Next Article in Journal
Combining Soft Polysilazanes with Melt-Shear Organization of Core–Shell Particles: On the Road to Polymer-Templated Porous Ceramics
Next Article in Special Issue
Scalable Preparation of Enantioenriched (S)-5-methylhept-2-en-4-one. Synthesis and Aroma Properties of Achiral Analogues Thereof
Previous Article in Journal
A Selective, Dual Emission β-Alanine Aminopeptidase Activated Fluorescent Probe for the Detection of Pseudomonas aeruginosa, Burkholderia cepacia, and Serratia marcescens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Site Selectivity in Pd-Catalyzed Reactions of α-Diazo-α-(methoxycarbonyl)acetamides: Effects of Catalysts and Substrate Substitution in the Synthesis of Oxindoles and β-Lactams

1
Laboratori de Química Orgànica, Facultat de Farmàcia i Ciències de l’Alimentació, Universitat de Barcelona, Av. Joan XXIII 27-31, 08028 Barcelona, Spain
2
Departamento de Química Orgánica I and Centro de Innovación en Química Avanzada (ORFEO-CINQA), Facultad de Ciencias Químicas, Universidad Complutense de Madrid, 28040 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(19), 3551; https://doi.org/10.3390/molecules24193551
Submission received: 19 September 2019 / Revised: 27 September 2019 / Accepted: 27 September 2019 / Published: 30 September 2019
(This article belongs to the Special Issue New Insights in Diversity Oriented Synthesis)

Abstract

:
The Pd-catalyzed intramolecular carbene C–H insertion of α-diazo-α-(methoxycarbonyl)acetamides to prepare oxindoles as well as β-lactams was studied. In order to identify what factors influence the selectivity of the processes, we explored how the reactions are affected by the catalyst type, using two oxidation states of Pd and a variety of ligands. It was found that, in the synthesis of oxindoles, ((IMes)Pd(NQ))2 can be used as an alternative to Pd2(dba)3 to catalyze the carbene CArsp2–H insertion, although it was less versatile. On the other hand, it was demonstrated that the Csp3–H insertion leading to β-lactams can be effectively promoted by both Pd(0) and Pd(II) catalysts, the latter being most efficient. Insight into the reaction mechanisms involved in these transformations was provided by DFT calculations.

Graphical Abstract

1. Introduction

In recent years, the development of new methodologies for the selective functionalization of unactivated C–H bonds has become a very active area of research [1,2]. As part of this exciting field, the transition metal-catalyzed intramolecular carbene C–H insertion by decomposition of α-diazocarbonyl compounds has emerged as a powerful methodology for the construction of carbocyclic and heterocyclic frameworks [3,4,5,6,7,8]. A number of transition metal complexes have been used as effective catalysts to generate reactive metallacarbenes starting from α-diazocarbonyl compounds [9,10,11,12,13,14,15,16]. Among them, rhodium(II) [17,18], copper(I) [19,20], and more recently ruthenium(II) catalysts [21,22,23,24,25,26,27] have been proven to be especially useful for the development of highly selective carbene C–H insertion methodologies via a variety of reaction modes. Interestingly, palladium, one of the most commonly employed metals in homogeneous catalysis, remains underexploited in this type of carbene C–H insertion processes. Thus, while the effectiveness of palladium complexes in catalyzing carbene C–H insertion reactions from α-diazo carbonyl compounds was demonstrated some time ago [28], their use has been restricted to a couple of examples of insertion into CArsp2–H bonds [29,30,31]. This fact is highly surprising if we take into account the great success of palladium catalysis in cross-coupling reactions of diazo compounds with organic halides, pseudohalides, or arylboronic acids [32,33,34,35] and that the palladium-catalyzed cyclopropanation of olefins with diazomethane is a widely used synthetic methodology [36].
The longstanding research on transition metal-catalyzed carbene C–H insertion has generated an extensive literature on the use of dirhodium(II) catalysts to promote the intramolecular C–H insertion of α-diazoacetamides [37,38], as the insertion products, namely β- and γ-lactams as well as 2-oxindoles, are common scaffolds found in numerous natural products. In recent years, some ruthenium(II) catalysts have also been applied to promote this kind of C–H insertion process [21,22,25,39,40,41,42,43]. It has been shown that the site selectivity of these reactions not only depends on the type of diazocarbonyl compound but also is governed by conformational, steric, as well as electronic factors [44,45,46]. Moreover, some dramatic ligand effects have also been observed. For example, the use of carboxylate and, in particular, carboxamide ligands in dirhodium(II) catalysts has resulted in highly chemo-, regio- and stereoselective transformations [47,48] despite a variety of potentially competitive carbene-mediated processes (Scheme 1).
As part of our research program on the synthesis of nitrogen heterocycles, we have been exploring different ways to increase the versatility of palladium catalysis in C–C bond-forming reactions [49,50,51], for example, by controlling the ambiphilic character of the organopalladium intermediates in the intramolecular coupling reactions with carbonyl derivatives [52]. Continuing our search for methodologies to enhance the synthetic potential of organopalladium chemistry, we have also investigated the versatility of palladium as a catalyst for the carbene C–H insertion. In this context, we reported that palladium catalysts are able to promote Csp3–H insertion of carbenes derived from α-diazoesters to form pyrrolidines through intramolecular Csp3–Csp3 bond formation [53,54].
We have also demonstrated that the carbene CArsp2–H functionalization of α-diazo-α-(methoxycarbonyl)acetanilides to give oxindoles can also be promoted by using Pd2(dba)3 as the catalyst [55]. This allowed us to develop a one-pot methodology to prepare 3-(chloroethyl)oxindoles by means of a sequential C–H insertion/alkylation process (Equation (a) in Scheme 2). More recently, we have described the synthesis of β-lactams by Pd(II)-catalyzed carbene Csp3–H insertion of α-diazo-α-(methoxycarbonyl)acetamides (Equation (b) in Scheme 2) [56].
The aim of the current work was to gain more insight into the Pd-catalyzed carbene insertion reactions of α-diazo-α-(methoxycarbonyl)acetamides leading to oxindoles and β-lactams. In order to ascertain the factors governing the selectivity of the processes, we explored how the reactions are affected by the substituents on α-diazoacetamide and the type of catalyst, using complexes with two oxidation states of Pd and a variety of ligands (Scheme 3). Herein, we present a full account of our experimental and computational studies on the intramolecular insertion of α-diazo-α-(methoxycarbonyl)acetamides using Pd(0) and Pd(II) complexes, focused on identifying differences in the catalyst reactivities and selectivities.

2. Results and Discussion

The oxindole system is a common structural motif in pharmaceuticals and natural products, and the development of efficient procedures for their preparation is, therefore, of considerable interest [57,58,59,60].
Our previous studies on the carbene reactions of α-diazo-α-(methoxycarbonyl)acetanilides showed that, when using palladium catalysts, the C–H insertion giving the oxindole occurs selectively in the CArsp2–H rather than the Csp3–H bonds (Equation (a) of Scheme 2) [55]. The optimization process with α-diazoacetanilide 1a revealed that the carbene CArsp2–H insertion can be selectively promoted by both Pd(0) and Pd(II), the best result being obtained when using Pd2(dba)3 in the presence of Cs2CO3 in dichloroethane at reflux for 96 h, which afforded oxindole 2a in 66% yield (Table 1, entry 1). However, this reaction required a high catalyst loading and long reaction time. All attempts to increase the efficiency of the Pd(0)-catalyzed reaction by adding different phosphine ligands failed, always resulting in a slower reaction rate.
In order to improve the efficiency of the carbene insertion of α-diazoacetanilide 1a, in the present study, we tested some palladium catalysts bearing non-phosphine ligands. (Pd(allyl)Cl)2 was first explored for the insertion/alkylation sequential process (Table 1, entries 2–3), but significant amounts of the starting material were recovered even after 60 h. In contrast, the use of ((IMes)Pd(NQ))2 required a notably shorter reaction time and lower catalyst loading (Table 1, entry 4).
With these data in hand, we then explored the generality of the NHC-Pd(0) catalyst in the sequential C–H insertion/alkylation process, leading to 3-(chloroethyl)-3-(methoxycarbonyl)oxindoles. Table 2 gathers the results of the reactions of different α-diazoacetanilides (1b1j) when using ((IMes)Pd(NQ))2 as the catalyst and compares them with those previously obtained with Pd2(dba)3 [55].
As can be seen in Table 2, like Pd2(dba)3, ((IMes)Pd(NQ))2 also selectively promoted the carbene insertion into the arylic C–H bond in substrates having substituents at the nitrogen atom with primary, secondary, as well as tertiary Csp3–H bonds. Notably, ((IMes)Pd(NQ))2 was more efficient than Pd2(dba)3 in promoting the insertion of N-alkylacetanilides lacking substituents on the arylic ring (1b1d) (Table 2, entries 1–6) or bearing an electron-donating substituent (1g) (Table 2, entries 11–12). In contrast, the introduction of electron-withdrawing groups dramatically diminished the catalytic efficiency of ((IMes)Pd(NQ))2. Thus, for example, acetanilide 1f, bearing an ortho-bromo substituent, was recovered unchanged when using ((IMes)Pd(NQ))2, whereas in the presence of Pd2(dba)3, it afforded oxindole 2f in 23% yield (Table 2, entries 9 and 10). Oxindole 2h, with a fluoro group, was isolated in a modest 39% yield when using ((IMes)Pd(NQ))2 but obtained in an acceptable 64% yield in the presence of Pd2(dba)3 (Table 2, entries 13–14). No insertion product was formed in the ((IMes)Pd(NQ))2-catalyzed reaction of acetanilide 1i, which bears a methoxycarbonyl substituent, while oxindole 2i was isolated in 22% yield when using Pd2(dba)3 (Table 2, entries 15–16). ((IMes)Pd(NQ))2 also gave worse results in the sequential insertion/alkylation sequence from N,N-diphenylacetamide 1e (Table 2, entries 7 and 8) and N-naphthylacetamide 1j (Table 2, entries 17–18).
At this point, we studied the decomposition of α-diazo-N-pyridinylacetamide 3, which, upon treatment with a catalytic amount of ((IMes)Pd(NQ))2 under the optimized reaction conditions, afforded mesoionic compound 4 (77%), resulting from the interception of the transient metallacarbene by the pyridine nitrogen (Scheme 4). This kind of mesoionic imidazopyridines have been previously prepared [61] and show interesting insecticidal activity [62].
Regarding the palladium-catalyzed reactions shown in Table 1 and Table 2, it should be noted that the formation of β-lactam products, resulting from the possible competitive carbene Csp3–H insertion, was not observed. This sharply contrasts with the competition between the carbene CArsp2–H and Csp3–H insertions observed in Rh(II)-promoted processes (see, for instance, Scheme 1(a)) [48].
Our previously reported DFT calculations suggested that the Pd(0)-catalyzed CArsp2–H insertion of α-diazo-α-(methoxycarbonyl)acetanilides to give oxindoles proceeds via a genuine stepwise mechanism involving a Pd-mediated 1,5-hydrogen migration from the initially generated pallada(0)carbene complex, followed by a reductive elimination [55]. It was also shown that the complete chemoselectivity of the process, which exclusively produces oxindoles over β-lactams, takes place mainly under kinetic control. We now decided to check the generality of this unprecedented Pd(0)-mediated mechanism using the model Pd(NHC) (NHC = 1,3-bis(phenyl)-imidazil-2-ylidene) catalyst.
Figure 1 shows the computed reaction profile for the formation of oxindole 2b’, precursor of the experimentally observed oxindole 2b (see Table 2). As expected, based on our previous report [55], the process begins from pallada(0)carbene INT0, formed upon reaction of α-diazoamide 1b and the active catalyst species Pd(NHC). This species evolves into intermediate INT1 in a highly exergonic process (∆GR = –25.3 kcal/mol) through the transition state TS1 with a computed activation barrier of 16.2 kcal/mol, which is fully compatible with the reaction conditions used in the experiment. Closer inspection of TS1 indicates that this reaction step can be viewed as a Pd-mediated 1,5-hydrogen migration, which involves the formal oxidation of the transition metal. The alternative Csp3–H activation reaction via TS1’, which would involve an analogous Pd-mediated 1,4-hydrogen migration, is not competitive in view of the much higher activation barrier (∆G = 8.2 kcal/mol) computed for this alternative transformation. Therefore, it is once again found that the complete selectivity of the process takes place under kinetic control. The transformation ends with an exergonic (∆GR = –6.8 kcal/mol) reductive elimination reaction that directly transforms INT1 into oxindole 2b’ with the concomitant release of the active catalytic species Pd(NHC) through the transition state TS2 (with a feasible activation barrier of 18.7 kcal/mol).
On the other hand, we have also reported that the intramolecular carbene C–H insertion of α-diazo-α-(methoxycarbonyl)acetamides to form β-lactams can be effectively catalyzed by Pd(II) complexes (Scheme 2b) [56]. Due to its ubiquitous presence in the molecular structure of natural products and biologically active compounds, β-lactam synthesis has attracted considerable attention over the years [63,64,65]. To further understand the impact of the electronic nature of the catalyst in the aforementioned insertion reaction, we decided to explore the use of some Pd(0)-precatalysts. Table 3 shows the results of the reactions of diversely substituted N-benzyl-N-tbutyl-α-diazoacetamides (5a5m) when using either Pd2(dba)3 or ((IMes)Pd(NQ))2 as the catalyst and compares them with those previously obtained in the Pd(II)-catalyzed reactions [56].
The examples in Table 3 confirm the generality and functional group tolerance of Pd2(dba)3 and ((IMes)Pd(NQ))2 as catalysts for these insertion processes. Although both Pd(0) and Pd(II) catalysts can be used to promote Csp3–H insertion, some significant differences were identified. On the whole, the Pd(II) complexes proved to be more versatile for β-lactam formation despite not always giving the highest yield. In general, the resulting β-lactams were obtained in moderate to good overall yields (53–90%), usually as mixtures of cis and trans isomers, in transformations proceeding with high site selectivity, except when using 5f. In this particular case, the use of Pd(II) catalysts to promote its decomposition resulted in the formation of trans-7f (10–14%), together with major amounts of aldehyde 10. On the other hand, when ((IMes)Pd(NQ))2 was used, a complex reaction mixture was obtained, from which cis-7f (9%), trans-7f (10%), and aldehyde 10 (15%) were isolated, while no cyclization product was obtained in the presence of Pd2(dba)3.
Notably, partial or complete isomerization of the cis β-lactam to the more stable trans isomer took place during the chromatographic purification of the reaction mixtures, which explains the observed differences in cis/trans ratios before and after purification.
The effect of adding phosphine ligands to the Pd(0) catalyst was also explored. However, similarly to the oxindole-forming reactions, the use of Pd2(dba)3 in the presence of phosphine ligands resulted in slower reaction rates. Thus, for example, treatment of 5a with Pd2(dba)3 (10 mol%) in the presence of (o-tolyl)3P (20 mol%) under otherwise the same reaction conditions gave a mixture of 6 (12%), cis-7 (22%), and trans-7 (26%) together with some unreacted α-diazoamide (10%) (result not included in the table).
Interestingly, no product from the potentially competitive palladium-catalyzed cross-coupling of the aryl halide with the α-diazoamide moiety [32,33,34,35] was observed in the reactions of 5k (Table 3, entries 41–42) and 5l (Table 3, entries 45–46) when using Pd(0) catalysts.
As can be seen in Table 3, the effect of the substituent at the benzyl group on the outcome of the process varied according to its electronic nature as well as its position on the aromatic ring. The introduction of electron-donor groups led to an increased formation of the cycloheptapyrrolone product, especially when using Pd(0) catalysts. The increase was lower when the substituent was located at the ortho-position, probably due to steric interactions. In contrast, the electron-withdrawing groups generally diverted the palladacarbene away from the Buchner reaction in favor of the Csp3–H insertion. Similar electronic effects have been observed in related Rh(II)-catalyzed transformations [48]. On the other hand, the use of ((IMes)Pd(NQ))2 as the catalyst in the reaction of those substrates bearing electron-withdrawing groups resulted in the formation of minor amounts of the corresponding γ-lactam (8), which arises from the tbutyl Csp3–H insertion. Notably, the γ-lactams were not observed when using Pd(II) catalysts.
The above results also confirm the significant impact of the electronic nature of the Pd catalyst and ensue electrophilicity of the carbene intermediate in the reaction pathway. Thus, whereas benzylic Csp3–H insertion is strongly favored over the Buchner reaction when using Pd(II), an increased cycloheptapyrrolone product formation is usually observed with the more electron-rich Pd(0) complexes. Interestingly, Rh(II)-catalyzed transformations show opposite reactivity trends, in which highly electrophilic Rh(II) complexes favor Buchner reactions over benzylic Csp3–H insertion [44,45,46].
Finally, we explored the transition metal-catalyzed decomposition of α-diazoacetamide 11, which bears a (4-pyridinyl)methyl group instead of the N-benzyl substituent (Table 4). Treatment of 11 with a catalytic amount of ((IMes)Pd(NQ))2 under the optimized reaction conditions resulted in the complete decomposition of the material (Table 4, entry 1). When the reaction was performed at lower temperatures (refluxing dichloromethane), 11 was recovered unchanged (Table 4, entry 2). In contrast, the use of Pd(II) catalysts (Table 4, entries 3 and 4) resulted in the formation of β-lactam 12 (cis/trans mixture), with (SIPr)Pd(allyl)Cl affording the highest yield. Interestingly, α-diazoacetamide 11 was recovered unchanged when using the well-known (Rh(OAc)2)2 catalyst (Table 4, entry 5).
In our previous work on the synthesis of β-lactams, we studied computationally the Pd(II)-catalyzed insertion reaction of α-diazo-α-(methoxycarbonyl)acetamides [56]. According to our DFT calculations, this insertion reaction also occurs stepwise and involves an unprecedented Pd(II)-promoted Mannich-type reaction through a pallada(II)carbene-induced zwitterionic intermediate. To shed light on the reaction mechanism and the influence of the Pd(0) catalyst on the selectivity of the C–H insertion described above, DFT calculations were also carried out. To this end, the process involving 5a in the presence of the model Pd(NHC) (NHC = 1,3-bis(phenyl)-imidazil-2-ylidene) catalyst was explored.
Figure 2 shows the computed reaction profile for the transformation of the initially formed pallada(0)carbene intermediate INT0 into the observed β-lactams cis-7a and trans-7a. Similar to the reaction profile computed for the formation of oxindoles discussed above (Figure 1), the initial INT0 readily evolves into the corresponding five-membered pallada(II)cycle INT1 via the transition state TS1 in a highly exergonic transformation (∆GR ≈ –40 kcal/mol). Once again, this saddle point is associated with a 1,4-hydrogen migration reaction mediated by the transition metal fragment, thus confirming the generality of this type of C–H insertion reaction. Then, species INT1 is transformed into the final β–lactam through a reductive elimination reaction via TS2, which releases the active catalyst able to enter into a new catalytic cycle. Interestingly, the computed cis/trans activation barrier differences for either the initial 1,4-H migration (∆G = 1.3 kcal/mol) or for the subsequent elimination reaction (∆G = 0.2 kcal/mol) indicates that the formation of the cis-7a β-lactam should be only slightly favored over the formation of its trans counterpart, which nicely matches the experimental findings (see Table 3, entry 2).

3. Materials and Methods

3.1. General Information

All commercially available reagents were used without further purification. 1H- and 13C-NMR spectra were recorded using Me4Si as the internal standard with a Varian Mercury 400 instrument (Oxford, United Kingdom). Chemical shifts are reported in ppm downfield (δ) from Me4Si for 1H and 13C-NMR. TLC was carried out on SiO2 (silica gel 60 F254, Merck), and the spots were located with UV light or 1% aqueous KMnO4. Flash chromatography was carried out on SiO2 (silica gel 60, SDS, 230–400 mesh ASTM). Organic extracts were dried over anhydrous Na2SO4 during workup of reactions. Evaporation of solvents was accomplished with a rotatory evaporator (Büchi, Flavill, Switzerland). α-Diazoacetamides 1a1j and oxindoles 2a2j are known compounds previously prepared by us [55], as are α-diazoacetamides 5a5m and β-lactams 7a7m [56].

3.2. Synthesis of N-Benzyl-N-(2-pyridinyl)-α-(ethoxycarbonyl)-α-diazoacetamide (3).

To a solution of 2-benzylaminopyridine (0.5 g, 2.71 mmol) and Et3N (0.39 mL, 2.71 mmol) in CH2Cl2 (10 mL), cooled at 0 °C, ethyl malonyl chloride (0.38 mL, 2.71 mmol) was added slowly. The mixture was stirred at room temperature for 24 h. After the reaction was completed, the mixture was poured into water and extracted with CH2Cl2. The organic extracts were washed with saturated NaHCO3 aqueous solution, dried, filtered, and concentrated. The residue was purified by chromatography (SiO2, from CH2Cl2 to CH2Cl2/MeOH 97:3) to give N-benzyl-N-(2-pyridinyl)-α-(ethoxycarbonyl) acetamide (0.77 g, 95%).
To a solution of N-benzyl-N-(2-pyridinyl)-α-(ethoxycarbonyl)acetamide (0.25 g, 0.84 mmol) and Et3N (0.14 mL, 1.0 mmol) in dry acetonitrile (15 mL), p-toluenesulfonylazide (2.6 mL of a 11% solution in toluene, 1.45 mmol) was added dropwise. The mixture was stirred at room temperature for 90 h. The solvent was removed in vacuo, and the resulting residue was partitioned between CH2Cl2 and 10% NaOH aqueous solution. The organic extracts were dried, filtered, and concentrated. The residue was purified by chromatography (SiO2, from CH2Cl2 to CH2Cl2/MeOH 99:1) to give N-benzyl-N-(2-pyridinyl)-α-(ethoxycarbonyl)-α-diazoacetamide (3, 150 mg; 55%) as a brown oil. 1H-NMR (CDCl3, 400 MHz) δ 1.07 (t, J = 7.2 Hz, 3H), 3.92 (q, J = 7.2 Hz, 2H), 5.22 (s, 2H), 7.04 (ddd, J = 7.4, 4.8, and 0.8 Hz, 1H), 7.10 (dt, J = 8.4 and 0.8 Hz, 1H), 7.20 (tt, J = 7.2 and 1.2 Hz, 1H), 7.23–7.29 (m, 2H), 7.32–7.36 (m, 2H), 7.60 (ddd, J = 8.4, 7.4 and 2.0 Hz, 1H), and 8.40 (ddd, J = 4.8, 2.0 and 0.8 Hz, 1H). 13C-NMR (CDCl3, 100.6 MHz) δ 14.3 (CH3), 52.3 (CH2), 61.4 (CH2), 70.2 (C), 118.3 (CH), 120.8 (CH), 127.4 (CH), 127.9 (2 CH), 128.6 (2 CH), 137.5 (C), 138.0 (CH), 148.6 (CH), 156.1 (C), 161.4 (C), and 162.3 (C).

3.3. Synthesis of N-tert-Butyl-N-(4-pyridinylmethyl)-α-(methoxycarbonyl)-α-diazoacetamide (11).

To a solution of N-tert-butyl-N-(4-pyridinylmethyl)amine (0.67 g, 4.1 mmol) and Et3N (0.58 mL, 4.1 mmol) in CH2Cl2 (20 mL), cooled at 0 °C, methyl malonyl chloride (0.57 mL, 5.3 mmol) was added slowly. The mixture was stirred at room temperature for 24 h. After the reaction was completed, the mixture was poured into water and extracted with CH2Cl2. The organic extracts were washed with saturated NaHCO3 aqueous solution, dried, filtered, and concentrated to give N-tert-butyl-N-(4-pyridinylmethyl)-α-(methoxycarbonyl)acetamide as an orange oil (0.98 g, 90%), which was used in the next reaction without purification.
To a solution of N-tert-butyl-N-(4-pyridinylmethyl)-α-(methoxycarbonyl)acetamide (0.5 g, 1.89 mmol) and DBU (0.45 mL, 2.85 mmol) in dry acetonitrile (6 mL), a solution of p-ABSA (515 mg, 2.1 mmol) in dry acetonitrile (2 mL) was added dropwise. The mixture was stirred at room temperature overnight. The solvent was removed in vacuo, and the resulting residue was partitioned between CH2Cl2 and 10% NaOH aqueous solution. The organic extracts were dried, filtered, and concentrated. The residue was purified by chromatography (SiO2, from CH2Cl2 to CH2Cl2/MeOH 98:2) to give N-tert-butyl-N-(4-pyridinylmethyl)-α-(methoxycarbonyl)-α-diazoacetamide (11, 145 mg; 26%) as an orange oil. 1H-NMR (CDCl3, 400 MHz) δ 1.40 (s, 9H), 3.77 (s, 3H), 4.63 (s, 2H), 7.15 (d, J = 6.0 Hz, 2H), and 7.57 (d, J = 6.0 Hz, 2H). 13C-NMR (CDCl3, 100.6 MHz) δ 29.0 (3 CH3), 50.6 (CH2), 52.4 (CH3), 59.5 (C), 68.7 (C), 121.9 (2 CH), 149.3 (C), 150.2 (2 CH), 162.8 (C), and 163.5 (C).

3.4. Characterization Data for New Compounds of Scheme 4 and Tables 3 and 4

1-Benzyl-3-(ethoxycarbonyl)-2-oxo-2,3-dihydro-1H-imidazo (1,2-a)pyridin-4-ium-3-ylide (4). Amorphous orange solid. 1H-NMR (CDCl3, 400 MHz) δ 1.45(t, J = 7.2 Hz, 3H), 4.43 (q, J = 7.2 Hz, 2H), 5.20 (s, 2H), 6.99 (ddd, J = 8.8, 1.6, and 0.8 Hz, 1H), 7.08 (ddd, J = 7.6, 6.8, and 1.2 Hz, 1H), 7.25–7.37 (m, 5H), 7.40 (ddd, J = 8.8, 7.6, and 1.2 Hz, 1H), 9.65 (d, J = 6.8 Hz, 1H). 13C-NMR (CDCl3, 100.6 MHz) δ 14.9 (CH3), 43.3 (CH2), 60.0 (CH2), 94.0 (C), 106.4 (CH), 115.7 (CH), 127.8 (2 CH), 128.2 (CH), 129.0 (CH), 129.1 (2 CH), 129.9 (CH), 135.4 (C), 135.6 (C), 157.0 (C), and 162.6 (C).
Methyl 2-tert-butyl-6-chloro-3-oxo-1H-2,3-dihydrocyclohepta(c)pyrrole-3a-carboxylate (6d). 1H-NMR (CDCl3, 400 MHz, signals from a 4.5:1 mixture of trans-7d and 6d) δ 1.45 (s, 9H), 3.64 (s, 3H), 4.21 (dd, J = 14.8 and 1.2 Hz, 1H), 4.44 (d, J = 14.8 Hz, 1H), 5.63 (d, J = 10.2 Hz, 1H), 6.17 (d, J = 6.8 Hz, 1H), 6.39 (d, J = 10.2 Hz, 1H), and 6.62 (d, J = 6.8 Hz, 1H).
Methyl 1-(4-cyanobenzyl)-5,5-dimethyl-2-oxopyrrolidine-3-carboxylate (8e). Amorphous orange solid. 1H-NMR (CDCl3, 400 MHz) δ 1.16 (s, 3H), 1.22 (s, 3H), 2.19 (dd, J = 13.2 and 9.2 Hz, 1H), 2.33 (dd, J = 13.2 and 9.2 Hz, 1H), 3.62 (t, J = 9.2 Hz, 1H), 3.82 (s, 3H), 4.33 (d, J = 16.0 Hz, 1H), 4.60 (d, J = 16.0 Hz, 1H), 7.39 (d, J = 8.4 Hz, 2H), 7.60 (d, J = 8.4 Hz, 2H). 13C-NMR (CDCl3, 100.6 MHz) δ 27.1 (CH3), 28.1 (s, CH3), 38.3 (CH2), 43.2 (CH2), 47.4 (CH), 53.0 (CH3), 60.0 (C), 111.5 (C), 118.8 (C), 128.3 (2 CH), 132.6 (2 CH), 143.9 (C), 170.2 (C), and 170.9 (C).
Methyl cis-1-tert-butyl-4-[4-(dimethylamino)phenyl]-2-oxoazetidine-3-carboxylate (cis-7f). Amorphous orange solid. 1H-NMR (CDCl3, 400 MHz) δ 1.30 (s, 9H), 2.95 (s, 6H), 3.40 (s, 3H), 4.18 (d, J = 6.0 Hz, 1H), 4.83 (d, J = 6.0 Hz, 1H), 6.65 (d, J = 8.8 Hz, 2H), and 7.21 (d, J = 8.8 Hz, 2H).
Methyl 2-tert-butyl-7-chloro-3-oxo-1H-2,3-dihydrocyclohepta[c]pyrrole-3a-carboxylate (6g).1H-NMR (CDCl3, 400 MHz, signals from a 11:1 mixture of trans-7g and 6g) δ 1.45 (s, 9H), 3.64 (s, 3H), 4.22 (dd, J = 15.6 and 2.0 Hz, 1H), 4.47 (dd, J = 15.6 and 2.4 Hz, 1H), 5.60 (d, J = 9.6 Hz, 1H), 6.27–6.30 (m, 1H), 6.32 (dd, J = 9.6 and 7.2 Hz, 1H), and 6.61 (dd, J = 7.2 and 1.2 Hz, 1H).
Methyl 2-tert-butyl-5-chloro-3-oxo-1H-2,3-dihydrocyclohepta[c]pyrrole-3a-carboxylate (6g’).1H-NMR (CDCl3, 400 MHz, signals from a 12:4:1 mixture of trans-7g, 6g, and 6g’) δ 1.46 (s, 9H), 3.64 (s, 3H), 4.23 (dd, J = 15.2 and 2.0 Hz, 1H), 4.42 (dd, J = 15.2 and 2.4 Hz, 1H), 5.76 (t, J = 1.2 Hz, 1H), 6.18–6.21 (m, 1H), and 6.38–6.40 (m, 2H).
Methyl 1-(3-cyanobenzyl)-5,5-dimethyl-2-oxopyrrolidine-3-carboxylate (8h).1H-NMR (CDCl3, 400 MHz, significant signals from a 2:1 mixture of 8h and cis-7h) δ 1.16 (s, 3H), 1.24 (s, 3H), 2.19 (dd, J = 12.8 and 9.2 Hz, 1H), 2.33 (dd, J = 12.8 and 9.2 Hz, 1H), 3.62 (t, J = 9.2 Hz, 1H), 3.80 (s, 3H), 4.33 (d, J = 16.0 Hz, 1H), and 4.55 (d, J = 16.0 Hz, 1H).
Methyl 2-tert-butyl-8-methoxy-3-oxo-1H-2,3-dihydrocyclohepta[c]pyrrole-3a-carboxylate (6i).1H-NMR (CDCl3, 400 MHz, signals from a 1.5:1 mixture of 6i and cis-7i) δ 1.47 (s, 9H), 3.55 (s, 3H), 3.69 (s, 3H), 4.23 (dd, J = 14.4 and 1.6 Hz, 1H), 4.48 (dd, J = 14.4 and 1.6 Hz, 1H), 5.65 (d, J = 7.2 Hz, 1H), 6.29–6.33 (m, 1H), 6.31 (d, J = 7.6 Hz, 1H), and 6.38–6.45 (m, 1H). 13C-NMR (CDCl3, 100.6 MHz, significant signals from a 1:8 mixture of 6i and cis-7i) δ 27.6 (3 CH3), 49.4 (CH2), 52.8 (CH3), 55.1 (C), 57.5 (CH3), 99.8 (CH), 120.6 (CH), 122.5 (CH), and 126.9 (CH).
Methyl 1-(2-fluorobenzyl)-5,5-dimethyl-2-oxopyrrolidine-3-carboxylate (8j).1H-NMR (CDCl3, 400 MHz, signals from a 8:1 mixture of 8j and cis-7j) δ 1.14 (s, 3H), 1.25 (s, 3H), 2.15 (dd, J = 12.8 and 9.2 Hz, 1H), 2.31 (dd, J = 12.8 and 9.2 Hz, 1H), 3.61 (t, J = 9.2 Hz, 1H), 3.82 (s, 3H), 4.46 (d, J = 16.0 Hz, 1H), 4.55 (d, J = 16.0 Hz, 1H), 7.00 (ddd, J = 10.4, 8.4, and 1.2 Hz, 1H), 7.09 (td, J = 7.6 and 1.2 Hz, 1H), 7.19–7.25 (m, 1H), and 7.39 (td, J = 7.6 and 1.2 Hz, 1H).
Methyl 1-(2-bromobenzyl)-5,5-dimethyl-2-oxopyrrolidine-3-carboxylate (8k). Amorphous orange solid. 1H-NMR (CDCl3, 400 MHz) δ 1.18 (s, 3H), 1.23 (s, 3H), 2.20 (dd, J = 12.8 and 9.2 Hz, 1H), 2.35 (dd, J = 12.8 and 9.2 Hz, 1H), 3.65 (t, J = 9.2 Hz, 1H), 3.83 (s, 3H), 4.45 (d, J = 16.4 Hz, 1H), 4.64 (d, J = 16.4 Hz, 1H), 7.10 (ddd, J = 8.0, 6.8 and 2.4 Hz, 1H), 7.24–7.30 (m, 2H), 7.51 (dd, J = 7.6 and 1.2 Hz, 1H). 13C-NMR (CDCl3, 100.6 MHz) δ 26.9 (s, CH3), 27.8 (s, CH3), 38.3 (CH2), 43.1 (CH2), 47.6 (CH), 52.9 (CH3), 60.0 (C), 122.6 (C), 128.0 (CH), 128.9 (CH), 129.2 (CH), 132.7 (CH), 137.1 (C), 170.1 (C), and 171.1 (C).
Methyl 1-(2-iodobenzyl)-5,5-dimethyl-2-oxopyrrolidine-3-carboxylate (8l). Amorphous orange solid. 1H-NMR (CDCl3, 400 MHz) δ 1.20 (s, 3H), 1.22 (s, 3H), 2.20 (dd, J = 13.2 and 9.2 Hz, 1H), 2.35 (dd, J = 13.2 and 9.2 Hz, 1H), 3.65 (t, J = 9.2 Hz, 1H), 3.83 (s, 3H), 4.35 (d, J = 16.4 Hz, 1H), 4.59 (d, J = 16.4 Hz, 1H), 6.93 (td, J = 8.0 and 1.6 Hz, 1H), 7.23 (dd, J = 8.0 and 1.6 Hz, 1H), 7.30 (td, J = 8.0 and 1.2 Hz, 1H), and 7.79 (dd, J = 8.0 and 1.2 Hz, 1H).
Methyl cis-1-tert-butyl-2-oxo-4-(pyridin-4-yl)azetidine-3-carboxylate (cis-12).1H-NMR (CDCl3, 400 MHz, signals from a 2.6:1 mixture of trans-12 and cis-12) δ 1.32 (s, 9H), 3.36 (s, 3H), 4.28 (d, J = 6.4 Hz, 1H), 4.88 (d, J = 6.4 Hz, 1H), 7.32-7.35 (m, 2H), and 8.61–8.65 (m, 2H). 13C-NMR (CDCl3, 100.6 MHz, signals from a 2.6:1 mixture of trans-12 and cis-12) δ 28.2 (3 CH3), 52.2 (CH3), 55.4 (CH), 55.5 (C), 59.0 (CH), 121.1 (2 CH), 146.2 (C), 150.2 (2 CH), 162.5 (C), and 165.9 (C).
Methyl trans-1-tert-butyl-2-oxo-4-(pyridin-4-yl)azetidine-3-carboxylate (trans-12).1H-NMR (CDCl3, 400 MHz) δ 1.27 (s, 9H), 3.67 (d, J = 2.4 Hz, 1H), 3.78 (s, 3H), 4.84 (d, J = 2.4 Hz, 1H), 7.33 (dd, J = 4.4 and 1.6 Hz, 2H), and 8.64 (dd, J = 4.4 and 1.6 Hz, 2H). 13C-NMR (CDCl3, 100.6 MHz) δ 28.2 (3 CH3), 53.0 (CH3), 55.2 (CH), 55.7 (C), 62.1 (CH), 121.6 (2 CH), 148.5 (C), 150.7 (2 CH), 161.7 (C), and 167.0 (C).

4. Computational Details

All the calculations reported in this paper were performed with the Gaussian 09 suite of programs [66]. Electron correlation was partially taken into account using the hybrid functional usually denoted as B3LYP [67,68,69] in conjunction with the D3 dispersion correction suggested by Grimme et al. [70] using the standard double-ζ quality def2-SVP [71,72] basis set for all atoms. The Polarizable Continuum Model (PCM) [73,74,75] was used to model the effects of the solvent. This level is denoted PCM(solvent)-B3LYP-D3/def2-SVP. Geometries were fully optimized in solution without any geometry or symmetry constraints. Reactants, intermediates, and products were characterized by frequency calculations and have positive definite Hessian matrices. Transition structures (TSs) show only one negative eigenvalue in their diagonalized force constant matrices, and their associated eigenvectors were confirmed to correspond to the motion along the reaction coordinate under consideration using the Intrinsic Reaction Coordinate (IRC) method [76]. Frequency calculations were also used to determine the difference between the potential (E) and Gibbs (G) energies, G − E, which contains the zero-point, thermal, and entropy energies. Potential energies were refined, Esol, by means of single point (SP) calculations at the same level with a larger basis set, def2-TZVPP [71,72], where all elements were described with a triple-ζ plus polarization quality basis set. This level is denoted PCM(solvent)-B3LYP-D3/def2-TZVPP//PCM(solvent)-B3LYP-D3/def2-SVP. The ΔG and ΔG values given in the text were obtained from the Gibbs energy in solution, Gsol, which was calculated by adding the thermochemistry corrections, G − E, to the refined SP energies, Esol, i.e., Gsol = Esol + G − E. See Supplementary Materials.

5. Conclusions

In summary, in the present paper, we report a full account of our experimental and computational studies on the Pd-catalyzed intramolecular carbene C–H insertion of α-diazo-α-(methoxycarbonyl)acetamides to prepare oxindoles and β-lactams. We have explored how the reactions are affected by the substituents on the α-diazoamide moiety and by the catalyst type, exploring the use of both Pd(0) and Pd(II) catalysts.
The chemoselectivity of the Pd-catalyzed C–H insertion reactions of α-diazo acetamides is mainly governed by the nature of the substrates. Thus, while N-aryl acetamides selectively afforded oxindoles, starting from N,N-dialkyl acetamides, the corresponding β-lactams were obtained.
The results obtained in the annulation reactions to form oxindoles show that Pd(0) catalysts were much more efficient than Pd(II) catalysts and that ((IMes)Pd(NQ))2 can be used as an alternative to Pd2(dba)3 to catalyze the carbene CArsp2–H insertion.
On the other hand, both Pd(0) and Pd(II) catalysts can be used to promote Csp3–H insertion leading to β-lactams. However, the Pd(II) complexes proved to be more versatile for β-lactam formation since the use of Pd(0) catalysts resulted in increased amounts of Buchner products and, in some cases, of the corresponding γ-lactams.
Our computational studies show that when using ((IMes)Pd(NQ))2 as the catalyst, both CArsp2–H and Csp3–H insertions involve similar stepwise reaction mechanisms, in which a palladium-mediated hydrogen migration is followed by a reductive elimination.

Supplementary Materials

The following are available online, Copies of 1H and 13C-NMR spectra of new compounds and Cartesian coordinates and total energies of all species described in the manuscript.

Author Contributions

D.S. designed and planned the research; D.S. and M.-L.B. supervised the experimental work; F.P.-J. and A.A. performed the experimental work and characterized the compounds; I.F. carried out the computational studies; D.S. and I.F. prepared the manuscript for publication.

Funding

This research was funded by MINECO/FEDER, Spain (Projects CTQ2015-64937-R, CTQ2016-78205-P, CTQ2016-81797-REDC, and RTI2018-09394-B-I00).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Godula, K.; Sames, D. C-H Bond functionalization in complex organic synthesis. Science 2006, 312, 67–72. [Google Scholar] [CrossRef]
  2. Yamaguchi, J.; Yamaguchi, A.D.; Itami, K. C-H Bond functionalization: emerging synthetic tools for natural products and pharmaceuticals. Angew. Chem. Int. Ed. 2012, 51, 8960–9009. [Google Scholar] [CrossRef]
  3. Davies, H.M.L.; Manning, J.R. Catalytic C-H functionalization by metal carbenoid and nitrenoid insertion. Nature 2008, 451, 417–424. [Google Scholar] [CrossRef]
  4. Doyle, M.P.; Duffy, R.; Ratnikov, M.; Zhou, L. Catalytic carbene insertion into C-H bonds. Chem. Rev. 2010, 110, 704–724. [Google Scholar] [CrossRef]
  5. Zheng, C.; You, S.-L. Recent development of direct asymmetric functionalization of inert C-H bonds. RSC Adv. 2014, 4, 6173–6214. [Google Scholar] [CrossRef]
  6. Ford, A.; Miel, H.; Ring, A.; Slattery, C.N.; Maguire, A.R.; McKervey, M.A. Modern organic synthesis with α-diazocarbonyl compounds. Chem. Rev. 2015, 115, 9981–10080. [Google Scholar] [CrossRef]
  7. Lombard, F.J.; Coster, M.J. Rhodium(II)-catalysed intramolecular C-H insertion α- to oxygen: reactivity, selectivity and applications to natural product synthesis. Org. Biomol. Chem. 2015, 13, 6419–6431. [Google Scholar] [CrossRef]
  8. Hu, F.; Xia, Y.; Ma, C.; Zhang, Y.; Wang, J. C-H bond functionalization based on metal carbene migratory insertion. Chem. Commun. 2015, 51, 7986–7995. [Google Scholar] [CrossRef]
  9. Cai, Y.; Zhu, S.-F.; Wang, G.-P.; Zhou, Q.-L. Iron-catalyzed C-H functionalization of indoles. Adv. Synth. Catal. 2011, 353, 2939–2944. [Google Scholar] [CrossRef]
  10. Yao, T.; Hirano, K.; Satoh, T.; Miura, M. Nickel- and cobalt-catalyzed direct alkylation of azoles with N-tosylhydrazones bearing unactivated alkyl groups. Angew. Chem. Int. Ed. 2012, 51, 775–779. [Google Scholar] [CrossRef]
  11. Yu, Z.; Ma, B.; Chen, M.; Wu, H.-H.; Liu, L.; Zhang, J. Highly site-selective direct C-H bond functionalization of phenols with α-aryl-α-diazoacetates and diazooxindoles via gold catalysis. J. Am. Chem. Soc. 2014, 136, 6904–6907. [Google Scholar] [CrossRef]
  12. Liu, X.-G.; Zhang, S.-S.; Wu, J.-Q.; Li, Q.; Wang, H. Cp*Co(III)-catalyzed direct functionalization of aromatic C-H bonds with α-diazomalonates. Tetrahedron Lett. 2015, 56, 4093–4095. [Google Scholar] [CrossRef]
  13. Zhao, D.; Kim, J.H.; Stegemann, L.; Strassert, C.A.; Glorius, F. Cobalt(III)-catalyzed directed C-H coupling with diazo compounds: straightforward access towards extended π-systems. Angew. Chem. Int. Ed. 2015, 54, 4508–4511. [Google Scholar] [CrossRef]
  14. Fructos, M.R.; Díaz-Requejo, M.M.; Pérez, P.J. Gold and diazo reagents: a fruitful tool for developing molecular complexity. Chem. Commun. 2016, 52, 7326–7335. [Google Scholar] [CrossRef]
  15. Liu, L.; Zhang, J. Gold-catalyzed transformations of α-diazocarbonyl compounds: selectivity and diversity. Chem. Soc. Rev. 2016, 45, 506–516. [Google Scholar] [CrossRef] [PubMed]
  16. Conde, A.; Sabenya, G.; Rodríguez, M.; Postils, V.; Luis, J.M.; Díaz-Requejo, M.M.; Costas, M.; Pérez, P.J. Iron and manganese catalysts for the selective functionalization of arene C(sp2)-H bonds by carbene insertion. Angew. Chem. Int. Ed. 2016, 55, 6530–6534. [Google Scholar] [CrossRef] [PubMed]
  17. Davies, H.M.L.; Parr, B.T. Rhodium carbenes. In Contemporary Carbene Chemistry; Wiley: Hoboken, NJ, USA, 2013; pp. 363–403. [Google Scholar]
  18. Yakura, T.; Nambu, H. Recent topics in application of selective Rh(II)-catalyzed C-H functionalization toward natural product synthesis. Tetrahedron Lett. 2018, 59, 188–202. [Google Scholar] [CrossRef]
  19. Díaz-Requejo, M.M.; Pérez, P.J. Coinage metal catalyzed C-H bond functionalization of hydrocarbons. Chem. Rev. 2008, 108, 3379–3394. [Google Scholar] [CrossRef]
  20. Zhao, X.; Zhang, Y.; Wang, J. Recent developments in copper-catalyzed reactions of diazo compounds. Chem. Commun. 2012, 48, 10162–10173. [Google Scholar] [CrossRef]
  21. Choi, M.K.-W.; Yu, W.-Y.; Che, C.-M. Ruthenium-catalyzed stereoselective intramolecular carbenoid C-H insertion for β- and γ-lactam formations by decomposition of α-diazoacetamides. Org. Lett. 2005, 7, 1081–1084. [Google Scholar] [CrossRef]
  22. Grohmann, M.; Buck, S.; Schäffler, L.; Maas, G. Diruthenium(I,I) catalysts for the formation of β- and γ-lactams via carbenoid C-H insertion of α-diazoacetamides. Adv. Synth. Catal. 2006, 348, 2203–2211. [Google Scholar] [CrossRef]
  23. Choi, M.K.-W.; Yu, W.-Y.; So, M.-H.; Zhou, C.-Y.; Deng, Q.-H.; Che, C.-M. A Non-cross-linked soluble polystyrene-supported ruthenium catalyst for carbenoid transfer reactions. Chem. Asian J. 2008, 3, 1256–1265. [Google Scholar] [CrossRef]
  24. Reddy, A.R.; Zhou, C.-Y.; Guo, Z.; Wei, J.; Che, C.-M. Ruthenium-porphyrin-catalyzed diastereoselective intramolecular alkyl carbene insertion into C-H bonds of alkyl diazomethanes generated in situ from N-tosylhydrazones. Angew. Chem. Int. Ed. 2014, 53, 14175–14180. [Google Scholar] [CrossRef] [PubMed]
  25. Nakagawa, Y.; Chanthamath, S.; Liang, Y.; Shibatomi, K.; Iwasa, S. Regio- and enantioselective intramolecular amide carbene insertion into primary C-H bonds using Ru(II)-Pheoux catalyst. J. Org. Chem. 2019, 84, 2067–2618. [Google Scholar] [CrossRef] [PubMed]
  26. Solé, D.; Amenta, A.; Bennasar, M.-L.; Fernández, I. Grubbs catalysts in intramolecular carbene C(sp3)-H insertion reactions from α-diazoesters. Chem. Commun. 2019, 55, 1160–1163. [Google Scholar] [CrossRef] [PubMed]
  27. Solé, D.; Amenta, A.; Bennasar, M.-L.; Fernández, I. Pd- and Ru-Catalyzed intramolecular carbene CAr-H functionalization of γ-amino-α-diazoesters for the synthesis of tetrahydroquinolines. Chem. Eur. J. 2019, 25, 10239–10245. [Google Scholar] [CrossRef] [PubMed]
  28. Taber, D.F.; Amedio, J.C.; Sherill, R.G. Palladium-mediated diazo insertion: Preparation of 3-alkyl-2-carbomethoxycyclopentenones. J. Org. Chem. 1986, 51, 3382–3384. [Google Scholar] [CrossRef]
  29. Matsumoto, M.; Watanabe, N.; Kobayashi, H. Metal-catalyzed intramolecular cyclization of 2-diazo-4-(4-indolyl)-3-oxobutanoic acid esters. Heterocycles 1987, 26, 1479–1482. [Google Scholar] [CrossRef]
  30. Rosenberg, M.L.; Aasheim, J.H.F.; Trebbin, M.; Uggerud, E.; Hansen, T. Synthesis of a 1,3,4,5-tetrahydrobenzindole β-ketoester. Tetrahedron Lett. 2009, 50, 6506–6508. [Google Scholar] [CrossRef]
  31. Goll, J.M.; Fillion, E. Tuning the reactivity of palladium carbenes derived from diphenylketene. Organometallics 2008, 27, 3622–3625. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Wang, J. Recent developments in Pd-catalyzed reactions of diazo compounds. Eur. J. Org. Chem. 2011, 1015–1026. [Google Scholar] [CrossRef]
  33. Barluenga, J.; Valdés, C. Tosylhydrazones: new uses for classic reagents in palladium-catalyzed cross-coupling and metal-free reactions. Angew. Chem. Int. Ed. 2011, 50, 7486–7500. [Google Scholar] [CrossRef] [PubMed]
  34. Shao, Z.; Zhang, H. N-Tosylhydrazones: versatile reagents for metal-catalyzed and metal-free cross-coupling reactions. Chem. Soc. Rev. 2012, 41, 560–572. [Google Scholar] [CrossRef] [PubMed]
  35. Xiao, Q.; Zhang, Y.; Wang, J. Diazo compounds and N-tosylhydrazones: novel cross-coupling partners in transition-metal-catalyzed reactions. Acc. Chem. Res. 2013, 46, 236–247. [Google Scholar] [CrossRef] [PubMed]
  36. Reiser, O. Cyclopropanation and other reactions of palladium-carbene (and carbine) complexes. In Handbook of Organopalladium Chemistry for Organic Synthesis; Negishi, E., Ed.; Wiley-Interscience: New York, NY, USA, 2002; Volume 1, pp. 1561–1577. [Google Scholar]
  37. Gois, P.M.P.; Afonso, C.A.M. Stereo- and regiocontrol in the formation of lactams by rhodium-carbenoid C-H insertion of α-diazoacetamides. Eur. J. Org. Chem. 2004, 3773–3788. [Google Scholar] [CrossRef]
  38. Ring, A.; Ford, A.; Maguire, A.R. Substrate and catalyst effects in C-H insertion reactions of α-diazoacetamides. Tetrahedron Lett. 2016, 57, 5399–5406. [Google Scholar] [CrossRef]
  39. Grohmann, M.; Maas, G. Ruthenium catalysts for carbenoid intramolecular C-H insertion of 2-diazoacetoacetamides and diazomalonic ester amides. Tetrahedron 2007, 63, 12172–12178. [Google Scholar] [CrossRef]
  40. Large, T.; Müller, T.; Kunkel, H.; Buck, S.; Maas, G. Ruthenium- and rhodium-catalyzed carbenoid reactions of diazoesters in hexaalkylguanidinium-based ionic liquids. Z. Naturforsch. B 2012, 67, 347–353. [Google Scholar] [CrossRef]
  41. Chan, W.-W.; Kwong, T.-L.; Yu, W.-Y. Ruthenium-catalyzed intramolecular cyclization of diazo-β-ketoanilides for the synthesis of 3-alkylideneoxindoles. Org. Biomol. Chem. 2012, 10, 3749–3755. [Google Scholar] [CrossRef]
  42. Liu, N.; Tian, Q.-P.; Yang, Q.; Yang, S.-D. Ruthenium-catalyzed intramolecular cyclization and fluorination to form 3-fluorooxindoles. Synlett 2016, 27, 2621–2625. [Google Scholar]
  43. Yamamoto, K.; Qureshi, Z.; Tsoung, J.; Pisella, G.; Lautens, M. Combining Ru-catalyzed C-H functionalization with Pd-catalyzed asymmetric allylic alkylation: synthesis of 3-allyl-3-aryl oxindoles derivatives from aryl α-diazoamides. Org. Lett. 2016, 18, 4954–4957. [Google Scholar] [CrossRef] [PubMed]
  44. Merlic, C.A.; Zechman, A.L. Selectivity in rhodium(II) catalyzed reactions of diazo compounds: effects of catalyst electrophilicity, diazo substitution, and substrate substitution. From chemoselectivity to enantioselectivity. Synthesis 2003, 1137–1156. [Google Scholar] [CrossRef]
  45. Davies, H.M.L.; Morton, D. Guiding principles for site selective and stereoselective intermolecular C-H functionalization by donor/acceptor rhodium carbenes. Chem. Soc. Rev. 2011, 40, 1857–1869. [Google Scholar] [CrossRef]
  46. DeAngelis, A.; Panish, R.; Fox, J.M. Rh-catalyzed intermolecular reactions of α-alkyl-α-diazo carbonyl compounds with selectivity over β-hydride migration. Acc. Chem. Res. 2016, 49, 115–127. [Google Scholar] [CrossRef] [PubMed]
  47. Brown, D.S.; Elliott, M.C.; Moody, C.J.; Mowlem, T.J.; Marino, J.P.; Padwa, A. Ligand effects in the rhodium(II)-catalyzed reactions of α-diazoamides. Oxindole formation is promoted by the use of rhodium(II) perfluorocarboxamide catalysts. J. Org. Chem. 1994, 59, 2447–2455. [Google Scholar] [CrossRef]
  48. Miah, S.; Slawin, A.M.Z.; Moody, C.J.; Sheedan, S.M.; Marino, J.P., Jr.; Semones, M.A.; Padwa, A.; Richards, I.C. Ligand effects in the rhodium(II) catalyzed reactions of diazoamides and diazoimides. Tetrahedron 1996, 52, 2489–2514. [Google Scholar] [CrossRef]
  49. Solé, D.; Pérez-Janer, F.; Mancuso, R. Pd0-Catalyzed intramolecular α-arylation of sulfones: domino reactions in the synthesis of functionalized tetrahydroquinolines. Chem. Eur. J. 2015, 21, 4580–4584. [Google Scholar] [CrossRef]
  50. Solé, D.; Pérez-Janer, F.; Zulaica, E.; Guastavino, J.F.; Fernández, I. Pd-Catalyzed α-arylation of sulfones in a three-component synthesis of 3-[2-(phenyl/methylsulfonyl)ethyl]indoles. ACS Catal. 2016, 6, 1691–1700. [Google Scholar] [CrossRef]
  51. Solé, D.; Pérez-Janer, F.; García-Rodeja, Y.; Fernández, I. Exploring partners for the domino α-arylation/Michael addition reaction leading to tetrahydroquinolines. Eur. J. Org. Chem. 2017, 2017, 799–805. [Google Scholar] [CrossRef]
  52. Solé, D.; Fernández, I. Controlling the ambiphillic nature of σ-arylpalladium intermediates in intramolecular cyclization reactions. Acc. Chem. Res. 2014, 47, 168–179. [Google Scholar] [CrossRef]
  53. Solé, D.; Mariani, F.; Bennasar, M.-L.; Fernández, I. Palladium-catalyzed intramolecular carbene insertion into C(sp3)-H bonds. Angew. Chem. Int. Ed. 2016, 55, 6467–6470. [Google Scholar] [CrossRef] [PubMed]
  54. Solé, D.; Amenta, A.; Mariani, F.; Bennasar, M.-L.; Fernández, I. Transition metal-catalysed intramolecular carbenoid C-H insertion for pyrrolidine formation by decomposition of α-diazoesters. Adv. Synth. Catal. 2017, 359, 3654–3664. [Google Scholar] [CrossRef]
  55. Solé, D.; Pérez-Janer, F.; Fernández, I. Palladium-catalysed intramolecular carbenoid insertion of α-diazo-α-(methoxycarbonyl)acetanilides for oxindoles synthesis. Chem. Commun. 2017, 53, 3110–3113. [Google Scholar] [CrossRef] [PubMed]
  56. Solé, D.; Pérez-Janer, F.; Bennasar, M.-L.; Fernández, I. Palladium catalysis in the intramolecular carbene C-H insertion of α-diazo-α-(methoxycarbonyl)acetamides to forma β-lactams. Eur. J. Org. Chem. 2018, 4446–4455. [Google Scholar] [CrossRef]
  57. Gabriele, B.; Salerno, G.; Veltri, L.; Costa, M.; Massera, C. Stereoselective synthesis of (E)-3-(methoxycarbonyl)methylene-1,3-dihydroindol-2-ones by palladium-catalyzed oxidative carbonylation of 2-ethynylanilines. Eur. J. Org. Chem. 2001, 4607–4613. [Google Scholar] [CrossRef]
  58. Kischkewitz, M.; Daniliuc, C.-G.; Studer, A. 3-Alkyl-3-cyano-oxindoles from 2-cyano-2-diazo-N-phenylacetamides via cyclizing carbene insertion and subsequent radical oxidation. Org. Lett. 2016, 18, 1206–1209. [Google Scholar] [CrossRef]
  59. Ma, C.; Xing, D.; Hu, W. Catalyst-free halogenation of α-diazocarbonyl compounds with N-halosuccinimides: synthesis of 3-halooxindoles or vinyl halides. Org. Lett. 2016, 18, 3134–3137. [Google Scholar] [CrossRef]
  60. Mo, S.; Xu, C.; Xu, J. Expeditious and convenient synthesis of polycyclic difluoroboron complexes of 2-oxoindoline-3-carboxamides by tandem reaction. Adv. Synth. Catal. 2016, 358, 1767–1777. [Google Scholar] [CrossRef]
  61. Moody, C.J.; Miah, S.; Slawin, A.M.Z.; Mansfield, D.J.; Richards, I.C. Ligands effects in the metal catalyzed reactions of N-aryldiazoamides: ylide formation vs. insertion reactions. Tetrahedron 1998, 54, 9689–9700. [Google Scholar] [CrossRef]
  62. Heil, M.; Hoffmeister, L.; Webber, M.; Ilg, K.; Goergens, U.; Turberg, A. Preparation of mesoionic imidazopyridines for use as insecticides. PCT Int. Appl. WO 2018192872 A1 20181025.
  63. Bonardi, A.; Costa, M.; Gabriele, B.; Salerno, G.; Chiusoli, G.P. Versatile synthesis of beta-lactams, gamma-lactams or oxazolidines by palladium-catalysed oxidative carbonylation of 1-substituted prop-2-ynylamines. Tetrahedron Lett. 1995, 36, 7495–7498. [Google Scholar] [CrossRef]
  64. Hosseyni, S.; Jarrahpour, A. Recent advances in β-lactam synthesis. Org. Biomol. Chem. 2018, 16, 6840–6852. [Google Scholar] [CrossRef] [PubMed]
  65. Synofzik, J.; Dar’in, D.; Novikov, M.S.; Kantin, G.; Bakulina, O.; Krasavin, M. α-Acyl-α-diazoacetates in transition-metal-free β-lactam synthesis. J. Org. Chem. 2019, 84, 12101. [Google Scholar] [CrossRef] [PubMed]
  66. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  67. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  68. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1998, 37, 785–789. [Google Scholar] [CrossRef] [PubMed]
  69. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: a critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef] [Green Version]
  70. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements. J. Chem. Phys. 2010, 132, 154104–154119. [Google Scholar] [CrossRef]
  71. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef]
  72. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  73. Miertuš, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilization of ab-initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  74. Pascual-Ahuir, J.L.; Silla, E.; Tuñón, I. GEPOL: An improved description of molecular surfaces. III. A new algorithm for the computation of a solvent-excluding surface. J. Comp. Chem. 1994, 15, 1127–1138. [Google Scholar] [CrossRef]
  75. Barone, V.; Cossi, M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar] [CrossRef]
  76. Gonzalez, C.; Schlegel, H.B. Reaction path following in mass-weighted internal coordinates. J. Phys. Chem. 1990, 94, 5523–5527. [Google Scholar] [CrossRef]
Sample Availability: Not available.
Scheme 1. Typical Rh(II)-catalyzed reactions of α-diazoamides.
Scheme 1. Typical Rh(II)-catalyzed reactions of α-diazoamides.
Molecules 24 03551 sch001
Scheme 2. Pd catalysts in C–H insertion reactions of α-diazo-α-(methoxycarbonyl)acetamides.
Scheme 2. Pd catalysts in C–H insertion reactions of α-diazo-α-(methoxycarbonyl)acetamides.
Molecules 24 03551 sch002
Scheme 3. Catalyst and substituent effects in Pd-catalyzed C–H insertion reactions of α-diazo-α-(methoxycarbonyl)acetamides.
Scheme 3. Catalyst and substituent effects in Pd-catalyzed C–H insertion reactions of α-diazo-α-(methoxycarbonyl)acetamides.
Molecules 24 03551 sch003
Scheme 4. Pd-catalyzed cyclization of α-diazoamide 3.
Scheme 4. Pd-catalyzed cyclization of α-diazoamide 3.
Molecules 24 03551 sch004
Figure 1. Computed reaction profile for the transformation of pallada(0)carbene INT0 into oxindole 2b’: Relative free energies and bond distances are given in kcal/mol and angstroms, respectively. All data have been computed at the PCM(dichloroethane)-B3LYP-D3/def2-TZVPP//PCM(dichloroethane)-B3LYP-D3/def2-SVP level.
Figure 1. Computed reaction profile for the transformation of pallada(0)carbene INT0 into oxindole 2b’: Relative free energies and bond distances are given in kcal/mol and angstroms, respectively. All data have been computed at the PCM(dichloroethane)-B3LYP-D3/def2-TZVPP//PCM(dichloroethane)-B3LYP-D3/def2-SVP level.
Molecules 24 03551 g001
Figure 2. Computed reaction profile for the transformation of pallada(0)carbene INT0 into β-lactams cis-7a and trans-7a: Relative free energies and bond distances are given in kcal/mol and angstroms, respectively. All data have been computed at the PCM(dichloroethane)-B3LYP-D3/def2-TZVPP//PCM(dichloroethane)-B3LYP-D3/def2-SVP level.
Figure 2. Computed reaction profile for the transformation of pallada(0)carbene INT0 into β-lactams cis-7a and trans-7a: Relative free energies and bond distances are given in kcal/mol and angstroms, respectively. All data have been computed at the PCM(dichloroethane)-B3LYP-D3/def2-TZVPP//PCM(dichloroethane)-B3LYP-D3/def2-SVP level.
Molecules 24 03551 g002
Table 1. Palladium-catalyzed reactions of α-diazoamide 1a 1.
Table 1. Palladium-catalyzed reactions of α-diazoamide 1a 1.
Molecules 24 03551 i001
EntryCatalyst (mol%)TimeProducts (Yield (%)) 2
1Pd2(dba)3 (10)96 h 32a (66)
2(Pd(allyl)Cl)2 (5)24 h1a/2a (1:2.2) 4
3(Pd(allyl)Cl)2 (5)60 h1a/2a (1:5.5) 4
4((IMes)Pd(NQ))2 (4)24 h2a (65)
1 All reactions were conducted with 1a (0.2 mmol), catalyst (see table), and Cs2CO3 (1 equiv.) in dichloroethane (0.2 M) at reflux. Pd2(dba)3 = Tris(dibenzylideneacetone)dipalladium(0). ((IMes)Pd(NQ))2 = 1,3-Bis(2,4,6-trimethylphenyl)imidazol-2-ylidene (1,4-naphthoquinone) palladium(0) dimer. 2 Isolated yield. 3 Cs2CO3 (2 equiv.). 4 1H-NMR ratio; yields were not quantified.
Table 2. Palladium-catalyzed cyclization reactions of α-diazoamides 1b1j.
Table 2. Palladium-catalyzed cyclization reactions of α-diazoamides 1b1j.
Molecules 24 03551 i002
Entry Substance 1Catalyst 1,2 Product (Yield (%)) 3
1 Molecules 24 03551 i0031b (R:Me)Pd2(dba)3 Molecules 24 03551 i0042b (71)
21b (R:Me)((IMes)Pd(NQ))22b (79)
31c (R:Et)Pd2(dba)32c (54)
41c (R:Et)((IMes)Pd(NQ))22c (78)
51d (R:i-Pr)Pd2(dba)32d (61)
61d (R:i-Pr)((IMes)Pd(NQ))22d (64)
71e (R:Ph)Pd2(dba)3 42e (55) 5
81e (R:Ph)((IMes)Pd(NQ))22e (35)
9 Molecules 24 03551 i0051fPd2(dba)3 Molecules 24 03551 i0062f (23) 6
101f((IMes)Pd(NQ))22f (0) 7
11 Molecules 24 03551 i0071g (R1:OMe)Pd2(dba)3 Molecules 24 03551 i0082g (36)
121g (R1:OMe)((IMes)Pd(NQ))22g (76)
131h (R1F)Pd2(dba)32h (64)
141h (R1:F)((IMes)Pd(NQ))22h (39) 8
151i (R1:CO2Me)Pd2(dba)3 42i (22)
161i (R1:CO2Me)((IMes)Pd(NQ))22i (0) 9
17 Molecules 24 03551 i0091jPd2(dba)3 4 Molecules 24 03551 i0102j (27)
181j((IMes)Pd(NQ))22j (23) 10
1 Reactions were conducted with 1 (0.2 mmol), Pd2(dba)3 (10 mol%), and Cs2CO3 (1 equiv.) in dichloroethane (0.2 M) at reflux for 96 h. 2 Reactions were conducted with 1 (0.2 mmol), ((IMes)Pd(NQ))2 (4 mol%), and Cs2CO3 (1 equiv.) in dichloroethane (0.2 M) at reflux for 48 h. 3 Isolated yield. 4 Catalyst loading: 15%. 5 20% of unreacted 1e was recovered. 6 1H-NMR analysis of the reaction mixture showed a 1:1 mixture of 1f and 2f. 7 1f was recovered. 8 1H-NMR analysis of the reaction mixture showed a 1:4.8 mixture of 1h and 2h. 9 Complex reaction mixture. 10 1H-NMR analysis of the reaction mixture showed a 1:1.5 mixture of 1j and 2j.
Table 3. Palladium-catalyzed cyclization reactions of α-diazoamides 5a5m 1.
Table 3. Palladium-catalyzed cyclization reactions of α-diazoamides 5a5m 1.
Molecules 24 03551 i011
Entry5 (X)Catalyst (mol%)6/7/8 2,3cis-7/trans-7 2,3Products (Yield (%)) 4
15a (H)Pd2(dba)3 (10)26/74/046/286a (20)
cis-7a (9), trans-7a (42)
25a (H)((IMes)Pd(NQ))2 (2.5)35/65/040/256a (28)
cis-7a (35), trans-7a (23)
35a (H)(Pd(allyl)Cl)2 (5)0/100/047/53cis-7a (25), trans-7a (65)
45a (H)(SIPr)Pd(allyl)Cl (15)0/100/029/71cis-7a (17), trans-7a (59)
55b (4-MeO)Pd2(dba)3 (10)50/50/021/296b (24)
cis-7b (9), trans-7b (20)
65b (4-MeO)((IMes)Pd(NQ))2 (2.5)34/66/046/206b (22)
cis-7b (21), trans-7b (24)
75b (4-MeO)(Pd(allyl)Cl)2 (5)14/86/029/576b (8)
cis-7b (18), trans-7b (39)
85b (4-MeO)(SIPr)Pd(allyl)Cl (15)16/84/042/426b (10)
cis-7b (8), trans-7b (52)
95c (4-MeS)Pd2(dba)3 (10)18/82/036/466c (6)
cis-7c (4), trans-7c (30)
105c (4-MeS)((IMes)Pd(NQ))2 (2.5)11/89/068/216c (7)
cis-7c (30), trans-7c (50)
115c (4-MeS)(Pd(allyl)Cl)2 (5)5/95/042/53cis-7c (10), trans-7c (48)
125c (4-MeS)(SIPr)Pd(allyl)Cl (15)12/88/048/406c (7)
cis-7c (4), trans-7c (60)
135d (4-Cl)Pd2(dba)3 (10)15/85/08/776d (9)
cis-7d (5), trans-7d (50)
145d (4-Cl)((IMes)Pd(NQ))2 (2.5)5/95/069/266d (5)
cis-7d (53), trans-7d (19)
155d (4-Cl)(Pd(allyl)Cl)2 (5)0/100/052/48cis-7d (42), trans-7d (40)
165d (4-Cl)(SIPr)Pd(allyl)Cl (15)0/100/062/38cis-7d (27), trans-7d (42)
175e (4-CN)Pd2(dba)3 (10)3/97/067/30cis-7e (22), trans-7e (38)
185e (4-CN)((IMes)Pd(NQ))2 (2.5)5/75/2025/50cis-7e (15), trans-7e (33)
8e (15)
195e (4-CN)(Pd(allyl)Cl)2 (5)0/100/071/29cis-7e (34), trans-7e (36)
205e (4-CN)(SIPr)Pd(allyl)Cl (15)0/100/038/62cis-7e (27), trans-7e (51)
215f (4-NMe2)Pd2(dba)3 (10)CM------
225f (4-NMe2)((IMes)Pd(NQ))2 (2.5)CM---cis-7f (9), trans-7f (10)
10 (15) 5
235f (4-NMe2)(Pd(allyl)Cl)2 (5)CM---trans-7f (14), 10 (36)
245f (4-NMe2)(SIPr)Pd(allyl)Cl (15)CM---trans-7f (10), 10 (45)
255g (3-Cl)Pd2(dba)3 (10)8/92/054/386g6 (7)
cis-7g (30), trans-7g (43)
265g (3-Cl)((IMes)Pd(NQ))2 (2.5)6/94/073/216g6 (5)
cis-7g (50), trans-7g (20)
275g (3-Cl)(Pd(allyl)Cl)2 (5)0/100/064/36cis-7g (44), trans-7g (36)
285g (3-Cl)(SIPr)Pd(allyl)Cl (15)0/100/050/50cis-7g (30), trans-7g (31)
295h (3-CN)Pd2(dba)3 (10)0/100/033/67cis-7h (25), trans-7h (48)
305h (3-CN)((IMes)Pd(NQ))2 (2.5)4/74/2230/44cis-7h (15), trans-7h (37)
8h (6)
315h (3-CN)(Pd(allyl)Cl)2 (5)0/100/047/53cis-7h (22), trans-7h (43)
325h (3-CN)(SIPr)Pd(allyl)Cl (15)0/100/029/71cis-7h (15), trans-7h (55)
335i (2-MeO)Pd2(dba)3 (10)14/86/043/436i (8)
cis-7i (12), trans-7i (19)
345i (2-MeO)((IMes)Pd(NQ))2 (2.5)20/80/057/236i (10)
cis-7i (45), trans-7i (24)
355i (2-MeO)(Pd(allyl)Cl)2 (5)5/95/042/53cis-7i (23), trans-7i (31)
365i (2-MeO)(SIPr)Pd(allyl)Cl (15)6/94/039/55cis-7i (10), trans-7i (63)
375j (2-F)Pd2(dba)3 (10)8/92/08/84cis-7j (5), trans-7j (66)
385j (2-F)((IMes)Pd(NQ))2 (2.5)5/83/1224/59cis-7j (10), trans-7j (62)
8j (11)
395j (2-F)(Pd(allyl)Cl)2 (5)0/100/069/31cis-7j (26), trans-7j (56)
405j (2-F)(SIPr)Pd(allyl)Cl (15)0/100/00/100trans-7j (66)
415k (2-Br)Pd2(dba)3 (10)7/93/070/236k (6)
cis-7k (46), trans-7k (15)
425k (2-Br)((IMes)Pd(NQ))2 (2.5)0/85/1566/19cis-7k (43), trans-7k (20)
8k (12)
435k (2-Br)(Pd(allyl)Cl)2 (5)14/86/052/346k (9)
cis-7k (34), trans-7k (31)
445k (2-Br)(SIPr)Pd(allyl)Cl (15)8/92/021/716k (4)
cis-7k (8), trans-7k (42)
455l (2-I)Pd2(dba)3 (10)0/100/09/91cis-7l (5), trans-7l (45)
465l (2-I)((IMes)Pd(NQ))2 (2.5)0/78/2266/12cis-7l (64), trans-7l (12)
8l (15)
475l (2-I)(Pd(allyl)Cl)2 (5)8/92/050/426l (9)
cis-7l (49), trans-7l (38)
485l (2-I)(SIPr)Pd(allyl)Cl (15)0/100/077/23cis-7l (53), trans-7l (15)
495m (3-MeO, 4-MeO)Pd2(dba)3 (10)50/50/021/296m (25)
cis-7m (11), trans-7m (20)
505m (3-MeO, 4-MeO)((IMes)Pd(NQ))2 (2.5)32/68/044/246m (13)
cis-7m (23), trans-7m (30)
515m (3-MeO, 4-MeO)(Pd(allyl)Cl)2 (5)15/85/08/776m (6)
trans-7m (39)
525m (3-MeO, 4-MeO)(SIPr)Pd(allyl)Cl (15)24/76/043/336m (9)
cis-7m (6), trans-7m (38)
1 Reaction conditions: Catalyst (see table) in DCE at reflux for 24 h. 2 Ratios determined by integration of characteristic 1H-NMR absorptions from the spectrum of the reaction mixture. 3 The majority of reactions were performed twice, while the Buchner:β-lactam:γ-lactam ratio (6/7/8) was essentially the same in the two runs and the cis:trans ratio was quite different due to the partial isomerization of cis β-lactams to the more stable trans isomers during the work-up or even when recording the 1H-NMR spectra. 4 Yields refer to products isolated by chromatography. 5 4-(Dimethylamino)benzaldehyde (10). 6 Mixture of regioisomers.
Table 4. Transition metal-catalyzed reactions of α-diazoamide 11 1.
Table 4. Transition metal-catalyzed reactions of α-diazoamide 11 1.
Molecules 24 03551 i012
EntryCatalyst (mol%)SolventTemp.TimeProducts (Yield (%)) 2
1((IMes)Pd(NQ))2 (2.5)DCEreflux24 h---
2((IMes)Pd(NQ))2 (2.5)CH2Cl2reflux24 h11
3(Pd(allyl)Cl)2 (5)DCEreflux24 hcis-12/trans-12 (1:10, 30%)
4(SIPr)Pd(allyl)Cl (15)DCEreflux48 hcis-12/trans-12 (1:2.2, 54%)
5(Rh(OAc)2)2 (3)CH2Cl2r.t.24 h11
1 All reactions were conducted with 11 (0.2 mmol). 2 Isolated yield.

Share and Cite

MDPI and ACS Style

Solé, D.; Pérez-Janer, F.; Amenta, A.; Bennasar, M.-L.; Fernández, I. Site Selectivity in Pd-Catalyzed Reactions of α-Diazo-α-(methoxycarbonyl)acetamides: Effects of Catalysts and Substrate Substitution in the Synthesis of Oxindoles and β-Lactams. Molecules 2019, 24, 3551. https://doi.org/10.3390/molecules24193551

AMA Style

Solé D, Pérez-Janer F, Amenta A, Bennasar M-L, Fernández I. Site Selectivity in Pd-Catalyzed Reactions of α-Diazo-α-(methoxycarbonyl)acetamides: Effects of Catalysts and Substrate Substitution in the Synthesis of Oxindoles and β-Lactams. Molecules. 2019; 24(19):3551. https://doi.org/10.3390/molecules24193551

Chicago/Turabian Style

Solé, Daniel, Ferran Pérez-Janer, Arianna Amenta, M.-Lluïsa Bennasar, and Israel Fernández. 2019. "Site Selectivity in Pd-Catalyzed Reactions of α-Diazo-α-(methoxycarbonyl)acetamides: Effects of Catalysts and Substrate Substitution in the Synthesis of Oxindoles and β-Lactams" Molecules 24, no. 19: 3551. https://doi.org/10.3390/molecules24193551

Article Metrics

Back to TopTop