Skip to content
BY 4.0 license Open Access Published by De Gruyter April 7, 2021

One-pot synthesis of benzofurans via heteroannulation of benzoquinones

  • Maryam Pirouz , M. Saeed Abaee EMAIL logo , Pernille Harris and Mohammad M. Mojtahedi

Abstract

Three different reactions were explored leading to the synthesis of various benzofurans. All reactions took place under AcOH catalysis in a one-pot manner. As a result, benzoquinone derivatives underwent heteroannulation with either itself or cyclohexanones to produce furanylidene-benzofuran or benzofuran structures, respectively.

Introduction

Heterocyclic compounds possessing the benzofuran core [1], of either natural or synthetic origin [2], are very important due to their exhibition of various biological activities. This has led to numerous investigations to design procedures for the synthesis of benzofuran based structures [3,4] and to study their biological behavior [5] as anti-oxidant [6], anticancer [7], antimicrobial [8], antitumor [9], and immunomodulatory [10] agents. As a result, the extensive physiological properties and the high natural occurrence of benzofuran derivatives have resulted in their use as versatile biodynamic and useful therapeutic agents. Important natural examples include moracins [11], cicerfuran [12], and conocarpan [13], while bufuralol [14], amiodarone [15], and ailanthoidol [16] are representative derivatives of biologically active synthetic benzofuran molecules (Fig. 1).

Figure 1 Important bioactive benzofuran containing structures
Figure 1

Important bioactive benzofuran containing structures

The broad spectrum of biochemical activities of various benzofurans has led to many investigations to prepare specific target structures from natural sources or via synthetic methods. Several arylbenzofurans were extracted from the stem bark [17], root bark [18], and leaves [19] of various mulberry trees, some of which were also prepared via multistep syntheses [20,21]. Notable methods for the synthesis of benzofurans include metal-free cyclization of ortho-hydroxystilbenes mediated by hypervalent iodine reagents [22], Ru-catalyzed isomerization [23] of appropriate precursors [24], cross-coupling of alkali-metal salts of silanols with aromatic halides [25], Pd-catalyzed addition of potassium aryltrifluoroborates to aliphatic nitriles [26], Sonogashira coupling of O-iodoanisoles with terminal alkynes followed by electrophilic cyclizations [27], and one-pot Pd-catalyzed coupling of ortho-bromophenols with enolizable ketones [28].

In the framework of our studies on the synthesis of various heterocyclic systems [29,30,31], and in continuation of our program on the use of cyclohexenone derivatives 1a–c in the synthesis [32,33,34], we decided to take advantage of the reactivity of cyclohexenone derivatives for possible coupling with benzoquinone (BQ), as shown in Scheme 1 for the heteroannulation of 1a with BQ. This work resulted in the one-pot synthesis of benzofuran and furanylidenebenzofuran systems 2–4 in PhMe/AcOH (4:1) medium.

Scheme 1 One-pot synthesis of 2
Scheme 1

One-pot synthesis of 2

Results and discussion

Confident in the application of [3+2] heteroannulations in the preparation of benzofuran structures, we first targeted the synthesis of 2. For this purpose, we reacted 1a with BQ in a refluxing PhMe/AcOH mixture, in which 81% of 2 was obtained after 24 h (Scheme 1). It is noteworthy that the synthesis of 2 was previously reported via a three-step process taking 6 days and achieving an overall yield of about 80% [35], whereas we could access 2 through a single step reaction (Scheme 2). Mechanistically, the reaction is probably going through a formal [3+2] process, and presumably a cyclic transition state (1a-BQH+) is involved. Protonation of BQ in acidic conditions and formation of BQH+ [36] and its participation in hetrocycloadditions have precedence [37].

Scheme 2 Stepwise vs one-pot synthesis of 2
Scheme 2

Stepwise vs one-pot synthesis of 2

To gain some understanding of the process, we replaced 1a with 1b. Treatment of BQ with 1b (Table 1) in refluxing PhMe/AcOH solution for 3 days gave a complex mixture of products from which m-cresol, hydroquinone (HQ), and low yields of product 3 were identified (entry 1). The presence of a HQ skeleton in the structure of 3 and the formation of m-cresol as a side product suggested that a redox process could be involved. Therefore, similar reactions were performed replacing BQ with HQ. When HQ and 1b were used, no formation of 3 was detected (entry 2), while a repeat of each reaction in the absence of 1b was also unsuccessful (entries 3–4). In contrast, a mixture of BQ and HQ produced a 70% yield of 3 after an 18 h reflux in PhMe/AcOH (entry 5). It is noteworthy that 3 belongs to the group of furanylidene-benzofuran heterocycles which contain both benzofuran and additional lactone moieties and are important for their pigment [38] and biochemical [39] properties.

Table 1

Optimization of the synthesis of 3 using 1b

Table 1 Optimization of the synthesis of 3 using 1b
Entry 1b (equiv.) BQ (equiv.) HQ (equiv.) Conditions Time (h) Yield (%)a
1 1.0 2.5 0.0 AcOH, PhMe 72 15
2 1.0 0.0 2.5 AcOH, PhMe 72 0
3 0.0 1.0 0.0 AcOH, PhMe 72 0
4 0.0 0.0 1.0 AcOH, PhMe 72 0
5 0.0 2.0 1.0 AcOH, PhMe 18 70
6 0.0 2.0 1.0 CD3CO2D, PhMe 18 70
7 0.0 2.0 1.0 CHD, AcOH, PhMe 72 0
8 0.0 2.0 1.0 AcOH, CHCl3 18 0
9 0.0 2.0 1.0 AcOH, H2O 18 0
10 0.0 2.0 1.0 AcOH, mesitylene 18 55
11 0.0 2.0 1.0 AcOH, xylene 18 53
12 0.0 2.0 1.0 AcOH, benzene 18 33
13 0.0 2.0 1.0 CF3CO2H, PhMe 18 0
14 0.0 2.0 1.0 HCO2H, PhMe 18 67
15 0.0 2.0 1.0 PTSA, PhMe 18 0
16 0.0 2.0 1.0 HCl, PhMe 18 0
17 0.0 2.0b 1.0 AcOH, PhMe 18 0
18 0.0 2.0c 1.0 AcOH, PhMe 18 0
  1. a

    Isolated yields.

  2. b

    BQ was replaced with 2,5-DMBQ.

  3. c

    BQ was replaced with 2,6-DMBQ.

To further understand the process, we repeated the reactions in the presence of CD3CO2D, in which the same results were obtained (entry 6), proving the lack of the participation of acetic acid in the structure of the product. Similarly, use of cyclohexa-1,4-diene (CHD), as a radical scavenger [40], halted the reaction completely, supporting the inclusion of a redox process in the reaction (entry 7). Changing the solvent (entries 8–12) or the acid reagent (entries 13–16) did not improve the results. The use of two other BQ derivatives (2,5-dimethylcyclohexa-2,5-diene-1,4-dione, 2,5-DMBQ, entry 17 and 2,6-dimethylcyclohexa-2,5-diene-1,4-dione, 2,6-DMBQ, entry 18) gave no respective products, as would be expected from their more electron rich nature.

To confirm that a hydrogen donor source such as 1b or HQ is needed for the reaction to proceed, we replaced 1b with either 1c or 1d and repeated the reactions in the absence of HQ (Scheme 3). Thus, with 2:1 mixtures of BQ and 1c (or 1d) no detectable reactions were observed.

Scheme 3 Reactions in the presence of 1c–d
Scheme 3

Reactions in the presence of 1c–d

Based on the results, a pathway was proposed in which BQ is first protonated to BQH+ [36]. This hypothetical intermediate is stabilized via ring opening to a conjugated oxo-bis-enone moiety, which in turn is attacked by water to form the hydroxyethylidene-furanone intermediate. Further oxidation of this intermediate by BQ and its coupling with the resulting HQ, followed by the final lactonization, gives 3 (Scheme 4).

Scheme 4 Mechanism of the synthesis of 3
Scheme 4

Mechanism of the synthesis of 3

The structure of 3 was identified based on its NMR spectra. In 1H NMR, the presence of two doublets and one doublet of doublets signals at about 6.8–7.2 ppm was in accordance with the HQ unit, while two doublet signals at 8.4 and 7.0 ppm were indicative of the unsaturated lactone ring. Similarly, DEPT-135, COSY, HSQC, and APT experiments supported the formation of the suggested structure. To confirm, a single crystal of 3 was prepared and subjected to X-ray crystallographic experiments. Fig. 2 clearly supports the assignment.

Figure 2 ORTEP representation of 3, thermal ellipsoids set at the 40% probability level
Figure 2

ORTEP representation of 3, thermal ellipsoids set at the 40% probability level

Based on these results, we decided to design a pathway in which a similar combination of BQ structures would occur via an elimination reaction as opposed to a redox process. For this purpose, we selected 2,6-dichlorobenzoquinone (BQCl2), which is more susceptible to protonation [41], and has the potential to produce dibenzofurans, when reacted with ethyl 2-methyl-4-oxocyclohex-2-ene-1-carboxylate (Hagmann ester, 1e). This would occur by taking advantage of a possible facile HCl elimination to couple the two reactants and reach a tricyclic dibenzofuran target structure (Scheme 5). As a result, a facile regioselective coupling was observed, giving 4.

Scheme 5 Synthesis of 4
Scheme 5

Synthesis of 4

The mechanism of the reaction presumably goes through the formation of BQCl2H+. This species is formed upon treatment of the reactants with AcOH, while 1e is also enolized to dienol 1e′ under the acidic conditions. Then, the enol 1e′ attacks the electron-poor BQCl2H+ to form 4′. Rearrangement of the double bond to a more stable conjugated position promotes the ring closure and the process is followed by HCl elimination and a final dehydration/aromatization step to form 4 (Scheme 6). The structure of the product was elucidated with NMR spectroscopy techniques and was confirmed with X-ray crystallographic analysis (Fig. 3).

Scheme 6 Mechanism of the synthesis of 4
Scheme 6

Mechanism of the synthesis of 4

Figure 3 ORTEP representation of 4, thermal ellipsoids set at the 40% probability level
Figure 3

ORTEP representation of 4, thermal ellipsoids set at the 40% probability level

Conclusion

In summary, we successfully conducted the synthesis of benzofuran and furanylidene-benzofuran systems via the one-pot coupling of BQ with either itself or cyclohexenone under refluxing acidic conditions, using no coupling reagent. The benzofuran formation occurs either through [3+2] heteroannulation of the starting cyclohexenone moiety with BQ derivatives (in the case of 2 and 4) or the “dimerization” of BQ (in the case of 3), followed by either spontaneous oxidation (aromatization) or elimination (dehydration) steps. The procedures are clearly effective in producing the target compounds under inexpensive conditions, in shorter time periods and with fewer reaction steps when compared to the previous reports for known products [35]. Based on these results, we are currently developing the procedures to use a broader spectrum of reactants, such as phenols and thiophenols.

Experimental

Melting points are uncorrected. FT-IR spectra were recorded using KBr disks on a Bruker Vector-22 spectrometer. NMR spectra were recorded at 400 MHz for 1H NMR and 101 MHz for 13C NMR on a Bruker Ascend 400 MHz spectrometer in DMSO-d6 solutions using TMS as an internal standard reference. Chemical shift values (δ) are reported in ppm relative to the residual solvent signal in DMSO, while coupling constants (J) are given in Hz. Multiplicities are reported as s (singlet), d (doublet), dd (doublet of doublets), m (multiplet) and etc. Flash column chromatography was performed using silica gel 60 (0.035–0.070 mm particle size). Elemental analyses were performed using a Thermo Finnigan Flash EA 1112 instrument. LCMS analysis was carried out on a Waters ACQUITY UPLC system with a PDA and SQD2 electrospray detector and a Thermo Accucore C18 2.6 μm, 2.1 × 50 mm column. TLC experiments were carried out on pre-coated silica gel plates using hexanes/EtOAc as the eluent. Chemicals and starting materials were purchased from commercial sources. Product 2 [35] was known and identified by NMR spectra. Products 3 and 4 were new and were characterized by analysis of their 1H NMR, 13C NMR, IR, mass spectra, and X-ray crystallography.

Synthesis of 8-hydroxy-3,4-dihydrodibenzo[b,d]furan-1(2H)-one (2)

To a solution of PhMe (4.0 mL) and glacial AcOH (1.0 mL) was added 1a (100 μL, 1.0 mmol) and benzoquinone (216 mg, 2.0 mmol) and the mixture was refluxed for 24 hours. Saturated aqueous NaHCO3 (excess) was added to the mixture and the organic layer was diluted with EtOAc (10.0 mL). The organic layer was dried over Na2SO4 and after evaporation of the volatiles, the residue was purified with flash chromatography over silica gel using EtOAc/hexanes (1:5) to obtain 2.

White crystals (81%); 1H NMR (400 MHz, DMSO-d6) δ 9.40 (s, 1H), 7.43 (d, J = 9.0 Hz, 1H), 7.26 (d, J = 2.5 Hz, 1H), 6.75 (dd, J = 2.5, 9.0 Hz, 1H), 3.03-3.00 (m, 2H), 2.49-2.48 (m, 2H), 2.19-2.12 (m, 2H); 13C NMR (101 MHz, DMSO-d6) δ 194.8, 172.5, 155.1, 148.3, 124.6, 115.9, 113.6, 112.2, 106.1, 33.8, 23.7, 22.4; MS: m/z 202 (M+).

Synthesis of (E)-5-hydroxy-3-(5-oxofuran-2(5H)-ylidene) benzofuran-2(3H)-one (3)

To a solution of PhMe (4.0 mL) and glacial AcOH (1.0 mL) was added benzoquinone (216 mg, 2.0 mmol) and hydroquinone (110 mg, 1.0 mmol) and the mixture was refluxed for 18 hours. Saturated aqueous NaHCO3 (excess) was added to the mixture and the organic layer was diluted with EtOAc (10.0 mL). The organic layer was dried over Na2SO4 and after evaporation of the volatiles, the residue was purified with flash chromatography over silica gel using EtOAc/hexanes (1:5) to obtain 3.

Red crystals (70%); 210 °C (decomposes); 1H NMR (400 MHz, DMSO-d6) δ 9.65 (s, 1H), 8.43 (d, J = 5.5 Hz, 1H), 7.18 (d, J = 2.5 Hz, 1H), 7.10 (d, J = 8.5 Hz, 1H), 7.01 (d, J = 5.5 Hz, 1H), 6.85 (dd, J = 2.5, 8.5 Hz, 1H); 13C NMR (101 MHz, DMSO-d6) δ 168.0, 166.9, 155.4, 154.5, 146.9, 141.2, 125.2, 122.0, 119.2, 120.4, 111.1, 106.0; IR (KBr) ν 1064, 1466, 1766, 2922, 3467 cm−1; MS: m/z 230 (M+). Anal. Calcd for C12H6O5: C, 62.62; H, 2.63. Found: C, 62.50; H, 2.75.

X-ray data for 3

C12H6O5, M = 230.17 g/mol, triclinic system, space group P-1, a = 7.0075(7), b = 9.9741(7), c = 14.5706(12) Å, α=81.108(6), β = 79.856(8), γ=70.532(8), V = 940.10(15) Å3, Z = 4, Dc = 1.626 g/cm−3, μ(Mo-Kα) = 0.129 mm−1, crystal dimension of 0.1 × 0.1 × 0.1 mm. The X-ray data collection for 3 was performed on an Agilent Supernova Diffractometer. Data processing was done using CrysAlisPro (Agilent Technologies). The structure was solved by using SHELXS, and structure refinement was carried out with SHELXL [42]. The non-hydrogen atoms were refined anisotropically by full matrix least-squares on F2 values to final R1 = 0.0822, wR2 = 0.2630, and S = 1.009 with 309 parameters using 4647 independent reflection (θ range = 3.19–29.65°). Hydrogen atoms were included on ideal positions using riding coordinates. Crystallographic data for 3 have been deposited with the Cambridge Crystallographic Data Centre. Copies of the data can be obtained, free of charge, on application to The Director, CCDC-2046512, Union Road, Cambridge CB2 1EZ, UK. Fax: +44 1223 336033 or e-mail: deposit@ccdc.cam.ac.uk.

Synthesis of ethyl 6-chloro-8-hydroxy-1-methyldibenzo[b,d]furan-2-carboxylate (4)

To a solution of PhMe (4.0 mL) and glacial AcOH (1.0 mL) was added 1e (160 μL, 1.0 mmol) and BQCl2 (264 mg, 1.5 mmol) and the mixture was refluxed for 3 hours. Saturated aqueous NaHCO3 (excess) was added to the mixture and the organic layer was diluted with EtOAc (10.0 mL). The organic layer was dried over Na2SO4 and after evaporation of the volatiles, the residue was purified with flash chromatography over silica gel using EtOAc/hexanes (1:5) to obtain 4.

White crystals (83%); mp = 183–184 °C; 1H NMR (400 MHz, DMSO-d6) δ 9.99 (s, 1H), 7.91 (d, J = 9.0, 1H), 7.58 (d, J = 9.0 Hz, 1H), 7.45 (d, J = 2.0 Hz, 1H), 7.06 (d, J = 2.0 Hz, 1H), 4.32 (q, J = 7.0 Hz, 2H), 2.88 (s, 3H), 1.35 (t, J = 7.0 Hz, 3H); 13CNMR (101 MHz, DMSO-d6) δ 167.3, 157.6, 154.7, 145.6, 136.8, 130.6, 126.2, 125.9, 123.9, 115.9, 116.0, 109.8, 107.9, 61.2, 17.3, 14.6 ; IR (KBr) ν 3344, 2853, 1681, 1258, 1073, 779 cm−1; MS: m/z = 304 [M]+. Anal. Calcd for C16H13ClO4: C, 63.07; H, 4.30. Found: C, 63.25; H, 4.52.

X-ray data for 4

C16H13ClO4, M = 304.71 g/mol, monoclinic system, space group P21/c, a = 12.0554(3), b = 7.2442(1), c = 15.3414(3) Å, β = 95.954(2), V = 1332.56(5) Å3, Z = 4, Dc = 1.519 g/cm-3, μ(Cu-Kα) = 2.672 mm−1, crystal dimension of 0.25 × 0.20 × 0.18 mm. The structure was solved using SHELXS and refined with SHELXL [40]. The non-hydrogen atoms were refined anisotropically by full matrix least-squares on F2 values to final R1 = 0.0470, wR2 = 0.1360, and S = 1.025 with 193 parameters using 2792 independent reflection (θ range = 3.69–76.56°). Hydrogen atoms were inserted at ideal positions.

  1. Research funding: The Research Council at CCERCI is acknowledged for financial support of this work.

  2. Conflict of interest: Authors state no conflicts of interest.

  3. Data availability statement: Crystallographic data for 4 have been deposited with the Cambridge Crystallographic Data Centre. Copies of the data can be obtained, free of charge, on application to The Director, CCDC-2046513, Union Road, Cambridge CB2 1EZ, UK. Fax: +44 1223 336033 or e-mail: . All other data generated or analyzed during this study are included in this published article and its supplementary information file.

References

[1] Khanam H, Shamsuzzaman. Bioactive Benzofuran derivatives: A review. Eur J Med Chem. 2015 Jun;97:483–504.10.1016/j.ejmech.2014.11.039Search in Google Scholar PubMed

[2] Chand K, Rajeshwari, Hiremathad A, Singh M, Santos MA, Keri RS. A review on antioxidant potential of bioactive heterocycle benzofuran: natural and synthetic derivatives. Pharmacol Rep. 2017 Apr;69(2):281–95.10.1016/j.pharep.2016.11.007Search in Google Scholar PubMed

[3] Kadieva MG, Oganesyan ÉT. Methods for the synthesis of benzofuran derivatives [review]. Chem Heterocycl Compd. 1997;33(11):1245–58.10.1007/BF02320323Search in Google Scholar

[4] More KR. Review on synthetic routes for synthesis of benzofuran-based compounds. J Chem Pharm Res. 2017;9:210–20.Search in Google Scholar

[5] Naik R, Harmalkar DS, Xu X, Jang K, Lee K. Bioactive benzofuran derivatives: moracins A-Z in medicinal chemistry. Eur J Med Chem. 2015 Jan;90:379–93.10.1016/j.ejmech.2014.11.047Search in Google Scholar PubMed

[6] Rindhe SS, Rode MA, Karale BK. New benzofuran derivatives as an antioxidant agent. Indian J Pharm Sci. 2010 Mar;72(2):231–5.10.4103/0250-474X.65022Search in Google Scholar PubMed PubMed Central

[7] Coşkun D, Tekin S, Sandal S, Coşkun MF. Synthesis, characterization, and anticancer activity of new benzofuran substituted chalcones. J Chem. 2016;2016:7678486.10.1155/2016/7678486Search in Google Scholar

[8] Ashok D, Kumar RS, Gandhi DM, Sarasija M, Jayashree A, Adam S. Solvent-free microwave-assisted synthesis and biological evaluation of 2,2-dimethylchroman-4-one based benzofurans. Heterocycl Commun. 2016;22(6):363–8.10.1515/hc-2016-0147Search in Google Scholar

[9] Othman DI, Abdelal AM, El-Sayed MA, El Bialy SA. Novel benzofuran derivatives: synthesis and antitumor activity. Heterocycl Commun. 2013;19(1):29–35.10.1515/hc-2012-0119Search in Google Scholar

[10] Khaleghi F, Jantan I, Din LB, Yaacob WA, Khalilzadeh MA, Bukhari SN. Immunomodulatory effects of 1-(6-hydroxy-2-isopropenyl-1-benzofuran-5-yl)-1-ethanone from Petasites hybridus and its synthesized benzoxazepine derivatives. J Nat Med. 2014 Apr;68(2):351–7.10.1007/s11418-013-0805-9Search in Google Scholar PubMed

[11] Mitsuo T, Shigemitsu N, Shoji U, Tadashi M, Akira S, Kokichi T. Moracin C and D, new phytoalexins from diseased mulberry. Chem Lett. 1978;7(11):1239–40.10.1246/cl.1978.1239Search in Google Scholar

[12] Damodar K, Kim JK, Jun JG. Unified syntheses of gramniphenols F and G, cicerfuran, morunigrol C and its derivative. Tetrahedron Lett. 2016;57(10):1183–6.10.1016/j.tetlet.2016.02.006Search in Google Scholar

[13] Silva AR, Polo EC, Martins NC, Correia CR. Enantioselective oxy-Heck–matsuda arylations: expeditious synthesis of dihydrobenzofuran systems and total synthesis of the neolignan (−)-conocarpan. Adv Synth Catal. 2018;360(2):346–65.10.1002/adsc.201701278Search in Google Scholar

[14] Weerawarna SA, Geisshüsler SM, Murthy SS, Nelson WL. Enantioselective and diastereoselective hydroxylation of bufuralol. Absolute configuration of the 7-(1-hydroxyethyl)-2-[1-hydroxy-2-(tert-butylamino)ethyl]benzofurans, the benzylic hydroxylation metabolites. J Med Chem. 1991 Oct;34(10):3091–7.10.1021/jm00114a019Search in Google Scholar PubMed

[15] Iqbal N, Iqbal N, Maiti D, Cho EJ. Access to multifunctionalized benzofurans by aryl nickelation of alkynes: efficient synthesis of the anti-arrhythmic drug amiodarone. Angew Chem Int Ed Engl. 2019 Oct;58(44):15808–12.10.1002/anie.201909015Search in Google Scholar PubMed

[16] Lin SY, Chen CL, Lee YJ. Total synthesis of ailanthoidol and precursor XH14 by Stille coupling. J Org Chem. 2003 Apr;68(7):2968–71.10.1021/jo020653tSearch in Google Scholar PubMed

[17] Wang SK, Duh CY. Cytotoxic cyclopenta[b]benzofuran derivatives from the stem bark of Aglaia formosana. Planta Med. 2001 Aug;67(6):555–7.10.1055/s-2001-16471Search in Google Scholar PubMed

[18] Piao SJ, Qiu F, Chen LX, Pan Y, Dou DQ. New Stilbene, Benzofuran, and coumarin glycosides from Morus alba. Helv Chim Acta. 2009;92(3):579–87.10.1002/hlca.200800275Search in Google Scholar

[19] Chan EW, Wong SK, Tangah J, Inoue T, Chan HT. Phenolic constituents and anticancer properties of Morus alba (white mulberry) leaves. J Integr Med. 2020 May;18(3): 189–95.10.1016/j.joim.2020.02.006Search in Google Scholar PubMed

[20] Kinoshita T, Ichinose K. A short step convenient synthesis of 2-phenylbenzofuran form 3-phenylcoumarin. Heterocycles. 2005;65(7):1641–54.10.3987/COM-05-10424Search in Google Scholar

[21] Kunyane P, Sonopo MS, Selepe MA. Synthesis of isoflavones by tandem demethylation and ring-opening/cyclization of methoxybenzoylbenzofurans. J Nat Prod. 2019 Nov;82(11):3074–82.10.1021/acs.jnatprod.9b00681Search in Google Scholar PubMed

[22] Singh FV, Wirth T. Hypervalent iodine mediated oxidative cyclization of o-hydroxystilbenes into benzo- and naphthofurans. Synthesis. 2012;44(8):1171–7.10.1002/chin.201232105Search in Google Scholar

[23] van Otterlo WA, Morgans GL, Madeley LG, Kuzvidza S, Moleele SS, Thornton N, et al. An isomerization-ring-closing metathesis strategy for the synthesis of substituted benzofurans. Tetrahedron. 2005;61(32):7746–55.10.1016/j.tet.2005.05.090Search in Google Scholar

[24] Varela-Fernández A, González-Rodríguez C, Varela JA, Castedo L, Saá C. Cycloisomerization of aromatic homo- and bis-homopropargylic alcohols via catalytic Ru vinylidenes: formation of benzofurans and isochromenes. Org Lett. 2009 Nov;11(22):5350–3.10.1021/ol902212hSearch in Google Scholar PubMed

[25] Denmark SE, Smith RC, Chang WT, Muhuhi JM. Cross-coupling reactions of aromatic and heteroaromatic silanolates with aromatic and heteroaromatic halides. J Am Chem Soc. 2009 Mar;131(8):3104–18.10.1021/ja8091449Search in Google Scholar PubMed PubMed Central

[26] Wang X, Liu M, Xu L, Wang Q, Chen J, Ding J, et al. Palladium-catalyzed addition of potassium aryltrifluoroborates to aliphatic nitriles: synthesis of alkyl aryl ketones, diketone compounds, and 2-arylbenzo[b]furans. J Org Chem. 2013 Jun;78(11):5273–81.10.1021/jo400433mSearch in Google Scholar PubMed

[27] Yue D, Yao T, Larock RC. Synthesis of 2,3-disubstituted benzo[b] furans by the palladium-catalyzed coupling of o-iodoanisoles and terminal alkynes, followed by electrophilic cyclization. J Org Chem. 2005 Dec;70(25):10292–6.10.1021/jo051299cSearch in Google Scholar PubMed PubMed Central

[28] Eidamshaus C, Burch JD. One-pot synthesis of benzofurans via palladium-catalyzed enolate arylation with o-bromophenols. Org Lett. 2008 Oct;10(19):4211–4.10.1021/ol801510nSearch in Google Scholar PubMed

[29] Mojtahedi MM, Mehraban M, Darvishi K, Abaee MS. Ultrasound mediated synthesis of dihydropyrano [3,2-d][1,3]dioxin-7-carbonitrile derivatives in H2O/EtOH medium. Heterocycl Commun. 2017;23(2):91–5.10.1515/hc-2017-0014Search in Google Scholar

[30] Abaee MS, Akbarzadeh E, Shockravi A, Mojtahedi MM, Mehraki E, Khavasi HR. Three-component anti selective Mannich reactions in a tetrahydro-4-pyranone system by using PDAG-Co catalyst. Heterocycl Commun. 2014;20(2):123–8.10.1515/hc-2014-0001Search in Google Scholar

[31] Poursharifi MJ, Mojtahedi MM, Abaee MS, Hashemi MM. The first in situ synthesis of 1,3-dioxan-5-one derivatives and their direct use in Claisen-Schmidt reactions. Heterocycl Commun. 2019;25(1):85–90.10.1515/hc-2019-0013Search in Google Scholar

[32] Abaee MS, Mojtahedi M, Zahedi MM, Sharifi R, Mesbah AW, Massa W. Synthesis and structural elucidation of novel bisarylmethylidenes of cyclic enones. Synth Commun. 2007;37(17):2949–57.10.1080/07370650701471756Search in Google Scholar

[33] Mojtahedi MM, Abaee MS, Zahedi MM, Jalali MR, Mesbah AW, Massa W, et al. Facile solvent-free synthesis and structural elucidation of styrylcyclohex-2-enone derivatives. Monatsh Chem. 2008;139(8):917–21.10.1007/s00706-007-0847-3Search in Google Scholar

[34] Abaee MS, Mohammadi M, Mojtahedi M, Halvagar MR. Halvagar. M. R. Synthesis and Diels-Alder reactions of novel styrylcyclohexenol dienes: a general pathway for the synthesis of tricyclic lactone and octahydronaphthalene systems. Tetrahedron Lett. 2016;57(36):4094–7.10.1016/j.tetlet.2016.07.096Search in Google Scholar

[35] Craven PG, Taylor RJ. Total synthesis and structural confirmation of (±)-cuevaene A. Tetrahedron Lett. 2012;53(40):5422–5.10.1016/j.tetlet.2012.07.121Search in Google Scholar

[36] Bhongale PV, Joshi SS, Mali NA. Selective monoalkylation of hydroquinone in the presence of SO3H-functionalized ionic liquids as catalysts. Chem Pap. 2020;74(12): 4461–71.10.1007/s11696-020-01230-1Search in Google Scholar

[37] Pushkarskaya E, Wong B, Han C, Capomolla S, Gu C, Stoltz BM, et al. Single-step synthesis of 3-hydroxycarbazoles by annulation of electron-rich anilines and quinones. Tetrahedron Lett. 2016;57(50):5653–7.10.1016/j.tetlet.2016.11.009Search in Google Scholar

[38] von Nussbaum F, Rüth M, Spiteller P, Hübscher-Weissert T, Löbermann F, Polborn K, et al. Coloured artefacts formed by oxidation of benzene-1,2,4-triol and β-dopa during the extraction of cortinarius violaceus (Agaricales) with alcohols. Eur J Org Chem. 2012;2012(2):380–90.10.1002/ejoc.201101215Search in Google Scholar

[39] Dang PH, Nguyen HX, Nguyen NT, Le HN, Nguyen MT. α-Glucosidase inhibitors from the stems of Embelia ribes. Phytother Res. 2014 Nov;28(11):1632–6.10.1002/ptr.5175Search in Google Scholar PubMed

[40] Walton JC, Studer A. Evolution of functional cyclohexadiene-based synthetic reagents: the importance of becoming aromatic. Acc Chem Res. 2005 Oct;38(10):794–802.10.1021/ar050089jSearch in Google Scholar PubMed

[41] Kepler S, Zeller M, Rosokha SV. Anion–π complexes of halides with p-benzoquinones: structures, thermodynamics, and criteria of charge transfer to electron transfer transition. J Am Chem Soc. 2019 Jun;141(23):9338–48.10.1021/jacs.9b03277Search in Google Scholar PubMed

[42] Sheldrick GM. A short history of SHELX. Acta Crystallogr A. 2008 Jan;64(Pt 1):112–22.10.1107/S0108767307043930Search in Google Scholar PubMed

Received: 2020-12-02
Accepted: 2021-02-15
Published Online: 2021-04-07

© 2021 Maryam Pirouz et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/hc-2020-0120/html
Scroll to top button