Next Article in Journal
Design, Synthesis and Fungicidal Activities of Some Novel Pyrazole Derivatives
Next Article in Special Issue
Enantiomerically Pure Phosphonated Carbocyclic 2'-Oxa-3'-Azanucleosides: Synthesis and Biological Evaluation
Previous Article in Journal
Study of Leaf Metabolome Modifications Induced by UV-C Radiations in Representative Vitis, Cissus and Cannabis Species by LC-MS Based Metabolomics and Antioxidant Assays
Previous Article in Special Issue
Cycloaddition of 1,3-Butadiynes: Efficient Synthesis of Carbo- and Heterocycles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis with Perfect Atom Economy: Generation of Furan Derivatives by 1,3-Dipolar Cycloaddition of Acetylenedicarboxylates at Cyclooctynes

1
Organic Chemistry, Technische Universität Chemnitz, Strasse der Nationen 62, 09111 Chemnitz, Germany
2
Inorganic Chemistry, Technische Universität Chemnitz, 09107 Chemnitz, Germany
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(9), 14022-14035; https://doi.org/10.3390/molecules190914022
Submission received: 1 August 2014 / Revised: 28 August 2014 / Accepted: 1 September 2014 / Published: 5 September 2014
(This article belongs to the Special Issue Cycloaddition Chemistry)

Abstract

:
Cyclooctyne and cycloocten-5-yne undergo, at room temperature, a 1,3-dipolar cycloaddition with dialkyl acetylenedicarboxylates 1a,b to generate furan-derived short-lived intermediates 2, which can be trapped by two additional equivalents of 1a,b or alternatively by methanol, phenol, water or aldehydes to yield polycyclic products 3bd, orthoesters 4ac, ketones 5 or epoxides 6a,b, respectively. Treatment of bis(trimethylsilyl) acetylenedicarboxylate (1c) with cyclooctyne leads to the ketone 7 via retro-Brook rearrangement of the dipolar intermediate 2c. In all cases, the products are formed with perfect atom economy.

Graphical Abstract

1. Introduction

Esters of acetylene dicarboxylic acid and especially dimethyl acetylenedicarboxylate (DMAD, 1a) are highly versatile tools for organic chemists [1,2,3,4,5,6,7]. These compounds are successfully utilized as dienophiles in Diels–Alder reactions, as dipolarophiles in 1,3-dipolar cycloadditions and also as components in [2 + 2] or other cycloaddition reactions. Furthermore, they can be used as powerful Michael acceptors, and in several cases, nucleophilic addition and formation of zwitterions are combined with other addition or (formal) cycloaddition steps to yield a variety of products by multicomponent reactions [1].
Tetramerization of 1a, that has been performed by storing the neat substance at room temperature for several years or by heating it neatly or in solution, indicates an unusual reaction course of acetylenedicarboxylates (Scheme 1) [8,9,10]. Formation of the crystalline product 3a was explained by a dimerization step, in which 1a functions as dipolarophile and, also, as 1,3-dipole [8,10,11]. The resulting carbene intermediate 2a is trapped by a third molecule of 1a, and, finally, Diels–Alder reaction of the furan unit with a fourth molecule of 1a leads to 3a. The structure of this tetramer was confirmed by X-ray single crystal structure analysis [10,12]. The compound 3a has been utilized in several transformations, such as thermal retro-Diels–Alder reactions [9,13] or [4 + 2]-cycloaddition in the presence of cyclopentadienes [14], as well as photolysis [8,9] or treatment with triphenylphosphine [10]. The tetramer of diethyl acetylenedicarboxylate has also been prepared by a procedure, which is analogous to the synthesis of 3a [9]. However, formation of products with structures similar to that of 3a is very limited when 1a was heated with other alkynes or alkenes [13,15]. For example, the carbene intermediate 2a can also be trapped in the presence of electron-poor dimethyl fumarate to generate the cyclopropane-derived compound dihydro-3a, but with an excess of tolane, the corresponding interception product was obtained only in trace amounts. This is remarkable because it is well known that carbenes, which were produced by photolysis of methyl aryl(diazo)acetate, react with acceptor-substituted and also electron-rich π-systems to form three-membered rings [16]. In a rhodium-catalyzed transformation, 1a and terminal alkenes led to cyclopropane derivatives that might be explained by trapping of carbene 2a with the help of the olefin [17,18]. However, such a mechanism was left out of consideration.
Scheme 1. Tetramerization of dimethyl acetylenedicarboxylate (1a).
Scheme 1. Tetramerization of dimethyl acetylenedicarboxylate (1a).
Molecules 19 14022 g003

2. Results and Discussion

We accidentally found that the alkynes 1a and cyclooctyne [19] undergo an exothermic reaction at room temperature [20,21,22,23]. Thus, the transformation was conveniently performed in dichloromethane (20 h/20 °C) and led to the crystalline product 3b with 79% yield (Scheme 2). The structure of 3b was confirmed not only by NMR spectroscopic data but also by single crystal X-ray diffraction analysis (Figure 1). Obviously, the 1,3-dipolar cycloaddition of 1a at the ring-strained dipolarophile cyclooctyne to generate the intermediate 2b is much more rapid than the dimerization of 1a to produce 2a. This allows the synthesis of the interception product 3b without heating or long reaction times and also formation of several other trapping products of 2b.
Scheme 2. Reaction of 1a with cyclooctyne in the absence or in the presence of other reagents.
Scheme 2. Reaction of 1a with cyclooctyne in the absence or in the presence of other reagents.
Molecules 19 14022 g004
Figure 1. ORTEP representation (50% probability level) of the molecular structures of products 3b (left) and 4a (right); hydrogen atoms are omitted for clarity.
Figure 1. ORTEP representation (50% probability level) of the molecular structures of products 3b (left) and 4a (right); hydrogen atoms are omitted for clarity.
Molecules 19 14022 g001
After treatment of cyclooctyne with 1a in methanol instead of dichloromethane, the orthoester 4a was isolated in 60% yield. The surprising structure and especially the stereochemistry of this product were proved by spectroscopic data and X-ray diffraction analysis (Figure 1). When 1a and cyclooctyne were analogously reacted in an excess of deuterated methanol (CD3OD), the product 4b, which indicated the selective incorporation of exactly one equivalent of the deuterated reagent, was obtained with 62% yield. In the presence of phenol instead of methanol, the conversion of 1a and cyclooctyne only led to a low yield (10%) of the corresponding orthoester 4c. We did not get any orthoester trapping product with a structure similar to that of 4 after heating 1a alone in pure methanol (up to 80 °C) because addition of the solvent at the C≡C bond to form dimethyl 2-methoxybut-2-enedioates dominated.
In the case of 4ac, we exclusively isolated the depicted (E)-stereoisomer, whereas a mixture of (E)-5 and (Z)-5 resulted after exposure of cyclooctyne to 1a in aqueous tetrahydrofuran. The latter products were separated by chromatography to yield (E)-5 (21%) and (Z)-5 (20%), which were assigned with the help of NOE-NMR experiments. The genesis of 5 can be explained through interception of dipolar intermediate 2b by water. The product of this step is similar to orthoesters 4 but includes the substructure of a cyclic hemiacetal, which is transformed into 5 by ring opening followed by tautomerism.
When aldehydes such as acetaldehyde or benzaldehyde were used as solvents for the reaction of cyclooctyne with 1a, epoxides 6a and 6b, respectively, were formed (Scheme 2). In the case of 6a, the 1H-NMR spectrum of the crude reaction mixture indicated the generation of two diastereomers in a ratio of about 3.5:1 although four diastereomers are possible. Only the main product, however, could be isolated by chromatography (22% yield), and the relative configurations of its stereocenters were determined by single crystal X-ray diffraction analysis (Figure 2). The epoxide 6b was obtained in 21% yield as a mixture of two diastereomers, which could be separated by chromatography.
Figure 2. ORTEP representation (30% probability level) of the molecular structure of 6a; hydrogen atoms are omitted for clarity; only one of the two enantiomers of the asymmetric unit is shown.
Figure 2. ORTEP representation (30% probability level) of the molecular structure of 6a; hydrogen atoms are omitted for clarity; only one of the two enantiomers of the asymmetric unit is shown.
Molecules 19 14022 g002
The synthesis of 3b from 1a and cyclooctyne can be transferred to other acetylenedicarboxylates and cycloalkynes. Thus, 1a underwent a similar cascade of cycloaddition reactions in the presence of cycloocten-5-yne [24,25,26] to afford the product 3c in 37% yield (Scheme 3). On the other hand, diethyl acetylenedicarboxylate (1b) reacted only slightly slower than 1a with cyclooctyne to give the polycycle 3d with 61% yield. We attempted to perform the analogous transformation with di-tert-butyl acetylenedicarboxylate, however, not any characterizable product was obtained, even on heating the reaction mixture to 40 °C or on utilizing methanol as solvent. We assume that the sterically hindered diester is not able to enter into the rate-determining first step, which prevents not only the generation of the intermediate 2 but also the formation of products of type 3 or 4. However, treatment of bis(trimethylsilyl) acetylenedicarboxylate (1c) with cyclooctyne in anhydrous dichloromethane at 40 °C led to the oily 1:1 adduct 7 (85% yield). The 13C, 1H long-range correlation 2D-NMR spectra and especially the 29Si-NMR data indicated that the structure of 7 includes an oxygen-bound trimethylsilyl group and also such a group directly connected with carbon (Scheme 3). The genesis of 7 is easily explained by a 1,3-dipolar cycloaddition of the substrates and subsequent retro-Brook rearrangement of the intermediate 2c. This means that the corresponding vinylic carbon atom of 2c should possess nucleophilic properties to allow an intramolecular attack at silicon. Thus, the question arises whether intermediates of type 2 have to be generally described by dipolar resonance structures such as 2b and 2c or by carbene structures like 2a and 2b'. The formation of the three-membered ring in the products 3ad and 6a,b, respectively, can be interpreted via nucleophilic attack of the negatively charged carbon atom of 2 at the π-system of 1 or the aldehydes followed by ring closure of the resulting dipolar species or alternatively by a cheletropic reaction of the carbene version of 2 [27]. Furthermore, generation of orthoesters 4ac is possible through trapping of dipole 2b and is less easily explained via carbene 2b'. Finally, the retro-Brook rearrangement 2c7 and also the fact that interception of 2 to prepare products with three-membered rings is successful only in case of electron-poor π-systems but not with simple alkenes or alkynes, are strong arguments to prefer the dipolar resonance structure of intermediates 2 [28,29,30,31,32].
Scheme 3. Furan derivatives from acetylenedicarboxylates and cyclooctynes.
Scheme 3. Furan derivatives from acetylenedicarboxylates and cyclooctynes.
Molecules 19 14022 g005

3. Experimental

3.1. General Information

Melting points were determined with a Pentakon Dresden Boetius apparatus. FTIR spectra were recorded with a Nicolet iS5 spectrophotometer (Thermo Scientific) and solutions in KBr cuvettes. Alternatively, a FTS-165 spectrometer (BioRad) was used. 1H-NMR spectra were recorded with a Unity Inova 400 spectrometer operating at 400 MHz. By using the same spectrometer, 13C-NMR data were recorded at 100.6 MHz, 2H-NMR spectra were measured at 61.4 MHz and 29Si-NMR spectra at 79.5 MHz. NMR signals were referenced to TMS (δ = 0) or solvent signals and recalculated relative to TMS. The multiplicities of 13C-NMR signals were determined with the aid of DEPT135 experiments. HRMS (ESI) spectra were recorded with a Bruker micrOTOF-QII spectrometer. Elemental analyses were performed with a Vario Micro Cube from Elementar. HPLC was carried out with HPLC Pump 64 and Variable Wavelength Monitor (Knauer). TLC was performed with Macherey-Nagel Polygram SIL G/UV254 polyester sheets. Diffraction data for 3b, 4a and 6a were collected with an Oxford Gemini S diffractometer, with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å) (3b) or Cu Kα radiation (λ = 1.54184 Å) (4a, 6a). The structures were solved by direct methods and refined by full-matrix least-squares procedures on F2 [33,34]. Graphics of the molecular structures have been created by using ORTEP [35]. All non-hydrogen atoms were refined anisotropically, and a riding model was employed in the treatment of the hydrogen atom positions.

3.2. Synthesis of Trimethyl 3-[2-Methoxy-3,4-bis(methoxycarbonyl)-5,6,7,8,9,10-hexahydro-2H-2,4a-epoxybenzo[8]annulen-1-yl]cycloprop-1-ene-1,2,3-tricarboxylate (3b)

To a solution of 1a (1.00 g, 7.04 mmol, freshly distilled) in anhydrous CH2Cl2 (2.5 mL), cyclooctyne (189 mg, 1.75 mmol) was added with stirring at 0 °C. After 20 h at ambient temperature, the solvent and the excess of 1a were removed at reduced pressure (finally 1 h at 40 °C and 0.1 mbar), and the residue was purified by flash chromatography (SiO2, CHCl3/ethyl acetate 15:1) to give 3b (0.74 g, 79%) as a pale yellow solid, which was repeatedly crystallized from cyclohexane to yield colorless crystals with m.p. 90 °C that are appropriate for X-ray diffraction analysis. 1H-NMR (CDCl3): δ = 1.15–1.55 (m, 4H), 1.58–1.82 (m, 4H), 1.97–2.08 (m, 2H, 10'-H or 5'-H), 2.42‒2.62 (m, 2H, 10'-H or 5'-H), 3.52 (s, 3H, O-C-OCH3), 3.64 (s, 3H, OCH3), 3.75 (s, 3H, OCH3), 3.76 (s, 3H, OCH3), 3.84 (s, 3H, OCH3), 3.85 (s, 3H, OCH3). 13C-NMR (CDCl3): δ = 23.12 (t), 23.26 (t, 2C), 25.33 (t), 25.61 (t), 29.03 (t), 32.27 (s, C-3), 52.15 (q, 2C, OCH3), 52.75 (q, OCH3), 53.14 (q, OCH3), 53.17 (q, OCH3), 54.87 (q, O-C-OCH3), 89.42 (s, C-4a'), 114.19 (s), 115.65 (s, O-C-OCH3), 116.49 (s), 140.36 (s), 150.82 (s), 157.01 (s), 157.12 (s), 157.36 (s), 160.62 (s), 163.23 (s), 164.18 (s), 170.19 (s). IR (CDCl3): = 2954 (m), 1732 (s), 1268 (s) cm−1. Elemental Analysis calcd. for C26H30O12: C: 58.42%, H: 5.66%. Found: C: 58.45%, H: 5.63%.
Crystal Data for 3b: C26H30O12, M = 534.50 g·mol−1, monoclinic, C2/c, λ = 0.71073 Å, a = 31.6876 (13) Å, b = 10.1200 (3) Å, c = 16.1580 (6) Å, β = 101.909 (4) °, V = 5070.0 (3) (4) Å3, Z = 8, ρcalcd = 1.400 Mg∙m−3, μ = 0.112 mm−1, T = 100 (2) K, θ range 3.23°–25.00°, 9688 reflections collected, 4448 independent reflections (Rint = 0.0204), R1 = 0.0462, wR2 = 0.0865 (I > 2σ(I)).

3.3. Synthesis of Trimethyl 3-[2-Methoxy-3,4-bis(methoxycarbonyl)-5,6,9,10-tetrahydro-2H-2,4a-epoxybenzo[8]annulen-1-yl]cycloprop-1-ene-1,2,3-tricarboxylate (3c)

To cycloocten-5-yne (74 mg, 0.70 mmol) in CDCl3 (0.7 mL), 1a (400 mg, 2.79 mmol) was added. The mixture was allowed to stand at room temperature for 20 h. After removal of the solvent at reduced pressure, the residue was purified by chromatography (silica gel, dichloromethane/ethyl acetate 15:1) to give 3c (138 mg, 37%) as a slightly yellow oil. 1H-NMR (CDCl3): δ = 1.99–2.32 (m, 4H, CH2), 2.64–2.94 (m, 4H, CH2), 3.45 (s, 3H, O-C-OCH3), 3.61 (s, 3H, OCH3), 3.73 (s, 3H, OCH3), 3.77 (s, 3H, OCH3), 3.83 (s, 3H, OCH3), 3.84 (s, 3H, OCH3), 5.47–5.54 (m, 2H, CH=CH). 13C-NMR (CDCl3): δ = 24.03 (t), 24.07 (t), 27.55 (t), 27.64 (t), 31.53 (s), 52.03 (q, OCH3), 52.16 (q, OCH3), 52.63 (q, OCH3), 52.96 (q, OCH3), 53.07 (q, OCH3), 54.62 (q, O-C-OCH3), 91.06 (s, C-4a'), 113.51 (s), 115.23 (s, O-C-OCH3), 115.47 (s), 128.18 (d, CH=CH), 129.35 (d, CH=CH), 142.64 (s), 150.62 (s), 156.81 (s), 156.90 (s), 157.25 (s), 159.43 (s), 162.99 (s), 164.16 (s), 170.07 (s). IR (CDCl3): = 2954 (m), 1732 (s), 1270 (s) cm−1. HRMS: m/z calcd. for C26H28O12 ([M+H]+): 533.1659, found: 533.1592; calcd. for ([M+Na]+): 555.1478, found: 555.1470; calcd. for ([M+K]+): 571.1218, found: 571.1169.

3.4. Synthesis of Triethoxy 3-[2-Ethoxy-3,4-bis(ethoxycarbonyl)-5,6,7,8,9,10-hexahydro-2H-2,4a-epoxybenzo[8]annulen-1-yl]cycloprop-1-ene-1,2,3-tricarboxylate (3d)

To cyclooctyne (20 mg, 0.19 mmol) in CDCl3 (0.7 mL), 1b (0.13 g, 0.74 mmol) was added. The mixture was allowed to stand at room temperature for 20 h. After removal of the solvent at reduced pressure, the residue was purified by chromatography (silica gel, dichloromethane/ethyl acetate 15:1) to give 3d (70 mg, 61%) as a slightly yellow oil. 1H-NMR (CDCl3): δ = 1.13–1.20 (m, 6H, CH3), 1.25–1.35 (m, 12H, CH3), 1.39–1.83 (m, 8H, CH2), 2.01–2.12 (m, 2H, CH2), 2.48–2.67 (m, 2H, CH2), 3.70–3.78 (m, 1H, OCH2), 3.84–3.91 (m, 1H, OCH2), 3.99–4.08 (m, 1H, OCH2), 4.11–4.25 (m, 5H, OCH2), 4.27–4.36 (m, 4H, OCH2). 13C-NMR (CDCl3): δ = 13.99 (q, 2 C, CH3), 13.94 (q, CH3), 13.98 (q, CH3), 14.07 (q, CH3), 15.02 (q, CH3), 23.27 (t, CH2), 23.31 (t, CH2), 23.33 (t, CH2), 25.46 (t, CH2), 25.68 (t, CH2), 29.02 (t, CH2), 32.36 (s, C-3), 61.01 (t, OCH2), 61.08 (t, OCH2), 61.53 (t, OCH2), 62.22 (t, OCH2), 62.37 (t, OCH2), 63.59 (t, OCH2), 89.29 (s, C-4a'), 114.65 (s), 115.40 (s), 116.20 (s), 140.68 (s), 151.17 (s), 156.87 (s), 156.91 (s), 157.17 (s), 160.07 (s), 162.96 (s), 163.91 (s), 169.68 (s). IR (CDCl3): = 2984 (m), 1727 (s), 1260 (s) cm−1. HRMS: m/z calcd. for C32H42O12 ([M+H]+): 619.2755, found: 619.2697; calcd. for ([M+Na]+): 641.2574, found: 641.2608; calcd. for ([M+K]+): 657.2313, found: 657.2263.

3.5. Synthesis of Methyl (E)-2-(2,2-Dimethoxy-4,5,6,7,8,9-hexahydrocycloocta[b]furan-3(2H)-ylidene)acetate (4a)

To cyclooctyne (271 mg, 2.51 mmol) in anhydrous MeOH (2.5 mL), 1a (1.40 g, 9.85 mmol) was added. The mixture was stirred at room temperature for 20 h. After removal of the solvent at reduced pressure, the residue was purified by chromatography (silica gel, cyclohexane/tert-butyl methyl ether 14:1) to give 4a (426 mg, 60%) as a colorless solid with m.p. 60–61 °C (from cyclohexane/tert-butyl methyl ether). 1H-NMR (CDCl3): δ = 1.39–1.46 (m, 2H, 6'-CH2), 1.49–1.57 (m, 2H, 7'-CH2), 1.61–1.69 (m, 2H, 5'-CH2), 1.68–1.75 (m, 2H, 8'-CH2), 2.46 (m, 2H, 9'-CH2), 2.73 (t, J = 6.3, 2H, 4'-CH2), 3.33 (s, 6H, 2'-(OCH3)2), 3.69 (s, 3H, CO2CH3), 5.50 (s, 1H, 2-CH). 13C-NMR (CDCl3): δ = 22.2 (t, C-4'), 25.6 (t, C-6'), 26.5 (t, C-7'), 27.1 (t, C-9'), 28.7 (t, C-8'), 29.3 (t, C-5'), 51.0 (q, 1-OCH3), 51.3 (q, 2'-(OCH3)2), 106.3 (d, C-2), 112.5 (s, C-3a'), 121.7 (s, C-2'), 152.1 (s, C-3'), 166.3 (s, C-1), 170.3 (s, C-9a'). IR (film): = 2928 (s), 2853 (m), 1718 (s), 1642 (m), 1595 (s), 1458 (m), 1441 (m), 1353 (m), 1324 (m), 1277 (s), 1226 (s), 1167 (s), 1126 (s), 1086 (m), 1050 (s), 1022 (m), 1009 (m), 969 (m), 931 (m), 911 (m), 886 (w), 843 (m), 778 (w) cm−1. HRMS: m/z calcd. for C15H22O5 ([M+H]+): 283.1521, found: 283.1540; calcd. for ([M+Na]+): 305.1360, found: 305.1359. Elemental Analysis calcd. for C15H22O5: C: 63.81%, H: 7.85%, found: C: 63.82%, H: 7.90%.
Crystal Data for 4a: C15H22O5, M = 282.32 g·mol−1, monoclinic, I2/a, λ = 1.54184 Å, a = 18.1382 (3) Å, b = 11.3690 (2) Å, c = 15.9193 (4) Å, β = 118.261 (2) °, V = 2891.44(12) Å3, Z = 8, ρcalcd = 1.297 Mg∙m−3, μ = 0.798 mm−1, T = 110 (2) K, θ range 3.10°–61.986°, 4314 reflections collected, 2243 independent reflections (Rint = 0.0233), R1 = 0.0452, wR2 = 0.1172 (I > 2σ(I)).

3.6. Synthesis of Methyl d4-(E)-2-(2,2-Dimethoxy-4,5,6,7,8,9-hexahydrocycloocta[b]furan-3(2H)-ylidene)acetate (4b)

As described for 4a, cyclooctyne was analogously treated with CD3OD and 1a, and workup led to 4b (447 mg, 62%) as a colorless solid with m.p. 63–64 °C (from cyclohexane/tert-butyl methyl ether). 1H-NMR (CDCl3): δ = 1.35–1.42 (m, 2H, 6'-CH2), 1.46–1.53 (m, 2H, 7'-CH2), 1.57–1.65 (m, 2H, 5'-CH2), 1.64–1.71 (m , 2H, 8'-CH2), 2.42 (m, 2H, 9'-CH2), 2.69 (t, J = 6.3, 2H, 4'-CH2), 3.27 (s, 3H, 2'-OCH3), 3.63 (s, 3H, CO2CH3). 2H-NMR (CHCl3:CDCl3 = 9:1): δ = 3.30 (s, 3D, 2'-OCD3), 5.53 (s, 1D, 2-CD). 13C-NMR (CDCl3): δ = 22.0 (t, C-4'), 25.5 (t, C-6'), 26.4 (t, C-7'), 26.9 (t, C-9'), 28.6 (t, C-8'), 29.2 (t, C-5'), 50.4 (s (DEPT), sept with 1JCD = 22.0, 2'-OCD3), 50.9 (q, OCH3), 51.1 (q, OCH3), 105.8 (s (DEPT), t with 1JCD = 24.5, C-2), 112.4 (s, C-3a'), 121.6 (s, C-2'), 152.0 (s, C-3'), 166.1 (s, C-1), 170.2 (s, C-9a'). HRMS: m/z calcd. for C15H18D4O5 ([M+H]+): 287.1791, found: 287.1772; calcd. for ([M+Na]+): 309.1611, found: 309.1633; calcd. for ([M+K]+): 325.1350, found: 325.1367.

3.7. Synthesis of Methyl (E)-2-(2-methoxy-2-phenoxy-4,5,6,7,8,9-hexahydrocycloocta[b]furan-3(2H)-ylidene)acetate (4c)

To a solution of phenol (3.00 g, 31.9 mmol) and cyclooctyne (0.40 g, 3.7 mmol) in anhydrous THF (4 mL), 1a (1.05 g, 7.4 mmol) was added with the help of a syringe within 2 h. Thereafter, the mixture, which changed its color to deep red, was stirred at ambient temperature for 16 h. After removal of the solvent at reduced pressure, the residue was dissolved in Et2O (40 mL) and washed with aqueous sodium hydroxide (10%, 2 L). After drying of the organic phase (MgSO4) and removal of the solvent, the residue was purified by chromatography (first SiO2, CH2Cl2, then basic Al2O3, CH2Cl2) to furnish a green oil that was treated with hexane. This led to the precipitation of a colorless solid, which was removed by filtration. After removal of the solvent, 4c (247 mg, 10%) was obtained as a green oil that can be further purified by HPLC (LiChrospher Si 60 (5μ), 20 × 2 cm, CH2Cl2, 20 mL/min). 1H-NMR (CDCl3, relaxation delay d1 = 15 s): δ = 1.00–1.72 (m, 8H), 2.38–2.41 (m, 2H), 2.53–2.60 (m, 1H), 2.77–2.83 (m, 1H), 3.44 (s, 3H, 2'-OCH3), 3.70 (s, 3H, CO2CH3), 5.74 (s, 1H, 2-H), 7.06–7.25 (m, 5H, O-Ph). 13C-NMR (CDCl3): δ = 22.0 (t, C-4'), 25.1 (t, C-6'), 26.2 (t, C-7'), 26.9 (t, C-9'), 28.3 (t, C-8'), 29.1 (t, C-5'), 51.1 (q, 1-OCH3), 51.5 (q, 2'-OCH3), 107.5 (d, C-2), 112.7 (s, C-3a'), 121.1 (d, Ph), 122.3 (s, C-2'), 124.4 (d, Ph), 128.8 (d, Ph), 152.3 (s, C-3'), 152.4 (s, Ph), 166.2 (s, C-1), 169.7 (s, C-9a'). IR (CDCl3): = 2931 (s), 1705 (s, C=O) 1506 (s), 1490 (s) cm−1.

3.8. Synthesis of Dimethyl (E)-2-(2-oxocyclooctylidene)succinate [(E)-5] and Dimethyl (Z)-2-(2-oxocyclooctylidene)succinate [(Z)-5]

To THF (2 mL) and water (1 mL), cyclooctyne (0.41 g, 3.84 mmol) and 1a (0.56 g, 3.94 mmol) were added with stirring at 0 °C. Thereafter, the mixture was stirred at ambient temperature for 21 h. This led to the formation of an inhomogeneous mixture, which was diluted with diethyl ether (50 mL) and dried (MgSO4). After removal of the solvent under reduced pressure, the resulting yellow liquid was analyzed by 1H-NMR, which indicated a 5:1 ratio of (E)-5 and (Z)-5. By using flash chromatography (SiO2, Et2O/hexane 1:2), (E)-5 (0.22 g, 21%) and (Z)-5 (0.20 g, 20%) were isolated as colorless liquids.
(E)-5: 1H-NMR (CDCl3): δ = 1.42‒1.48 (m, 2H), 1.52‒1.58 (m, 2H), 1.61‒1.68 (m, 2H), 1.78‒1.86 (m, 2H), 2.46‒2.51 (m, 2H, 8'-H), 2.72‒2.76 (m, 2H, 3'-H), 3.25 (s, 2H, 3-H), 3.64 (s, 3H, OCH3), 3.74 (s, 3H, OCH3). 13C-NMR (CDCl3): δ = 23.64 (t, CH2), 25.51 (t, CH2), 25.77 (t, CH2), 27.56 (t, CH2), 32.44 (t, C-8'), 36.04 (t, C-3), 43.46 (t, C-3'), 52.03 (q), 52.14 (q), 122.34 (s), 154.13 (s), 167.33 (s), 170.79 (s), 212.43 (s, C-2'). IR (CCl4): = 1745 (s) cm−1. Elemental Analysis calcd. for C14H20O5: C: 62.67%, H: 7.51%. found: C: 62.37%, H: 7.46%.
(Z)-5: 1H-NMR (CDCl3): δ = 1.44–1.63 (m, 6H), 1.75‒1.82 (m, 2H), 2.40‒2.44 (m, 2H, 8'-H), 2.62‒2.66 (m, 2H, 3'-H) 3.38 (s, 2H, 3-H), 3.68 (s, 6H, 2 × OCH3). 13C-NMR (CDCl3): δ = 23.62 (t, CH2), 25.49 (t, CH2), 25.74 (t, CH2), 27.53 (t, CH2), 32.40 (t, C-8'), 36.00 (t, C-3), 43.43 (t, C-3'), 51.99 (q, OCH3), 52.09 (q, OCH3), 122.32 (s), 154.08 (s), 167.28 (s), 170.74 (s), 212.36 (s, C-2'). IR (CCl4): = 1745 (s) cm‒1. Elemental Analysis calcd. for C14H20O5: C: 62.67%, H: 7.51%, found: C: 62.43%, H: 7.36%.

3.9. Synthesis of Dimethyl 2-Methoxy-1-[(2-methoxycarbonyl)-3-methyloxiran-2-yl]-5,6,7,8,9,10-hexahydro-2H-2,4a-epoxybenzo[8]annulene-3,4-dicarboxylate (6a)

To a solution of cyclooctyne (410 mg, 3.79 mmol) in acetaldehyde (3 mL, freshly distilled), a solution of 1a (540 mg, 3.79 mmol) in anhydrous THF (2 mL) was added with the help of a syringe within 2 h. The yellow mixture was stirred at room temperature overnight. Thereafter, the solvent and the excess of acetaldehyde were removed at reduced pressure, and the resulting yellow oil was purified by flash chromatography (SiO2, Et2O/hexane 1:1). The oily product was recrystallized twice from Et2O/hexane (1:3) to give 6a (180 mg, 22%) as colorless crystals with m.p. 66–67 °C. Single crystals, which were appropriate for X-ray diffraction analysis, were obtained by slow evaporation of a solution of 6a in cyclohexane. 1H-NMR (CDCl3): δ = 1.38 (d, 3J = 5.6, 3H, C–CH3), 1.25–1.91 (m, 8H), 2.11 (m, 2H, 5-H), 2.51 (ddd, 3J10-H(a),9-H(a) = 15, 2J10-H(a),10-H(b) = 13.2, 3J10-H(a),9-H(b) = 4.8, 1H, 10-H(a)), 2.66 (dt, 2J10-H(a),10-H(b) = 13.2, 3J10-H(b),9-H(a) = 4, 3J10-H(b),9-H(b) = 4, 1H, 10-H(b)), 3.61 (s, 3H, OCH3), 3.68 (q, 3J3'-H,CH3 = 5.6, 1 H, 3'-H), 3.73 (s, 3H, OCH3), 3.779 (s, 3H, OCH3), 3.783 (s, 3H, OCH3). 13C-NMR (CDCl3): δ = 13.74 (q, CH3), 23.01 (t, CH2), 23.21 (t, CH2), 23.37 (t, CH2), 25.13 (t, CH2), 25.58 (t, CH2), 29.60 (t, CH2), 52.24 (q, 2 OCH3), 52.54 (q, OCH3), 55.01 (q, OCH3), 57.60 (s, C-2'), 57.86 (d, C-3'), 89.53 (s, C-4a), 115.48 (s, C-2), 139.33 (s), 150.29 (s), 156.52 (s), 163.22 (s), 163.91 (s), 164.18 (s), 168.35 (s). IR (CCl4): = 1709 (s) cm−1. Elemental Analysis calcd. for C22H28O9: C: 60.54%, H: 6.47%. found: C: 60.55%, H: 6.45%.
Crystal Data for 6a: C22H28O9, M = 436.44 g∙mol−1, monoclinic, P21/c, λ = 1.54184 Å, a = 8.8577 (4) Å, b = 28.5456 (16) Å, c = 17.3589 (9) Å, β = 90.063 (4) °, V = 4389.2 (4) Å3, Z = 8, ρcalcd = 1.321 Mg∙m−3, μ = 0.863 mm−1, T = 100 (2) K, θ range 3.10°−61.77°, 28823 reflections collected, 6811 independent reflections (Rint = 0.0557), R1 = 0.0548, wR2 = 0.1478 (I > 2σ (I)).

3.10. Synthesis of Dimethyl 2-Methoxy-1-[(2-methoxycarbonyl)-3-phenyloxiran-2-yl]-5,6,7,8,9,10-hexahydro-2H-2,4a-epoxybenzo[8]annulene-3,4-dicarboxylate (6b)

To a solution of cyclooctyne (270 mg, 2.53 mmol) in benzaldehyde (4 mL, freshly distilled), a solution of 1a (1.10 g, 7.58 mmol) in anhydrous THF (2 mL) was added with the help of a syringe within 2 h. The orange mixture was stirred overnight at room temperature. Thereafter, the solvent was evaporated and the excess of benzaldehyde was removed at 50 °C and 9 × 10−2 mbar. The resulting yellow oil was purified by flash chromatography (SiO2, Et2O/hexane 1:1). The oily product was crystallized from Et2O/hexane (ratio 1:3) to give 6b (260 mg, 21%) as colorless crystals. 1H NMR data indicated that the substance consisted of two diastereomers in a 4:1 ratio. IR (mixture of isomers): = 1711 (s) cm−1. Elemental Analysis (mixture of isomers) calcd. for C27H30O9: C: 65.05%, H: 6.07%, found: C: 64.77%, H: 6.02%. The diastereomers of 6b were (partly) separated by flash chromatography (SiO2, CH2Cl2); the minor isomer was eluted before the main isomer.
Minor isomer: 1H-NMR (CDCl3): δ = 1.25–2.10 (m, 8H), 2.18 (m, 2H, 5-H), 2.54 (ddd, 3J10-H(a),9-H(a) = 15.2, 2J10-H(a),10-H(b) = 13, 3J10-H(a),9-H(b) = 4.8, 1H, 10-H(a)), 2.72 (dt, 2J10-H(a),10-H(b) = 13, 3J10-H(b), 9-H(a) = 4, 3J10-H(b),9-H(b) = 4, 1H, 10-H(b)), 3.39 (s, 3H, OCH3), 3.75 (s, 3H), 3.779 (s, 3H), 3.783 (s, 3H), 4.32 (s, 1H, 3'-H), 7.28–7.41 (m, 5 H, Ph). 13C-NMR (CDCl3): δ = 23.05 (t, CH2), 23.44 (t, CH2), 23.69 (t, CH2), 25.37 (t, CH2), 25.77 (t, CH2), 30.11 (t, CH2), 52.08 (q, CH3), 52.24 (q, 2C), 55.40 (q, CH3), 60.66 (s, C-2'), 63.11 (d, C-3'), 89.18 (s, C-4a), 115.82 (s, C-2), 126.51 (d, 2C), 128.00 (d, 2C), 128.29 (d, CH), 133.07 (s), 139.82 (s), 150.77 (s), 157.34 (s), 162.24 (s), 163.01 (s), 164.35 (s), 166.51 (s).
Major isomer: M.p. 116–118 °C. 1H-NMR (CDCl3): δ = 1.31–2.22 (m, 10 H), 2.55 (ddd, 3J10-H(a),9-H(a) = 15.2, 2J10-H(a),10-H(b) = 13, 3J10-H(a),9-H(b) = 4.8, 1 H, 10-H(a)), 2.85 (dt, 2J10-H(a),10-H(b) = 13, 3J10-H(b),9-H(a) = 4, 3J10-H(b),9-H(b) = 4, 1H, 10-H(b)), 3.42 (s, 3H, OCH3), 3.70 (s, 3H), 3.79 (s, 6H), 4.71 (s, 1H, 3'-H), 7.28–7.41 (m, 5H, Ph). 13C-NMR (CDCl3): δ = 22.94 (t, CH2), 23.20 (t, CH2), 23.45 (t, CH2), 25.08 (t, CH2), 25.72 (t, CH2), 29.82 (t, CH2), 52.16 (q, CH3), 52.28 (q, 2C), 55.10 (q, CH3), 60.54 (s, C-2'), 61.89 (d, C-3'), 89.67 (s, C-4a), 115.54 (s, C-2), 126.28 (d, 2C), 128.06 (d, 2C), 128.34 (d, CH), 133.25 (s), 138.99 (s), 150.43 (s), 156.48 (s), 163.21 (s), 163.80 (s), 165.26 (s), 166.65 (s).

3.11. Synthesis of (Z)-Trimethylsilyl 2-(2-Oxo-4,5,6,7,8,9-hexahydrocycloocta[b]furan-3(2H)-ylidene)-2-(trimethylsilyl)acetate (7)

To a solution of bis(trimethylsilyl) but-2-ynedioate (0.10 g, 0.37 mmol) in anhydrous methylene chloride (2 mL), cyclooctyne (0.06 g, 0.55 mmol) was added under nitrogen atmosphere. The mixture was stirred for 48 h at 40 °C. The solvent and the excess of cyclooctyne were evaporated under reduced pressure to give 0.17 g (85%) of an unstable yellow oil. 1H-NMR (CDCl3): δ = 0.23 (s, 9H, CSi(CH3)3), 0.30 (s, 9H, CO2Si(CH3)3), 1.44–1.60 (m, 6H, CH2), 1.69–1.75 (m, 2 H, CH2), 2.31 (t, J = 6.4, 2 H, CH2), 2.48 (t, J = 7.1, 2 H, CH2). 13C-NMR(CDCl3): δ = −1.48 (q, CSi(CH3)3), −0.23 (q, OSi(CH3)3), 21.11 (t, CH2), 25.30 (t, CH2), 25.85 (t, CH2), 26.00 (t, CH2), 27.02 (t, CH2), 28.15 (t, CH2), 114.48 (s, C-3a'), 133.46 (s, C=C-COOC), 146.84 (s, C=CCOOC), 158.85 (s, C-9a'), 168.35 (s, CO), 171.09 (s, CO). Signal assignment and determination of the stereochemistry were supported by HMBC and NOESY experiments, respectively. 29Si-NMR (CDCl3): δ = −3.86 (CSi(CH3)3), 25.59 (OSi(CH3)3). IR (CCl4): = 3010 (w), 2933 (s), 1785 (s), 1688 (s), 1249 (s) cm−1. HRMS: m/z calcd. for: C18H30O4Si2 ([M+Na]+): 389.1580, found: 389.1643.

4. Conclusions

In summary, it was shown in this study that cyclooctynes and acetylenedicarboxylates react under mild conditions to generate short-lived dipolar intermediates of type 2 via an unusual 1,3-dipolar cycloaddition. These intermediates can be trapped by molecules with electron-deficient π-systems to give complex polycyclic products through a cascade of cycloaddition reactions. On the other hand, interception of 2 in the presence of protic reaction partners leads to 2,3-dihydrofuran derivatives with the rare substitution pattern of an orthoester functionality and a 3-methylidene group [36,37,38]. In all cases, the products are formed with perfect atom economy, that means that all atoms of the substrate molecules are found in the product.
Currently, we investigate scope and limitations of the presented method to prepare furan derivatives and whether acetylenedicarboxylates can be substituted by prop-2-ynoic esters with different electron-withdrawing substituents in 3-position or by the analogous amides to generate intermediates similar to 2 and the corresponding subsequent products.

Supplementary Materials

Supplementary materials can be accessed at: www.mdpi.com/1420-3049/19/9/14022/s1.
CCDC 1016789 (3b), 1016790 (4a) and 1016788 (6a) 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 (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44 1223 336033; E-mail: [email protected]).

Supplementary Files

Supplementary File 1

Acknowledgments

The publication costs of this article were funded by the German Research Foundation/DFG (Geschäftszeichen INST 270/219-1) and the Technische Universität Chemnitz in the funding programme Open Access Publishing. We are grateful to the Fonds der Chemischen Industrie for financial support. M.K. thanks the FCI for a Ph.D. fellowship.

Author Contributions

S.B., O.P., F.T., and T.W. performed the experiments. A.I. measured NMR spectra and drew the schemes. M.K., T.R. and H.L. performed X-ray diffraction studies. K.B. supervised the work and wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References and notes

  1. Neochoritis, C.G.; Zarganes-Tzitzikas, T.; Stephanidou-Stephanatou, J. Dimethyl Acetylenedicarboxylate: A versatile tool in organic synthesis. Synthesis 2014, 46, 537–585. [Google Scholar] [CrossRef]
  2. Sahoo, M.K. Dimethyl acetylene dicarboxylate. Synlett 2007, 2142–2143. [Google Scholar] [CrossRef]
  3. Baumgarth, M. Synthesen mit Acetylendicarbonsäureestern, II. Chem.-Ztg. 1976, 100, 515–528. [Google Scholar]
  4. Bastide, J.; Hamelin, J.; Texier, F.; vo Quang, Y. Cycloaddition dipolaire-1,3 aux alcynes. Bull. Soc. Chim. Fr. 1973, 2871–2887. [Google Scholar]
  5. Bastide, J.; Hamelin, J.; Texier, F.; vo Quang, Y. Cycloaddition dipolaire-1,3 aux alcynes. Bull. Soc. Chim. Fr. 1973, 2555–2579. [Google Scholar]
  6. Baumgarth, M. Synthesen mit Acetylendicarbonsäureestern. Chem.-Ztg. 1972, 96, 361–372. [Google Scholar]
  7. Stelmach, J.E.; Winkler, J.D. Dimethyl acetylenedicarboxylate. In Encyclopedia of Reagents for Organic Synthesis; Paquette, L., Ed.; Wiley: New York, NY, USA, 2004. [Google Scholar]
  8. LeGoff, E.; LaCount, R.B. A thermal tetramer of dimethyl acetylenedicarboxylate. Tetrahedron Lett. 1967, 8, 2333–2335. [Google Scholar] [CrossRef]
  9. Kauer, J.C.; Simmons, H.E. The tetramers of acetylenedicarboxylic esters. J. Org. Chem. 1968, 33, 2720–2726. [Google Scholar] [CrossRef]
  10. Hocking, M.B.; van der Voort Maarschalk, F.W. X-ray structures of triphenylphosphine and 1,2,5-triphenylphosphole products with dimethyl acetylenedicarboxylate tetramer. Can. J. Chem. 1994, 72, 2428–2442. [Google Scholar] [CrossRef]
  11. Winterfeldt, E.; Giesler, G. Formation of tetramethyl furantetracarboxylate from dimethyl acetylenedicarboxylate. Angew. Chem. Int. Edit. 1966, 5, 579. [Google Scholar] [CrossRef]
  12. Hashmi, A.S.K.; Grundl, M.A.; Rivas Nass, A.; Naumann, F.; Bats, J.W.; Bolte, M. Photochemical synthesis of prochiral dialkyl 3,3-dialkylcyclopropene-1,2-dicarboxylates with facial shielding substituents and related substrates. Eur. J. Org. Chem. 2001, 4705–4732. [Google Scholar]
  13. Winterfeldt, E.; Giesler, G. Die thermische Tetramerisierung des Acetylendicarbonsäuredimethylesters. Chem. Ber. 1968, 101, 4022–4031. [Google Scholar] [CrossRef]
  14. Marchand, A.P.; Reddy S.P., Sharma, R.; Gadgil, V.R.; Watson, W.H.; Kashyap, R.P. Diels–Alder cycloadditions of cyclopentadienes to DMAD tetramer. Tetrahedron 1993, 49, 987–994. [Google Scholar]
  15. Gericke, R.; Winterfeldt, E. Additionen an die Dreifachbindung XVI, Bildung und Reaktionen des tetrameren Acetylendicarbonsäuremethylesters. Tetrahedron 1971, 27, 4109–4116. [Google Scholar] [CrossRef]
  16. Tomioka, H.; Hirai, K.; Tabayashi, K.; Murata, S.; Izawa, Y; Inagaki, S.; Okajima, T. Neighboring group participation in carbene chemistry. Effect of neighboring carboxylate group on carbene reactivities. J. Am. Chem. Soc. 1990, 112, 7692–7702. [Google Scholar]
  17. Shibata, Y.; Noguchi, K.; Hirano, M.; Tanaka, K. Rhodium-catalyzed highly enantio- and diastereoselective cotrimerization of alkenes and dialkyl acetylenedicarboxylates leading to furylcyclopropanes. Org. Lett. 2008, 10, 2825–2828. [Google Scholar] [CrossRef]
  18. Kobayashi, M.; Tanaka, K. Rhodium-catalyzed linear cross-trimerization of two different alkynes with an alkene and two different alkenes with an alkyne. Chem. Eur. J. 2012, 18, 9225–9229. [Google Scholar] [CrossRef]
  19. Tietze, L.F.; Eicher, T. Reaktionen und Synthesen im Organisch-Chemischen Praktikum und Forschungslaboratorium, 2nd ed.; Thieme: Stuttgart, Germany, 1991; p. 40. [Google Scholar]
  20. Heber, D.; Rösner, P.; Tochtermann, W. Cyclooctyne and 4-cyclooctyn-1-ol—Versatile building blocks in organic synthesis. Eur. J. Org. Chem. 2005, 4231–4247. [Google Scholar]
  21. In the presence of aluminum trichloride, treatment of 1a with cyclooctyne leads to completely different products, see ref. [22,23].
  22. Gleiter, R.; Treptow, B. Doubly bridged prismanes, Dewar benzenes and benzene derivatives from cyclooctyne and 1,8-cyclotetradecadiyne: En route to propella[n3]prismanes. Angew. Chem. Int. Edit. 1990, 29, 1427–1429. [Google Scholar]
  23. Gleiter, R.; Treptow, B. Synthesis of doubly bridged prismanes: An investigation of the photochemical behavior of doubly bridged Dewar benzenes. J. Org. Chem. 1993, 58, 7740–7750. [Google Scholar] [CrossRef]
  24. For the synthesis of cycloocten-5-yne, see ref. [25,26].
  25. Petersen, H.; Meier, H. Synthese von Cycloocteninen. Chem. Ber. 1980, 113, 2383–2397. [Google Scholar] [CrossRef]
  26. Schmitt, M. Synthesen, Photolysen und Blitzvakuumpyrolysen gespannter Cycloalkine. Ph.D. Thesis, Universität Mainz, Mainz, Germany, July 1992. [Google Scholar]
  27. Carbene intermediate 2a was exclusively discussed in previous literature [8,10,11] to explain the formation of 3a. However, it remains inexplicable why 2a cannot be trapped by simple alkenes or alkynes without electron-withdrawing substituents.
  28. For a review on Brook rearrangement and recent work on retro-Brook rearrangement, see ref. [29] and [30,31,32], respectively.
  29. Brook, A.G. Some molecular rearrangements of organosilicon compounds. Accounts Chem. Res. 1974, 7, 77–84. [Google Scholar] [CrossRef]
  30. Gan, Z.; Wu, Y.; Gao, L.; Sun, X.; Lei, J.; Song, Z.; Li, L. Studies on retro-[1,4] Brook rearrangement of 3-silyl allyloxysilanes. Observation of the formation of unusual 3,3-bissilyl enols. Tetrahedron 2012, 68, 6928–6934. [Google Scholar] [CrossRef]
  31. Bailey, W.F.; Jiang, X. Stereochemistry of the cyclization of 4-(t-butyldimethyl)siloxy-5-hexenyllithium: Cis-selective ring-closure accompanied by retro-[1,4]-Brook rearrangement. ARKIVOC 2005, 6, 25–32. [Google Scholar] [CrossRef]
  32. Mori, Y.; Futamura, Y.; Horisaki, K. Regioselective aliphatic retro-[1,4]-Brook rearrangements. Angew. Chem. Int. Edit. 2008, 47, 1091–1093. [Google Scholar] [CrossRef]
  33. Sheldrick, G.M. SHELXL97; Program for the Refinement of Crystal Structures; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  34. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. A 2008, 64, 112–122. [Google Scholar]
  35. Farrugia, L.J. WinGX and ORTEP for Windows: An update. J. Appl. Crystallogr. 2012, 45, 849–854. [Google Scholar] [CrossRef]
  36. To the best of our knowledge, this rare substitution pattern is found in a few compounds only, see ref. [37,38].
  37. Saalfrank, R.W.; Röß, U.; Mehling, A. Synthese und Strukturbestimmung verschiedener phenanthrokondensierter O-Heterocyclen. Chem. Ber. 1984, 117, 666–671. [Google Scholar] [CrossRef]
  38. Saalfrank, R.W.; Ackermann, E.; Winkler, H.; Paul, W.; Böhme, R. Synthese und Eigenschaften o-chinoider Donor/Akzeptor-substituierter Allene. Kristall- und Molekülstruktur von E-2,2,2',2'-Tetraethoxy-Δ3,3'(2H,2'H)-bi[phenanthro[9,10-b]furan]. Chem. Ber. 1980, 113, 2950–2958. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds 3b, 6a and 6b are available from the authors.

Share and Cite

MDPI and ACS Style

Banert, K.; Bochmann, S.; Ihle, A.; Plefka, O.; Taubert, F.; Walther, T.; Korb, M.; Rüffer, T.; Lang, H. Synthesis with Perfect Atom Economy: Generation of Furan Derivatives by 1,3-Dipolar Cycloaddition of Acetylenedicarboxylates at Cyclooctynes. Molecules 2014, 19, 14022-14035. https://doi.org/10.3390/molecules190914022

AMA Style

Banert K, Bochmann S, Ihle A, Plefka O, Taubert F, Walther T, Korb M, Rüffer T, Lang H. Synthesis with Perfect Atom Economy: Generation of Furan Derivatives by 1,3-Dipolar Cycloaddition of Acetylenedicarboxylates at Cyclooctynes. Molecules. 2014; 19(9):14022-14035. https://doi.org/10.3390/molecules190914022

Chicago/Turabian Style

Banert, Klaus, Sandra Bochmann, Andreas Ihle, Oliver Plefka, Florian Taubert, Tina Walther, Marcus Korb, Tobias Rüffer, and Heinrich Lang. 2014. "Synthesis with Perfect Atom Economy: Generation of Furan Derivatives by 1,3-Dipolar Cycloaddition of Acetylenedicarboxylates at Cyclooctynes" Molecules 19, no. 9: 14022-14035. https://doi.org/10.3390/molecules190914022

Article Metrics

Back to TopTop