Next Article in Journal
[6-(Thiophen-2-yl)-2,2′-bipyridine]bis(triphenylphosphine) Copper(I) Tetrafluoroborate
Previous Article in Journal
6-[(2S,3R)-3-(2,4-Difluorophenyl)-3-hydroxy-4-(1H-1,2,4-triazol-1-yl)butan-2-yl]-5-fluoropyrimidine-4-carbaldehyde
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis of 2-[2-(Ethoxymethoxy)phenyl]spiro[cyclopropane-1,2′-indene]-1′,3′-dione

1
Department of Chemistry, M.V. Lomonosov Moscow State University, Leninskie Gory 1-3, Moscow 119991, Russia
2
N.D. Zelinsky Institute of Organic Chemistry Russian Academy of Sciences, Leninsky Pr. 47, Moscow 119991, Russia
3
Dmitry Rogachev National Medical Research Center of Pediatric Hematology, Oncology and Immunology, Samory Mashela St. 1, Moscow 117997, Russia
*
Author to whom correspondence should be addressed.
Molbank 2023, 2023(1), M1604; https://doi.org/10.3390/M1604
Submission received: 13 February 2023 / Revised: 9 March 2023 / Accepted: 11 March 2023 / Published: 14 March 2023
(This article belongs to the Section Organic Synthesis)

Abstract

:
An 1,3-indanedione-derived donor–acceptor cyclopropane, bearing the ethoxymethyl-protected phenolic group at the ortho-position of the donor aryl substituent, has been synthesized using a reaction sequence involving the Knoevenagel condensation of 1,3-indanedione with the corresponding protected salicylaldehyde followed by the Corey–Chaykovsky cyclopropanation of the obtained adduct with dimethylsulfoxonium methylide. The structure of the synthesized cyclopropane was unambiguously proved by single-crystal X-ray diffraction data.

Graphical Abstract

1. Introduction

Cyclopropanes, bearing donor and acceptor substituents at vicinal carbon atoms have attracted considerable attention from organic chemists in recent decades due to their high reactivity with respect to various classes of reagents, such as nucleophiles, electrophiles, diverse compounds with multiple carbon–carbon, carbon–heteroatom, and heteroatom–heteroatom bonds, 1,3-dipoles, 1,3-dienes, etc. Because of this unique reactivity, such substrates have become known as donor–acceptor (DA) cyclopropanes [1,2,3,4,5,6,7]. Typically, DA cyclopropanes react as synthetic equivalents of a 1,3-dipole, in which the carbon atom connected to the donor group serves as the electrophile, while the carbon atom bonded to the acceptor group(s) appears for the nucleophilic center. However, we and others have shown that DA cyclopropanes, in which the donor is an electron-rich (hetero)aromatic group, can also react with the involvement of the ortho-position of the aromatic group in the process as the nucleophile, affording various annulation products [5,8,9,10,11,12].
Further extension of the multifaceted reactivity of DA cyclopropanes can be achieved by using substrates in which the donor aryl substituent contains a reactive functional group at the ortho-position to the three-membered ring [13,14,15,16,17,18,19,20,21,22,23,24,25]. For example, domino transformations of 2-hydroxyaryl-derived DA cyclopropanes, including the small ring opening and a new ring closure with the participation of phenolic oxygen, afforded various heterocyclic products [13,14,15]. However, the study of the reactivity of such cyclopropanes has usually been limited to the corresponding 2-(2-hydroxyaryl)cyclopropane-1,1-diesters. Therefore, the synthesis of related substrates with other acceptor substituents should be important for the development of original processes leading to new polycyclic products. Herein, we describe a simple approach to 1,3-indanedione-derived DA cyclopropane bearing a protected phenolic moiety at the ortho-position of the donor phenyl group.

2. Results and Discussion

The title cyclopropane was synthesized from the commercial starting compound by simple procedures. Salicylaldehyde 1 was protected with ethoxymethyl chloride (EOMCl) according to the procedure previously described [14]. The Knoevenagel reaction of the resulting product 2 with 1,3-indanedione was carried out similarly to related condensations of other aldehydes, which were previously reported [26]. The cyclopropanation of the synthesized alkene 3 to obtain the cyclopropane 4 was performed by adding dimethylsulfoxonium methylide, generated by the treatment of trimethylsulfoxonium iodide with sodium hydride, to the solution of 3 in DMF at 0 °C (Scheme 1). It is worth noting that both the order of addition and the reaction temperature are crucial to the efficiency of this process. The addition of the solution of alkene 3 to the generated ylide led to the formation of the complex mixture of products. Similarly, the yield of the desired compound 4 was much lower due to the formation of various side products when cyclopropanation was performed at room temperature.
1H NMR spectrum of compound 4 contains three doublets of doublets in the upfield region. These signals correspond to protons of the three-membered ring. Geminal coupling constant 2J for CH2 protons is 4.2 Hz, which is typical (from 4 to 5 Hz) for the methylene group of cyclopropane. The structure of indanedione-derived cyclopropane 4 was unambiguously confirmed by single-crystal X-ray analysis (Figure 1). This structure closely matches the structures of the related DA cyclopropane without ortho-substituent [26]. The small difference is in some rotation of the phenyl group, providing the most efficient interaction of ether oxygen with the closest carbonyl group. This interaction is possibly responsible for some shortening of the C(1)–C(2) bond between the atoms linked to donor and acceptor substituents in 4 (1.542 Å) compared to the analogue without the 2-ethoxymethoxy group (1.561 Å, [26]). NMR and IR spectral data for compound 4 resemble the corresponding data for other 1,3-indanedione-derived DA cyclopropanes [26,27,28]. For copies of spectra, see Supplementary Materials.

3. Materials and Methods

NMR spectra were acquired on a Bruker Avance 400 spectrometer (Bruker, Billerica, MA, USA) at room temperature; the chemical shifts δ were measured in ppm with respect to the solvent (1H: CDCl3, δ = 7.26 ppm; 13C: CDCl3, δ = 77.00). The splitting patterns were designated as s, singlet; d, doublet; m, multiplet; dd, double doublet; br., broad. The coupling constants (J) were in Hertz. Infrared spectra were recorded on an Infralum FT-801 spectrometer (Simex, Novosibirsk, Russian Federation) and a Thermo Nicolet IR-200 FT-IR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). High resolution and accurate mass measurements were carried out using a micrOTOF-QTM ESI-TOF (Electro Spray Ionization/Time of Flight, Bruker, Billerica, MA, USA). The melting points (m.p.) were determined using a 9100 capillary melting point apparatus (Electrothermal, Stone, UK). Analytical thin layer chromatography (TLC) was carried out with silica gel plates (silica gel 60, F254, supported on aluminum); the revelation was conducted by UV lamp (365 nm). Column chromatography was performed on silica gel 60 (230–400 mesh, Merck, Darmstadt, Germany). All reactions were carried out using freshly distilled and dry solvents. Commercial reagents employed in the synthesis were analytical grade, obtained from Aldrich (St. Louis, MI, USA) or Alfa Aesar (Ward Hill, MO, USA). CCDC 1889833 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (accessed on 9 September 2019) (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44-1223-336033; E-mail: [email protected]). Compound 2 was synthesized by the reported procedure [14].

3.1. 2-[2-(Ethoxymethoxy)benzylidene]indene-1,3(2H)-dione (3)

To the solution of 2-(ethoxymethoxy)benzaldehyde 2 (2.0 g, 11.1 mmol) and indane-1,3-dione (1.76 g, 12.0 mmol) in benzene (12 mL), piperidine (0.11 mL, 1.1 mmol) and acetic acid (126 μL, 2.2 mmol) were added. The mixture was refluxed with the Dean–Stark trap until water separation was finished (4 h). Upon cooling, the precipitate was formed. It was filtered and recrystallized by the dissolution of product in the minimal quantity of the boiling mixture of petroleum ether and ethyl acetate (4:1) followed by cooling the solution to between 0 and 5 °C. Product 3 was isolated as a brown solid in 68% yield (2.33 g); m.p. = 105–106 °C.
1H NMR (CDCl3, 400 MHz): δ = 1.25 (t, 3J = 7.1 Hz, 3H, CH3), 3.80 (q, 3J = 7.1 Hz, 2H, CH2), 5.38 (s, 2H, OCH2O), 7.13–7.16 (m, 1H, Ar), 7.25 (dd, 3J = 8.3 Hz, 4J = 0.8 Hz, 1H, Ar), 7.49–7.53 (m, 1H, Ar), 7.80–7.82 (m, 2H, Ar), 8.00–8.03 (m, 2H, Ar), 8.51 (s, 1H, CH=), 8.89 (dd, 3J = 8.0 Hz, 4J = 1.6 Hz, 1H, Ar). 13C NMR (CDCl3, 100 MHz): δ = 15.3 (CH3), 65.0 (CH2O), 93.7 (OCH2O), 114.5 (CH), 121.6 (CH), 122.8 (C), 123.3 (CH), 123.4 (CH), 128.7 (C), 134.0 (CH), 135.2 (CH), 135.36 (CH), 135.38 (CH), 140.3 (C), 141.6 (CH), 142.6 (C), 158.8 (C), 189.3 (CO), 190.7 (CO). IR (KBr): ν = 3088, 2985, 2908, 1726, 1687, 1608, 1582, 1480, 1459, 1373, 1347, 1216, 1199, 1171, 1150, 1114, 1075, 1017, 974 cm−1. HRMS ESI-TOF: m/z = 309.1126 [M + H]+ (309.1121 calculated for C19H17O4+).

3.2. 2-[2-(Ethoxymethoxy)phenyl]spiro[cyclopropane-1,2′-indene]-1′,3′-dione (4)

To the suspension of NaH (60% suspension in mineral oil, 62 mg, 1.55 mmol) in dry DMF (2 mL), trimethylsulfoxonium iodide (341 mg, 1.55 mmol) was added in a single portion under argon atmosphere at room temperature. The reaction mixture was stirred for 45 min. The obtained solution was added dropwise to the solution of alkene 3 (400 mg, 1.30 mmol) in DMF (4 mL) at 0 °C under argon atmosphere. The reaction mixture was stirred at 0 °C for 50 min and poured into the mixture of the saturated aq. NH4Cl solution and ice (10 mL). The product was extracted with ethyl acetate (5 × 5 mL) and the combined organic layers were dried with Na2SO4. The solvent was evaporated under the reduced pressure. The residue was purified by column chromatography on silica gel using the mixture of petroleum ether and ethyl acetate (1:1) as an eluent. Cyclopropane 4 was isolated as yellow crystals in 73% yield (305 mg). Rf = 0.44 (petroleum ether-ethyl acetate, 4:1); m.p. = 85–86 °C. 1H NMR (CDCl3, 400 MHz): δ = 0.93 (t, 3J = 7.3 Hz, 3H, CH3), 2.30 (dd, 2J = 4.2 Hz, 3J = 8.7 Hz, 1H, CH2), 2.37 (dd, 2J = 4.2 Hz, 3J = 9.0 Hz, 1H, CH2), 3.11 (dq, 2J = 9.6 Hz, 3J = 7.3 Hz, 1H, CH2), 3.17 (dq, 2J = 9.6 Hz, 3J = 7.3 Hz, 1H, CH2), 3.31 (dd, 3J = 9.0 Hz, 3J = 8.7 Hz, 1H, CH2), 4.76 (d, 2J = 7.0 Hz, 1H, OCH2O), 4.86 (d, 2J = 7.0 Hz, 1H, OCH2O), 6.98 (br. d, 3J = 8.3 Hz, 1H, Ar); 7.01–7.04 (m, 1H, Ar), 7.22–7.26 (m, 1H, Ar), 7.36 (br. d, 3J = 7.3 Hz, 1H, Ar); 7.72–7.80 (m, 3H Ar), 7.99 (br. d, 3J = 7.4 Hz, 1H, Ar). 13C NMR (CDCl3, 100 MHz): δ = 14.9 (CH3), 21.7 (CH2), 36.9 (CH), 41.7 (CH), 63.8 (CH2O), 92.7 (OCH2O), 113.1 (CH), 121.3 (CH), 122.2 (CH), 122.4 (CH), 123.6 (C), 129.3 (CH), 130.0 (CH), 134.4 (CH), 134.6 (CH), 141.7 (C), 142.7 (C), 156.3 (C), 195.7 (CO), 198.9 (CO). IR (KBr): ν = 2970, 2904, 2864, 1732, 1703, 1599, 1491, 1454, 1429, 1379, 1335, 1313, 1234, 1163, 1120, 1097, 1051, 995 cm−1. HRMS ESI-TOF: m/z = 323.1277 [M + H]+. (323.1278 calculated for C20H19O4+). Crystal Data for C20H18O4 (M = 322.34 g/mol): triclinic, space group P-1 (no. 2), a = 7.9274(4) Å, b = 8.4765(4) Å, c = 14.3237(8) Å, α = 77.297(4)°, β = 75.805(4)°, γ = 63.850(4)°, V = 830.66(8) Å3, Z = 2, T = 295 K, μ(CuKα) = 0.729 mm−1, Dcalc = 1.289 g/cm3, 8,517 reflections measured (3.930° ≤ Θ ≤ 70.282°), 2988 unique (Rint = 0.0315, Rsigma = 0.0252), which were used in all calculations. The final R1 was 0.0355 (I > 2σ (I)) and wR2 was 0.0989 (all data).

4. Conclusions

Here, we have described the synthesis of donor–acceptor cyclopropane, which may participate in various domino reactions due to the presence of the 1,3-indanedione moiety as an acceptor group, the effect of spiro activation, as well as the possible participation of an additional functionality in the donor substituent. The reactivity of this substrate is under investigation.

Supplementary Materials

The following supporting information can be downloaded. Figure S1: 1H NMR (CDCl3, 400 MHz) spectrum of compound 3; Figure S2: 13C NMR (CDCl3, 100 MHz) spectrum of compound 3; Figure S3: HRMS spectrum of compound 3; Figure S4: IR spectrum of compound 3; Figure S5: 1H NMR (CDCl3, 400 MHz) of compound 4; Figure S6: 13C NMR (CDCl3, 100 MHz) of compound 4; Figure S7: HRMS spectrum of compound 4; Figure S8: IR spectrum of 4.

Author Contributions

Conceptualization, O.A.I. and I.V.T.; methodology, I.V.T.; software, I.A.A. and N.K.R.; validation, O.A.I. and I.V.T.; formal analysis, I.A.A.; investigation, V.V.S., N.K.R., V.B.R. and E.D.S.; resources, I.V.T.; data curation, O.A.I.; writing—original draft preparation, I.V.T.; writing—review and editing, O.A.I. and I.V.T.; supervision, I.V.T.; project administration, I.V.T.; funding acquisition, I.V.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Russian Science Foundation (grant 21-13-00395).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Any data can be obtained from authors by request.

Acknowledgments

The X-ray studies were fulfilled using a STOE STADIVARI PILATUS 100-K diffractometer purchased as a part of the MSU development program.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Augustin, A.U.; Werz, D.B. Exploiting Heavier Organochalcogen Compounds in Donor−Acceptor Cyclopropane Chemistry. Acc. Chem. Res. 2021, 54, 1528–1541. [Google Scholar] [CrossRef] [PubMed]
  2. Ghosh, K.; Das, S. Recent advances in ring-opening of donor acceptor cyclopropanes using C-nucleophiles. Org. Biomol. Chem. 2021, 19, 965–982. [Google Scholar] [CrossRef]
  3. Pirenne, V.; Muriel, B.; Waser, J. Catalytic Enantioselective Ring-Opening Reactions of Cyclopropanes. Chem. Rev. 2021, 121, 227–263. [Google Scholar] [CrossRef] [PubMed]
  4. Singh, P.; Varshnaya, R.K.; Dey, R.; Banerjee, P. Donor−Acceptor Cyclopropanes as an Expedient Building Block Towards the Construction of Nitrogen-Containing Molecules: An Update. Adv. Synth. Catal. 2020, 362, 1447–1484. [Google Scholar] [CrossRef]
  5. Ivanova, O.A.; Trushkov, I.V. Donor−Acceptor Cyclopropanes in the Synthesis of Carbocycles. Chem. Rec. 2019, 19, 2189–2208. [Google Scholar] [CrossRef] [PubMed]
  6. Tomilov, Y.V.; Menchikov, L.G.; Novikov, R.A.; Ivanova, O.A.; Trushkov, I.V. Methods for the synthesis of donor–acceptor cyclopropanes. Russ. Chem. Rev. 2018, 87, 201–250. [Google Scholar] [CrossRef]
  7. Schneider, T.F.; Kaschel, J.; Werz, D.B. A New Golden Age for Donor-Acceptor Cyclopropanes. Angew. Chem. Int. Ed. 2014, 53, 5504–5523. [Google Scholar] [CrossRef]
  8. Belaya, M.A.; Knyazev, D.A.; Borisov, D.D.; Novikov, R.A.; Tomilov, Y.V. GaCl3-Mediated Cascade [2 + 4]-Cycloaddition/[4 + 2]-Annulation of Donor–Acceptor Cyclopropanes with Conjugated Dienes: Strategy for the Construction of Benzobicyclo[3.3.1]nonane Skeleton. J. Org. Chem. 2021, 86, 8089–8100. [Google Scholar] [CrossRef]
  9. Belaya, M.A.; Knyazev, D.A.; Novikov, R.A.; Tomilov, Y.V. “Diels-Alder reaction” in the ionic version: GaCl3-promoted formation of substituted cyclohexenes from donor–acceptor cyclopropanes and dienes. Tetrahedron Lett. 2020, 61, 151990. [Google Scholar] [CrossRef]
  10. Borisov, D.D.; Novikov, R.A.; Tomilov, Y.V. GaCl3-Mediated Reactions of Donor–Acceptor Cyclopropanes with Aromatic Aldehydes. Angew. Chem. Int. Ed. 2016, 55, 12233–12237. [Google Scholar] [CrossRef]
  11. Novikov, R.A.; Tarasova, A.V.; Korolev, V.A.; Timofeev, V.P.; Tomilov, Y.V. A New Type of Donor–Acceptor Cyclopropane Reactivity: The Generation of Formal 1,2- and 1,4-Dipoles. Angew. Chem. Int. Ed. 2014, 53, 3187–3191. [Google Scholar] [CrossRef] [PubMed]
  12. Novikov, R.A.; Korolev, V.A.; Timofeev, V.P.; Tomilov, Y.V. New dimerization and cascade oligomerization reactions of dimethyl 2-phenylcyclopropan-1,1-dicarboxylate catalyzed by Lewis acids. Tetrahedron Lett. 2011, 52, 4996–4999. [Google Scholar] [CrossRef]
  13. Fadeev, A.A.; Makarov, A.S.; Ivanova, O.A.; Uchuskin, M.G.; Trushkov, I.V. Extended Corey-Chaykovsky reactions: Transformations of 2-hydroxychalcones to benzannulated 2,8-dioxabicyclo[3.2.1]octanes and 2,3-dihydrobenzofurans. Org. Chem. Front. 2022, 9, 737–744. [Google Scholar] [CrossRef]
  14. Ivanova, O.A.; Andronov, V.A.; Vasin, V.S.; Shumsky, A.N.; Rybakov, V.B.; Voskressensky, L.G.; Trushkov, I.V. Expanding the Reactivity of Donor–Acceptor Cyclopropanes: Synthesis of Benzannulated Five-Membered Heterocycles via Intramolecular Attack of a Pendant Nucleophilic Group. Org. Lett. 2018, 20, 7947–7952. [Google Scholar] [CrossRef] [PubMed]
  15. Ivanov, K.L.; Bezzubov, S.I.; Melnikov, M.Y.; Budynina, E.M. Donor–acceptor cyclopropanes as ortho-quinone methide equivalents in formal (4+2)-cycloaddition to alkenes. Org. Biomol. Chem. 2018, 16, 3897–3909. [Google Scholar] [CrossRef]
  16. Xiao, J.-A.; Peng, H.; Liang, J.-S.; Meng, R.-F.; Su, W.; Xiao, Q.; Yang, H. Gold/scandium bimetallic relay catalysis of formal [5+2]- and [4+2]-annulations: Access to tetracyclic indole scaffolds. Chem. Commun. 2021, 57, 13369–13372. [Google Scholar] [CrossRef]
  17. Unnava, R.; Chahal, K.; Reddy, K.R. Synthesis of substituted 1,2-dihydroisoquinolines via Ni(II) and Cu(I)/Ag(I) catalyzed double nucleophilic addition of arylamines to ortho-alkynyl donor–acceptor cyclopropanes (o-ADACs). Org. Biomol. Chem. 2021, 19, 6025–6029. [Google Scholar] [CrossRef]
  18. Wang, D.; Zhao, J.; Chen, J.; Xu, Q.; Li, H. Intramolecular Arylative Ring Opening of Donor–Acceptor Cyclopropanes in the Presence of Triflic Acid: Synthesis of 9H-Fluorenes and 9,10-Dihydrophenanthrenes. Asian J. Org. Chem. 2019, 8, 2032–2036. [Google Scholar] [CrossRef]
  19. Mikhaylov, A.A.; Dilman, A.D.; Novikov, R.A.; Khoroshutina, Y.A.; Struchkova, M.I.; Arkhipov, D.E.; Nelyubina, Y.V.; Tabolin, A.A.; Ioffe, S.L. Tandem Pd-catalyzed C–C coupling/recyclization of 2-(2-bromoaryl)cyclopropane-1,1-dicarboxylates with primary nitro alkanes. Tetrahedron Lett. 2016, 57, 11–14. [Google Scholar] [CrossRef]
  20. Ma, W.; Fang, J.; Ren, J.; Wang, Z. Lewis Acid Catalyzed Formal Intramolecular [3 + 3] Cross-Cycloaddition of Cyclopropane 1,1-Diesters for Construction of Benzobicyclo[2.2.2]octane Skeletons. Org. Lett. 2015, 17, 4180–4183. [Google Scholar] [CrossRef]
  21. Wang, L.-F.; Shi, Z.-F.; Cao, X.-P.; Li, B.-S.; An, P. Construction of fused- and spiro-oxa-[n.2.1] skeletons by a tandem epoxide rearrangement/intramolecular [3+2] cycloaddition of cyclopropanes with carbonyls. Chem. Commun. 2014, 50, 8061–8064. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, W.; Ren, J.; Wang, Z. Acid-Catalyzed Domino Meinwald Rearrangement of Epoxides/Intramolecular [3+2] Cross-Cycloaddition of Cyclopropane-1,1-dicarboxylates. Eur. J. Org. Chem. 2014, 2014, 3561–3564. [Google Scholar] [CrossRef]
  23. Flisar, M.E.; Emmett, M.R.; Kerr, M.A. Catalyst-Free Tandem Ring-Opening/Click Reaction of Acetylene-Bearing Donor–Acceptor Cyclopropanes. Synlett 2014, 25, 2297–2300. [Google Scholar] [CrossRef]
  24. Wang, Z. Polar Intramolecular Cross-Cycloadditions of Cyclopropanes toward Natural Product Synthesis. Synlett 2012, 23, 2311–2327. [Google Scholar] [CrossRef]
  25. Xia, X.-F.; Song, X.-R.; Liu, X.-Y.; Liang, Y.-M. Lewis Acid-Catalyzed Intramolecular [3+2] Cycloaddition of Cyclopropane 1,1-Diesters with Alkynes for the Synthesis of Cyclopenta[c]chromene Skeletons. Chem. Asian J. 2012, 7, 1538–1541. [Google Scholar] [CrossRef]
  26. Chagarovskiy, A.O.; Strel’tsova, E.D.; Rybakov, V.B.; Levina, I.I.; Trushkov, I.V. Synthesis of 2,3-diaryl-2,3,4,4a-tetrahydro-5H-indeno[1,2-c]pyridazine-5-ones. Chem. Heterocycl. Compd. 2019, 55, 240–245. [Google Scholar] [CrossRef]
  27. Qian, P.; Du, B.; Song, R.; Wu, X.; Mei, H.; Han, J.; Pan, Y. N-Iodosuccinimide-Initiated Spirocyclopropanation of Styrenes with 1,3-Dicarbonyl Compound for the Synthesis of Spirocyclopropanes. J. Org. Chem. 2016, 81, 6546–6553. [Google Scholar] [CrossRef]
  28. Nambu, H.; Fukumoto, M.; Hirota, W.; Ono, N.; Yakura, T. An efficient synthesis of cycloalkane-1,3-dione-2-spirocyclopropanes from 1,3-cycloalkanediones using (1-aryl-2-bromoethyl)dimethylsulfonium bromides: Application to a one-pot synthesis of tetrahydroindol-4(5H)-one. Tetrahedron Lett. 2015, 56, 4312–4315. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Synthesis of cyclopropane 4. DIPEA = N,N-diisopropylethylamine.
Scheme 1. Synthesis of cyclopropane 4. DIPEA = N,N-diisopropylethylamine.
Molbank 2023 m1604 sch001
Figure 1. Molecular structure (ORTEP-3) from single-crystal X-ray study of 4.
Figure 1. Molecular structure (ORTEP-3) from single-crystal X-ray study of 4.
Molbank 2023 m1604 g001
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

Ivanova, O.A.; Shorokhov, V.V.; Andreev, I.A.; Ratmanova, N.K.; Rybakov, V.B.; Strel’tsova, E.D.; Trushkov, I.V. Synthesis of 2-[2-(Ethoxymethoxy)phenyl]spiro[cyclopropane-1,2′-indene]-1′,3′-dione. Molbank 2023, 2023, M1604. https://doi.org/10.3390/M1604

AMA Style

Ivanova OA, Shorokhov VV, Andreev IA, Ratmanova NK, Rybakov VB, Strel’tsova ED, Trushkov IV. Synthesis of 2-[2-(Ethoxymethoxy)phenyl]spiro[cyclopropane-1,2′-indene]-1′,3′-dione. Molbank. 2023; 2023(1):M1604. https://doi.org/10.3390/M1604

Chicago/Turabian Style

Ivanova, Olga A., Vitaly V. Shorokhov, Ivan A. Andreev, Nina K. Ratmanova, Victor B. Rybakov, Elena D. Strel’tsova, and Igor V. Trushkov. 2023. "Synthesis of 2-[2-(Ethoxymethoxy)phenyl]spiro[cyclopropane-1,2′-indene]-1′,3′-dione" Molbank 2023, no. 1: M1604. https://doi.org/10.3390/M1604

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