Abstract
The Ru-catalyzed intramolecular oxidative amidation (lactamization) of aromatic alkynylamines with 4-picoline N-oxide as an external oxidant has been developed. This chemoselective process is very efficient to achieve medium-sized ε- and ζ-lactams (seven- and eight-membered rings) but not for the formation of common δ-lactams (six-membered rings). DFT studies unveiled the capital role of the chain length between the amine and the alkyne functionalities: the longer the connector, the more favored the lactamization process vs hydroamination.
Introduction
Amines and their derivatives are present in a myriad of bioactive natural products and pharmaceuticals.1 Among all the synthetic strategies to these compounds,2 transition metal-catalyzed addition of amines (hydroamination) and amides (hydroamidation) to unsaturated C–C bonds have emerged as powerful sustainable tools to access such functionalities.3 In the particular case of alkynes, their electrophilic activation by π-coordination to transition metals make them prone to receive nucleophilic attacks, generating enamines or enamides.4 Apart from chemoselectivity, efficient strategies to control the regioselectivity of terminal alkynes have been developed to produce Markovnikov-type products.3d On the contrary, generation of catalytic metal-vinylidenes from the initially coordinated metal-alkyne allows the polarization of the unsaturated bond in such a way that the carbon α to the metal center is now electrophilic affording exclusively the anti-Markovnikov addition products.5 In cases where an oxidizing nucleophile is present, the oxygen atom is transferred to the carbon-metal bond of the vinylidene complex, entailing a very reactive ketene intermediate, which could be subsequently trapped by other nucleophiles (e.g., amine) present in the media (Scheme 1a).6 A pioneer work by Kim and Lee and Zhang et al. showed that ketenes generated from Rh- and Ru-vinylidenes in the presence of oxidants could be intermolecularly trapped with heteronucleophiles (e.g., amines, alcohols) or be involved in [2 + 2] cycloadditions, respectively.7 Recently, our group has successfully developed a general Ru-catalyzed intermolecular oxidative amidation of alkynes8 with all types of aliphatic and aromatic amines, which allowed for the synthesis of a variety of primary and secondary amides in good to excellent yields.
Scheme 1. (a) Metal-Catalyzed Oxidative Amidation of Alkynes and (b) Ru-Catalyzed Oxidative Lactamization of Alkynylamines.
More challenging would be the intramolecular processes since the desired intramolecular oxidative amidation (oxidative lactamization) would have to compete with the intramolecular hydroamination reaction (capture of the electrophilic vinylidene).9
Herein, we describe the chemoselective Ru-catalyzed oxidative lactamization10 of aromatic 1,5- and 1,6-alkynylamines, which provides efficient access to valuable seven-membered 1-, 2-, and 3-benzazepin-2-, 3-, and 4-ones,11 privileged scaffolds in natural products and pharmaceuticals,12 and eight-membered 4-benzazocin-5-ones, respectively13 (Scheme 1b).
Results and Discussion
We started our investigation by testing the oxidative lactamization of N-(2-ethynylphenethyl)propan-1-amine 1a (a 1,5-alkynylamine) under our previously optimized intermolecular oxidative amidation conditions for secondary amines (Table 1).8
Table 1. Optimization of the Reaction Conditionsa.
| entry | substrate | cat (%) | N-oxide (equiv) | KPF6 (equiv) | 2a (% yield) |
|---|---|---|---|---|---|
| 1b | 1a | 5 | 2 | 1 | dec |
| 2bc | 1a | 5 | 2 | dec | |
| 3 | 1a·HCl | 10 | 2 | 30 | |
| 4 | 1a·HCl | 10 | 2 | 0.3 | 50 |
| 5 | 1a·HCl | 10 | 2 | 1 | 58 |
| 6 | 1a·HCl | 10 | 1.1 | 1 | 92 (90)d |
| 7 | 1a·HCl | 3 | 1.1 | 1 | (97)def |
Typical conditions: 0.2 mmol of substrate, 2 mL of DCE, sealed tube, NMR yields.
Without K2CO3.
Similar results were obtained at 70 and 50 °C. At rt, the starting material was recovered.
Isolated yield.
100 °C, 30 min.
60 °C, 1 h. dec: decomposition.
To our initial surprise, heating a solution of 1a in DCE at 100 °C for 4 h in the presence of 5 mol % CpRuCl(PPh3)2 and 2 equiv of 4-picoline N-oxide with or without 1 equiv of KPF6 (entries 1 and 2) led to decomposition. Variation of temperature conditions from 100 °C to rt resulted again in decomposition14 or starting material recovery (entry 2 and Table S1a).15 Interestingly, using the ammonium salt 1a·HCl, to favor a controlled release of the amine during the reaction, allowed us to obtain the desired seven-membered 3-benzazepin-4-one 2a albeit in a low yield (entry 3 and Table S1b).16 We then analyzed the use of salt additives in the course of the reaction (Table S1b).8 Thus, addition of KPF6 (0.3 equiv) gave a moderate yield of 2a (entry 4), which could be slightly increased to 58% yield when 1 equiv of KPF6 was added (entry 5).17 These results would suggest that KPF6 might be involved not only in the formation of a cationic ruthenium catalyst but also as a counterion of the ammonium substrate.18 Interestingly, the amount of oxidant employed had a significant effect in the efficiency of the reaction (Table S1e). Thus, when the amount of N-oxide was reduced to 1.1 equiv (entry 6) an excellent 90% yield of 3-benzazepin-4-one 2a was obtained.19 To our delight, excellent yields of 2a were also obtained when loadings of 3% of catalyst were used even at 60 °C (entry 7 and Table S1f).20
Thus, the ammonium salts of (o-propynylphenyl)methanamines 3a·HCl (R = Me) and 3b·HCl (R = Bn) smoothly underwent the oxidative lactamization to give the corresponding 2-benzazepin-3-ones 4a and 4b in reasonable good yields (Scheme 2). As in the case of intermolecular oxidative amidation of anilines,8 the free primary and secondary o-butynylanilines 5a and 5b (R = H and Me, X = CH2) were readily converted into the corresponding 1-benzazepin-2-ones 6a and 6b in fairly good yields. The presence of a heteroatom in the linker is well tolerated. Thus, the free primary and secondary (o-propynyloxy)anilines 5c (R = H, X = O), 5d (R = Me, X = O), and 5e (R = Bn, X = O) smoothly cyclized to give the corresponding 1,4-benzoxazepinones 6c–e in moderate to good yields. In the case of the free primary (o-propynylthio)aniline 5f (R = H, X = S), a slow conversion was observed to give the 1,4-benzothiazepinone 6f in a moderate 43% yield.
Scheme 2. 1- and 2-Benzazepin-2- and -3-ones by Ru-Catalyzed Oxidative Lactamization of Aromatic 1,5-Alkynylamines.
Ammonium salt 3a·HCl, 0.3 equiv K2CO3.
Ammonium salt 3b·HCl, 0.3 equiv K2CO3.
Pleasingly, oxidative lactamization seems a very reliable process for the construction of medium-sized lactams since the higher homolog 1,6-alkynylamine, as ammonium salt 1c·HCl, smoothly afforded the corresponding eight-membered 3-benzazocinone 7 in an excellent 90% yield (Scheme 3). Unfortunately, the next homolog member 4-(2-ethynylphenyl)-N-propylbutan-1-amine failed to react due probably to an unfavorable Thorpe–Ingold effect. Conversely, the ammonium salt of the lower homolog 1,4-alkynylamine 1b·HCl slowly decomposed under the same conditions (Scheme 3). This last unexpected result led us to examine the competitive intramolecular hydroamination reaction of alkynylamines 1a·HCl and 1b·HCl (Scheme 3). Thus, while the 1,5-alkynylamine 1a·HCl gave the desired 3-benzazepine 8a but in a low 14% yield, the lower homolog 1,4-alkynylamine 1b·HCl gave, as somewhat expected, decomposition.9a
Scheme 3. Ru-Catalyzed Oxidative Lactamization of Aromatic 1,4- and 1,6-Alkynylamines 1b·HCl and 1c·HCl and Hydroamination of Aromatic 1,5- and 1,4-Alkynylamines 1a·HCl and 1b·HCl.
Same result without 4-Pic-N-oxide.
We speculate that the benzylamine derivatives could undergo competitive intramolecular C–H activation processes that lead to complex mixtures or decomposition. For this reason, we turned our attention to the evaluation of alkynylamide derivatives in the cyclization processes (Scheme 4). Thus, the 1,5-alkynylamide 9a underwent an oxidative lactamization to afford the imide 10 (benzazepine-2,4(3H)-dione) in a quite good yield, and the lower homolog 1,4-alkynylamide 9b gave the hydroamination product 11 (isoquinolin-1-one) in a moderate 48% yield under the same oxidative conditions or a 83% yield in the absence of 4-Pic-N-oxide (Scheme 4).21 By contrast, the reaction of 1,5-alkynylamide 9a in the absence of 4-Pic-N-oxide gave a complex mixture.
Scheme 4. Ru-Catalyzed Oxidative Lactamization and Intramolecular Hydroamination of 1,5-Alkynylamide 9a and 1,4-Alkynylamide 9b.
In the absence of 4-Pic-N-oxide.
The novelty of the transformation prompted us to undertake a DFT mechanistic analysis to unveil the origin of the chemoselectivity observed toward oxidative lactamization vs intramolecular hydroamination of alkynylamines.22 For this study, both 1,5-alkynylamine 1a and its lower homolog 1,4-alkynylamine 1b have been used as model substrates. Following our previous mechanistic proposal,8 the catalytic cycle for the oxidative lactamization can be divided in three main steps (Scheme 5, blue): (i) Ru(II)-catalyzed vinylidene formation, (ii) Ru(II)-vinylidene oxidation by external 4-picoline N-oxide, and (iii) metal-free lactamization of the generated ketene. On the other hand, the competitive Ru(II)-hydroamination processes toward the formation of dihydro-3-benzazepine 8a and dihydroisoquinoline 8b (not experimentally observed) would proceed through trapping of the initially formed Ru-vinylidenes by intramolecular hydroamination (Scheme 5, green).
Scheme 5. Competitive Mechanisms for the Ru(II)-Catalyzed Oxidative Lactamization (blue) and Ru(II)-Catalyzed Alkyne Intramolecular Hydroamination (Green) of 1,5- and 1,4-Alkynylamines.
The computed Gibbs energy profile for the formation of Ru-vinylidenes from 1a (black color) and 1b (red color) is depicted in Figure 1.
Figure 1.
Gibbs energy profiles (DLPNO-CCSD(T)/def2-TZVP//B3LYP-D3/6-31G(d,p)-SDD(Ru)DCE(SMD) 373 K) for the Ru(II)-vinylidenes IVa and IVb. Energies are relative to [CpRu(PPh3)2]+I and alkynylamines 1a and 1b, respectively, and are mass balanced.
The DFT computed mechanism starts with the chloride release forming the active catalyst [CpRu(PPh3)2]+, I, entailing the formation of the η2-alkyne intermediate IIa and IIb at −2.4 and −0.2 kcal·mol–1, respectively. The activation of the substrates goes through the well-established 1,2-hydrogen migration favored by the agostic interaction with the methylidyne C–H bond in intermediates III.23,24 The energy barrier, TSII–III, is quite affordable requiring 6.6 and 6.7 kcal·mol–1 for the formation of the agostic intermediates IIIa and IIIb falling at 2.9 and 5.3 kcal·mol–1, respectively. Intermediates III evolve through a 1,2-hydrogen migration overcoming energy barriers of 11.3 and 13.1 kcal·mol–1 to the Ru(II)-vinylidene intermediates IVa and IVb at −13.4 and −12.7 kcal mol–1, respectively. Experimentally, in a deuteration study of the Ru-catalyzed heterocyclization of aromatic bis-homopropargyl alcohols, we had already shown results that indicate the formation of ruthenium vinylidene species as key intermediates.25 In addition, other competitive cyclization experiments also showed the involvement of vinylidene intermediates. Thus, heterocyclization of 1,4-alkynylanilide with catalytic CpRuCl(PPh3)2 gave rise to the corresponding 1,4-dihydroquinolinone, a product derived from hydroamination through a Ru-vinylidene intermediate, while with catalytic RuCl2(p-cymene)2/PBu3 gave the indol derivative, most likely through a typical alkyne activation/aromatization.9a
We next turned our attention to the Ru(II)-catalyzed vinylidene oxidation from IVa and IVb. The Gibbs energy profiles for IVa (black pathway) and IVb (red pathway) are depicted in Figure 2 (right side of the profile).
Figure 2.

Gibbs energy profiles (DLPNO-CCSD(T)/def2-TZVP//B3LYP-D3/6-31G(d,p)-SDD(Ru)DCE(SMD) 373 K) for the Ru(II)-catalyzed oxidation with picoline N-oxide (right) and for the intramolecular Ru(II)-catalyzed hydroamination of Ru(II)-vinylidenes IVa and IVb (left). Truncated structure of TSIVa–Va (inset draw). Energies are relative to [CpRu(PPh3)2]+I and alkynylamines 1a and 1b, respectively, and are mass balanced.
The nucleophilic attack of picoline N-oxide to the Cα of the Ru-vinylidene (ΔG‡ = 22.2 and 20.0 kcal mol–1 for TSIVa–Va and TSIVb–Vb, respectively) takes place in a coplanar approach to the Ru-vinylidene (Figure 2, inset drawing), affording intermediate Va and Vb at −0.5 and −1.8 kcal mol–1, respectively. Then, release of 4-picoline occurs (ΔG‡ = 6.7 and 7.3 kcal mol–1 for both TSVa–VIa and TSVb–VIb, respectively) to afford η2-coordinated ruthenium ketene complexes VIa and VIb,26 which further evolve to the more stable free ketenes VIIa and VIIb with the recovery of the catalytic species I in a global exergonic process.
Competitive hydroamination pathways for both Ru(II)-vinylidenes IVa and IVb (Figure 2, left) were then analyzed. The initial step is the intramolecular nucleophilic attack of the amine to the Cα of the Ru(II)-vinylidenes IV to yield the cationic alkenylruthenium complexes VIIIa and VIIIb. The transition state of this step resulted in being significantly more favorable for vinylidene IVb (ΔΔG‡ = (27.3 – 14.6) = 12.7 kcal·mol–1) presumably due to the less entropic penalty caused by the closer proximity of the reactive centers and the formation of a more stable six-membered ring intermediate. Similarly, from the formed cationic heterocycles VIII, the subsequent bicarbonate assisted deprotonation followed by protonolysis of the Ru–C bonds and Ru-decoordination displays a lower energy barrier (ΔΔG‡ = (13.2 – 11.2) = 2.0 kcal mol–1, TSVIIIa–8avsTSVIIIb–8b) for the formation of the dihydroisoquinoline 8b rather than dihydrobenzazepine 8a.27
Although from vinylidene IVb the intramolecular hydroamination to 8b is more energetically favored than oxidative amidation with overall Gibbs energy barriers of 14.6 and 20.0 kcal mol–1, respectively, the cyclization of benzylamine 1b·HCl gave decomposition products through other non-analyzed competitive pathways.9a By contrast, from vinylidene IVa, oxidative lactamization to 2a is more favorable than hydroamination with overall barriers of 22.2 and 27.3 kcal mol–1, respectively.28
Finally, to have a complete picture of the oxidative lactamization process, pathways involving metal-bound ketenes VI and metal-free ketenes VII as intermediates were evaluated.29 It was found that the most favorable pathways involve the metal-free ketenes VII with a water molecule acting as a proton carrier (Figure 3),30 whereas that of metal-bound ketene is higher in energy (Figures S2–S5, Supporting Information).27 Thus, intramolecular nucleophilic attack of the amine to the central carbon in ketene VIIa (ΔG‡ = 12.2 kcal mol–1) affords enol–water complex IXa-H2O (ΔG° = 3.5 kcal mol–1),31 which after keto-enol tautomerization (ΔG‡ = 17.8 kcal mol–1) and water release gives rise to the ε-lactam 2a (ΔG° = −34.4 kcal mol–1).
Figure 3.
Gibbs energy profile (DLPNO-CCSD(T)/def2-TZVP//B3LYP-D3/6-31G(d,p)-SDD(Ru)DCE(SMD) 373 K) for the lactamization of metal-free ketene VIIa with water acting as a proton carrier. Energies are relative to [CpRu(PPh3)2]+I and alkynylamine 1a and are mass balanced.
Once the complete cyclization energetic profiles of alkynylamines were analyzed, we calculate the cyclization rate-determining steps of alkynylamides 9a and 9b (Figure 4). Starting from vinylidene XVIa, oxidative lactamization to 10 (Scheme 4) is more favorable than hydroamination with overall barriers of 22.5 and 33.0 kcal mol–1, respectively. Conversely, from vinylidene XVIb, the intramolecular hydroamination to 11 (Scheme 4) is more favorable than oxidative amidation with overall Gibbs energy barriers of 23.2 and 28.8 kcal mol–1, respectively. According to these results, the chain length between the alkyne and amide functionalities is crucial for the chemoselectivity found. The effects caused by the longer carbon-tethered in 9a (1,5-alkynylamide) vs9b (1,4-alkynylamide) slows down the intramolecular direct nucleophilic attack of the nitrogen to the vinylidene making the intermolecular addition of the picoline N-oxide a competitive process.
Figure 4.
Gibbs energy profiles (DLPNO-CCSD(T)/def2-TZVP//B3LYP-D3/6-31G(d,p)-SDD(Ru)DCE(SMD) 373 K) for the Ru(II)-catalyzed nucleophilic addition of picoline N-oxide (right) and the intramolecular nucleophilic addition of nitrogen to vinylidene (left) of Ru(II)-vinylidenes species XVIa and XVIb derived from alkynylamides 9a and 9b, respectively. Energies are relative to [CpRu(PPh3)2]+I and alkynylamines 9a and 9b, respectively, and are mass balanced.
Conclusions
Medium-sized benzofused seven-membered lactams (1-, 2-, 3-benzazepin-2-, -3-, and -4-ones) and eight-membered lactams (3-benzazocinone) could be efficiently assembled by chemoselective Ru-catalyzed oxidative lactamization of aromatic 1,5-alkynylamines and 1,6-alkynylamines with 4-picoline N-oxide as an external oxidant. DFT mechanistic studies revealed that the oxidation of the catalytic Ru-vinylidene intermediate occurs via a coplanar intermolecular nucleophilic attack of the 4-picoline N-oxide to the electrophilic α carbon. The chain length between the alkyne and the amine functionalities (1,5- and 1,6-alkynylamines) proved to be crucial for the chemoselective oxidation since only ε- and ζ-lactams could be obtained but not δ-lactams. In the case of 1,4-alkynylamides, the hydroamination product was obtained on being more favorable the intramolecular nucleophilic attack to the Ru-vinylidene.
Experimental Section
All reactions were performed under an argon atmosphere, and the glassware was oven dried at 80 °C or flame dried unless otherwise stated. All dry solvents were stored under an argon atmosphere and over 4 Å molecular sieves. All chemicals were purchased from Acros Organics, TCI Chemicals, Sigma-Aldrich, Alfa Aesar, or Strem Chemicals chemical companies and used without further purification. The catalyst for Sonogashira coupling, PdCl2(PPh3)2, was prepared according to previously published procedure.
For the catalytic reactions, the CpRuCl(PPh3)2 catalyst used was purchased from Strem; the 4-picoline-N-oxide was purchased from Aldrich, and the DCE anhydrous was purchased from Acros Organics.
Analytical thin-layer chromatography was carried out on silica-coated aluminum plates (silica gel 60 F254 Merck) using UV light as a visualizing agent (254 nm) and KMnO4 (solution of 1.5 g of potassium permanganate, 10 g of potassium bicarbonate and 1.25 mL of 10% sodium hydroxide in 200 mL of water) or p-anisaldehyde (solution of 3.7 mL of p-anisaldehyde, 1.5 mL of glacial acetic acid, 5 mL of conc. sulfuric acid in 135 mL of absolute ethanol) with heat as developing agents. Flash column chromatography was performed on silica gel 60 (Merck, 230–400 mesh) with the indicated eluent.
1H and 13C nuclear magnetic resonance experiments were carried out using a Varian Inova 400 MHz or a Varian Mercury 300 MHz. Coupling constants J are given in Hertz (Hz). Multiplicities are reported as follows: s = singlet, bs = broad singlet, d = doublet, t = triplet, q = quartet, sxt = sextet, m = multiplet, or as a combination of them. Multiplicities of 13C NMR signals were determined by DEPT experiments. Yields refer to isolated compounds estimated to be >95% pure as determined by 1H NMR.
Ru-Catalyzed Oxidative Lactamizations
General Procedure A
A suspension of the corresponding alkynylamine (1 equiv), CpRuCl(PPh3)2 (0.03 equiv), 4-Pic-N-oxide (1.1 equiv), and KPF6 (1 equiv) in DCE was heated in a screw-cap vial until complete disappearance of a starting material (TLC monitoring). The resulting mixture was washed with saturated solution of CuSO4 and extracted with DCM (3 × 10 mL). The combined organic layers were dried (Na2SO4) and concentrated in vacuo. The residue was purified by silica gel flash column chromatography to yield the corresponding lactam.
General Procedure B
A suspension of the corresponding alkynylamine hydrochloride (1 equiv), CpRuCl(PPh3)2 (0.03 equiv), K2CO3 (0.3 equiv), 4-Pic-N-oxide (1.1 equiv), and KPF6 (1 equiv) in DCE was heated in a screw-cap vial until complete disappearance of the starting material (TLC monitoring). The resulting mixture was washed with saturated solution of CuSO4 and extracted with DCM (3 × 10 mL). The combined organic layers were dried (Na2SO4) and concentrated in vacuo. The residue was purified by silica gel flash column chromatography to yield the corresponding lactam.
3-Propyl-1,3,4,5-tetrahydro-2H-benzo[d]azepin-2-one (2a)
Procedure B. The product was purified by silica gel chromatography (EtOAc/Hex 1:1). Compound 2a, 0.04 g (97%), yellow oil. 1H NMR (500 MHz, CDCl3): δ 7.18–7.14 (m, 1H), 7.13–7.10 (m, 1H), 7.10–7.06 (m, 2H), 3.89 (s, 2H), 3.77–3.61 (m, 2H), 3.44–3.30 (m, 2H), 3.15–3.09 (m, 2H), 1.58 (sxt, J = 7.4 Hz, 2H), 0.89 (t, J = 7.4 Hz, 3H). 13C{1H} NMR, DEPT (126 MHz, CDCl3): δ 171.8 (CO), 135.9(C), 131.8(C), 131,1 (CH), 130.3 (CH), 127.1 (CH), 126.6 (CH), 48.7 (CH2), 46.5 (CH2), 43.4 (CH2), 32.9 (CH2), 21.6 (CH2), 11.4 (CH3). MS (CI), m/z (%): 204 (M + 1, 100). HRMS (EI-TOF) m/z: [M]+ calcd for C13H17NO: 203.1310; found: 203.1315.
2-Methyl-1,2,4,5-tetrahydro-3H-benzo[c]azepin-3-one (4a)
Known compound.32 Procedure B. The product was purified by silica gel chromatography (EtOAc/Hex 7:3). Compound 4a, 0.024 g (69%), yellow oil. 1H NMR (500 MHz, CDCl3): δ 7.25–7.22 (m, 1H), 7.15–7.06 (m, 3H), 4.48 (s, 2H), 3.16 (t, J = 6.8 Hz, 2H), 3.04 (s, 3H), 2.91 (t, J = 6.8 Hz, 2H). 13C{1H} NMR, DEPT (126 MHz, CDCl3): δ 173.6 (CO), 137.7 (C), 134.2 (C), 130.5 (CH), 128.8 (CH), 128.1 (CH), 125.9 (CH), 54.4 (CH2), 35.2 (CH3), 33.5 (CH2), 28.7 (CH2).
2-Benzyl-1,2,4,5-tetrahydro-3H-benzo[c]azepin-3-one (4b)
Procedure B. The product was purified by silica gel chromatography (EtOAc/Hex 3:7). Compound 4b, 0.027 g (53%), colorless oil. 1H NMR (300 MHz, CDCl3): δ 7.63–7.53 (m, 4H), 7.21–7.13 (m, 3H), 7.06–7.04 (m, 1H), 6.85–6.82 (m, 1H), 4.66 (s, 2H), 4.40 (s, 2H), 3.22 (t, J = 6.8 Hz, 2H), 3.00 (t, J = 6.8 Hz, 2H). 13C{1H} NMR, DEPT (75 MHz, CDCl3): δ 173.1 (CO), 140.4 (C), 138.2 (C), 136.5 (C), 129.5 (CH), 129.3 (CH), 128.4 (2× CH), 127.9 (2× CH), 127.8 (2× CH), 127.7 (CH), 50.0 (CH2), 49.1 (CH2), 34.4 (CH2), 28.8 (CH2). HRMS (EI-TOF) m/z: [M]+ calcd for C17H17NO: 251.1310; found: 251.1316.
1,3,4,5-Tetrahydro-2H-benzo[b]azepin-2-one (6a)
Known compound.33 Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 3:7). Compound 6a, 0.017 (53%), yellow oil. 1H NMR (400 MHz, CDCl3): δ 7.70 (bs, 1H), 7.27–7.19 (m, 2H), 7.13 (td, J = 7.3, 1.3 Hz, 1H), 7.01–6.93 (m, 1H), 2.81 (t, J = 7.2 Hz, 2H), 2.36 (t, J = 7.2 Hz, 2H), 2.29–2.17 (m, 2H). 13C{1H} NMR, DEPT (101 MHz, CDCl3): δ 175.2 (CO), 137.9 (C), 134.5 (C), 130.0 (CH), 127.6 (CH), 125.9 (CH), 121.9 (CH), 32.9 (CH2), 30.5 (CH2), 28.6 (CH2). MS (CI), m/z (%): 162 (M + 1, 100).
1-Methyl-1,3,4,5-tetrahydro-2H-benzo[b]azepin-2-one (6b)
Known compound.34 Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 3:7). Compound 6b, 0.022 g (62%), yellow oil. 1H NMR (500 MHz, CDCl3): δ 7.29 (ddd, J = 7.9, 7.0, 1.8 Hz, 1H), 7.21–7.13 (m, 3H), 3.35 (s, 3H), 2.72 (t, J = 7.2 Hz, 2H), 2.30 (t, J = 7.2 Hz, 2H), 2.19–2.14 (m, 2H). 13C{1H} NMR, DEPT (126 MHz, CDCl3): δ 173.5 (CO), 143.9 (C), 135.3 (C), 129.4 (CH), 127.6 (CH), 126.2 (CH), 122.4 (CH), 35.3 (CH3), 33.3 (CH2), 30.2 (CH2), 29.0 (CH2). MS (CI), m/z (%): 176 (M + 1, 100). HRMS (CI-TOF) m/z: [M + H]+ calcd for C11H14NO: 176.1070; found: 176.1064.
2,3-Dihydrobenzo[b][1,4]oxazepin-4(5H)-one (6c)
Known compound.35 Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 1:1 to EtOAc). Compound 6c, 0.017 g (53%), yellow oil. 1H NMR (400 MHz, CDCl3): δ 8.28 (bs, 1H), 7.08–7.00 (m, 3H), 6.99–6.94 (m, 1H), 4.46 (t, J = 5.7 Hz, 2H), 2.86 (t, J = 5.7 Hz, 2H). 13C{1H} NMR, DEPT (101 MHz, CDCl3): δ 172.9 (CO) 148.6 (C), 128.9 (C), 125.5 (CH), 123.8 (CH), 122.2 (CH), 121.7 (CH), 68.8 (CH2), 36.9 (CH2). MS (CI), m/z (%): 164 (M + 1, 100).
5-Methyl-2,3-dihydrobenzo[b][1,4]oxazepin-4(5H)-one (6d)
Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 1:1 to EtOAc). Compound 6d, 0.021 g (58%), yellow oil. 1H NMR (400 MHz, CDCl3): δ 7.21–7.07 (m, 4H), 4.57 (t, J = 6.6 Hz, 2H), 3.35 (s, 3H), 2.64 (t, J = 6.6 Hz, 2H). 13C{1H} NMR, DEPT (101 MHz, CDCl3): δ 171.0 (CO), 149.5 (C), 138.3 (C), 127.0 (CH), 125.2 (CH), 123.1 (CH), 122.9 (CH), 74.5 (CH2), 35.2 (CH2), 35.0 (CH3). MS (CI), m/z (%): 178 (M + 1, 100). HRMS (CI-TOF) m/z: [M + H]+ calcd for C10H12NO2: 178.0863; found: 178.0860.
5-Benzyl-2,3-dihydrobenzo[b][1,4]oxazepin-4(5H)-one (6e)
Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 3:7). Compound 6e, 0.035 g (70%), yellow oil. 1H NMR (500 MHz, CDCl3): δ 7.30–7.07 (m, 9H), 5.07 (s, 2H), 4.62 (t, J = 6.6 Hz, 2H), 2.74 (t, J = 6.6 Hz, 2H). 13C{1H} NMR (126 MHz, CDCl3): δ 171.2 (CO), 149.9 (C), 137.5 (C), 137.3 (C), 128.7 (2× CH), 127.30 (CH), 127.26 (CH), 127.1 (2× CH), 125.2 (CH), 123.2 (CH), 123.0 (CH), 74.6 (CH2), 51.0 (CH2), 35.3 (CH2). MS (CI), m/z (%): 254 (M + 1, 100). HRMS (CI-TOF) m/z: [M + H]+ calcd for C16H16NO2: 254.1176; found: 254.1176.
2,3-Dihydrobenzo[b][1,4]thiazepin-4(5H)-one (6f)
Known compound.36 Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 4:6). Compound 6f, 0.031 g (43%), brown semisolid. 1H NMR (500 MHz, CDCl3): δ 8.41 (bs, 1H), 7.60 (dd, J = 7.7, 1.5 Hz, 1H), 7.35 (td, J = 7.7, 1.5 Hz, 1H), 7.16 (td, J = 7.7, 1.5 Hz, 1H), 7.11 (dd, J = 7.7, 1.5 Hz, 1H), 3.48–3.41 (m, 2H), 2.63 (t, J = 6.9 Hz, 2H). 13C{1H} NMR, DEPT (126 MHz, CDCl3): δ 174.0 (CO), 141.6 (C), 135.6 (CH), 129.9 (CH), 127.0 (C), 126.6 (CH), 123.4 (CH), 34.5 (CH2), 33.7 (CH2). MS (CI), m/z (%): 180 (M + 1, 100).
3-Propyl-3,4,5,6-tetrahydrobenzo[d]azocin-2(1H)-one (7)
Known compound.37 Procedure B. The product was purified by silica gel chromatography (EtOAc/Hex 3:7 to 1:1). Compound 7, 0.039 g (90%), yellow oil. 1H NMR (500 MHz, CDCl3): δ (243 K): 7.41 (dd, J = 6.9, 2.0 Hz, 1H), 7.21–7.08 (m, 3H), 4.07 (d, J = 11.7 Hz, 1H), 3.87 (dd, J = 15.8, 11.3 Hz, 1H), 3.46–3.24 (m, 3H), 3.07–2.83 (m, 3H), 2.16–2.08 (m, 1H), 1.64–1.53 (m, 1H), 1.53–1.33 (m, 2H), 0.75 (t, J = 7.3 Hz, 3H). 13C{1H} NMR (126 MHz, CDCl3): δ (243 K): 173.0 (CO), 139.9 (C), 136.0 (C), 130.0 (CH), 129.8 (CH), 127.2 (CH), 127.2 (CH), 51.4 (CH2), 49.5 (CH), 40.2 (CH), 36.7 (CH), 30.2 (CH), 20.9 (CH), 11.3 (CH3). MS (CI), m/z (%): 218 (M + 1, 100). HRMS (EI-TOF) m/z: [M]+ calcd for C14H19NO: 217.1467; found: 217.1469.
3-Methyl-1,5-dihydro-2H-benzo[d]azepine-2,4(3H)-dione (10)
Procedure A. The product was purified by silica gel chromatography (EtOAc/Hex 3:7). Compound 10, 0.028 g (73%), orange needles. 1H NMR (500 MHz, CDCl3): δ 7.30–7.27 (m, 4H), 4.11 (s, 4H), 3.13 (s, 3H). 13C{1H} NMR, DEPT (126 MHz, CDCl3): δ 171.0 (CO), 131.9 (2× C), 128.6 (2× CH), 128.5 (2× CH), 45.0 (CH2), 29.8 (CH3). MS (CI), m/z (%): 190 (M + 1, 100), 163 (3), 162 (28). HRMS (CI-TOF) m/z: [M + H]+ calcd for C11H12NO2: 190.0863; found: 190.0862.
Isoquinolin-1(2H)-one (11)
Procedure A in the absence of oxidant. Known compound.9a The product was purified by silica gel chromatography (EtOAc/Hex 1:1). Compound 11, 0.042 g (83%), brown solid. 1H NMR (300 MHz, CDCl3): δ 11.49 (s, 1H), 8.43 (d, J = 8.00 Hz, 1H), 7.68 (t, J = 6.9 Hz, 1H), 7.59–7.51 (m, 2H), 7.20 (d, J = 7.1 Hz, 1H), 6.58 (d, J = 7.1 Hz, 1H). 13C{1H} NMR, DEPT (75 MHz, CDCl3): δ 164.4 (C=O), 138.1(C), 132.5 (CH), 127.6 (CH), 127.3 (CH), 126.8 (CH), 126.2 (CH), 126.1(C), 106.7 (CH). MS (EI) (m/z, %): 145 (M+, 100), 118 (36), 90 (29). HRMS (EI-TOF) m/z: [M]+ calcd for C9H7NO: 145.0528; found: 145.0528.
Acknowledgments
This work has received financial support from MICINN (projects PID2020-118048GB-I00, PID2020-116861GB-I00, and ORFEO-CINQA network RED2018-102387-T), the Xunta de Galicia (project ED431C 2022/27, Centro singular de investigación de Galicia accreditation 2019-2022, ED431G 2019/03), and the European Union (European Regional Development Fund). We are also grateful to the CESGA (Xunta de Galicia) for computational time.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.joc.2c02770.
General experimental procedures, NMR spectra, and DFT calculations (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- a Hager A.; Vrielink N.; Hager D.; Lefranc J.; Trauner D. Synthetic approaches towards alkaloids bearing α-tertiary amines. Nat. Prod. Rep. 2016, 33, 491–522. 10.1039/C5NP00096C. [DOI] [PubMed] [Google Scholar]; b Afanasyev O. I.; Kuchuk E.; Usanov D. L.; Chusov D. Reductive Amination in the Synthesis of Pharmaceuticals. Chem. Rev. 2019, 119, 11857–11911. 10.1021/acs.chemrev.9b00383. [DOI] [PubMed] [Google Scholar]
- a de Figueiredo R. M.; Suppo J.-S.; Campagne J.-M. Nonclassical Routes for Amide Bond Formation. Chem. Rev. 2016, 116, 12029–12122. 10.1021/acs.chemrev.6b00237. [DOI] [PubMed] [Google Scholar]; b Chen W.; Ma L.; Paul A.; Seidel D. Direct α-C-H bond functionalization of unprotected cyclic amines. Nat. Chem. 2018, 10, 165–169. 10.1038/nchem.2871. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Trowbridge A.; Walton S. M.; Gaunt M. J. New Strategies for the Transition-Metal Catalyzed Synthesis of Aliphatic Amines. Chem. Rev. 2020, 120, 2613–2692. 10.1021/acs.chemrev.9b00462. [DOI] [PubMed] [Google Scholar]
- a Mueller T. E.; Beller M. Metal-Initiated Amination of Alkenes and Alkynes. Chem. Rev. 1998, 98, 675–704. 10.1021/cr960433d. [DOI] [PubMed] [Google Scholar]; b Yudha S. S.; Kuninobu Y.; Takai K. Rhenium-Catalyzed Hydroamidation of Unactivated Terminal Alkynes: Synthesis of (E)-Enamides. Org. Lett. 2007, 9, 5609–5611. 10.1021/ol702564e. [DOI] [PubMed] [Google Scholar]; c Herrero M. T.; de Sarralde J. D.; SanMartin R.; Bravo L.; Dominguez E. Cesium Carbonate-Promoted Hydroamidation of Alkynes: Enamides, Indoles and the Effect of Iron(III) Chloride. Adv. Synth. Catal. 2012, 354, 3054–3064. 10.1002/adsc.201200430. [DOI] [Google Scholar]; d Huang L.; Arndt M.; Gooßen K.; Heydt H.; Gooßen L. J. Late Transition Metal-Catalyzed Hydroamination and Hydroamidation. Chem. Rev. 2015, 115, 2596–2697. 10.1021/cr300389u. [DOI] [PubMed] [Google Scholar]
- a Kondo T.; Okada T.; Suzuki T.; Mitsudo T. A. Ruthenium-catalyzed intramolecular hydroamination of aminoalkynes. J. Organomet. Chem. 2001, 622, 149–154. 10.1016/S0022-328X(00)00869-X. [DOI] [Google Scholar]; b Patil N. T.; Huo Z.; Bajracharya G. B.; Yamamoto Y. Lactam Synthesis via the Intramolecular Hydroamidation of Alkynes Catalyzed by Palladium Complexes. J. Org. Chem. 2006, 71, 3612–3614. 10.1021/jo060142x. [DOI] [PubMed] [Google Scholar]; c Patil N. T.; Lutete L. M.; Wu H.; Pahadi N. K.; Gridnev I. D.; Yamamoto Y. Palladium-Catalyzed Intramolecular Asymmetric Hydroamination, Hydroalkoxylation, and Hydrocarbonation of Alkynes. J. Org. Chem. 2006, 71, 4270–4279. 10.1021/jo0603835. [DOI] [PubMed] [Google Scholar]; d Shu C.; Liu M.-Q.; Wang S.-S.; Li L.; Ye L.-W. Gold-Catalyzed Oxidative Cyclization of Chiral Homopropargyl Amides: Synthesis of Enantioenriched γ-Lactams. J. Org. Chem. 2013, 78, 3292–3299. 10.1021/jo400127x. [DOI] [PubMed] [Google Scholar]; e Kuai C.; Wang L.; Cui H.; Shen J.; Feng Y.; Cui X. Efficient and Selective Synthesis of (E)-Enamides via Ru(II)-Catalyzed Hydroamidation of Internal Alkynes. ACS Catal. 2016, 6, 186–190. 10.1021/acscatal.5b01791. [DOI] [Google Scholar]; f Hojo R.; Short S.; Jha M. Synthesis of 1,2-fused tricyclic indoles via Cu–/base-mediated hydroamination of alkynes. J. Org. Chem. 2019, 84, 16095–16104. 10.1021/acs.joc.9b02591. [DOI] [PubMed] [Google Scholar]; g Chiu W.-J.; Chen J.-Y.; Liu S.-I.; Barve I. J.; Huang W.-W.; Sun C.-M. One-pot Synthesis of Isoquinoline-Fused Isoquinolines via Intramolecular Hydroamination/Aza-Claisen Type Rearrangement Cascade. Adv. Synth. Catal. 2021, 363, 2834–2842. 10.1002/adsc.202001576. [DOI] [Google Scholar]; h Nie X.-D.; Han X.-L.; Sun J.-T.; Si C.-M.; Wei B.-G.; Lin G.-Q. Nickel-Catalyzed Regioselective Hydroamination of Ynamides with Secondary Amines. J. Org. Chem. 2021, 86, 3433–3443. 10.1021/acs.joc.0c02807. [DOI] [PubMed] [Google Scholar]
- Roh S. W.; Choi K.; Lee C. Transition Metal Vinylidene and Allenylidene Mediated Catalysis in Organic Synthesis. Chem. Rev. 2019, 119, 4293–4356. 10.1021/acs.chemrev.8b00568. [DOI] [PubMed] [Google Scholar]
- Álvarez-Pérez A.; Saá C.; Varela J. A. Oxidation of Alkynes via Catalytic Metal-Vinylidenes. Synthesis 2020, 52, 2639–2649. 10.1055/s-0040-1707860. [DOI] [Google Scholar]
- a Kim I.; Lee C. Rhodium-Catalyzed Oxygenative Addition to Terminal Alkynes for the Synthesis of Esters, Amides, and Carboxylic Acids. Angew. Chem., Int. Ed. 2013, 52, 10023–10026. 10.1002/anie.201303669. [DOI] [PubMed] [Google Scholar]; b Wang Y.; Zheng Z.; Zhang L. Ruthenium-Catalyzed Oxidative Transformations of Terminal Alkynes to Ketenes By Using Tethered Sulfoxides: Access to β-Lactams and Cyclobutanones. Angew. Chem., Int. Ed. 2014, 53, 9572–9576. 10.1002/anie.201403796. [DOI] [PubMed] [Google Scholar]
- Álvarez-Pérez A.; Esteruelas M. A.; Izquierdo S.; Varela J. A.; Saá C. Ruthenium-Catalyzed Oxidative Amidation of Alkynes to Amides. Org. Lett. 2019, 21, 5346–5350. 10.1021/acs.orglett.9b01993. [DOI] [PubMed] [Google Scholar]
- a Varela-Fernández A.; Varela J. A.; Saá C. Ruthenium-Catalyzed Cycloisomerization of Aromatic Homo- and Bis-Homopropargylic Amines/Amides: Formation of Indoles, Dihydroisoquinolines and Dihydroquinolines. Adv. Synth. Catal. 2011, 353, 1933–1937. 10.1002/adsc.201100095. [DOI] [Google Scholar]; b Álvarez-Pérez A.; González-Rodríguez C.; García-Yebra C.; Varela J. A.; Oñate E.; Esteruelas M. A.; Saá C. Catalytic Cyclization of o-Alkynyl Phenethylamines via Osmacyclopropene Intermediates: Direct Access to Dopaminergic 3-Benzazepines. Angew. Chem., Int. Ed. 2015, 54, 13357–13361. 10.1002/anie.201505782. [DOI] [PubMed] [Google Scholar]
- For a related macrolactonization process, see:; Zhang W. W.; Gao T. T.; Xu L. J.; Li B. J Macrolactonization of Alkynyl Alcohol through Rh(I)/Yb(III) Catalysis. Org. Lett. 2018, 20, 6534–6538. 10.1021/acs.orglett.8b02858. [DOI] [PubMed] [Google Scholar]
- a Fischer C.; Zultanski S. L.; Zhou H.; Methot J. L.; Shah S.; Hayashi I.; Hughes B. L.; Moxham C. M.; Bays N. W.; Smotrov N.; Hill A. D.; Pan B.-S.; Wu Z.; Moy L. Y.; Tanga F.; Kenific C.; Cruz J. C.; Walker D.; Bouthillette M.; Nikov G. N.; Deshmukh S. V.; Jeliazkova-Mecheva V. V.; Diaz D.; Michener M. S.; Cook J. J.; Munoz B.; Shearman M. S. Preparation of triazolobenzazepinone derivatives as γ-secretase modulators. Bioorg. Med. Chem. Lett. 2015, 25, 3488–3494. 10.1016/j.bmcl.2015.07.003. [DOI] [PubMed] [Google Scholar]; b Velasco-Rubio Á.; Varela J. A.; Saá C. Recent Advances in Transition-Metal-Catalyzed Oxidative Annulations to Benzazepines and Benzodiazepines. Adv. Synth. Catal. 2020, 362, 4861–4875. 10.1002/adsc.202000808. [DOI] [Google Scholar]
- a Evans P. A.; Holmes A. B. Medium ring nitrogen heterocycles. Tetrahedron 1991, 47, 9131–9166. 10.1016/S0040-4020(01)96203-9. [DOI] [Google Scholar]; b In Asymmetric Synthesis of Nitrogen Heterocycles, Royer J., Ed.; Wiley-VCH: Weinheim, 2009. [Google Scholar]; c Ouyang W.; Rao J.; Li Y.; Liu X.; Huo Y.; Chen Q.; Li X. Recent Achievements in the Rhodium-Catalyzed Concise Construction of Medium N-Heterocycles Azepines and Azocines. Adv. Synth. Catal. 2020, 362, 5576–5600. 10.1002/adsc.202001040. [DOI] [Google Scholar]; d Meyer A. G.; Bissember A. C.; Hyland C. J. T.; Williams C. C.; Szabo M.; Wales S. M.; Constable G. E. O.; Olivier W. J.Seven-membered rings. In Progress in Heterocyclic Chemistry, Gribble G. W.; Joule J. A., Eds.; Elsevier, 2021; Vol. 32, pp. 565–614. [Google Scholar]; e Moorefield C. N.; Newkome G. R.. 8 - Eight-membered and larger rings. In Progress in Heterocyclic Chemistry, Gribble G. W.; Joule J. A., Eds.; Elsevier, 2021; Vol. 32, pp. 615–635. [Google Scholar]
- a Zhou B.; Li L.; Zhu X.-Q.; Yan J.-Z.; Guo Y.-L.; Ye L.-W. Yttrium-Catalyzed Intramolecular Hydroalkoxylation/Claisen Rearrangement Sequence: Efficient Synthesis of Medium-Sized Lactams. Angew. Chem., Int. Ed. 2017, 56, 4015–4019. 10.1002/anie.201700596. [DOI] [PubMed] [Google Scholar]; b Wang N.; Gu Q.-S.; Li Z.-L.; Li Z.; Guo Y.-L.; Guo Z.; Liu X.-Y. Direct Photocatalytic Synthesis of Medium-Sized Lactams by C-C Bond Cleavage. Angew. Chem., Int. Ed. 2018, 57, 14225–14229. 10.1002/anie.201808890. [DOI] [PubMed] [Google Scholar]
- N-oxide remained unaffected.
- See Table 1a, Supporting Information, for other temperatures and conditions tried.
- The use of other non-Cp, non-phosphine, Cp* and cationic Ru(II) catalysts led to traces of 2a or decomposition. See Supporting Information for details. However, the use of the bisphosphine Ru(II) catalyst, CpRuCl(dppe), gave 2a in a similar 54% yield.; Watanabe T.; Mutoh Y.; Saito S. Ruthenium-Catalyzed Cycloisomerization of 2-Alkynylanilides: Synthesis of 3-Substituted Indoles by 1,2-Carbon Migration. J. Am. Chem. Soc. 2017, 139, 7749–7752. 10.1021/jacs.7b04564. [DOI] [PubMed] [Google Scholar]
- See Table 1b, Supporting Information, for the effect of additives.
- By contrast, decomposition or traces of products were again obtained when the Rh catalytic conditions developed by Lee were used. See Table 1c Supporting Information.
- By contrast, when the oxidant employed was (methylsulfinyl)benzene, a complex mixture was obtained. See Table 1e Supporting Information.
- See Tables 1f and 1g, Supporting Information, for catalyst loading and temperature optimization.
- a The N-oxide may be detrimental for the hydroamination process.; b Chapple D. E.; Hoffer M. A.; Boyle P. D.; Blacquiere J. M. Alkyne Hydrofunctionalization Mechanism Including an Off-Cycle Alkoxycarbene Deactivation Complex. Organometallics 2022, 41, 1532–1542. 10.1021/acs.organomet.2c00170. [DOI] [Google Scholar]
- See Supporting Information for computational details.
- a Wakatsuki Y.; Koga N.; Yamazaki H.; Morokuma K. Acetylene .pi.-Coordination, Slippage to .sigma.-Coordination, and 1,2-Hydrogen Migration Taking Place on a Transition Metal. The case of a Ru(II) Complex as Studied by Experiment and ab Initio Molecular Orbital Simulations. J. Am. Chem. Soc. 1994, 116, 8105–8111. 10.1021/ja00097a017. [DOI] [Google Scholar]; b De Angelis F.; Sgamellotti A.; Re N. Density Functional Study of Alkyne to Vinylidene Rearrangements in [(Cp)(PMe3)2Ru(HC⋮CR)]+ (R = H, Me). Organometallics 2002, 21, 5944–5950. 10.1021/om020723m. [DOI] [Google Scholar]; c De Angelis F.; Sgamellotti A.; Re N. Acetylene to vinylidene rearrangements on electron rich d6 metal centers: a density functional study. Dalton Trans. 2004, 3225–3230. 10.1039/b408452g. [DOI] [PubMed] [Google Scholar]; d Jiménez-Tenorio M.; Puerta M. C.; Valerga P.; Ortuño M. A.; Ujaque G.; Lledós A. Counteranion and Solvent Assistance in Ruthenium-Mediated Alkyne to Vinylidene Isomerizations. Inorg. Chem. 2013, 52, 8919–8932. 10.1021/ic401119p. [DOI] [PubMed] [Google Scholar]
- IIIb is lower in electronic energy than TSIIb-IIIb , but becomes higher than TSIIb-IIIb when thermal effects are included.
- Varela-Fernández A.; González-Rodríguez C.; Varela J. A.; Castedo L.; Saá C. Cycloisomerization of Aromatic Homo- and Bis-homopropargylic Alcohols via Catalytic Ru Vinylidenes: Formation of Benzofurans and Isochromenes. Org. Lett. 2009, 11, 5350–5353. 10.1021/ol902212h. [DOI] [PubMed] [Google Scholar]
- For the structural features of isolated Ir- and Rh-ketenes complexes, see:; Grotjahn D. B.; Collins L. S. B.; Wolpert M.; Bikzhanova G. A.; Lo H. C.; Combs D.; Hubbard J. L. First Direct Structural Comparison of Complexes of the Same Metal Fragment to Ketenes in Both C,C- and C,O-Bonding Modes. J. Am. Chem. Soc. 2001, 123, 8260–8270. 10.1021/ja004324z. [DOI] [PubMed] [Google Scholar]
- See Supporting Information for the whole evaluated pathways.
- Microkinetic analysis was also performed. See Supporting Information for details.
- For a proposed rhodium-bound ketene intermediate in the synthesis of β-lactams from alkynes and imines, see:; Kim I.; Roh S. W.; Lee D. G.; Lee C. Rhodium-Catalyzed Oxygenative [2 + 2] Cycloaddition of Terminal Alkynes and Imines for the Synthesis of β-Lactams. Org. Lett. 2014, 16, 2482–2485. 10.1021/ol500856z. [DOI] [PubMed] [Google Scholar]
- Water is formed in situ by the reaction of the ammonium salt with the base used.
- a Raspoet G.; Nguyen M. T.; Kelly S.; Hegarty A. F. Amination of Ketenes: Evidence for a Mechanism Involving Enols of Amides as Intermediates. J. Org. Chem. 1998, 63, 9669–9677. 10.1021/jo980642t. [DOI] [Google Scholar]; b Sung K.; Tidwell T. T. Amination of Ketene: A Theoretical Study. J. Am. Chem. Soc. 1998, 120, 3043–3048. 10.1021/ja972200p. [DOI] [Google Scholar]; c Allen A. D.; Tidwell T. T. Amination of Ketenes: Kinetic and Mechanistic Studies. J. Org. Chem. 1999, 64, 266–271. 10.1021/jo982054l. [DOI] [PubMed] [Google Scholar]
- Knobloch K.; Keller M.; Eberbach W. The Synthesis of Annulated Azepin-3-one Derivatives from 1,3,4-Pentatrienyl Nitrones by a Heterocyclization–Rearrangement Sequence. Eur. J. Org. Chem. 2001, 2001, 3313–3332. . [DOI] [Google Scholar]
- Chen W.-Y.; Gilman N. W. Synthesis of 7-phenylpyrimido[5,4-d][1]benzazepin-2-ones. J. Heterocycl. Chem. 1983, 20, 663–666. 10.1002/jhet.5570200330. [DOI] [Google Scholar]
- Donets P. A.; Cramer N. Diaminophosphine Oxide Ligand Enabled Asymmetric Nickel-Catalyzed Hydrocarbamoylations of Alkenes. J. Am. Chem. Soc. 2013, 135, 11772–11775. 10.1021/ja406730t. [DOI] [PubMed] [Google Scholar]
- Tandon V. K.; Awasthi A. K.; Maurya H. K.; Mishra P. InBr3- and AgOTf-catalyzed beckmann rearrangement of (E)-benzoheterocyclic oximes. J. Heterocycl. Chem. 2012, 49, 424–427. 10.1002/jhet.438. [DOI] [Google Scholar]
- Becerra-Figueroa L.; Ojeda-Porras A.; Gamba-Sánchez D. Transamidation of Carboxamides Catalyzed by Fe(III) and Water. J. Org. Chem. 2014, 79, 4544–4552. 10.1021/jo500562w. [DOI] [PubMed] [Google Scholar]
- For a study of conformations of eight-membered lactams, see:; Witosińska A.; Musielak B.; Serda P.; Owińska M.; Rys B. Conformation of Eight-Membered Benzoannulated Lactams by Combined NMR and DFT Studies. J. Org. Chem 2012, 77, 9784–9794. 10.1021/jo301859h. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.










