302919. [Google Scholar] [CrossRef]
  19. Motiwala, H.F.; Armaly, A.M.; Cacioppo, J.G.; Coombs, T.C.; Koehn, K.R.; Norwood IV, V.M.; Aube, J. HFIP in organic synthesis. Chem. Rev. 2022, 122, 12544–12747. [Google Scholar] [CrossRef] [PubMed]
  20. Listratova, A.V.; Titov, A.A.; Obydennik, A.Y.; Varlamov, A.V. N-propargyl aza-Claisen rearrangement in the synthesis of heterocycles. Tetrahedron 2022, 121, 132914. [Google Scholar] [CrossRef]
  21. Westermaier, M.; Mayr, H. Regio- and Stereoselective Ring-Opening Reactions of Epoxides with Indoles and Pyrroles in 2,2,2-Trifluoroethanol. Chem.-Eur. J. 2008, 14, 1638–1647. [Google Scholar] [CrossRef] [PubMed]
  22. Li, G.X.; Qu, J. Friedel–Crafts alkylation of arenes with epoxides promoted by fluorinated alcohols or water. Chem. Commun. 2010, 46, 2653–2655. [Google Scholar] [CrossRef] [PubMed]
  23. Lee, J.Y.; Lee, Y.S.; Chung, B.Y.; Park, H. Asymmetric synthesis of both enantiomers of novel tetracyclic heterocycle, furo [3′,2′:2,3]pyrrolo[2,1-a] isoquinoline derivative via a diastereoselective N-acyliminium ion cyclization. Tetrahedron 1997, 53, 2449–2458. [Google Scholar] [CrossRef]
  24. Titov, A.A.; Kobzev, M.S.; Borisova, T.N.; Listratova, A.V.; Evenko, T.V.; Varlamov, A.V.; Voskressensky, L.G. Facile Methods for the Synthesis of 8-Ylidene-1,2,3,8-tetrahydrobenzazecines. Eur. J. Org. Chem. 2020, 2020, 3041–3049. [Google Scholar] [CrossRef]
  25. Titov, A.A.; Purgatorio, R.; Obydennik, A.Y.; Listratova, A.V.; Borisova, T.N.; De Candia, M.; Catto, M.; Altomare, C.D.; Varlamov, A.V.; Voskressensky, L.G. Synthesis of Isomeric 3-Benzazecines Decorated with Endocyclic Allene Moiety and Exocyclic Conjugated Double Bond and Evaluation of Their Anticholinesterase Activity. Molecules 2022, 27, 6276. [Google Scholar] [CrossRef]
  26. Dantignana, V.; Milan, M.; Cussó, O.; Company, A.; Bietti, M.; Costas, M. Chemoselective Aliphatic C–H Bond Oxidation Enabled by Polarity Reversal. ACS Cent. Sci. 2017, 3, 1350–1358. [Google Scholar] [CrossRef]
  27. Call, A.; Cianfanelli, M.; Besalú-Sala, P.; Olivo, G.; Palone, A.; Vicens, L.; Ribas, X.; Luis, J.M.; Bietti, M.; Costas, M. Carboxylic Acid Directed γ-Lactonization of Unactivated Primary C–H Bonds Catalyzed by Mn Complexes: Application to Stereoselective Natural Product Diversification. J. Am. Chem. Soc. 2022, 144, 19542–19558. [Google Scholar] [CrossRef]
Figure 1. Biologically active natural lactone and spirolactone molecules.
Figure 1. Biologically active natural lactone and spirolactone molecules.
Ijms 25 01085 g001
Scheme 1. Synthesis of 3ag in HFIP in the presence of glacial AcOH. a Reaction conditions: a mixture of 1ag (0.3 mmol), glacial AcOH (0.9 mmol, 3.0 equiv) in HFIP (7.0 mL) was stirred at rt. Formation of pyrrolo[2,1-b][3]benzazepines 4, 5 instead furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinolines 3ef. c Pyrido[2,1-a]isoquinoline 2g (47%) and 1-ylidene-pyrrolo[2,1-a]isoquinoline 6b (19%) were isolated in addition of 3g. Formation of mixture pyrido[2,1-a]isoquinoline 2g (43%) and 1-ylidene-pyrrolo[2,1-a]isoquinoline 6b (31%).
Scheme 1. Synthesis of 3ag in HFIP in the presence of glacial AcOH. a Reaction conditions: a mixture of 1ag (0.3 mmol), glacial AcOH (0.9 mmol, 3.0 equiv) in HFIP (7.0 mL) was stirred at rt. Formation of pyrrolo[2,1-b][3]benzazepines 4, 5 instead furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinolines 3ef. c Pyrido[2,1-a]isoquinoline 2g (47%) and 1-ylidene-pyrrolo[2,1-a]isoquinoline 6b (19%) were isolated in addition of 3g. Formation of mixture pyrido[2,1-a]isoquinoline 2g (43%) and 1-ylidene-pyrrolo[2,1-a]isoquinoline 6b (31%).
Ijms 25 01085 sch001
Scheme 2. Transformations of isoquinolines 1ef.
Scheme 2. Transformations of isoquinolines 1ef.
Ijms 25 01085 sch002
Figure 2. Key correlations in the 2D NOESY (blue), ¹H-¹³C HMQC (red) and ¹H-¹³C HMBC (purple) of compound 6b.
Figure 2. Key correlations in the 2D NOESY (blue), ¹H-¹³C HMQC (red) and ¹H-¹³C HMBC (purple) of compound 6b.
Ijms 25 01085 g002
Scheme 3. Plausible reaction mechanism for the formation of furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinolines 3ag from 1ag.
Scheme 3. Plausible reaction mechanism for the formation of furo[2′,3′:2,3]pyrrolo[2,1-a]isoquinolines 3ag from 1ag.
Ijms 25 01085 sch003
Table 1. Synthesis of pyrido[2,1-a]isoquinolines 2ag.
Table 1. Synthesis of pyrido[2,1-a]isoquinolines 2ag.
Ijms 25 01085 i001
EntryRProductYield, %
1Me2a95 a
2i-Pr2b55
3Bn2c56
4Ph2d71
54-OMe-C6H4-2e79
64-F-C6H4-2f80
74-NO2-C6H4-2g68
a 78% for reaction in C2H5OH.
Table 2. Optimization of the reaction conditions.
Table 2. Optimization of the reaction conditions.
Ijms 25 01085 i002
Entryglacial AcOH (Equiv.)Yield 3a, %Yield 2a, %
1-2571
20.54343
33.055- a
45.056- a
a no traces of pyridoisoquinoline. Optimal conditions are bold in the table.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Obydennik, A.Y.; Titov, A.A.; Listratova, A.V.; Borisova, T.N.; Rybakov, V.B.; Voskressensky, L.G.; Varlamov, A.V. Concise and Free-Metal Access to Lactone-Annelated Pyrrolo[2,1-a]isoquinoline Derivatives via a 1,2-Rearrangement Step. Int. J. Mol. Sci. 2024, 25, 1085. https://doi.org/10.3390/ijms25021085

AMA Style

Obydennik AY, Titov AA, Listratova AV, Borisova TN, Rybakov VB, Voskressensky LG, Varlamov AV. Concise and Free-Metal Access to Lactone-Annelated Pyrrolo[2,1-a]isoquinoline Derivatives via a 1,2-Rearrangement Step. International Journal of Molecular Sciences. 2024; 25(2):1085. https://doi.org/10.3390/ijms25021085

Chicago/Turabian Style

Obydennik, Arina Y., Alexander A. Titov, Anna V. Listratova, Tatiana N. Borisova, Victor B. Rybakov, Leonid G. Voskressensky, and Alexey V. Varlamov. 2024. "Concise and Free-Metal Access to Lactone-Annelated Pyrrolo[2,1-a]isoquinoline Derivatives via a 1,2-Rearrangement Step" International Journal of Molecular Sciences 25, no. 2: 1085. https://doi.org/10.3390/ijms25021085

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