首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
A dual catalytic chemo-selective cross-coupling reaction of two enals is developed. One enal (without α-substitution) is activated by an NHC catalyst to form an acylazolium enolate intermediate that undergoes Michael-type addition to another enal molecule bearing an alkynyl substituent. Mechanistic studies indicate that non-covalent interactions between the alkynyl enal and the NHC·HX catalyst play important roles in substrate activation and enantioselectivity control. Many of the possible side reactions are not observed. Our reaction provides highly chemo- and diastereo-selective access to chiral lactones containing functionalizable 1,3-enyn units with excellent enantioselectivities (95 to >99% ee).

An NHC-catalyzed dual activation of two different enals is disclosed with both covalent and non-covalent activation pathways involved.

The development of chemo-selective reactions of two or more substrates bearing similar functional groups remains a classic challenge in organic synthesis.1 Enals (α,β-unsaturated aldehydes) are common building blocks that offer multiple useful modes of reactions. For instance, enals are readily used as Michael acceptors in many reactions including organic catalytic reactions mediated by amines.2 In the area of N-heterocyclic carbene (NHC) organocatalysis,3 enals are used as precursors of several NHC-bound intermediates, including Breslow acyl anion intermediates,4 homoenolate intermediates,5 enolate intermediates,6 and acylazolium intermediates.7 Somewhat surprisingly, on the other hand, there is little success in using enals as Michael acceptors to react with any of these NHC-bound intermediates.8 Elegant studies in this direction are from Scheidt, in which they showed that in the presence of an NHC catalyst, a homo coupling reaction of enals (with one enal molecule as the Michael acceptor) occurred effectively (Fig. 1a, top side).8a,c Berkessel reported an intramolecular reaction of two enal moieties (in one molecule) to form a bicyclic lactone adduct in the presence of an achiral NHC catalyst (Fig. 1a, bottom side).8b To the best of our knowledge, the intermolecular Michael addition reaction of two different enal substrates mediated by NHC catalysts has not been reported.9 Possible reasons for the difficulties of enals to behave as effective Michael acceptors likely include: (a) the relatively low electrophilicity of the α,β-unsaturated bonds of enals under the typical NHC catalytic conditions and (b) the presence of competing reactions involving both the alkene and aldehyde moieties of enals.Open in a separate windowFig. 1NHC-catalyzed reactions (a) with enals as Michael acceptors, (b) via cross intermolecular reactions of two enals, and (c) bio-active molecules bearing alkyne units.Here we disclose the first cross intermolecular reaction of two enals catalyzed by NHC catalysts (Fig. 1b). We envisioned that installation of an alkynyl substituent at the α-position of an enal can likely promote its reactivity as a Michael acceptor.10 The presence of an α-substituent can interrupt π-conjugations and thus minimize its reactivity via the corresponding enal-derived enolate/homoenolate intermediate formed with NHC, as shown by Bode, Glorius and others.6b,11 In addition, the alkynyl substituent can promote hydrogen-bonding interactions to increase the electrophilicity of the enal to react as a Michael acceptor, as observed in Jørgensen''s amine-catalyzed reactions.12 In our present study, a non-linear effect was observed regarding enantiomeric excesses of the NHC catalyst and the catalytic reaction product. The reaction enantioselectivity was also found to be sensitive to solvents and bases. These results suggested that the NHC and its azolium salt pre-catalyst (NHC·HX) played dual roles in our reaction: one is to activate the α-unsubstituted enal via the formation of the NHC-bound enolate intermediate,6 the other is to activate the α-alkynyl substituted enal via the acidic proton of the chiral NHC·HX (Fig. 1b, intermediate I & transition state TS-I).13 With respect to applications, carbon–carbon triple bonds are found in a good number of bioactive molecules such as cleviolide, (+)-prelaureatin, and oxamflatin (Fig. 1c).14 We demonstrated that our products containing these alkynyl units could be readily transformed into a diverse set of molecules.Cinnamaldehyde 1a and α-alkynyl enal 2a were chosen as the model substrates to search for suitable cross coupling reaction conditions (Table 1). The reactions were first carried out with Et3N as the base and THF as the solvent. When aminoindanol derived azoium salt A15 was used as the NHC pre-catalyst, the desired formal [4 + 2] product (3a) was obtained in a very encouraging yield (52%) with excellent ee and dr values (entry 1). The reactions appeared to be very sensitive to the structure of the NHC pre-catalysts, as similar azolium salts with N-phenyl or N–C6F5 substituents (B16 and C17) were completely ineffective, leading to no product formation (entries 2 & 3). Additional studies on the NHC pre-catalysts finally revealed that introduction of a Br substituent in the indane phenyl ring of the catalyst (D)18 led to 3a in 85% yield with 99% ee as nearly a single diastereomer (entry 4). Replacing Et3N with DIEA led to similar results (entry 5). Very interestingly, when the bases were replaced with DABCO or K3PO4, a significant drop in the enantioselectivity was observed (entries 6 & 7; see the ESI for more details). Changing the solvent from THF to CHCl3 or EtOAc has moderate effects on reaction yields (entries 8 & 9).Optimization of reaction conditionsa
EntryNHCBaseSolventYieldb (%)Eec (%)Drd
1 A Et3NTHF5299>20 : 1
2 B Et3NTHF0
3 C Et3NTHF0
4 D Et3N THF 85 99 >20 : 1
5 D DIEATHF8398>20 : 1
6 D DABCOTHF7267>20 : 1
7 D K3PO4THF8079>20 : 1
8 D Et3NCHCl36497>20 : 1
9 D Et3NEtOAc6899>20 : 1
Open in a separate windowaUnless otherwise specified, the reactions were carried using 1a (0.15 mmol), 2a (0.1 mmol), NHC (0.02 mmol), base (0.05 mmol) and solvent (1.0 mL) at rt for 24 h.bIsolated yield of 3a.cThe ee values were determined via HPLC on a chiral stationary phase.dDr values were determined via1H NMR of the crude reaction mixture.Our reactions are highly chemo-selective. Under all these conditions (Table 1), several possible side products were not formed. For example, possible adducts with enal 2a as the enolate precursor (to form 3a′ or 3a′′′) were not observed. This is not a complete surprise as α-substituted enals are unreactive azolium enolate intermediate precursors under NHC catalysis.11 Our results showed that mixing of enal 2a with highly reactive electrophiles (such as alkylidene diketone; see the ESI for more details) did not lead to any formal [2 + 4] addition product. Interestingly, the simple enal 1a did not behave as a Michael acceptor under our conditions, as homo-coupling adduct 3a′′ was not observed. In Scheidt''s elegant study, the introduction of a Lewis acid additive is necessary to activate one molecule of the enal to react as a Michael acceptor.8aOur further control experiments showed that when the α-alkynl substituent of 2a was replaced with an alkyl (e.g., Fig. 2, 2a1), vinyl (2a2), phenyl (2a3) or cyano (2a4) unit, the corresponding cross [2 + 4] reactions were not observed, with most of the starting materials remaining unchanged (for more details, see the ESI). It is clear that the alkynl unit present in enal 2a played more important roles than simply blocking the enal α-carbon to interrupt the π-conjugations. Although attempts to identify key intermediates (and possible non-covalent interactions) between the NHC catalysts and the two enals did not lead to conclusive mechanistic pictures, our experiments did show strong non-linear effects with respect to the optical purities of the NHC pre-catalyst and the reaction product (Fig. 3, see the ESI for more details).Open in a separate windowFig. 2Unsuccessful α-substituted enal substrates for the NHC catalytic chemo-selective cross [2 + 4] reactions.Open in a separate windowFig. 3Nonlinear effects with respect to the product ee and the catalyst ee values using different bases: (a) Et3N and (b) DABCO.Specifically, the reaction of 1a and 2a was studied by varying the enantiomeric purities of the NHC pre-catalyst D under the optimized reaction conditions as indicated in Table 1, entry 4 (Fig. 3). The ee values of the products and the ee values of the catalysts showed an obvious negative nonlinear effect (Fig. 3a). This nonlinear effect suggests that at least two catalysts are involved in the enantio-differentiating step of our reaction.19 It appears both of the enals (1a and 2a) are activated by NHC and/or its salt (NHC·HX) in our formal [2 + 4] reaction. It is well established that cinnamaldehyde (1a) can be activated by NHC to form an acylazolium enolate intermediate.6 We therefore propose that the other enal (2a) bearing an alkynyl unit is activated by the acidic proton from NHC·HX via non-covalent interactions. These non-covalent interactions between 2a and NHC·HX could be further supported by the “linear-effect” shown by the ee values of the products and the catalysts when using DABCO as the base (Fig. 3b). In this case, only one catalyst was involved in the enantio-differentiating step of our reaction, since the non-covalent H-bonding interactions between 2a and NHC·HX could be broken by a stronger base (e.g., DABCO, K3PO4, see the ESI for details) existing in the catalytic system. Similar activation of the α-alkynyl enal by a proton was proposed in Jørgensen''s amine-catalyzed reaction.11 In the field of NHC related catalysis, the use of NHC·HX as a H-bond donating catalyst has been demonstrated by Huang, Scheidt, Guin, and others.13The non-covalent interactions between the NHC pre-catalyst D and the alkynyl enal 2a can also be supported by 1H NMR analysis (Fig. 4). In the presence of the weak base Et3N, the acidic proton of the NHC pre-catalyst D shows an obvious change in the chemical shift after mixing with the alkynyl enal 2a (Fig. 4, a vs. b). Meanwhile, the chemical shift of the aldehyde proton of the substrate 2a is not changed in the same reaction system (a vs. c). These results support the existence of a non-covalent interaction between the NHC pre-catalyst D and the alkynyl enal 2a in our NHC organocatalytic reaction system (for more details, see the ESI).Open in a separate windowFig. 4Chemical shift of the acidic proton of the NHC pre-catalyst D under various conditions.We then examined the substrate scope using different enals (1) to react with 2a under the optimized reaction conditions indicated in Table 1, entry 4 (Scheme 1). Substituents could be installed at each position of the phenyl ring of the cinnamaldehyde 1a, with all the products afforded in moderate to excellent yields with excellent chemo-, enantio- and diastereoselectivities (3b to 3p). The β-phenyl rings of the enal substrates (1) could also be switched to a naphthyl group or heteroaromatic groups. The corresponding products were afforded in excellent enantioselectivities, although the yields or dr values dropped in these cases (3q to 3s). To our delight, aliphatic enals could also be used as the enolate precursors for this NHC catalyzed chemoselective reaction, with the desired products afforded in moderate yields with excellent dr and ee values (3t & 3u).Open in a separate windowScheme 1Scope of enals 1. aReaction conditions as stated in Table 1, entry 4. Yields are isolated yields after purification by column chromatography. Er values were determined via HPLC on a chiral stationary phase. bThe reaction was carried out on a 1.0 mmol scale based on 2a.The scope of the α-alkynyl enal substrates (2) was also examined (Scheme 2). Electron-donating substituents could be well tolerated on the β-phenyl rings of the α-alkynyl enals, with the desired products afforded in good yields with excellent ee values as single diastereomers (4a & 4b). The yields of the [2 + 4] products decreased when installing electron-withdrawing groups at any position of the β-phenyl rings, although the enantioselectivities were not affected (4c to 4f). The β-phenyl rings of the α-alkynyl enal substrates (2) could also be replaced with various heteroaromatic groups without obvious reduction in the product yields or stereoselectivities (4g & 4h). Substituents were also well tolerated on the phenyl rings attached to the alkynyl units of the enal substrates 2, with all the corresponding products afforded in moderate to good yields with excellent optical purities as single diastereomers (4i to 4p). Enal substrates 2 bearing heteroaromatic, aliphatic or terminal α-alkynyl groups also worked well in this reaction and gave the target products in moderate to good yields with excellent enantio- and diastereoselectivities (4q to 4w).Open in a separate windowScheme 2Scope of α-alkynyl enals 2. aReaction conditions as stated in Table 1, entry 4. Yields are isolated yields after purification by column chromatography. Er values were determined via HPLC on a chiral stationary phase. bThe reaction was carried out on a 6.4 mmol scale based on 2v (1.0 g).As a technical note, this chemo-selective reaction of α,β-unsaturated enals could be carried out on a large scale without reduction of the product ee or dr values, although the yields of the final products slightly dropped (e.g., Scheme 1, 3a & Scheme 2, 4v).Having examined the reaction scope with both of the enal reactants, we next seek to get additional insights into the reaction mechanism. Hammett studies20 were carried out using alkynyl enal substrates 2 bearing various p-substituents on the phenyl groups of the alkynyl units (Fig. 5). Alkynyl enal substrates 2 bearing 4-F (2i), 4-Cl (2j), 4-CF3 (2k), 4-CH3 (2l), and 4-OCH3 (2m) groups were chosen as the target substrates to evaluate their relative reaction rates compared with the alkynyl enal 2a. Kinetic studies showed that the substrates 2 bearing electron-withdrawing groups reacted faster than those bearing electron-donating groups (Fig. 5a). The Hammett plot of the relative reaction rates of the substrates 2i to 2m gave a positive slope (ρ = 1.0128). Therefore, a negatively charged transition state should be built up in the rate determining step of this [2 + 4] cycloaddition process. This is in accordance with the non-covalent H-bonding interactions that we have proposed to exist between the acidic proton of the NHC-precatalyst D and the alkynyl unit of the enal substrate 2 (Fig. 1b, TS-I, see the ESI for more details).Open in a separate windowFig. 5(a) Kinetic data and (b) Hammet plot for the competitive [2 + 4] cycloaddition reactions based on the σ values.Additionally, substrates 2x and 2y bearing steric bulky substituted phenyl groups on the alkynyl units were further examined for this NHC dual catalytic [2 + 4] cycloaddition reaction (Fig. 6). It is not surprising that the corresponding reaction products 4x and 4y were only afforded in poor yields with moderate ee values. Because the alkynly groups of the substrates 2x and 2y were shielded by the bulky mesityl and 2,6-diisopropylphenyl groups, the H-bonding interactions between the NHC pre-catalyst D and the alkynyl groups cannot be efficiently formed in these cases.Open in a separate windowFig. 6Reactions with enals 2 bearing bulky alkynyl substituents.The chiral alkynyl pyranone products obtained from this methodology are rich in functionalities for further synthetic transformations (Fig. 7). For instance, the alkynyl group in 3a could react with the adjacent phenyl group under the catalysis of Cu(OTf)2 to give tricyclic product 5 in a good yield without reduction of the optical purity.21 The terminal alkylnyl group in 4v could participate in various addition reactions and afford a variety of multi-functionalized alkene products in moderate to excellent yields with excellent ee values as single diastereomers (e.g., 6, 7, 8, 10).22 A click reaction between the alkynyl group in 4v and benzyl azide led to the formation of the chiral triazole product 9 in almost quantitative yield with excellent optical purity as a single diastereomer.22d The ethynyl group in 4v could be selectively reduced to an ethyl group with a Pd/C and CaCO3 catalyst in a hydrogen atmosphere (11). Pyranone 4w bearing a 2-trimethylsilylethynyl group could be coupled with 2-iodoaniline to give the indole product 12 in a moderate yield and diastereoselectivity with an excellent ee value.23Open in a separate windowFig. 7Synthetic transformations of the chiral pyranone products. aCuBr2, CH3CN, r.t.,1 h; bTogni reagent, TMSCN, Cu(OAc)2, terpyridine, CH3CN, 70 °C, 5 h; cTosNHNH2, FeCl3, TBHP, CH3CN, 80 °C, 8 h; dBnN3, sodium l-ascorbate, DCM/H2O (v/v = 1/1), r.t., 12 h; eNaI, TMSCl, H2O, CH3CN, r.t., 4 h; fPd/C, CaCO3, H2 (balloon), EtOH, r.t., 2 h.  相似文献   

2.
Simple α-(bromomethyl)styrenes can be processed to a variety of 1,1-difluorinated electrophilic building blocks via I(I)/I(III) catalysis. This inexpensive main group catalysis strategy employs p-TolI as an effective organocatalyst when combined with Selectfluor® and simple amine·HF complexes. Modulating Brønsted acidity enables simultaneous geminal and vicinal difluorination to occur, thereby providing a platform to generate multiply fluorinated scaffolds for further downstream derivatization. The method facilitates access to a tetrafluorinated API candidate for the treatment of amyotrophic lateral sclerosis. Preliminary validation of an enantioselective process is disclosed to access α-phenyl-β-difluoro-γ-bromo/chloro esters.

Simple α-(bromomethyl)styrenes can be processed to a variety of 1,1-difluorinated electrophilic building blocks via I(I)/I(III) catalysis.

Structural editing with fluorine enables geometric and electronic variation to be explored in functional small molecules whilst mitigating steric drawbacks.1 This expansive approach to manipulate structure–function interplay continues to manifest itself in bio-organic and medicinal chemistry.2 Of the plenum of fluorinated motifs commonly employed, the geminal difluoromethylene group3 has a venerable history.4 This is grounded in the structural as well as electronic ramifications of CH2 → CF2 substitution, as is evident from a comparison of propane and 2,2-difluoropropane (Fig. 1, upper). Salient features include localized charge inversion (C–Hδ+ to C–Fδ) and a widening of the internal angle from 112° to 115.4°.5 Consequently, geminal difluoromethylene groups feature prominently in the drug discovery repertoire6 to mitigate oxidation and modulate physicochemical parameters. Catalysis-based routes to generate electrophilic linchpins that contain the geminal difluoromethylene unit have thus been intensively pursued, particularly in the realm of main group catalysis.7–9 Motivated by the potential of this motif in contemporary medicinal chemistry, it was envisaged that an I(I)/I(III) catalysis platform could be leveraged to convert simple α-(bromomethyl)styrenes to gem-difluorinated linchpins: the primary C(sp3)–Br motif would facilitate downstream synthetic manipulations (Fig. 1, lower). To that end, p-TolI would function as a catalyst to generate p-TolIF2in situ in the presence of an external oxidant10 and an amine·HF complex. Alkene activation (I) with subsequent bromonium ion formation (II)11 would provide a pre-text for the first C–F bond forming process (III) with regeneration of the catalyst. A subsequent phenonium ion rearrangement12/fluorination sequence (III and IV) would furnish the geminal difluoromethylene group and liberate the desired electrophilic building block.Open in a separate windowFig. 1The geminal difluoromethylene group: bioisosterism, and catalysis-based access from α-(bromomethyl)styrenes via I(I)/I(III) catalysis.To validate this conceptual framework, a short process of reaction optimization (1a → 2a) was conducted to assess the influence of solvent, amine·HF ratio (Brønsted acidity)13 and catalyst loading (Table 1). Initial reactions were performed with p-TolI (20 mol%), Selectfluor® (1.5 equiv.) as an oxidant, and CHCl3 as the reaction medium. Variation of the amine : HF ratio was conducted to explore the influence of Brønsted acidity on catalysis efficiency (entries 1–4). An optimal ratio of 1 : 6 was observed enabling the product 2a to be generated in >95% NMR-yield. Although reducing the catalyst loading to 10 and 5 mol% (entries 5 and 6, respectively) led to high levels of efficiency (79% yield with 5 mol%), the remainder of the study was performed with 20 mol% p-TolI. Notably, catalytic vicinal difluorination was not observed at any point during this optimization, in contrast with previous studies from our laboratory.9d,i A solvent screen revealed the importance of chlorinated solvents (entries 7 and 8): in contrast, performing the reaction in ethyl trifluoroacetate (ETFA) and acetonitrile resulted in a reduction in yield (9 and 10). Finally, a control reaction in the absence of p-TolI confirmed that an I(I)/I(III) manifold was operational (entry 11). An expanded optimization table is provided in the ESI.Reaction optimizationa
EntrySolventAmine/HFCatalyst loading [mol%]Yieldb [%]
1CHCl31 : 4.52072
2 CHCl 3 1 : 6.0 20 >95
3CHCl31 : 7.52094
4CHCl31 : 9.232087
5CHCl31 : 6.01087
6CHCl31 : 6.0579
7DCM1 : 6.020>95
8DCE1 : 6.02093
9ETFA1 : 6.02084
10MeCN1 : 6.02050
11CHCl31 : 6.00<5
Open in a separate windowaStandard reaction conditions: 1a (0.2 mmol), Selectfluor® (1.5 equiv.), amine : HF source (0.5 mL), solvent (0.5 mL), p-TolI, 24 h, rt.bDetermined by 19F NMR using α,α,α-trifluorotoluene as internal standard.To explore the scope of this geminal difluorination, a series of α-(bromomethyl)styrenes were exposed to the standard reaction conditions (Fig. 2). Gratifyingly, product 2a could be isolated in 80% yield after column chromatography on silica gel. The parent α-(bromomethyl)styrene was smoothly converted to species 2b, as were the p-halogenated systems that furnished 2c and 2d (71 and 79%, respectively). The regioisomeric bromides 2e and 2f (70 and 62%, respectively) were also prepared for completeness to furnish a series of linchpins that can be functionalized at both termini by displacement and cross-coupling protocols (2a, 2e and 2f). Modifying the amine : HF ratio to 1 : 4.5 provided conditions to generate the tBu derivative 2g in 68% yield.14 Electron deficient aryl derivatives were well tolerated as is demonstrated by the formation of compounds 2h–2k (up to 91%). Disubstitution patterns (2l, 81%), sulfonamides (2m, 75%) and phthalimides (2n, 80%) were also compatible with the standard catalysis conditions. Gratifyingly, compound 2n was crystalline and it was possible to unequivocally establish the structure by X-ray crystallography (Fig. 2, lower).15 The C9–C8–C7 angle was measured to be 112.6° (cf. 115.4° for 2,2-difluoropropane).5 Intriguingly, the C(sp3)–Br bond eclipses the two C–F bonds rather than adopting a conformation in which dipole minimization is satisfied (F1–C8–C9–Br dihedral angle is 56.3°).Open in a separate windowFig. 2Exploring the scope of the geminal difluorinative rearrangement of α-(bromomethyl)styrenes via I(I)/I(III) catalysis. Isolated yields after column chromatography on silica gel are reported. X-ray crystal structure of compound 2n (CCDC 2055892). Thermal ellipsoids shown at 50% probability.Cognizant of the influence of Brønsted acidity on the regioselectivity of I(I)/I(III) catalyzed alkene difluorination,9d the influence of the amine : HF ratio on the fluorination of electronically non-equivalent divinylbenzene derivatives was explored (Fig. 3, top). Initially, compound 3 bearing an α-(trifluoromethyl)styrene motif was exposed to the standard catalysis conditions with a 1 : 4.5 amine : HF ratio. Exclusive, chemoselective formation of 4 was observed in 79% yield. Simple alteration of the amine : HF ratio to 1 : 7.5 furnished the tetrafluorinated product 5 bearing both the geminal and vicinal difluoromethylene16 groups (55% yield. 20% of the geminalgeminal product was also isolated. See ESI). Relocating the electron-withdrawing group (α-CF3 → β-CO2Me) and repeating the reaction with 1 : 4.5 amine : HF generated the geminal CF2 species 7 in analogy to compound 4. However, increasing the amine : HF ratio to 1 : 6.0 led exclusively to double geminal difluorination (8, 55%).Open in a separate windowFig. 3Exploring the synthetic versatility of this platform. (Top) Leveraging Brønsted acidity to achieve chemoselective fluorination. (Centre) Bidirectional functionalization. (Bottom) Preliminary validation of an enantioselective variant.Similarly, bidirectional geminal difluorination of the divinylbenzene derivatives 9 and 11 was efficient, enabling the synthesis of 10 (46%) and 12 (70%), respectively. This enables facile access to bis-electrophilic fluorinated linchpins for application in materials chemistry.Preliminary validation of an enantioselective variant8d was achieved using the trisubstituted alkene 13. To that end, a series of C2-symmetric resorcinol-based catalysts were explored (see Fig. 3, inset). This enabled the generation of product 15 in up to 18 : 82 e.r. and 71% isolated yield. It is interesting to note that this catalysis system was also compatible with the chlorinated substrate E-14. A comparison of geometric isomers revealed a matched-mismatched scenario: whilst E-14 was efficiently converted to 16 (75%, 14 : 86 e.r.), Z-14 was recalcitrant to rearrangement (<20%).To demonstrate the synthetic utility of the products, chemoselective functionalization of linchpin 2a was performed to generate 17 (57%) and 18 (87%), respectively (Fig. 4). Finally, this method was leveraged to generate an API for amyotrophic lateral sclerosis. Whereas the reported synthesis17 requires the exposure of α-bromoketone 19 to neat DAST over 7 days,18 compound 2h can be generated using this protocol over a more practical timeframe (24 h) on a 4 mmol scale. This key building block was then processed, via the amine hydrochloride salt 20, to API 21.Open in a separate windowFig. 4Selected modification of building blocks 2a and 2h. Conditions: (a) NaN3, DMF, 110 °C, 16 h. (b) Pd(OH)2/C (10 mol%), EtOH, 1 M HCl, rt, 24 h; (c) CDI, Et3N, THF, 60 °C, 16 h; (d) malonyl chloride, DCM, 0 °C, 2 h.  相似文献   

3.
Convergent paired electrosynthesis is an energy-efficient approach in organic synthesis; however, it is limited by the difficulty to match the innate redox properties of reaction partners. Here we use nickel catalysis to cross-couple the two intermediates generated at the two opposite electrodes of an electrochemical cell, achieving direct arylation of benzylic C–H bonds. This method yields a diverse set of diarylmethanes, which are important structural motifs in medicinal and materials chemistry. Preliminary mechanistic study suggests oxidation of a benzylic C–H bond, Ni-catalyzed C–C coupling, and reduction of a Ni intermediate as key elements of the catalytic cycle.

A direct arylation of benzylic C–H bonds is achieved by integrating Ni-catalyzed benzyl–aryl coupling into convergent paired electrolysis.

Electrochemical organic synthesis has drawn much attention in recent years.1 Compared to processes using stoichiometric redox agents, electrosynthesis can potentially be more selective and safe, generate less waste, and operate under milder conditions.1b In the majority of examples, the reaction of interest occurs at one electrode (anode for oxidation or cathode for reduction), while a sacrificial reaction occurs at the counter electrode to fulfil electron neutrality.1a,2 Paired electrolysis uses both anodic and cathodic reactions for the target synthesis, thereby maximizing energy efficiency.1a,3 However, there are comparatively few examples of paired electrolysis for organic synthesis.1a,3,4Paired electrolysis might be classified into three types: parallel, sequential, and convergent (Fig. 1).1a,3a In parallel paired electrolysis (Fig. 1a), the two half reactions are simultaneous but non-interfering. In sequential paired electrolysis (Fig. 1b), a substrate is oxidized and reduced (or vice versa) sequentially. In convergent paired electrolysis (Fig. 1c), intermediates generated by the anodic and cathodic processes react with one another to yield the product.1a,3a,4b,c,5 The activation mode of all three types of paired electrolysis is based on the innate redox reactivity of substrates. As a result, the types of reactions that could be conducted by paired electrolysis remain limited. We proposed a catalytic version of convergent paired electrolysis, where a catalyst is used to cross-couple the two intermediates generated at the two separated electrodes (Fig. 1d). Although mediators have been used in paired electrosynthesis,3a,4c,6 catalytic coupling of anodic and cathodic intermediates remains largely undeveloped. This mode of action will leverage the power of cross-coupling to electrosynthesis, opening up a wide substrate and product space. Here we report the development of such a process, where cooperative nickel catalysis and paired electrolysis enable direct arylation of benzylic C–H bonds (Fig. 1e).Open in a separate windowFig. 1Different types of paired electrolysis: (a) parallel paired electrolysis, (b) sequential paired electrolysis, (c) convergent paired electrolysis, (d) catalytic convergent paired electrolysis and (e) this work.Our method can be used to synthesize diarylmethanes, which are important structural motifs in bioactive compounds,7 natural products8 and materials.9 Direct arylation of benzylic C–H bonds has been recognized as an efficient strategy to synthesize diarylmethanes, and methods using metal catalysis10 and in particular combined photoredox and transition-metal catalysis have been reported.11 Electrosynthesis provides a complementary approach to these methods, with the potential advantages outlined above. The groups of Yoshida12 and Waldvogel13 previously developed synthesis of diarylmethanes via a Friedel–Crafts-type reaction of a benzylic cation and a nucleophile. The benzylic cations were generated by anodic oxidation of benzylic C–H bonds.14 To avoid the overoxidation of products and to stabilize the very reactive benzylic cations, the reactions had to be conducted in two steps, where the benzylic cations generated in the anodic oxidation step had to be trapped by a reagent. We thought a Ni catalyst could be used to trap the benzyl radical to form an organonickel intermediate, which is then prone to a Ni-catalyzed C–C cross-coupling reaction. Although such a coupling scheme was unprecedented, Ni-catalyzed electrochemical reductive coupling of aryl halides was well established.15 We were also encouraged by a few recent reports of combined Ni catalysis and electrosynthesis for C–N,16 C–S,17 and C–P18 coupling reactions.We started our investigations using the reaction between 4-methylanisole 1a and 4-bromoacetophenone 2a as a test reaction (Table 1). Direct arylation of benzylic C–H bonds was challenging and was typically conducted using toluene derivatives in large excess, e.g., as a solvent.11a,b,11df To improve the reaction efficiency, we decided to use only 3 equivalents of 4-methylanisole 1a relative to 2a. After some initial trials, we decided to conduct the reaction in an undivided cell using a constant current of 3 mA. These conditions are straightforward from a practical point of view. After screening various reaction parameters, we found that a combination of 4,4′-dimethoxy-2-2′-bipyridine (L1) and (DME)NiBr2 as a catalyst, THF/CH3CN (4 : 1) as a solvent, fluorine-doped tin oxide (FTO) coated glass as an anode and carbon fibre as a cathode gave a 50% GC yield of 1-(4-(4-methoxybenzyl)phenyl)ethanone 3a after 18 h (Table 1, entry 1). Extending the reaction time to 36 h improved the yield to 76% (isolated yield) (entry 2). The target products were formed in a diminished yield with other bipyridine type ligands (entries 3–5). Solvents commonly used in Ni-catalyzed cross-coupling reactions, such as DMA and DMF, were less effective (entries 7–8). Replacing carbon fibre by nickel foam or platinum foil as the cathode was detrimental to the coupling, but substantial yields were still obtained (entries 9–10). On the other hand, FTO could not be replaced as the anode. Using carbon fibre as the anode shut down the reaction (entry 11). Likewise, using Pt foil as the anode gave only a 7% GC yield (entry 12). The sensitivity of the reaction outcomes to the electrodes originates from the electrode-dependent redox properties of reaction components (see below). Additional data showing the influence of other reaction parameters such as nickel sources, current, concentration, and electrolytes are provided in the ESI (Table S1, ESI).Summary of the influence of key reaction parametersa
EntryLigandAnodeCathodeSolventYield (%)
1 L1 FTOCarbon fibreTHF/CH3CN = 4 : 156
2 L1 FTOCarbon fibreTHF/CH3CN = 4 : 176b
3 L2 FTOCarbon fibreTHF/CH3CN = 4 : 143
4 L3 FTOCarbon fibreTHF/CH3CN = 4 : 146
5 L4 FTOCarbon fibreTHF/CH3CN = 4 : 121
6 L1 FTOCarbon fibreCH3CN4
7 L1 FTOCarbon fibreDMA15
8 L1 FTOCarbon fibreDMF6
9 L1 FTONi foamTHF/CH3CN = 4 : 145
10 L1 FTOPt foilTHF/CH3CN = 4 : 128
11 L1 Carbon fibre (1 cm2)Carbon fibreTHF/CH3CN = 4 : 10
12 L1 Pt foil (cm2)Carbon fibreTHF/CH3CN = 4 : 17
Open in a separate windowaReaction conditions: 1a (0.6 mmol), 2a (0.2 mmol), (DME)NiBr2 (6 mol%), ligand (7.2 mol mol%), LutHClO4 (0.1 M), and lutidine (0.8 mmol) in solvent (2 mL) at 40 °C. GC yield.bReaction time: 36 h. Isolated yield.With the optimized reaction conditions in hand, we explored the substrate scope (Table 2). A large number of aryl and heteroaryl bromides could be coupled (3a–3x). These substrates may contain electron-rich, neutral, or poor groups. For aryl bromide bearing electron-donating groups, replacing (DME)NiBr2 by Ni(acac)2 gave higher yields (3k–3o). The method tolerates numerous functional groups in the (hetero)aryl bromides, including for example ketone (3a), nitrile (3b, 3u, and 3v), ester (3c, 3m, 3n, and 3s), amide (3d), aryl-Cl (3q), CF3(3i, 3t, and 3w), OCF3(3e), aryl-F(3x), pyridine (3w and 3x), and arylboronic ester (3g). We then probed the scope of benzylic substrates using 4-bromoacetophenone 2a as the coupling partner (3aa–3ai). Toluene and electron-rich toluene derivatives were readily arylated (3aa–3ac). Toluene derivatives containing an electron-withdrawing group such as fluoride (3ad) and chloride (3ae) could also be arylated, although a higher excess of them (10 equiv.) was necessary. More elaborated toluene derivatives containing an additional ester (3af, 3ai) or ether (3ag, 3ah, and 3ai) were also viable.Substrate scopea
Open in a separate windowaReaction conditions: 1 (0.6 mmol), 2 (0.2 mmol), (DME)NiBr2 (6 mol%), L1 (7.2 mol mol%), LutHClO4 (0.1 M), and lutidine (0.8 mmol) in THF/CH3CN (4 : 1, 2 mL) at 40 °C. Isolated yield.b(DME)NiBr2 (5 mol%) and L1 (6 mol%) were used as the catalysts.cNi(acac)2 was used instead of (DME)NiBr2.dSolvent: THF/CH3CN (3 : 1, 2 mL).e2 mmol toluene or its derivative was used as the substrate.fReaction time: 60 h.Linear sweep voltammetry (LSV) was applied to probe the possible processes at both the anode and cathode. The measurements were made in THF/CN3CN (4 : 1, 2 mL) using [LutH]ClO4 (0.1 M) as the electrolyte and lutidine (0.4 M) as an additional base to mimic the coupling conditions. The LSV curves of individual reaction components indicate that only 4-methylanisole 1a and the Ni catalyst may be oxidized at the anode (Fig. 2a). The current at 3 mA appears to be the sum of the oxidation currents of 1a and the Ni catalyst. Meanwhile, LSV curves indicate that only the Ni catalyst might be reduced at the cathode (Fig. 2b).Open in a separate windowFig. 2The LSV curves of different reaction components at the anode or cathode. The components were dissolved in THF/CN3CN (4 : 1, 2 mL); the solution also contained [LutH]ClO4 (0.1 M) and lutidine (0.4 M). Scan rate: 50 mV s−1. (a) The LSV curves of different reaction components at the FTO anode; (b) the LSV curves of different reaction components at the carbon fibre cathode; (c) the LSV curves of different reaction components at the carbon fibre anode.It was observed that FTO was an essential anode for the reactions. If FTO was replaced by a carbon fibre anode, no coupling product was obtained. LSV was performed to probe the oxidation of 1a and the Ni catalyst on a carbon fibre anode (Fig. 2c). The oxidation of the Ni catalyst was much easier on carbon fibre than on FTO. At 3 mA, the oxidation is exclusively due to the Ni catalyst. This result suggests that the absence of coupling on the carbon fibre anode is due to no oxidation of 1a. The different redox properties of 1a and the Ni catalyst observed on different electrodes might be attributed to the different nature of surface species which influence the electron transfer. Although FTO is rarely used in electrosynthesis, it is widely used in electrocatalysis and photoelectrocatalysis for energy conversion.19 FTO is stable, commercially available and inexpensive. In our reactions, the FTO anode could be reused at least three times.The LSV curves in Fig. 2 revealed the issue of “short-circuit” of catalyzed/mediated paired electrolysis in an undivided cell, as the catalyst or mediator can be reduced and oxidized at both the cathode and anode. When carbon fibre or graphite was used as the anode, the short-circuit problem was very severe so that nearly no current was used for electrosynthesis. However, by using an appropriate anode such as FTO, the short-circuit problem was alleviated and around half of the current was used to oxidize the substrate (1a) while the other half was used to oxidize the nickel complex. The remaining short-circuit is one of the reasons why the current efficiencies of the reactions are low (<10%). Another factor contributing to the low current efficiency is the instability of the benzyl radical, which can abstract hydrogen from the solvent to regenerate the substrate. Nevertheless, useful products could be obtained in synthetically useful yields under conditions advantageous to previous methods.For the test reaction (Table 1), a small amount of homo-coupling product bis(4-methoxyphenyl)methane (<2%) was detected by GC-MS under the optimized conditions. In the absence of ligand L1, the yield of the homo-coupling products increased (∼8%). In the presence of a radical acceptor, the electron-withdrawing alkene vinyl benzoate, the product originating from the addition of a benzyl radical to the olefin was obtained in about 12% GC yield (ESI, Scheme S1). These data support the formation of a benzyl radical intermediate. As bromide existed in our reaction system, it is possible to be oxidized to form a bromine radical. Previous studies showed that a bromine radical can react with a toluene derivative to give a benzyl radical.11b,g,20 To probe the involvement of the Br radical, we conducted a coupling of 4-methylanisole 1a with 4′-Iodoacetophenone, using Ni(acac)2 instead of (DME)NiBr2 as the Ni source. We obtained a GC yield of 24% for the coupling after 18 h (Scheme S2). This result suggests that a Br-free path exists for the coupling, although a non-decisive involvement of Br/Br˙ cannot be ruled out.Based on the data described above, we propose a mechanism for the coupling (Scheme 1). The oxidation of a toluene derivative at the anode gives a benzyl radical. This radical is trapped by a LNi(ii)(Ar)(Br) species (B) in the solution to give a LNi(iii)(Ar)(benzyl)(Br) intermediate (C). The latter undergoes reductive elimination to give a diarylmethane and a LNi(i)(Br) species (A). There are at least two ways A can be convert to B to complete the catalytic cycle: either by oxidative addition of ArBr followed by a 1-e reduction at the cathode or by first 1-e reduction to form a Ni(0) species followed by oxidation addition of ArBr. In addition to a toluene derivative, a Ni species is oxidized at the anode. We propose that this oxidation is an off cycle event, which reduces the faradaic and catalytic efficiency but does not shut down the productive coupling.Open in a separate windowScheme 1Proposed mechanism of the direct arylation of benzylic C–H bonds.  相似文献   

4.
The asymmetric total synthesis of (+)-xestoquinone and (+)-adociaquinones A and B was achieved in 6–7 steps using an easily accessible meso-cyclohexadienone derivative. The [6,6]-bicyclic decalin B–C ring and the all-carbon quaternary stereocenter at C-6 were prepared via a desymmetric intramolecular Michael reaction with up to 97% ee. The naphthalene diol D–E ring was constructed through a sequence of Ti(Oi-Pr)4-promoted photoenolization/Diels–Alder, dehydration, and aromatization reactions. This asymmetric strategy provides a scalable route to prepare target molecules and their derivatives for further biological studies.

The asymmetric total synthesis of (+)-xestoquinone and (+)-adociaquinones A and B was achieved in 6–7 steps using an easily accessible meso-cyclohexadienone derivative.

Various halenaquinone-type natural products with promising biological activity have been isolated from marine sponges of the genus Xestospongia1 from the Pacific Ocean. (+)-Halenaquinone (1),2,3 (+)-xestoquinone (2), and (+)-adociaquinones A (3) and B (4)4,5 bearing a naphtha[1,8-bc]furan core (Fig. 1) are the most typical representatives of this family. Naturally occurring (−)-xestosaprol N (5) and O (6)6,7 have the same structure as 3 and 4 except for a furan ring, while a naphtha[1,8-bc]furan core can also be found in fungus-isolated furanosteroids (−)-viridin (7) and (+)-nodulisporiviridin E (8)8,9 (Fig. 1). Halenaquinone (1) was first isolated from the tropical marine sponge Xestospongia exigua2 and it shows antibiotic activity against Staphylococcus aureus and Bacillus subtilis. Xestoquinone (2) and adociaquinones A (3) and B (4) were firstly isolated, respectively, from the Okinawan marine sponge Xestospongia sp.4a and the Truk Lagoon sponge Adocia sp.,4b and they show cardiotonic,4a,c cytotoxic,4b,i antifungal,4i antimalarial,4j and antitumor4l activities. These compounds inhibit the activity of pp60v-src protein tyrosine kinase,4d topoisomerases I4e and II,4f myosin Ca2+ ATPase,4c,g and phosphatases Cdc25B, MKP-1, and MKP-3.4h,kOpen in a separate windowFig. 1Structure of halenaquinone-type natural products and viridin-type furanosteroids.Owing to their diverse bioactivities, the synthesis of this family of natural compounds has been extensively studied, with published pathways making use of Diels–Alder,3a,d,e,5ac,e,g furan ring transfer,5b Heck,3b,c,5f,7,9b,d palladium-catalyzed polyene cyclization,5d Pd-catalyzed oxidative cyclization,3f and hydrogen atom transfer (HAT) radical cyclization9c reactions. In this study, we report the asymmetric total synthesis of (+)-xestoquinone (2), (−)-xestoquinone (2′), and (+)-adociaquinones A (3) and B (4) (Fig. 1).The construction of the fused tetracyclic B–C–D–E skeleton and the all carbon quaternary stereocenter at C-6 is a major challenge towards the total synthesis of xestoquinone (2) and adociaquinones A (3) and B (4). Based on our retrosynthetic analysis (Scheme 1), the all-carbon quaternary carbon center at C-6 of cis-decalin 12 could first be prepared stereoselectively from the achiral aldehyde 13via an organocatalytic desymmetric intramolecular Michael reaction.10,11 The tetracyclic framework 10 could then be formed via a Ti(Oi-Pr)4-promoted photoenolization/Diels–Alder (PEDA) reaction12–16 of 11 and enone 12. Acid-mediated cyclization of 10 followed by oxidation state adjustment could be subsequently applied to form the furan ring A of xestoquinone (2). Finally, based on the biosynthetic pathway of (+)-xestoquinone (2)4b,5c and our previous studies,7 the heterocyclic ring F of adociaquinones A (3) and B (4) could be prepared from 2via a late-stage cyclization with hypotaurine (9).Open in a separate windowScheme 1Retrosynthetic analysis of (+)-xestoquinone and (+)-adociaquinones A and B.The catalytic enantioselective desymmetrization of meso compounds has been used as a powerful strategy to generate enantioenriched molecules bearing all-carbon quaternary stereocenters.10,11 For instance, two types of asymmetric intramolecular Michael reactions were developed using a cysteine-derived chiral amine as an organocatalyst by Hayashi and co-workers,11a,b while a desymmetrizing secondary amine-catalyzed asymmetric intramolecular Michael addition was later reported by Gaunt and co-workers to produce enantioenriched decalin structures.11c Prompted by these pioneering studies and following the suggested retrosynthetic pathway (Scheme 1), we first screened conditions for organocatalytic desymmetric intramolecular Michael addition of meso-cyclohexadienone 13 (Table 1) in order to form the desired quaternary stereocenter at C-6. Compound 13 was easily prepared on a gram scale via a four-step process (see details in the ESI).Attempts of organocatalytic desymmetric intramolecular Michael additiona
EntryCat. (equiv.)Additive (equiv.)SolventTimeYield/d.r. at C2be.e.c
1(R)-cat.I (0.5)Toluene10.0 h52%/10.3 : 1 14a: 96%; 14b: 75%
2(R)-cat.I (1.0)Toluene4.0 h60%/10.0 : 1 14a: 93%; 14b: 75%
3(R)-cat.I (1.0)MeOH4.0 h47%/5.5 : 1 14a: 86%; 14b: −3%
4(R)-cat.I (1.0)DCM10.0 h28%/24.0 : 1 14a: 91%; 14b: 7%
5(R)-cat.I (1.0)Et2O10.0 h22%/22.0 : 1 14a: 91%; 14b: 65%
6(R)-cat.I (1.0)MeCN10.0 h12%/2.6 : 1 14a: 90%; 14b: 62%
7(R)-cat.I (1.0)Toluene/MeOH (2 : 1)4.0 h47%/10.0 : 1 14a: 87%; 14b: −38%
8d(R)-cat.I (1.0)AcOH (5.0)Toluene4.0 h60%e/2.1 : 1 14a: 96%; 14b: 95%
9d(R)-cat.I (0.5)AcOH (2.0)Toluene6.0 h75%e/4.0 : 1 14a: 97%; 14b: 91%
10d(R)-cat.I (0.5)AcOH (0.2)Toluene6.0 h73%e/4.3 : 1 14a: 96%; 14b: 92%
11f(R)-cat.I (0.5)AcOH (0.2)Toluene6.0 h75%e/8.0 : 1g 14a: 95%; 14b: 93%
12h(R)-cat.I (0.2)AcOH (0.2)Toluene9.0 h80%i/6.0 : 1j 14a: 97%; 14b: 91%
Open in a separate windowaAll reactions were performed using 13 (5.8 mg, 0.03 mmol, 1.0 equiv., and 0.1 M) and a catalyst at room temperature in analytical-grade solvents, unless otherwise noted.bThe yields and diastereoisomeric ratios (d.r.) were determined from the crude 1H NMR spectrum of 14 using CH2Br2 as an internal standard, unless otherwise noted.cThe enantiomeric excess (e.e.) values were determined by chiral high-performance liquid chromatography (Chiralpak IG-H).dCompound 13: 9.6 mg, 0.05 mmol, and 0.1 M.eIsolated combined yield of 14a + 14b.fCompound 13: 192 mg, 1.0 mmol, and 0.1 M.gThe d.r. values decreased to 1 : 1 after purification by silica gel column chromatography.hCompound 13: 1.31 g, 6.82 mmol, and 0.1 M.iIsolated combined yield of 12a + 12b.jThe d.r. values were determined from the crude 1H NMR spectrum of 12 obtained from the one-pot process.We initially investigated the desymmetric intramolecular Michael addition of 13 using (S)-Hayashi–Jørgensen catalysts,17 and found that the absolute configuration of the obtained cis-decalin was opposite to the required stereochemistry of the natural products (see Table S1 in the ESI). In order to achieve the desired absolute configuration of the angular methyl group at C-6, (R)-cat.I was used for further screening. In the presence of this catalyst, the intramolecular Michael addition afforded 14a (96% e.e.) and 14b (75% e.e.) in a ratio of 10.3 : 1 and 52% combined yield (entry 1, Table 1). We assumed that the enantioselectivity of the reaction was controlled by the more sterically hindered aromatic group of (R)-cat.I, which protected the upper enamine face and allowed an endo-like attack by the si-face of cyclohexadienone, as shown in the transition state TS-A (Table 1). In order to increase the yield of this reaction and improve the enantioselectivity of 14b, we further screened solvents and additives. Increasing the catalyst loading from 0.5 to 1.0 equivalents and screening various reaction solvents did not improve the enantiomeric excess of 14b (entries 2–7, Table 1). Therefore, based on previous studies,11d,e we added 5.0 equivalents of acetic acid (AcOH) to a solution of compound 13 and (R)-cat.I in toluene, which improved the enantiomeric excess of 14b to 95% with a 60% combined yield (entry 8, Table 1). And, the stability of (R)-cat.I has also been verified in the presence of AcOH (see Table S2 in the ESI). Further adjustment of the (R)-cat.I and AcOH amount and ratio (entries 9–12, Table 1) indicated that 0.2 equivalents each of (R)-cat.I and AcOH were the best conditions to achieve high enantioselectivity for both 14a and 14b, and it also increased the reaction yield (entry 12, Table 1). The enantioselectivity was not affected when the optimized reaction was performed on a gram scale: 14a (97% e.e.) and 14b (91% e.e.) were obtained in 80% isolated yield (entry 12, Table 1). We also found that the gram-scale experiments needed a longer reaction time which led a slight decrease of the diastereoselectivity. The purification of the cyclized products by silica gel flash column chromatography indicated that the major product 14a was epimerized and slowly converted to the minor product 14b (entry 11, Table 1). Both 14a and 14b are useful in the syntheses because the stereogenic center at C-2 will be converted to sp2 hybridized carbon in the following transformations. Therefore, the aldehyde group of analogues 14a and 14b was directly protected with 1,3-propanediol to give the respective enones 12a and 12b for use in the subsequent PEDA reaction.Afterward, we selected the major cyclized cis-decalins 12a and 12a′ (obtained by using (S)-cat.I in desymmetric intramolecular Michael addition, see Table S1 in the ESI) as the dienophiles to prepare the tetracyclic naphthalene framework 10 through a sequence of Ti(Oi-Pr)4-promoted PEDA, dehydration, and aromatization reactions (Scheme 2). When using 3,6-dimethoxy-2-methylbenzaldehyde (11) as the precursor of diene, no reaction occurred between 12a/12a′ and 11 under UV irradiation at 366 nm in the absence of Ti(Oi-Pr)4 (Scheme 2A). In contrast, the 1,2-dihydronaphthalene compounds 16a and 16a′ were successfully synthesized when 3.0 equivalents of Ti(Oi-Pr)4 were used. Based on our previous studies,13a,e the desired hydroanthracenol 15a was probably generated through the chelated intermediate TS-B and the cycloaddition occurred through an endo direction (Scheme 2B).18 The newly formed β-hydroxyl ketone groups in 15a and 15a′ could then be dehydrated with excess Ti(Oi-Pr)4 to form enones 16a and 16a′. These results confirmed the pivotal role of Ti(Oi-Pr)4 in this PEDA reaction: it stabilized the photoenolized hydroxy-o-quinodimethanes and controlled the diastereoselectivity of the reaction.Open in a separate windowScheme 2PEDA reaction of 11 and enone 12.Subsequent aromatization of compounds 16a and 16a′ with 2,3-dichloro-5,6-dicyanobenzoquinone (DDQ) at 80 °C afforded compounds 10a and 10a′ bearing a fused tetracyclic B–C–D–E skeleton. The stereochemistry and absolute configuration of 10a were confirmed by X-ray diffraction analysis of single crystals (Scheme 3). The synthesis of (+)-xestoquinone (2) and (+)-adociaquinones A (3) and B (4) was completed by forming the furan A ring. Compound 10 was oxidized using bubbling oxygen gas in the presence of t-BuOK to give the unstable diosphenol 17a, which was used without purification in the next step. The subsequent acid-promoted deprotection of the acetal group led to the formation of an aldehyde group, which reacted in situ with enol to furnish the pentacyclic compound 18 bearing the furan A ring. The stereochemistry and absolute configuration of 18 were confirmed by X-ray diffraction analysis of single crystals (Scheme 3). Further oxidation of 18 with ceric ammonium nitrate afforded (+)-xestoquinone (2) in 82% yield. Following the same reaction process, (−)-xestoquinone (2′) was also synthesized from 10a′ in order to determine in the future whether xestoquinone enantiomers differ in biological activity. Further heating of a solution of (+)-xestoquinone (2) with hypotaurine (9) at 50 °C afforded a mixture of (+)-adociaquinones A (3) (21% yield) and B (4) (63% yield). We also tried to optimize the selectivity of this condensation by tuning the reaction temperature and pH of reaction mixtures (see Table S3 in the ESI). The 1H and 13C NMR spectra, high-resolution mass spectrum, and optical rotation of synthetic (+)-xestoquinone (2), (+)-adociaquinones A (3) and B (4) were consistent with those data reported by Nakamura,4a,g Laurent,4j Schmitz,4b Harada5a,c and Keay.5dOpen in a separate windowScheme 3Total synthesis of (+)-xestoquinone and (+)-adociaquinones A and B.  相似文献   

5.
Palladium-catalyzed regioselective di- or mono-arylation of o-carboranes was achieved using weakly coordinating amides at room temperature. Therefore, a series of B(3,4)-diarylated and B(3)-monoarylated o-carboranes anchored with valuable functional groups were accessed for the first time. This strategy provided an efficient approach for the selective activation of B(3,4)–H bonds for regioselective functionalizations of o-carboranes.

B–H: site-selective B(3,4)–H arylations were accomplished at room temperature by versatile palladium catalysis enabled by weakly coordinating amides.

o-Carboranes, icosahedral carboranes – three-dimensional arene analogues – represent an important class of carbon–boron molecular clusters.1 The regioselective functionalization of o-carboranes has attracted growing interest due to its potential applications in supramolecular design,2 medicine,3 optoelectronics,4 nanomaterials,5 boron neutron capture therapy agents6 and organometallic/coordination chemistry.7 In recent years, transition metal-catalyzed cage B–H activation for the regioselective boron functionalization of o-carboranes has emerged as a powerful tool for molecular syntheses. However, the 10 B–H bonds of o-carboranes are not equal, and the unique structural motif renders their selective functionalization difficult, since the charge differences are very small and the electrophilic reactivity in unfunctionalized o-carboranes reduces in the following order: B(9,12) > B(8,10) > B(4,5,7,11) > B(3,6).8 Therefore, efficient and selective boron substitution of o-carboranes continues to be a major challenge.Recently, transition metal-catalyzed carboxylic acid or formyl-directed B(4,5)–H functionalization of o-carboranes has drawn increasing interest, since it provides an efficient approach for direct regioselective boron–carbon and boron–heteroatom bond formations (Scheme 1a),9 with major contributions by the groups of Xie,10 and Yan,11 among others.12 Likewise, pyridyl-directed B(3,6)–H acyloxylations (Scheme 1b),13 and amide-assisted B(4,7,8)–H arylations14 (Scheme 1c) have been enabled by rhodium or palladium catalysis, respectively.15,16 Despite indisputable progress, efficient approaches for complementary site-selective functionalizations of o-carboranes are hence in high demand.17 Hence, metal-catalyzed position-selective B(3,4)–H functionalizations of o-carboranes have thus far not been reported.Open in a separate windowScheme 1Chelation-assisted transition metal-catalyzed cage B–H activation of o-carboranes.Arylated compounds represent key structural motifs in inter alia functional materials, biologically active compounds, and natural products.18 In recent years, transition metal-catalyzed chelation-assisted arylations have received significant attention as environmentally benign and economically superior alternatives to traditional cross-coupling reactions.19 Within our program on sustainable C–H activation,20 we have now devised a protocol for unprecedented cage B–H arylations of o-carboranes with weak amide assistance, on which we report herein. Notable features of our findings include (a) transition metal-catalyzed room temperature B–H functionalization, (b) high levels of positional control, delivering B(3,4)-diarylated and B(3)-monoarylated o-carboranes, and (c) mechanistic insights from DFT computation providing strong support for selective B–H arylation (Scheme 1d).We initiated our studies by probing various reaction conditions for the envisioned palladium-catalyzed B–H arylation of o-carborane amide 1a with 1-iodo-4-methylbenzene (2a) at room temperature (Tables 1 and S1). We were delighted to observe that the unexpected B(3,4)-di-arylated product 3aa was obtained in 59% yield in the presence of 10 mol% Pd(OAc)2 and 2 equiv. of AgTFA, when HFIP was employed as the solvent, which proved to be the optimal choice (entries 1–5).21 Control experiments confirmed the essential role of the palladium catalyst and silver additive (entries 6–7). Further optimization revealed that AgOAc, Ag2O, K2HPO4, and Na2CO3 failed to show any beneficial effect (entries 8–11). Increasing the reaction temperature fell short in improving the performance (entries 12 and 13). The replacement of the amide group in substrate 1a with a carboxylic acid, aldehyde, ketone, or ester group failed to afford the desired arylation product (see the ESI). We were pleased to find that the use of 1.0 equiv. of trifluoroacetic acid (TFA) as an additive improved the yield to 71% (entry 14). To our delight, replacing the silver additive with Ag2CO3 resulted in the formation of B(3)–H mono-arylation product 4aa as the major product (entries 15–16).Optimization of reaction conditionsa
EntryAdditiveSolventYield of 3aa/%Yield of 4aa/%
1AgTFAPhMe00
2AgTFADCE00
3AgTFA1,4-Dioxane00
4AgTFATFE213
5AgTFAHFIP594
6AgTFAHFIP00b
7HFIP00
8AgOAcHFIP5<3
9Ag2OHFIP<3<3
10K2HPO4HFIP00
11Na2CO3HFIP00
12AgTFAHFIP534c
13AgTFAHFIP423d
14 AgTFA HFIP 71 <3 e
15Ag2CO3HFIP934f
16 Ag 2 CO 3 HFIP 5 55 f , g
Open in a separate windowaReaction conditions: 1a (0.20 mmol), 2 (0.48 mmol), Pd(OAc)2 (10 mol%), additive (0.48 mmol), solvent (0.50 mL), 25 °C, 16 h, and isolated yield.bWithout Pd(OAc)2.cAt 40 °C.dAt 60 °C.eTFA (0.2 mmol) was added.f 1a (0.20 mmol), 2a (0.24 mmol), Pd(OAc)2 (5.0 mol%), and Ag2CO3 (0.24 mmol).g 2a was added in three portions every 4 h. DCE = dichloroethane, TFE = 2,2,2-trifluoroethanol, HFIP = hexafluoroisopropanol, and TFA = trifluoroacetic acid.With the optimized reaction conditions in hand, we probed the scope of the B–H di-arylation of o-carboranes 1a with different aryl iodides 2 (Scheme 2). The versatility of the room temperature B(3,4)–H di-arylation was reflected by tolerating valuable functional groups, including bromo, chloro, and enolizable ketone substituents. The connectivity of the products 3aa and 3ab was unambiguously verified by X-ray single crystal diffraction analysis.22Open in a separate windowScheme 2Cage B(3,4)–H di-arylation of o-carboranes.Next, we explored the effect exerted by the N-substituent at the amide moiety (Scheme 3). Tertiary amides 1b–1f proved to be suitable substrates with optimal results being accomplished with substrate 1a. The effect of varying the cage carbon substituents R1 on the reaction''s outcome was also probed, and both aryl and alkyl substituents gave the B–H arylation products and the molecular structures of the products 3dd, 3ea and 3fa were fully established by single-crystal X-ray diffraction.Open in a separate windowScheme 3Effect of substituents on B–H diarylation. aAt 50 °C.The robustness of the palladium-catalyzed B–H functionalization was subsequently investigated for the challenging catalytic B–H monoarylation of o-carboranes (Scheme 4). The B(3)–H monoarylation, as confirmed by single-crystal X-ray diffraction analysis of products 4aa and 4ai, proceeded smoothly with valuable functional groups, featuring aldehyde and nitro substituents, which should prove invaluable for further late-stage manipulation.Open in a separate windowScheme 4Cage B(3)–H mono-arylation of o-carboranes.To elucidate the palladium catalysts'' working mode, a series of experiments was performed. The reactions in the presence of TEMPO or 1,4-cyclohexadiene produced the desired product 3aa, which indicates that the present B–H arylation is less likely to operate via radical intermediates (Scheme 5a). The palladium catalysis carried out in the dark performed efficiently (Scheme 5b). Compound 4aa could be converted to di-arylation product 3aa with high efficiency, indicating that 4aa is an intermediate for the formation of the diarylated cage 3aa (Scheme 5c).Open in a separate windowScheme 5Control experiments.To further understand the catalyst mode of action, we studied the site-selectivity of the o-carborane B–H activation for the first B–H activation at the B3 versus B4 position and for the second B–H activation at the B4 versus B6 position using density functional theory (DFT) at the PBE0-D3(BJ)/def2-TZVP+SMD(HFIP)//TPSS-D3(BJ)/def2-SVP level of theory (Fig. 1). Our computational studies show that the B3 position is 5.8 kcal mol−1 more favorable than the B4 position for the first B–H activation, while the B4 position is 3.4 kcal mol−1 more favorable than the B6 position for the second B–H activation. It is noteworthy that here the interaction between AgTFA and a cationic palladium(ii) complex was the key to success, being in good agreement with our experimental results (for more details, see the ESI).Open in a separate windowFig. 1Computed relative Gibbs free energies in kcal mol−1 and the optimized geometries of the transition states involved in the B–H activation at the PBE0-D3(BJ)/def2-TZVP+SMD(HFIP)//TPSS-D3(BJ)/def2-SVP level of theory. (a) First B–H activation transition states at the B3 and B4 positions. (b) Second B–H activation transition states at the B4 and B6 positions. Irrelevant hydrogen atoms in the transition states are omitted for clarity and the bond lengths are given in Å.A plausible reaction mechanism is proposed which commences with an organometallic B(3)–H activation of 1a with weak assistance of the amide group and assistance by AgTFA to form the cationic intermediate I (Scheme 6). Oxidative addition with the aryl iodide 2 affords the proposed cationic palladium(iv) intermediate II, followed by reductive elimination to give the B(3)-mono-arylation product 4aa. Subsequent B(4)-arylation occurs assisted by the weakly coordinating amide to generate the B(3,4)-di-arylation product 3aa. Due to the innate higher reactivity of the B(4)–H bond in intermediate 4aa – which is inherently higher than that of the B(6)–H bond – the B(3,6)-di-arylation product is not formed.Open in a separate windowScheme 6Proposed reaction mechanism.In summary, room temperature palladium-catalyzed direct arylations at cage B(3,4) positions in o-carboranes have been achieved with the aid of weakly coordinating, synthetically useful amides. Thus, palladium-catalyzed B–H activations enable the assembly of a wealth of arylated o-carboranes. This method features high site-selectivity, high tolerance for functional groups, and mild reaction conditions, thereby offering a platform for the design and synthesis of boron-substituted o-carboranes. Our findings offer a facile strategy for selective activations of B(3,4)–H bonds, which will be instrumental for future design of optoelectronics, nanomaterials, and boron neutron capture therapy agents.  相似文献   

6.
A novel and efficient desymmetrizing asymmetric ortho-selective mono-bromination of bisphenol phosphine oxides under chiral squaramide catalysis was reported. Using this asymmetric ortho-bromination strategy, a wide range of chiral bisphenol phosphine oxides and bisphenol phosphinates were obtained with good to excellent yields (up to 92%) and enantioselectivities (up to 98.5 : 1.5 e.r.). The reaction could be scaled up, and the synthetic utility of the desired P-stereogenic compounds was proved by transformations and application in an asymmetric reaction.

A highly efficient desymmetrizing asymmetric bromination of bisphenol phosphine oxides was developed, providing a wide range of chiral bisphenol phosphine oxides and bisphenol phosphinates with high yields and enantioselectivities.

P-Stereogenic compounds are a class of privileged structures, which have been widely present in natural products, drugs and biologically active molecules (Fig. 1a).1–4 In addition, they are also important chiral materials for the development of chiral catalysts and ligands (Fig. 1b), because the chirality of the phosphorus atom is closer to the catalytic center which can cause remarkable stereo-induction.5,6 Thus, the development of efficient methods for the synthesis of P-stereogenic compounds with novel structures and functional groups is very meaningful.5a Conventional syntheses of P-stereogenic compounds mainly depended on the resolution of diastereomeric mixtures and chiral-auxiliary-based approaches, in which stoichiometric amounts of chiral reagents are usually needed.7 By comparison, asymmetric catalytic strategies, including asymmetric desymmetric reactions of dialkynyl, dialkenyl, diaryl and bisphenol phosphine oxides,8–14 (dynamic) kinetic resolution of tertiary phosphine oxides,15 and asymmetric reactions of secondary phosphine oxides,16 can effectively solve the above-mentioned problems and have been considered as the most direct and efficient synthesis methods for constructing P-chiral phosphine oxides (Fig. 1c). Among them, organocatalytic asymmetric desymmetrization methods have been sporadic, in which the reaction sites were mainly limited to the hydroxyl group of bisphenol phosphine oxides that hindered their further transformation.8–11 It is worth mentioning that asymmetric desymmetrization methods, especially organocatalytic desymmetrization reactions, due to their unique advantages of mild reaction conditions and wide substrate scope, have become an important strategy for asymmetric synthesis. Accordingly, the development of efficient organocatalytic desymmetrization strategy for the synthesis of important functionalized P-stereogenic compounds which contain multiple conversion groups is very meaningful and highly desirable.Open in a separate windowFig. 1(a) Examples of natural products containing P-stereogenic centers. (b) P-Stereogenic compound type ligand and catalyst. (c) Typical P-stereogenic compounds'' synthetic strategies.On the other hand, asymmetric bromination has been demonstrated to be one of the most attractive approaches for chiral compound syntheses.17 Since the pioneering work on peptide catalyzed asymmetric bromination for the construction of biaryl atropisomers,18a the reports on constructing axially biaryl atropisomers,18 C–N axially chiral compounds,19 atropisomeric benzamides,20 axially chiral isoquinoline N-oxides,21 and axially chiral N-aryl quinoids22 by electrophilic aromatic bromination have been well developed (Scheme 1a). In comparison, the desymmetrization of phenol through asymmetric bromination to construct central chirality remains a daunting task. Miller discovered a series of tailor made peptide catalyzed enantioselective desymmetrizations of diarylmethylamide through ortho-bromination (Scheme 1b).23 Recently, Yeung realized amino-urea catalyzed desymmetrizing asymmetric ortho-selective mono-bromination of phenol derivatives to fix a new class of potent privileged bisphenol catalyst cores with excellent yields and enantioselectivities (Scheme 1b).24 Despite this elegant work, there is no report on the synthesis of P-centered chiral compounds using the desymmetrizing asymmetric bromination strategy.Open in a separate windowScheme 1(a) Constructing axially chiral compounds by asymmetric bromination. (b) Known synthesis of central chiral compounds via asymmetric bromination. (c) This work: access to P-stereogenic compounds via desymmetrizing enantioselective bromination.Taking into account the above-mentioned consideration, we speculated that bisphenol phosphine oxides and bisphenol phosphinates are potential substrate candidates for desymmetrizing asymmetric bromination to construct P-stereogenic centers. The advantages of using bisphenol phosphine oxides and bisphenol phosphinates as substrates are shown in two aspects. First, the ortho-position of electron rich phenol is easy to take place electrophilic bromination reaction. Second, the corresponding bromination product structure contains abundant synthetic conversion groups, including bromine, hydroxyl group, alkoxy group and phosphoryl group. To achieve this goal, two challenges need to be overcome: (i) finding a suitable chiral catalyst for the desymmetrization process to induce enantiomeric control is troublesome, due to the remote distance between the prochiral phosphorus center and the enantiotopic site; (ii) selectively brominating one phenol to inhibit the formation of an achiral by-product is difficult. Herein, we report a chiral squaramide catalyzed asymmetric ortho-bromination strategy to construct a wide range of chiral bisphenol phosphine oxides and bisphenol phosphinates with good to excellent yields and enantioselectivities (Scheme 1c). It is worth mentioning that the obtained P-stereogenic compounds can be further transformed at multiple sites.Our initial investigation was carried out with bis(2-hydroxyphenyl)phosphine oxide 1a and N-bromosuccinimide (NBS) 2a as the model substrates, 10 mol% chiral amino-thiourea 4a as the catalyst, and toluene as the solvent, which were stirred at −78 °C for 12 h. As a result, the reaction gave the desired desymmetrization product 3a in 65% yield with 56 : 44 e.r. (Table 1, entry 1). Then, thiourea 4b was tested, in which a little better result was obtained (Table 1, entry 2). To our delight, using the chiral squaramides 4c–4f as the catalysts, the enantiomeric ratios of the desymmetrization products had been significantly improved (Table 1, entries 3–6). Especially, when chiral squaramide catalyst 4c was applied to this reaction, the enantiomeric ratio of 3a was increased to 95 : 5 (Table 1, entry 3). To further improve the yield and enantioselectivity, we next optimized the reaction conditions by varying reaction media and additives. As shown in Table 1, the reaction was affected by the solvent dramatically. Product 3a was obtained with low yield and enantioselectivity in DCM (Table 1, entry 7). Also, when Et2O was used as the solvent, the yield and e.r. value of product 3a were all decreased (Table 1, entry 8). As a result, the initial used toluene was the optimal solvent. We also inspected the effect of different bromine sources, and found that the initially used NBS was the optimal one (Table 1, entries 3, 11 and 12). Fortunately, by adjusting the amount of bisphenol phosphine oxides to 1.5 equiv., the yield and the enantiomeric ratio of 3a were increased to 80% and 96.5 : 3.5, respectively (Table 1, entries 3, 13 and 14). Further increasing the amount of bisphenol phosphine oxides to 2.0 equiv. resulted in a reduced enantioselectivity (Table 1, entry 15).Optimization of the reaction conditionsa
EntryCat.Bromine sourceSolventYieldb (%)e.r.c
1 4a 2a Toluene6556 : 44
2 4b 2a Toluene4968 : 32
3 4c 2a Toluene6195 : 5
4 4d 2a Toluene4175 : 25
5 4e 2a Toluene5393 : 6
6 4f 2a Toluene3961 : 39
7 4c 2a DCM4789 : 11
8 4c 2a Et2O3967 : 33
9d 4c 2a Toluene6994 : 6
10e 4c 2a Toluene6193 : 7
11 4c 2b Toluene6394 : 6
12 4c 2c Toluene6587 : 13
13f 4c 2a Toluene7595 : 5
14g 4c 2a Toluene8096.5 : 3.5
15h 4c 2a Toluene7995 : 5
Open in a separate windowaReaction conditions: a mixture of 1a (0.05 mmol), 2a (0.05 mmol) and cat. 4 (10 mol%) in the solvent (0.5 mL) was stirred at −78 °C for 12 h.bIsolated yield.cDetermined by HPLC analysis.d3 Å MS (10.0 mg) was used as the additive.e4 Å MS (10.0 mg) was used as the additive.f 1a : 2a = 1.2 : 1.g 1a : 2a = 1.5 : 1.h 1a : 2a = 2.0 : 1.Under the optimized reaction conditions, the scope of the desymmetrizing asymmetric ortho-selective mono-bromination of phosphine oxides was examined. Firstly, the variation of the P-center substituted group was investigated. As shown in Table 2, a variety of P-aryl, P-alkyl substituted phosphine oxides and phosphinates (3a–3f) were well amenable to this reaction and the corresponding ortho-brominated products were obtained in good yield (up to 87%) with high enantiomeric ratios (up to 98.5 : 1.5 e.r.). Moreover, regardless of whether the R was a bulky group or a smaller one, the enantiomeric ratios of the products were maintained at excellent levels. Especially, when the P-center substituted group was ethoxyl (1e), the corresponding bromination product 3e was obtained in 80% yield with 98.5 : 1.5 e.r. When a P-methyl substituted phosphine oxide was used as the substrate, a moderate yield and enantiomeric ratio were obtained for 3g.The scope of bisphenol phosphine oxides with different substituents on the P-atoma,b,c
Open in a separate windowaReaction conditions: a mixture of 1a (0.15 mmol), 2a (0.1 mmol) and 4c (10 mol%) in toluene (1.0 mL) was stirred at −78 °C for 12 h.bIsolated yield.cDetermined by HPLC analysis.Next, using the ethoxyl substituted phosphinate as the template, a diversity of phosphinates with a 5-position substituent on the phenyl ring were examined (Table 3). To our delight, a range of phosphinates with different alkyl substituent on the phenyl ring was suitable for the currently studied reaction and the desired products 3h–3l were obtained with very good enantioselectivities (90.5 : 9.5–97.5 : 2.5 e.r.). Furthermore, substrates with aryl and alkoxy groups at the 5-position of the phenol moiety were also tolerated well under the reaction conditions, and gave the products 3m–3q with good to excellent yields (81–92%) and enantioselectivities (95 : 5–98.5 : 1.5 e.r.). Moreover, when a disubstituted phenol phosphinate substrate was used, the desired bromination product 3r was also delivered with a good yield and e.r. value.The scope of bisphenol phosphinatesa,b,c
Open in a separate windowaReaction conditions: a mixture of 1a (0.15 mmol), 2a (0.1 mmol) and 4c (10 mol%) in toluene (1.0 mL) was stirred at −78 °C for 12 h.bIsolated yield.cDetermined by HPLC analysis.Then, we turned our attention to inspect the scope of ortho-bromination of P-adamantyl substituted phosphine oxides. As exhibited in Table 4, 5-methyl, 5-ethyl and 4,5-dimethyl aryl substituted phosphine oxides could be transformed into the corresponding products (3s, 3t and 3u) with excellent yields (81–89%) and enantioselectivities (95 : 5–96 : 4 e.r.). Upon increasing the size of the 5-position substituent on the phenyl ring of phosphine oxides, the enantioselectivities of the products 3v–3y had a little decreasing tendency (81 : 19–93 : 7 e.r.). The absolute configuration of 3v was determined by X-ray diffraction analysis and those of other products were assigned by analogy.25The scope of adamantyl substituted bisphenol phosphine oxidesa,b,c
Open in a separate windowaReaction conditions: a mixture of 1a (0.15 mmol), 2a (0.1 mmol) and 4c (10 mol%) in toluene (1.0 mL) was stirred at −78 °C for 12 h.bIsolated yield.cDetermined by HPLC analysis.24d 1a : 2a = 1.2 : 1.To demonstrate the utility of this desymmetrizing asymmetric ortho-selective mono-bromination, the reaction was scaled up to 1.0 mmol, and the corresponding product 3a was obtained in 80% yield with 96.5 : 3.5 e.r. (98.5 : 1.5 e.r. after single recrystallization) (Scheme 2a). The encouraging results implied that this strategy had the potential for large-scale production. Additionally, the transformations of products 3a and 3e were also investigated (Scheme 2b). In the presence of Pd(OAc)2 and bulky electron-rich ligand S-Phos, 3a could react with phenylboronic acid effectively, in which the desired cross-coupling product 5 was generated in high yield with maintained enantioselectivity. In the presence of Lawesson''s reagent, 3a could be transformed into thiophosphine oxide 6 with a high yield and e.r. value. Furthermore, 3e could react with methyl lithium to afford the DiPAMP analogue 3g in 85% yield with 98.5 : 1.5 e.r. And 3e could also be converted to chiral bidentate Lewis base 7 by a straightforward alkylation reaction. It was encouraging to find that 7 could be used as a catalyst for the asymmetric reaction between trans-chalcone and furfural, in which the desired product 8 was furnished with moderate stereoselectivity (Scheme 2c).26Open in a separate windowScheme 2(a) Large-scale reaction. (b) Synthetic transformations. (c) Application of the transformed product.Since the mono-bromination product 3a could undergo further bromination to form the dibromo adduct, we wondered whether this second bromination is a kinetic resolution process. As shown in Scheme 3a, a racemic sample of 3a was subjected to the catalytic conditions ((±)-3a and 2a in a 2 : 1 molar ratio). Upon complete consumption of 2a (with the formation of a dibromo product in 49% yield), the mono-bromination product 3a was recovered in 51% yield with 99 : 1 e.r. This result indicated that the second bromination was indeed a kinetic resolution process and had a positive contribution to the enantioselectivity. Considering the excellent enantiomeric ratio of recovered 3a, we further investigated the reaction of rac-9 with 2a under kinetic resolution conditions (Scheme 3b). To our delight, the unreacted raw material 9 can be obtained in 51% yield with 99.5 : 0.5 e.r., and chiral dihalogenated product 10 can also be generated in 49% yield with 90 : 10 e.r.Open in a separate windowScheme 3Kinetic resolution process.To investigate the mechanism, we performed some control experiments. First, a mono-methyl protected phosphine oxide substrate was prepared and subjected to ortho-bromination under the optimal conditions. As shown in Scheme 4a, the corresponding product 11 was obtained with 72.5 : 27.5 e.r. When the same reaction conditions were applied to the dimethyl protected phosphine oxide substrate, no reaction occurred (Scheme 4b). These results indicated that the phenol moieties of the substrate were essential for the bromination reaction. In fact, hydrogen bonds formed between the two phenolic hydroxyl groups and P Created by potrace 1.16, written by Peter Selinger 2001-2019 O could be observed in the single crystal structure of the product 3w.25 Furthermore, when thiophosphine oxide, which had a weak hydrogen bond acceptor P Created by potrace 1.16, written by Peter Selinger 2001-2019 S group, was prepared and tested in the reaction, the corresponding product 6 was obtained with a lower yield and enantioselectivity than that of 3a (Scheme 4c). This result suggested that the intramolecular hydrogen bonds of the substrate might be beneficial for both the reactivity and the enantioselectivity.27 In light of the control experiments and previous studies,24 two possible mechanisms were proposed (see the ESI).Open in a separate windowScheme 4Control experiments: (a) mono-methyl protected phosphine oxide substrate was evaluated; (b) dimethyl protected phosphine oxide substrate was examined; (c) thiophosphine oxide substrate was investigated.In summary, a novel and efficient desymmetrizing asymmetric ortho-selective mono-bromination of bisphenol phosphine oxides under chiral squaramide catalysis was reported. Using this asymmetric ortho-bromination strategy, a wide range of chiral bisphenol phosphine oxides and bisphenol phosphinates were obtained with good to excellent yields and enantioselectivities. The reaction could be scaled up, and the synthetic utility of the desired P-stereogenic compounds was proved by transformations and application in an asymmetric reaction. Ongoing studies focus on the further mechanistic investigations and the potential applications of these chiral P-stereogenic compounds in other asymmetric transformations.  相似文献   

7.
8.
A transition metal free process for conjunctive functionalization of alkenylboron ate-complexes with electrophilic fluoroalkylthiolating reagents is described, affording β-trifluoroalkylthiolated and difluoroalkylthiolated boronic esters in good yield and excellent diastereoselectivity. The potential applicability of the method was demonstrated by the preparation of a difluoromethylthiolated mimic 12 of a potential drug molecule PF-4191834 for the treatment of asthma.

A transition metal free process for conjunctive functionalization of alkenylboron ate-complexes with electrophilic fluoroalkylthiolating reagents affords β-tri- and difluoroalkylthiolated boronic esters in good yield and diastereoselectivity.

An electrophile-induced 1,2-metalate migration of an alkenylboron “ate” complex and subsequent base-promoted β-elimination to form a functionalized cis-alkene, now the so-called Zweifel reaction, was first reported by Zweifel and co-workers in 1967 (Fig. 1A).1–3 The reaction was proposed to proceed via an initial attack of the π electron of the alkene moiety to iodine to generate a zwitterionic iodonium ion, which then undergoes a stereospecific 1,2-metalate to afford a β-iodoboronic ester, followed by anti-elimination upon treatment with a base to afford a cis-olefin. Thus, if the iodine is replaced by an alternative electrophilic reagent and the use of a base is omitted, an interrupted-Zweifel reaction for the preparation of a stereospecific β-functionalized boronic ester could be realized. Toward this end, Aggarwal reported the first example of such a reaction by employing PhSeCl as the electrophilic reagent.4 It was proposed that PhSeCl first reacts with an alkenylboronate complex to form a zwitterionic seleniranium ion. Subsequent diastereospecific 1,2-metalate migration affords the stereospecific β-seleno-alkylboronate (Fig. 1B). Likewise, shortly after, Denmark and co-workers reported an analogous Lewis-base catalysed enantioselective and diastereoselective carbosulfenylation of an alkenylboronate complex using N-arylthiosaccharin as the electrophile (Fig. 1C).5Open in a separate windowFig. 1The interrupted Zweifel reaction.In light of these discoveries and our recent success in the development of a toolbox of electrophilic fluoroalkylthiolating reagents including three trifluoromethylthiolating reagents α-cumyltrifluoromethane sulfenate,6N-trifluoromethylthio-saccharin7 and N-trifluoromethylthiodibenzenesulfonimide,8 and two difluoromethylthiolating reagents N-difluoromethylthiophthalimide9 and S-(difluoromethyl)benzenesulfonothioate,10 we wondered whether these electrophilic fluoroalkylthiolating reagents could also trigger the proposed stereospecific 1,2-metalatation of the alkenylboronate complex to afford β-fluoroalkylthiolated borane derivatives (Fig. 1D). The trifluoromethylthio (–SCF3) and the difluoromethylthio (–SCF2H) groups have gained great attention recently, partially because of their high and tuneable lipophilicity11 that might improve the drug candidate''s cell membrane permeability and consequently, its overall pharmacokinetics.12 Thus, the development of new efficient reactions for the incorporation of the trifluoromethylthio13 or difluoromethylthio groups14 would be of vital importance in facilitating medicinal chemists'' endeavours in new drug discovery. Herein, we report that by employing electrophilic difluoromethylthiolating reagent PhSO2SCF2H 2a as the electrophile, the proposed difluoromethylthiolating induced stereospecific 1,2-metalate migration of alkenyl boronate complexes occurred smoothly to afford β-difluoromethylthiolated boronic esters in good yields and excellent diastereoselectivity. Likewise, when electrophilic trifluoromethylthiolating reagent N-trifluoromethylthiosaccharin 7 was used, an analogous reaction for the diastereoselective formation of β-trifluoromethylthiolated boronic esters was successfully achieved.We began our study by examining the reaction of the electrophilic difluoromethylthiolating reagent 2a with the alkenylboronate complex which was generated in situ by mixing 1a and PhLi in diethyl ether. It was found that the reaction in CH3CN occurred in full conversion after 12 hours at room temperature, affording the corresponding product 3a in 53% yield (Table 1, entry 1). When the amount of PhLi was increased to 1.3 equivalents, the yield was increased to 76%, while the yield decreased to 66% when 2.0 equivalents of PhLi were used, likely due to the decomposition of the product under strong basic conditions (Table 1, entries 1–5). We then further investigated the effect of the reaction temperature and the solvent. It was found that the temperature did not affect the reaction significantly since the yields of the desired products were decreased slightly to 72% and 70%, respectively, when the reactions were conducted at 0 °C or −15 °C (Table 1, entries 6 and 7). Likewise, the reaction was not sensitive to the polarity of the solvent since reactions conducted in less polar solvents such as THF or CH2Cl2 or nonpolar solvents like toluene occurred in slightly lower 60–73% yields (Table 1, entries 9–11). We also found that reaction using N-difluoromethylthiophthalimide as the electrophilic difluoromethylthiolating reagent gave the same product in a slightly lower yield (Table 1, entry 8).Optimization of conditions for the reaction of the alkenyl boronate complex with PhSO2SCF2Ha
EntryEquiv. of PhLiSolventTemp (°C)Yielda (%)
11.0CH3CNrt53
21.1CH3CNrt60
31.2CH3CNrt72
41.3CH3CNrt76(72)b
52.0CH3CNrt66
61.3CH3CN072
71.3CH3CN−1570
81.3CH3CNrt56c
91.3THFrt73
101.3CH2Cl2rt64
111.3Toluenert60
Open in a separate windowaReaction conditions: vinyl boronate 1a (0.10 mmol) and reagent 2a (0.15 mmol), in CH3CN (1.0 mL) at room temperature for 12 h; Yields were determined by 19F NMR spectroscopy using PhCF3 as an internal standard.bIsolated yield.c N-Difluoromethylthiophthalimide was used.With optimum reaction conditions established, a range of different alkenylboronate complexes were tested under standard conditions (Scheme 1). Alkenylboronate complexes obtained by treating 3,6-dihydro-2H-pyran-4-boronic acid pinacol ester with diverse aryl lithiums reacted efficiently with reagent 2a to give the corresponding β-difluoroalkylthionated boronic esters 3b–e and 3g–m in good yield and excellent diastereoselectivity. A range of aryllithiums with both the electron-donating methoxy group (3c) and electron-withdrawing groups such as a fluoride (3d) or a trifluoromethyl group (3g) or a bulky tert-butyl group at meta-position (3i) worked well. The reaction can also proceed smoothly for naphthyllithium (3h) and n-butyllithium (3j). Moreover, organolithiums generated from heteroaromatics, such as indole (3k), benzothiophene (3l), benzofuran (3m), could also be used. Notably, it is well-known that bromine is not compatible with butyl lithium. Yet, 3f with a para-bromophenyl moiety was obtained from the reaction of the alkenylboronate complex in situ generated by treating (3,6-dihydro-2H-pyran-4-yl)lithium with 4-bromophenylboronic acid pinacol ester. However, the alkenylboronate complex generated by treating (E)-4,4,5,5-tetramethyl-2-(5-phenylpent-1-en-1-yl)-1,3,2-dioxaborolane with tert-butyllithium, failed to react with reagent 2a to give the corresponding β-difluoroalkylthionated boronic esters (3r). Next, the scope with respect to the alkenyl boronic ester component was explored. 3,6-Dihydro-2H-thiopyran-4-ylboronic acid pinacol ester (3n), or N-Ts-3,6-dihydro-2H-pyran-4-boronic acid pinacol ester (3o) and 1-phenylvinylboronic acid pinacol ester (3q) could react well to afford the corresponding products. To demonstrate the scalability of the reaction, 3p was prepared on a gram scale in 75% yield. Furthermore, bridged cyclic boronate 3s could also be obtained in moderate yield, and the anti diastereoselectivity of the reaction was confirmed by X-ray diffraction of its single crystals.Open in a separate windowScheme 1Scope of 1,2-metalate migration of alkenyl boronates with reagent 2a.a a Reaction conditions: alkenyl or aryl boronic ester (0.30 mmol, 1.0 equiv.), R3Li (0.33 mmol, 1.1 equiv.) in Et2O (1.5 mL) at −78 °C to room temperature for 30 min; then the solvent was swapped with CH3CN (3.0 mL); 2a (0.45 mmol, 1.5 equiv.) was added. Isolated yield. b R3Li (0.39 mmol) in Et2O (1.5 mL) at 0 °C to room temperature for 30 min. c The mixture was treated with NaBO3 (0.9 mmol, 3.0 equiv.) in THF/H2O (v/v = 1 : 1, 6 mL) at room temperature for 6 h.Furthermore, it was found that the resultant boronic esters could be easily oxidized to alcohols, with the difluoromethylthio group remaining intact, by treatment with 3.0 equivalents of NaBO3 at room temperature for 6 h. For example, difluoromethylthiolated β-alcohols 4a–4d were obtained in moderate to good yields under these conditions (Scheme 1).In general, it is a common practice to use E or Z-alkenes in the reaction to probe whether the reaction is stereo-specific. Thus, we examined the reaction of E-(3′-phenylpropyl)vinyl boronic acid pinacol ester and Z-(3′-phenylpropyl)vinyl boronic acid pinacol ester under standard conditions. It was found that the reaction is stereospecific since the reactions of E- and Z-alkenyl boronic esters specifically produced corresponding anti- and cis-difluoromethylthiolated alcohols (4e and 4f) with excellent diasteroselectivity (>20 : 1), respectively (Scheme 2).Open in a separate windowScheme 2Reactions of E- and Z-alkenyl boronate complexes with reagent 2a.To further expand the scope of the reaction, we studied the difluoromethylthiolative triggered stereospecific 1,2-metalate migration of in situ generated vinyl boronate complexes from enantio-enriched secondary alkyl boronic esters with vinyl lithium. The resulting crude alkyl boronic esters were then sequentially oxidized by NaBO3 and Jone''s oxidation to give α-chiral ketone derivatives. It was found that chirality of the secondary alkyl boronic esters was stereospecifically transferred to the final products 6a–c with 100% es (Scheme 3).Open in a separate windowScheme 3Synthesis of α-chiral ketones by stereospecific 1,2-migration.a a Reaction conditions: alkyl boronic ester (0.30 mmol, 1.0 equiv.), R3Li (0.36 mmol, 1.2 equiv.) in Et2O (1.5 mL) at −78 °C to room temperature for 30 min; then the solvent was swapped with CH3CN (3.0 mL); 2a (0.45 mmol, 1.5 equiv.) was added; and then NaBO3 (0.9 mmol, 3.0 equiv.) in THF/H2O (v/v = 1 : 1, 6 mL) was used; and then Jone''s reagent (0.45 mmol, 1.5 equiv.) was used. Isolated yield.Encouraged by the excellent diastereoselective difluoromethylthiolation of alkenyl boronic acid pinacol esters, we then extended this highly selective reaction to analogous trifluoromethylthiolation triggered 1,2-metalate migration of alkenylboronate (Scheme 4). It was found that when N-trifluoromethylthiosaccharin 7 was used as the electrophilic trifluoromethylthiolating reagent, the reaction of alkenylboronate derived from PhLi occurred smoothly in CH3CN after 12 h at 0 °C to give β-trifluoroalkylthionated boronic ester 8a in 76% yield (8a). Likewise, a variety of other aryllithiums could be successfully employed in this reaction to afford the corresponding β-trifluoroalkylthionated boronic esters (8b–h) in high yields. This reaction appears to be compatible with labile functional groups such as chlorine (8b), trifluoromethyl (8c), ketal (8d), and acetal (8e). In addition, organolithiums generated from heteroaromatics, such as benzofuran (8g) and benzothiophene (8h) could also be employed. Lastly, it was found that a single diastereoisomer with an anti configuration (8i) was isolated in 75% yield when the corresponding E-alkenyl boronic ester was used. Yet, the scope of alkenyoboronate complexes for the reaction with N-trifluoromethylthiosaccharin 7 is not as broad as that with PhSO2SCF2H since alkenylboronate complexes generated by treating 3,6-dihydro-2H-pyran-4-boronic acid pinacol ester with n-butyllithium or by treating 2,2-dimethylethenylboronic acid pinacol ester with lithium benzothiophene failed to produce the desired β-trifluoroalkylthionated boronic esters 8j and 8k under the standard conditions.Open in a separate windowScheme 4Scope of 1,2-metalate migration of alkenyl boronates with electrophilic trifluoromethylthiolating reagent 7.a a Reaction conditions: alkenyl boronic ester (0.30 mmol, 1.0 equiv.), R3Li (0.33 mmol, 1.1 equiv.) in Et2O (1.5 mL) at −78 °C to room temperature for 30 min; then the solvent was swapped with CH3CN (3.0 mL); reagent 5 (0.45 mmol) was added. b R3Li (0.39 mmol, 1.3 equiv.) in Et2O (1.5 mL) at 0 °C to room temperature for 30 min. Isolated yield.To further demonstrate the great potential of this reaction, we applied this protocol as a key step in the synthesis of a difluoromethylthiolated mimic of PF-4191834, which is a potent competitive inhibitor of the 5-lipoxygenase (5 LOX) enzyme for the treatment of mild to moderate asthma15 (Fig. 2). Firstly, arylsulfide 11 was synthesized efficiently by deborylthiolation of organoboron 9 with thiosulfonate 10 in the presence of 5 mol% CuSO4 as the catalyst. Lithium halide exchange of compound 11 with t-butyllithium at −78 °C for 30 min generated the corresponding aryl lithium species in situ, which was treated with 3,6-dihydro-2H-pyran-4-boronic acid pinacol ester to afford the alkenyl boronate complex. Switching the solvent from diether ether to CH3CN, followed by the addition of 1.5 equivalents of PhSO2SCF2H 2a, and further reaction at room temperature for 12 h produced the difluoromethylthiolated mimic of PF-4191834 12 in 70% yield. This example showed the potential of the current protocol in the preparation of biological active compounds.Open in a separate windowFig. 2Construction of PF-4191834 mimic by conjunctive cross-coupling.In summary, a method of conjunctive three-component coupling between alkenyl boronic esters, organolithiums and electrophilic fluoroalkylthiolating reagents was successfully developed, affording β-trifluoroalkylthionated and difluoroalkylthionated boronic esters in good yield and excellent diastereoselectivity. The reaction is stereospecific since the reaction of the E-alkenyl boronic ester specifically gave an anti-difluoromethylthiolated β-alcohol and the reaction of the Z-alkenyl boronic ester specifically gave cis-difluoromethylthiolated β-alcohol 4f with excellent diasteroselectivity (>20 : 1). The potential applicability of the method was demonstrated by the preparation of a difluoromethylthiolated derivative of a potential drug molecule for the treatment of asthma PF-4191834 12. The reactions of the alkenyl boronate complexes with other electrophilic fluoroalkylating reagents are currently actively underway in our laboratory.  相似文献   

9.
Brønsted acid catalyzed formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloadditions of donor–acceptor cyclobutenes, cyclopropenes, and siloxyalkynes with benzopyrylium ions are reported. [4 + 2]-cyclization/deMayo-type ring-extension cascade processes produce highly functionalized benzocyclooctatrienes, benzocycloheptatrienes, and 2-naphthols in good to excellent yields and selectivities. Moreover, the optical purity of reactant donor–acceptor cyclobutenes is fully retained during the cascade. The 1,3-dicarbonyl product framework of the reaction products provides opportunities for salen-type ligand syntheses and the construction of fused pyrazoles and isoxazoles that reveal a novel rotamer-diastereoisomerism.

Brønsted acid catalysis realizes formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloadditions of donor–acceptor cyclobutenes, cyclopropanes, and siloxyalkynes with benzopyrylium ions.

Medium-sized rings are the core skeletons of many natural products and bioactive molecules,1 and a growing number of strategies have been developed for their synthesis.2 Because of their enthalpic and entropic advantages, ring expansion is a highly efficient methodology for these constructions.3–6 For example, Sun and co-workers have developed acid promoted ring extensions of oxetenium and azetidinium species formed from siloxyalkynes with cyclic acetals and hemiaminals (Scheme 1a).4 Takasu and co-workers have reported an elegant ring expansion with a palladium(ii) catalyzed 4π-electrocyclic ring-opening/Heck arylation cascade with fused cyclobutenes (Scheme 1b).5 Each transformation is initiated by the formation of fused bicyclic units followed by ring expansion or rearrangement to give medium-sized rings.Open in a separate windowScheme 1Cycloaddition/ring expansion background and this work.Strategies for the formation of fused bicyclic compounds rely on cycloaddition of dienes or dipoles with unsaturated cyclic compounds7 and, if the reactant cyclic compound is strained and chiral, the resulting bicyclic compound is activated toward ring opening that results in retention of chirality. We have recently reported access to donor–acceptor cycloalkenes by [3 + n] cycloaddition that have the prerequisites of unsaturated cyclic compounds suitable for cycloaddition.8 Donor–acceptor (D–A) cyclopropenes9 and cyclobutenes10 have sufficient strain in the resulting bicyclic compounds to undergo ring opening. We envision that the selection of a diene or dipolar reactant and suitable reaction conditions could realize cycloaddition and subsequent ring expansion. Benzopyrylium species,11,12 which are generated by metal or acid catalysis, have attracted our attention. We anticipated that their high reactivity would overcome the conventional unfavorable kinetic and/or thermodynamic factors that typically impede medium-sized ring formation. From a mechanistic perspective, benzopyrylium species are often formed by transition metal catalyzed reactions with 2-alkynylbenzaldehydes.11 Recently, the reaction between 1H-isochromene acetal and Brønsted acid catalyst forms the 2-benzopyrylium salts that could react with functional alkenes to give same cycloaddition products.12 Consequently, we believed that the [4 + 2]-cyclization between benzopyrylium species and donor–acceptor cyclobutenes, cyclopropenes, or siloxyalkynes would give bridged oxetenium intermediates, which contain high strain energy that should provide the driving force for ring expansion. The “push and pull” electronic effect of donor–acceptor functional groups facilitates deMayo-type ring-opening of the cyclobutane or cyclopropane skeletons (Scheme 1c).13Here we report bis(trifluoromethanesulfonyl)imide (HNTf2) catalyzed formal [4 + 4]-, [4 + 3]- and [4 + 2]-cycloaddition reactions of D–A cyclobutenes, cyclopropenes, and siloxyalkynes with benzopyrylium salts. Polysubstituted benzo-cyclooctatrienes, benzocycloheptatrienes and 2-naphthols, are produced in good to excellent yields and selectivities. Complete retention of configuration occurs using chiral cyclobutenes, and opportunities for further functionalization are built into these constructions.Initially, we conducted transition metal catalyzed reactions with 2-alkynylbenzaldehyde intending to produce the corresponding benzopyrylium ion and explore the possibility of cycloaddition/ring opening with donor–acceptor cyclobutene 2a. Use of Ph3PAuCl/AgSbF6, Pd(OAc)2 and Cu(OTf)2, which were efficient catalysts in previous transformations,11 gave only a trace amount of cycloaddition product (Fig. 1). Spectral analysis showed that mostly starting material remained (Fig. 1-i and ii). Increasing the reaction temperature led only to decomposition of 2-alkynylbenzaldehyde (Fig. 1-iv) or donor–acceptor cyclobutene 2a (Fig. 1-iv).Open in a separate windowFig. 1Reaction of 2-alkynylbenzaldehyde with donor–acceptor cyclobutene 2a.From these disappointments we turned our attention to 1H-isochromene acetal 1a as the benzopyrylium ion precursor. Various Lewis acid catalysts were employed with limited success, but we were pleased to observe the formation of the desired formal [4 + 4]-cycloaddition benzocyclooctatriene product 3aa, albeit in low yields (Table 1, entries 1–6). Selection of the Brønsted super acid HNTf2 (ref. 14) proved to be the most promising, producing 3aa in 40% yield (Table 1, entry 7). Increasing the amount of D–A cyclobutene 2a by 30% gave a much higher yield of 3aa (Table 1, entry 8 vs. 7). Optimization of the stoichiometric reaction between 1a and 2a by increasing the reaction temperature from rt to 35 °C led to the formation of the desired product in 76% isolated yield (Table 1, entry 9 vs. 8). However, a further increase in the reaction temperature (Table 1, entry 10) or reducing the catalyst loading to 5 mol% did not improve the yield of 3aa (Table 1, entry 11).Optimization of reaction conditionsa
EntryCat (10 mol%)Temp. (°C)Yieldb
1Sc(OTf)3rt23
2Yb(OTf)3rtTrace
3In(OTf)3rt16
4TiCl4rtTrace
5BF3·OEt2rt20
6TMSOTfrt17
7HNTf2rt40
8cHNTf2rt67
9cHNTf23578(76)d
10c,eHNTf26062
11c,fHNTf23550
Open in a separate windowaReactions were performed by adding the catalyst (10 mol%) to 1a (0.1 mmol) and 2a (0.1 mmol) in CH2Cl2 (2 mL) at the corresponding temperature for 24 h.bYields were determined by 1H NMR spectroscopic analysis with CH2Br2 as the internal standard.c1.3 equiv. 2a was used.dIsolated yield.eReaction performed at 60 °C for 12 h.f5 mol% catalyst loading.With optimized conditions using 1a in hand, we examined the scope of the formal [4 + 4]-cycloaddition reactions of D–A cyclobutenes 2 with a diverse set of acetal compounds 1. As shown in Scheme 2, a wide range of acetal substrates (1a–1i) with different substituents at different positions all reacted smoothly with D–A cyclobutene 2a to form the corresponding benzocyclooctatriene products 3 in good to excellent yields. Structural variations in the acetals produced only modest changes in product yields which ranged from 55 to 87%. Similarly, both electron-withdrawing and electron-donating substituents at the 4-position of the cyclobutene phenyl ring produced the corresponding products (3db–3dd) in good yields, and 2-naphthyl (2e) and 2-thienyl (2f) substituted cyclobutenes were suitable substrates (85% and 51% product yields, respectively). trans-1,2,3,4-Tetrasubstituted (R2 = CH3) 2-sil-oxycyclobutenecarboxylate 2g also underwent [4 + 4]-cycloaddition with 1d in good yield and fully retained its diastereoselectivity. The structure of 3fa was confirmed by X-ray diffraction (Scheme 2).15Open in a separate windowScheme 2Scope of the [4 + 4]-cycloaddition reaction of D–A cyclobutenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10 mol%) to 1 (0.1 mmol) and 2 (0.13 mmol) in CH2Cl2 (2 mL) at 35 °C for 24 h. Isolated yields are reported.Chiral donor–acceptor cyclobutenes with high enantiomeric excess and diastereoselectivity are conveniently obtained by catalytic [3 + 1]-cycloaddition of enoldiazoacetates with acyl ylides of sulfur.10 To determine if optical purity is retained, chiral D–A cyclobutene 2a (80% ee) was reacted with 1d under the optimized conditions, and the corresponding benzocyclo-octatriene product 3da was obtained in good yield with complete retention of configuration (Scheme 3, eqn (1)). With trans-disubstituted 2g and 2h that have higher optical purity, however, 3dg and 3dh were obtained in moderate yields with full retention of diastereo- and enantioselectivities, but addition products 8dg and 8dh were formed competitively (Scheme 3, eqn (2)). These compounds resulted from initial addition then desilylation, indicating that the [4 + 2]-cyclization is a stepwise reaction. Attempts to suppress the competing pathway by changing solvents or using the isopropyl acetal substrate to form a bulky 2-propanol nucleophile (1j) failed (for details, see ESI). However, p-methoxy (R1) substituted 1g that would stabilize the incipient benzopyrylium ion gave higher selectivity (4 : 1 vs. 2 : 1), and the desired [4 + 4] cycloaddition product 3gg was isolated in 50% yield with 97% ee and >19 : 1 dr (Scheme 3, eqn (3)).Open in a separate windowScheme 3Stereochemical features of the [4 + 4]-cycloaddition of D–A cyclobutenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10 mol%) to 1 (0.1 mmol) and 2 (0.13 mmol) in CH2Cl2 (2 mL) at 35 °C for 24 h. Isolated yields are reported.To further expand the generality of this strategy, we investigated its use with donor–acceptor cyclopropenes 4. The desired formal [4 + 3]-cycloaddition products 5 were obtained in good to excellent yields (Scheme 4). Optimized conditions used 20 mol% HNTf2 catalyst with 4 Å molecular sieves at room temperature for 2 h. The scope of this [4 + 3]-cycloaddition reaction with cyclopropenes 4 showed that acetals bearing both electron-donating and electron-withdrawing substituents on the aromatic ring were tolerated. However, as with [4 + 4]-cycloaddition reactions, the [4 + 3] reactions of 1 with R1 = alkyl or H did not produce any of the desired products. Furthermore, 3-substituted cyclopropenes 4b–4d participated in this reaction, and their products (5ab–5ad) were obtained in 63%–81% yields.Open in a separate windowScheme 4Scope of the [4 + 3]-cycloaddition reaction of D–A cyclopropenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (20 mol%) to 1 (0.2 mmol), 4 (0.24 mmol) and 4 Å (50 mg) in CH2Cl2 (2 mL) at rt for 2 h. Isolated yields are reported.Siloxyalkynes 6, as electron-rich alkynes, have been widely used in diverse cyclization reaction.4,16 We expected that 6 could also participate in [4 + 2]-cyclization/ring-expansion cascade processes, giving substituted 2-naphthol products. Interestingly, substituent controlled diverse products were obtained in good to excellent reactivity and selectivity (Scheme 5). Aryl (R1) substituted acetals (1a–1d, 1f, 1h, 1i) reacted with siloxyalkynes 6a–6d, giving 2-naphthols (7aa–7ad and 7ba–7ia) in 35%–96% yields. With electron-donating group (EDG) substituents (7ba and 7ca) on the aromatic ring, higher reactivity was observed relative to those with electron-withdrawing groups (7da–7ia). In addition, the acetal with R1 = H (1m) reacted with siloxyalkynes 6 to form the 2-naphthol-1-carboxaldehyde derivative in good yield, and the structure of 7ma was confirmed by X-ray diffraction (Scheme 5b).15 Intriguingly, the n-butyl (R1) substituted acetal 1n reacted with siloxyalkynes via a [4 + 2]-cyclization with loss of methyl pentanoate (BuCO2CH3), affording siloxy naphthalenes 7na–7nd that are important precursors to the widely used axially chiral 2,2′-binols.17 The substrate scope of siloxyalkynes 6 for their formal [4 + 2]-cycloaddition reaction with n-butyl substituted acetal 1n was also explored (Scheme 5c). In all cases methyl pentanoate was eliminated to form 2,3-disubstituted naphthalene products (7na–7nd). Alkyl substituted siloxyalkynes (6a–6c) showed higher reactivity compared with phenyl substituted siloxyalkynes 6d. It should be mentioned that, recently, a similar transformation using BF3·OEt2 as catalyst or in excess (2 equiv.) with 2,4,6-collidine (1 equiv.) was reported,18 and HNTf2 was stated to be much less effective. To clarify this discrepancy, we carefully repeated these transformations (7ma and 7na) and found that all starting materials are completely consumed in less than 10 min to deliver [4 + 2]-cycloaddition products in good yields. Prolonging the reaction time to 12 hours, which was the reaction time used by the authors, results in their decomposition to a complex mixture of materials.Open in a separate windowScheme 5Scope of the [4 + 2]-cycloaddition reaction of siloxyalkynes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10–30 mol%) to 1 (0.2 mmol) and 6 (0.3–0.4 mmol) in CH2Cl2 (2 mL) at rt for 10 min. Isolated yields are reported. b40 mol% HNTf2 catalyst was used.To illustrate the utility of this process, a large scale catalytic [4 + 4] cycloaddition was performed, and adduct 3da was obtained in 87% yield. Further transformations were conducted for the synthesis of pyrazole and isoxazole structures based on its 1,3-dicarbonyl skeleton (Fig. 2a). Compound 3da reacted with hydrazine and hydroxylamine in refluxed ethanol, affording pyrazole 14da and isoxazole 15da in 89% yield or 66% yield, respectively. The structure of 15da was confirmed by X-ray diffraction.15 Interestingly, two NMR distinguishable interconvertible diastereoisomers were detected for each of these eight-membered cyclic products (2.5 : 1 and 5 : 1 dr for 14da and 15da, respectively, in CDCl3). These diastereoisomers are rotamers (for details, see ESI) that exist at equilibrium with each other in solution but form one crystalline product (X-ray structure of 15da). In addition, the cycloaddition product 7ma reacted with chiral 1,2-cyclohexanediamine and 1,2-diphenylethylenediamine to give salen-type ligands L1 and L2 in 53% yield and 74% yield, respectively, which provides new opportunities for ligand screening (Fig. 2b).19Open in a separate windowFig. 2Large scale reaction, further transformations, and ligands synthesis.In the mechanistic possibility considered for these HNTf2 catalyzed cycloaddition reactions (Fig. 3), protonation of acetal 1 with HNTf2 gives the corresponding highly reactive benzopyrylium intermediate int-I, which reacts with donor–acceptor cyclobutenes 2, cyclopropenes 4, or siloxyalkynes 6 affording addition intermediates int-II that undergo ring closure to int-III. Ring expansion then occurs to deliver 3, 5 and 7 in good to excellent yields with fully retained stereoselectivities. Furthermore, the formed TIPSNTf2 ([4 + 4]- and [4 + 3]-cycloaddition) or HNTf2 {[4 + 2] cycloaddition} are active acid catalysts for the conversion of 1 to benzopyrylium intermediate int-I that continues the catalytic cycle. With sterically larger D–A cyclobutenes or when a less ring-strained benzocyclopentene 12 is employed (for details, see ESI), the competing direct desilylation of int-II occurs, delivering addition byproducts 8 or 13. Compounds 7na–7nd arise from the analog to int-III from which ketene formation or methanol displacement effects 1,4-elimination.Open in a separate windowFig. 3Proposed mechanism.  相似文献   

10.
Aurone-derived azadienes are well-known four-atom synthons for direct [4 + n] cycloadditions owing to their s-cis conformation as well as the thermodynamically favored aromatization nature of these processes. However, distinct from this common reactivity, herein we report an unusual formal migrative annulation with siloxy alkynes initiated by [2 + 2] cycloaddition. Unexpectedly, this process generates benzofuran-fused nitrogen heterocyclic products with formal substituent migration. This observation is rationalized by less common [2 + 2] cycloaddition followed by 4π and 6π electrocyclic events. DFT calculations provided support to the proposed mechanism.

A HNTf2-catalyzed formal migrative cycloaddition of aurone-derived azadienes with siloxy alkynes has been developed to provide access to benzofuran-fused dihydropyridines.

Benzofuran is an important scaffold in biologically important natural molecules and therapeutic agents.1 Among them, benzofuran-fused nitrogen heterocycles are particularly noteworthy owing to their broad spectrum of bioactivities for the treatment of various diseases (Fig. 1).2 Consequently, the development of efficient methods for their assembly has been a topic receiving enthusiastic attention from synthetic chemists.3 Notably, aurone-derived azadienes (e.g., 1) have been extensively employed as precursors toward these skeletons owing to their easy availability and versatile reactivity (Scheme 1a).3 The polarized conjugation system, combined with the preexisting s-cis conformation, has enabled them to serve as ideal annulation partners for the synthesis of nitrogen heterocycles of variable ring sizes. Moreover, the aromatization nature of these processes by forming a benzofuran ring provides additional driving force for them to behave as a perfect four-atom synthon for [4 + n] cycloaddition.3 In contrast, the use of such species as a two-atom partner for [2 + n] cycloaddition has been less developed.3c,k,4 Herein, we report a new migrative annulation leading to benzofuran-fused dihydropyridines of unexpected topology (Scheme 1b, with formal R2 migration), which is initiated by the less common [2 + 2] cycloaddition.Open in a separate windowFig. 1Benzofuran-fused N-heterocyclic natural and bioactive molecules.Open in a separate windowScheme 1Synthesis of benzofuran-fused nitrogen heterocycles.Siloxy alkynes are another important family of building blocks in organic synthesis.5–8 The presence of a highly polarized C–C triple bond enables such molecules to serve as versatile two-carbon cycloaddition partners in various annulation reactions.5–7 In the above context and in continuation of our interest in the study of such electron-rich alkynes,7 we envisioned that the reaction between aurone-derived azadienes 1 and siloxy alkynes 2 should lead to facile electron-inversed [4 + 2] cycloaddition to form benzofuran-fused dihydropyridine products (Scheme 1b). Interestingly, the expected product 3′ from direct [4 + 2] cycloaddition was not observed. Instead, a dihydropyridine product 3 with formal R2 migration was observed. Careful analysis of the mechanism suggested that a [2 + 2] cycloaddition followed by 4π and 6π electrocyclic steps might be responsible for this unexpected product topology (vide infra).We began our investigation with the model substrates 1a and 2a, which were easily prepared in one step from aurone and 1-hexyne, respectively.8 Various Lewis acids were initially examined as potential catalysts for this cycloaddition (Table 1). Unfortunately, common Lewis acids (e.g., TiCl4, BF3·OEt2, Sc(OTf)3, In(OTf)3, and AgOTf) were all ineffective (entries 1–5). Substrate decomposition into an unidentifiable mixture was typically observed. However, further screening indicated that AgNTf2 served as an effective catalyst, leading to benzofuran-fused dihydropyridine 3a in 44% yield (entry 6). Careful analysis by X-ray crystallography confirmed that it was not formed by simple [4 + 2] cycloaddition, as the positions of the phenyl and the siloxy groups were switched (vs. the expected topology). The distinct catalytic performance of AgNTf2 (vs. AgOTf) suggested that the triflimide counter anion Tf2N might be important. However, further screening of various metal triflimide salts did not improve the reaction efficiency (entry 7). Instead, we were delighted to find that the corresponding Brønsted acid HNTf2 served as a better catalyst (57% yield, entry 8). However, triflic acid (TfOH) led to no desired product in spite of complete conversion (entry 9). After considerable efforts in the optimization of other reaction parameters, an improved yield of 75% was obtained with 2.5 mol% of HNTf2 and 2.5 equivalents of 2a at 60 °C (entry 10). Solvent screening indicated that the reaction proceeded faster in DCE with comparable yield (entry 11). However, other solvents were all inferior (entries 12–15). Finally, with a reversed order of addition of the two reactants, the yield was slightly improved (entry 16). We believe that this might be related to the relative decomposition rates of the substrates.Reaction conditionsa
EntryCatalystSolventTime (h)Yield (%)
1TiCl4DCM90
2BF3·OEt2DCM90
3Sc(OTf)3DCM90
4In(OTf)3DCM90
5AgOTfDCM90
6AgNTf2DCM944
7Sc(NTf2)3DCM90
8HNTf2DCM957
9HOTfDCM90
10bHNTf2DCM4275
11bHNTf2DCE1872
12bHNTf2CHCl31820
13bHNTf2THF180
14bHNTf2MeCN180
15bHNTf2EtOAc180
16b,cHNTf2DCE1881 (76)d
Open in a separate windowa 2a (0.06 mmol) was added to the solution of 1a (0.05 mol) and the catalyst (10 mol%). Yield was determined by analysis of the 1H NMR spectrum of the crude mixture using CH2Br2 as an internal standard.bRun with 2.5 mol% catalyst and 2.5 equiv. of 2a at 60 °C.c 1a was added into the solution of 2a and the catalyst.dYield in parentheses was isolated yield.With the optimized conditions, we examined the reaction scope. A range of aurone-derived azadienes with different electron-donating and electron-withdrawing substituents at various positions smoothly participated in this formal migrative cycloaddition process with siloxy alkyne 2a (Scheme 2). The corresponding benzofuran-fused dihydropyridine products 3 were formed with excellent selectivity and moderate to good efficiency. A thiophene unit was also successfully incorporated into the product (3h). However, substitution with a pyridinyl group shut down the reactivity, even with 1.1 equivalents of HNTf2. Other siloxy alkynes bearing different alkyl substituents on the triple bond were also good reaction partners, except that these reactions were more efficient when the catalyst loading was increased to 10 mol% (Table 2). Unfortunately, direct aryl substitution on the alkyne triple bond resulted in essentially no reaction (entry 7). Notably, in spite of the strong acidic conditions, various functional groups, such as TIPS-protected alcohol (3p) and acetal (3c), were tolerated. Moreover, increasing steric hindrance in close proximity to the reaction centers (e.g., tBu group in 3i and 3r) did not obviously affect the reaction efficiency.Scope of siloxyl alkynesa
EntryR 3 Yield (%)
1 3m 66
2 3n 74
3 3o 53b
4 3p 64
5 3q 58
6 3r 62
7 3s <5
Open in a separate windowaConditions: 1d (0.3 mmol), 2 (0.75 mmol), HNTf2 (10 mol%), DCE (3 mL), 60 °C. Isolated yield.bRun with 2.5 mol% of HNTf2.Open in a separate windowScheme 2Scope of aurone-derived azadienes. Conditions: 1 (0.3 mmol), 2a (0.75 mmol), HNTf2 (2.5 mol%), DCE (3.0 mL), 60 °C. Isolated yield.Owing to the electron-rich silyl enol ether motif, the benzofuran-fused dihydropyridine products can be transformed into other related heterocycles upon treatment with electrophiles. For example, deprotection of the silyl group in 3d with TBAF in the presence of water produced ketone 4a (eqn (1)). In the presence of NBS or NCS, the corresponding bromoketone 4b and chloroketone 4c were obtained, respectively (eqn (2)). These reactions were both efficient and highly diastereoselective. The structures of 4b and 4c were also confirmed by X-ray crystallography. Moreover, deprotection of the N-tosyl group with Li/naphthalene followed by air oxidation led to the highly-substituted benzofuran-fused pyridine 5, the core structure of a family of bioactive molecules (eqn (3)).2A possible mechanism is proposed to rationalize the unusual formal migrative process (Scheme 3). The reaction begins with LUMO-lowering protonation of the aurone-derived azadiene 1 by HNTf2.9 Then, the electron-rich alkyne attacks the resulting activated iminium intermediate I, leading to ketenium ion II after intermolecular C–C bond formation. Subsequent intramolecular cyclization from the electron-rich enamine motif to the electrophilic ketenium unit forms oxetene III. The formation of this oxetene can also be considered as a [2 + 2] cycloaddition of the two reactants.6ad,11 Subsequent 4π-electrocyclic opening of oxetene III affords azatriene IV. Further 6π-electrocyclic closing leads to the observed product 3. This observed product topology is fully consistent with this pathway. It is worth noting that the excellent performance with HNTf2 might be attributed to the low nucleophilicity and good compatibility of its counter anion with the highly electrophilic cationic intermediates (e.g., ketenium II) in this process. We have also carried out DFT studies. The results indicated that the proposed pathway is energetically viable and consistent with the experimental data (Scheme 3 and Fig. S1). Moreover, some other possible pathways that engage the nitrogen atom in intermediate II to directly attack the ketenium in a [4 + 2] mode were explored. However, no reasonable transition state could be located (Fig. S2). Thus, the origin of preference toward [2 + 2] cycloaddition remains unclear.Open in a separate windowScheme 3Proposed mechanism and free energies (in kcal mol−1) computed at the M06-2X(D3)/6-311G(d,p)-SMD//M06-2X/6-31G(d) level of theory.We also prepared TIPSNTf2 and examined its catalytic activity in this reaction since it is known that such a Lewis acid might be generated in situ.10 However, no reaction was observed when TIPSNTf2 was used in place of HNTf2, suggesting that it is unlikely the actual catalyst. Finally, in order to probe the nature of the substituent migration (intermolecular vs. intramolecular), we carried out a cross-over experiment (Scheme 4). Under the standard conditions, the reaction using a 1 : 1 mixture of 1d and 1k led to exclusive formation of 3d and 3k, without detection of any cross-over products. This result is consistent with the proposed intramolecular migration pathway.Open in a separate windowScheme 4Cross-over experiment.In conclusion, we have discovered an unusual formal migrative cycloaddition of aurone-derived azadienes with siloxy alkynes. In the presence of a catalytic amount of HNTf2, this reaction provided expedient access to a range of useful benzofuran-fused dihydropyridine products with unexpected topology, distinct from normal [4 + 2] cycloaddition. Although aurone-derived azadienes are ideal four-atom synthons for direct [4 + n] cycloaddition, the present process is initiated by less common [2 + 2] cycloaddition, which is critical for the observed product formation. Subsequent electrocyclic opening and cyclization steps provide a reasonable rationale. The heterocyclic products generated from this process are precursors toward other useful structures, such as benzofuran-fused pyridines.  相似文献   

11.
The direct catalytic asymmetric hydrogenation of (Z)-α-dehydroamino boronate esters was realized. Using this approach, a class of therapeutically relevant alkyl-substituted α-amidoboronic esters was easily synthesized in high yields with generally excellent enantioselectivities (up to 99% yield and 99% ee). The utility of the products has been demonstrated by transformation to their corresponding boronic acid derivatives by a Pd-catalyzed borylation reaction and an efficient synthesis of a potential intermediate of bortezomib. The clean, atom-economic and environment friendly nature of this catalytic asymmetric hydrogenation process would make this approach a new alternative for the production of alkyl-substituted α-amidoboronic esters of great potential in the area of organic synthesis and medicinal chemistry.

The direct catalytic asymmetric hydrogenation of (Z)-α-dehydroamino boronate esters was realized.

Since FDA approval of bortezomib1 for the treatment of multiple myeloma, chiral α-aminoboronic acids have been recognized as key pharmacophores for the design of proteasome inhibitors.2 The incorporation of chiral α-aminoboronic acid motifs at the C-terminal position of a peptide3 to develop potential clinical drug candidates has drawn increasing interest4 (Fig. 1). Meanwhile, chiral α-amidoboronic acids and their derivatives are useful synthetic building blocks for the stereospecific construction of chiral amine compounds.5 The biological and synthetic value of α-amidoboronates has led to considerable efforts for the development of efficient synthetic methods. However, up to now, limited transition-metal-catalyzed asymmetric approaches have been reported. The widely used strategies to synthesize these compounds are stepwise Matteson homologation/N-nucleophilic replacement,6 borylation of imines,7 and alkene functionalization.8 Recently, two other elegant approaches, Ni-catalyzed decarboxylative borylation of α-amino acid derivatives9 and enantiospecific borylation of lithiated α-N-Boc species,10 were reported by the Baran and Negishi groups, respectively. To the best of our knowledge, the majority of the methods relied on either stoichiometric amounts of chiral auxiliaries6,7a,b or substrate-control strategies9 and most of these methods enable the construction of aryl-substituted α-aminoboronates. Enantioselective methods to access unfunctionalized alkyl-substituted α-aminoboronic esters are still rarely developed and so far only two examples have been realized by the Miura8a and Scheidt7f groups, respectively. Considering that most therapeutically relevant α-amidoboronic acid fragments contain an alkyl subunit and the fact that the options for the synthesis of alkyl-substituted α-amidoboronic esters in an enantioselective manner are still rare, the development of other distinct approaches would be highly desirable. Herein, we report a new alternative to access these compounds by catalytic asymmetric hydrogenation of (Z)-α-dehydroamidoboronate esters. With this approach, the desired chiral alkyl-substituted α-amidoboronic esters could be obtained in high yields and generally excellent enantioselectivities (up to 99% yield and 99% ee) with simple purification.Open in a separate windowFig. 1Selected inhibitors containing chiral alkyl-substituted α-amidoboronic acids.Catalytic asymmetric hydrogenation of olefins is an atom-economic, environmentally friendly and clean process for the synthesis of valuable pharmaceuticals, agricultural compounds and feedstock chemicals.11 Recently, hydrogenation of vinylboronic compounds has emerged for the preparation of chiral boronic compounds in a regiodefined manner.12,13 However, surprisingly α-dehydroamido boronate esters and their derivatives, as elegant precursors to access alkyl-substituted α-amidoboronic compounds, have never been used as substrates in asymmetric hydrogenation and remain a challenging project. To our knowledge, only one efficient hydrogenation approach to (1-halo-1-alkenyl) boronic esters was reported for indirect synthesis of alkyl-substituted α-aminoboronic esters but it was accompanied by inevitable de-halogenated by-products14 (Scheme 1). Given the catalytic efficiency and atom economy of the hydrogenation method, the development of a new direct hydrogenation approach to construct these important chiral alkyl-substituted α-amidoboronic esters would be very appealing.Open in a separate windowScheme 1Approaches towards the synthesis of chiral alkyl-substituted α-aminoboronic esters.The inspiration for our approach to the hydrogenation of α-dehydroamido boronates came from the molecular structures of relevant biologically active inhibitors containing alkyl-substituted α-amidoboronic acid fragments. Due to the limited stability of free α-aminoboronic acids, an electron-withdrawing carboxylic N-substituent is often required.15 Thus, we envisaged that N-carboxyl protected α-dehydroamido boronate esters could serve as a potential precursor for the synthesis of alkyl-substituted α-amidoboronates through Rh-catalyzed asymmetric hydrogenation of the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond16 (Fig. 1), a strategically distinct approach to the construction of unfunctionalized alkyl-substituted α-amidoboronic esters. However, challenges still remain, including: (1) how to synthesize α-dehydroamido boronates; (2) the facile transmetalation process of the starting materials leading to deboronated by-products in the hydrogenation process;17 (3) the unknown stability of α-amidoboronic compounds in the presence of a transition-metal catalyst and hydrogen molecules. As part of our continuous efforts to develop efficient hydrogenation approaches to construct valuable motifs,18 here we present the results of the investigation to address the aforementioned challenges.The desired aryl-substituted (Z)-α-dehydroamido boronates could be obtained by Cu-catalyzed regioselective hydroborylation of ynamide according to a previous report.19 However, different α/β-regioselectivity was observed for the preparation of alkyl-substituted (Z)-α-dehydroamido boronate esters and a new synthetic route was developed (Scheme 2, see the ESI for details). Of note, (Z)-α-dehydroamido boronate esters should be purified with deactivated silica gel,7c or else protodeborylation would occur readily with flash chromatography.Open in a separate windowScheme 2Synthetic route to (Z)-α-dehydroamino boronates.In order to check the feasibility of our hypothesis, three substrates were prepared with Rh(NBD)2BF4 and examined and our group prepared (Rc,Sp)-DuanPhos under 50 atm hydrogen pressure (Table 1). Gratifyingly, substrate 1b reacted smoothly to provide the desired product 2b in high yield and enantioselectivity (>99% conv., 98% ee, entry 2) whilst the reaction with substrate 1a yielded a mixture of deborylation products and 1c did not work at all (entries 1 and 3). Of note, we did not observe deborylation products with 1b under the current reaction conditions and we did not select (Z)-α-dehydroamido boronic acid 1a as the model substrate because of its poor solubility in most solvents. Then, a variety of chiral diphosphine ligands were investigated along with Rh(NBD)2BF4 and the results are shown in Table 1. In most cases, the reaction proceeded smoothly to furnish the desired products and the best results were obtained when (Rc,Sp)-DuanPhos was used as the ligand (entries 2 and 4–12). Poor results were obtained with axially bidentate phosphine ligands (entries 5, 6 and 9). (R,R)-QuinoxP* and (R,R)-Ph-BPE also gave good conversion with a slightly decreased ee whilst (R,R)-iPr-DuPhos exhibited poor results (entries 4, 7 and 10). Subsequent solvent screening revealed that the desired products could be obtained in most of the solvents and 1,2-DCE was the best solvent. (Entry 13, see the ESI).Condition optimization for catalytic asymmetric hydrogenation of 1a
EntrySubLigandConv.b (%)eec (%)
1d 1a (Rc,Sp)-DuanPhos89n.d.
2e 1b (Rc,Sp)-DuanPhos>9998
3 1c (Rc,Sp)-DuanPhosn.r.n.d.
4 1b (R,R)-QuinoxP*>9997
5 1b (S)-SegPhos>9917
6 1b (S)-BINAP>9910
7 1b (R,R)-iPr-DuPhos>993
8 1b (R,S)-Cy-JosiPhos>9914
9 1b (R)-BIPHEP>99−30
10 1b (R,R)-Ph-BPE>99−86
11 1b (S,S)-f-Binaphane>9961
12 1b (2S,4S)-BDPP>9959
13e,f,g 1b (Rc,Sp)-DuanPhos>99(99)99
Open in a separate windowaUnless otherwise mentioned, the reactions were performed with 1 (0.1 mmol), Rh(NBD)2BF4 (10 mol%), and a ligand (11 mol%) in 1.0 mL THF at 50 °C for 15 h.bDetermined by crude 1H NMR.cDetermined with chiral HPLC.dThe reaction was performed in iPrOH.eRh(NBD)2BF4 (1.0 mol%) and ligand (1.05 mol%) were used.fIsolated yield in parentheses.g1,2-DCE was used as the solvent. Pin = 2,3-dimethyl-2,3-butanediol; dan = 1,8-diaminonaphthalene.With the optimized reaction conditions in hand, a series of (Z)-α-dehydroamido boronate esters were tested and the results are summarized in Table 2. All the substrates reacted smoothly to give the corresponding alkyl-substituted α-amidoboronates in high yields with good to excellent enantioselectivities (2b, 2d–2r, and 2u, 99% yield, 57–99% ee). Alkyl-substituted (Z)-α-dehydroamido boronate esters were well tolerated in the current reaction, providing the corresponding α-amidoboronates in high yields and excellent enantioselectivities (2d–2i, 99% yield, 96–99% ee). Aryl-substituted (Z)-α-dehydroamido boronate esters with electron-donating (2j–l, 2n and 2p–r) and withdrawing (2m and 2o) substituents could also give the desired products in excellent yield with excellent enantioselectivities (90–99% ee). The ortho-methyl-substituted substrate 1r reacted smoothly to give the desired product with excellent enantioselectivity, but the 2,6-dimethyl-substituted substrate 1z could not react at all. Functional groups such as ether, halo and benzyl were well tolerated in the current reaction (2k, 2l, 2m and 2o–q). Replacement of the N-substituents with acyclic carbamate was also tolerated but with a decreased ee (2u and 2z). Substrates containing a chiral oxazolidin-2-one unit bearing bulky Ph-substituents around the nitrogen and oxygen were also competent, yielding the desired products with good to excellent diastereoselectivities (2s, 2t, and 2v–y). Of note, the substrate 1s bearing an N-Ms substituent and the cyclic substrate 1t did not work in the current reaction. The absolute configuration of generated α-amidoboronates was assigned as (S) by X-ray crystallographic analysis of 2i (Scheme 3).20Substrate scope.a,b,c
Open in a separate windowaUnless otherwise mentioned, the reactions were performed with 1 (0.1 mmol), Rh(NBD)2BF4 (1.0 mol%), and a ligand (1.05 mol%) in 1.0 mL 1,2-DCE at 50 °C under 50 atm H2 for 15 h.bIsolated yield.cDetermined with chiral HPLC.dDetermined by crude 1H NMR.Open in a separate windowScheme 3Scale-up synthesis and synthetic utility.To demonstrate the utility of the products, a scale-up reaction (0.62 g) was successfully performed with 0.1 mol% catalytic loading, giving 2b in 99% yield and 98% ee, and 2b could be easily transformed to a more stable α-amidoborate 3b with KHF2,6d,21 followed by hydrolysis with TMSCl to yield α-amido boronic acid 4b in 46% yield,22 which could also be obtained from 2b by treating it with BCl3 in 84% yield, without loss of the optical purity.8b2m could easily be transformed to 4m in 68% yield by a Pd-catalyzed borylation reaction. Meanwhile, after hydrogenation of 1x to 2x′ and transformation of 2x′ to its trifluoroborate derivative 3x′, removal of the benzyl group of 3x′ with Pd/C under hydrogenation conditions23 yielded the primary α-aminoborate 4x in 62% yield in three steps, which could serve as a potential precursor15 to synthesize bortezomib.  相似文献   

12.
Despite the versatility of amphoteric molecules, stable and easily accessible ones are still limitedly known. As a result, the discovery of new amphoteric reactivity remains highly desirable. Herein we introduce 3-aminooxetanes as a new family of stable and readily available 1,3-amphoteric molecules and systematically demonstrated their amphoteric reactivity toward polarized π-systems in a diverse range of intermolecular [3 + 2] annulations. These reactions not only enrich the reactivity of oxetanes, but also provide convergent access to valuable heterocycles.

Despite the versatility of amphoteric molecules, stable and easily accessible ones are still limitedly known.

Amphoteric molecules, which bear both nucleophilic and electrophilic sites with orthogonal reactivity, represent an attractive platform for the development of chemoselective transformations.1 For example, isocyanides are well-established 1,1-amphoteric molecules, with the terminal carbon being both nucleophilic and electrophilic, and this feature has enabled their exceptional reactivity in numerous multi-component reactions.2 In the past few decades, substantial effort has been devoted to the search for new amphoteric molecules.1–5 Among them, 1,3-amphoteric molecules proved to be versatile. The Yudin and Beauchemin laboratories have independently developed two types of such molecules, α-aziridine aldehydes and amino isocyanates, respectively.4,5 With an electrophilic carbon and a nucleophilic nitrogen in relative 1,3-positions, these molecules are particularly useful for the chemoselective synthesis of heterocycles with high bond-forming efficiency without protective groups (Fig. 1). However, such elegant amphoteric systems still remain scarce. Therefore, the development of new stable amphoteric molecules with easy access remains highly desirable.Open in a separate windowFig. 1Representative [1,3]-amphoteric molecules versus 3-aminooxetanes.In this context, herein we introduce 3-aminooxetanes as a new type of 1,3-amphoteric molecules and systematically demonstrate their reactivity in a range of [3 + 2] annulations, providing rapid access to diverse heterocycles. Notably, 3-aminooxetanes are bench-stable and either commercially available or easily accessible. However, their amphoteric reactivity has not been appreciated previously.Oxetane is a useful functional group in both drug discovery and organic synthesis.6–9 Owing to the ring strain, it is prone to nucleophilic ring-opening, in which it serves as an electrophile (Scheme 1A).6–8 We envisioned that, if a nucleophilic group is installed in the 3-position (e.g., amino group), such molecules should exhibit 1,3-amphoteric reactivity due to the presence of both nucleophilic and electrophilic sites (Scheme 1B). Importantly, the 1,3-relative position is crucial for inhibiting self-destructive intra- or intermolecular ring-opening (i.e. the 3-nucleophilic site attack on oxetane itself) due to high barriers. Thus, such orthogonality is beneficial to their stability. In contrast, the nucleophilic site is expected to react with an external polarized π bond (e.g., X = Y, Scheme 1B), which enables a better-positioned nucleophile (Y) to attack the oxetane and cyclize. Thus, a formal [3 + 2] annulation should be expected. Unlike the well-known SN2 reactivity of oxetanes with simple bond formation, this amphoteric reactivity would greatly enrich the chemistry of oxetanes with multiple bond formations and provide expedient access to various heterocycles. In contrast to the conventional approaches that require presynthesis of advanced intermediates (e.g., intramolecular ring-opening),8 the exploitation of such amphoteric reactivity in an intermolecular convergent manner from simple substrates would be more practically useful. Moreover, more activation modes could be envisioned in addition to oxetane activation. In 2015, Kleij and coworkers reported an example of cyclization between 3-aminooxetane and CO2 in 55% yield, which provided a pioneering precedent.10 However, a systematic study to fully reveal such amphoteric reactivity in a broad context remains unknown in the literature.Open in a separate windowScheme 1Typical oxetane reactivity and the new amphoteric reactivity.To test our hypothesis, we began with the commercially available 3-aminooxetanes 1a and 1b as the model substrates. Phenyl thioisocyanate 2a and CS2 were initially employed as reaction partners, as they both have a polarized C Created by potrace 1.16, written by Peter Selinger 2001-2019 S bond as well as a relatively strong sulfur nucleophilic motif. Moreover, the resulting desired products, iminothiazolidines and mercaptothiazolidines, are both heterocycles with important biological applications (Fig. 2).11 To our delight, simple mixing these two types of reactants in DCM resulted in spontaneous reactions at room temperature without any catalyst. The corresponding [3 + 2] annulation products iminothiazolidine 3a and mercaptothiazolidine 4a were both formed with excellent efficiency (Scheme 2). It is worth mentioning that catalyst-free ring-opening of an oxetane ring is rarely known, particularly for intermolecular reactions.6–9 In this case, the high efficiency is likely attributed to the suitable choice and perfect position of the in situ generated sulfur nucleophile.Open in a separate windowFig. 2Selected bioactive molecules containing iminothiazolidine and mercaptothiazolidine motifs.Open in a separate windowScheme 2Initial results between 3-aminooxetanes and thiocarbonyl compounds.The catalyst-free annulation protocol is general with respect to various 3-aminooxetanes and isothiocyanates. A range of iminothiazolidines and mercaptothiazolidines were synthesized with high efficiency under mild conditions (Scheme 3). Many of them were obtained in quantitative yield. Quaternary carbon centers could also be generated from 3-substituted 3-aminooxetanes (e.g., 3j). The structure of product 3b was unambiguously confirmed by X-ray crystallography.Open in a separate windowScheme 3Formal [3 + 2] annulation with isothiocyanates and CS2. Reaction conditions: 1 (0.3–0.4 mmol), 2 (1.1 equiv.) or CS2 (1.5 equiv.), DCM (2 mL), RT, 3 h for 3 and 36 h for 4. Yields are for the isolated products.With the initial success of thiocarbonyl partners, we next turned our attention to isocyanates, in which the carbonyl group serves as the [3 + 2] annulation motif. Compared with sulfur as the nucleophilic site in the above cases, the oxygen atom is less nucleophilic. As expected, initial tests of the reactivity by mixing 1b and 5a resulted in no desired annulation product 6a in the absence of a catalyst (Table 1, entry 1). Next, Brønsted acids, including TsOH and the super acid HNTf2, were examined as catalysts, but with no success (entries 2 and 3). We then resorted to various Lewis acids, particularly those oxophilic ones, in hope of activating the oxetane unit. Unfortunately, many of them still remained ineffective (e.g., ZnCl2, AuCl, and FeCl3). However, to our delight, further screening of stronger Lewis acids helped identify Sc(OTf)3, Zn(OTf)2, and In(OTf)3 to be effective at room temperature, leading to the desired iminooxazolidine product 6a in good yield (entries 7–9). Its structure was confirmed by X-ray crystallography. Nevertheless, aiming to search for a cheaper catalyst, we continued to optimize this reaction at a higher temperature using previous ineffective catalysts. Indeed, FeCl3 was found to be effective at 80 °C (61% yield, entry 10), while Brønsted acid TsOH remained ineffective at this temperature (entry 11). Notably, decreasing the loading of FeCl3 to 1 mol% led to a higher yield (89% yield, entry 12). However, further decreasing to 0.5 mol% resulted in slightly diminished efficiency (entry 13).Reaction conditions for annulation with isocyanatesa
EntryCatalystYieldb (%)
10
2TsOH·H2O0
3HNTf20
4ZnCl20
5AuCl0
6FeCl30
7Sc(OTf)374
8Zn(OTf)278
9In(OTf)390
10FeCl3c61
11TsOH·H2Oc0
12FeCl3c (1 mol%)89(84)d
13FeCl3c (0.5 mol%)85
Open in a separate windowaReaction scale: 1b (0.1 mmol), 5a (0.1 mmol), catalyst (10 mol%), toluene (1 mL).bYield based on analysis of the 1H NMR spectra of the crude reaction mixture using trichloroethylene as an internal reference. For all the entries, the urea product from simple amine addition to isocyanate 5a accounts for the mass balance.cRun at 80 °C.dIsolated yield.While there are multiple effective catalysts, FeCl3 was selected for the scope study in view of its low price. Various substituted 3-aminooxetanes and isocyanates were subjected to this annulation protocol (Scheme 4). The corresponding iminooxazolidine products were all obtained in good to excellent yields. Isocyanates containing an electron-donating or electron-withdrawing group were both suitable reaction partners. Remarkably, a 1.5 mmol scale reaction of 6a also worked efficiently.Open in a separate windowScheme 4Formal [3 + 2] annulation between 3-aminooxetanes and isocyanates. Reaction scale: 1 (0.3 mmol), 5 (0.3 mmol), FeCl3 (1 mol%), toluene (2 mL).Although (thio)isocyanates and CS2 have been successfully utilized in the formal [3 + 2] annulation with 3-aminooxetanes, these partners are relatively reactive. We were curious about whether the C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bond in relatively inert molecules could react in a similar manner. For example, the C Created by potrace 1.16, written by Peter Selinger 2001-2019 O bond in CO2 is both thermodynamically and kinetically inert relative to typical organic carbonyl groups. However, as a cheap, abundant and green one-carbon source, CO2 has been a subject of persistent investigations owing to its versatility in various transformations leading to valuable materials.12 Specifically, if CO2 could be employed as a partner for the [3 + 2] annulation with 3-aminooxetanes, it would represent an attractive synthesis of oxazolidinones, a well-known heterocycle with applications in both organic synthesis and medicinal chemistry.13 In this context, we next studied the possibility of utilizing CO2 in our annulation.As expected, the reaction between 1b and CO2 at 1 atmospheric pressure did not proceed without a catalyst (Table 2, entry 1). Next, we examined representative Lewis acids, such as Sc(OTf)3, In(OTf)3 and FeCl3. Among them, Sc(OTf)3 exhibited the highest catalytic activity at room temperature (22% yield, entry 2). The reaction efficiency could be improved at 80 °C (65% yield, entry 6), but no further improvement could be made at a higher temperature or with other solvents. Next, we resorted to organic nitrogen bases, as they were known as effective activators of CO2.14 While Et3N and DABCO were completely ineffective for the reaction in MeCN at 80 °C, fortunately, TMG, TBD, and DBU were competent for the desired process (entries 7–11). Among them, DBU exhibited the best performance, leading to the desired product 7a in 89% yield (entry 11). It is worth noting that the polar solvent MeCN was found to be crucial for the base-catalyzed reactivity. Less polar solvents, such as toluene, DCE or THF, completely shut down the reaction. We believe that effective stabilization of certain polar intermediates involved here is critically beneficial to decreasing the reaction barrier. Finally, unlike the previous Lewis acid-catalyzed annulation with isocyanates, this base-catalyzed [3 + 2] annulation with CO2 proceeds via a different activation mode (i.e., to activate CO2 rather than oxetane). We believe that expansion of possible activation modes in this type of amphoteric reactivity will enrich the chemistry of oxetanes.Reaction conditions for annulation with CO2a
EntryCatalyst T Conv. (%)Yield (%)
1RT00
2Sc(OTf)3RT4822
3In(OTf)3RT339
4Zn(OTf)2RT70
5Sc(OTf)360 °C10061
6Sc(OTf)380 °C10065
7Et3N80 °C00
8DABCO80 °C50
9TMG80 °C7254
10TBD80 °C10088
11DBU80 °C10089
Open in a separate windowaReaction scale: 1b (0.1 mmol), CO2 (1 atm), solvent (0.5 mL). Yields based on analysis of the 1H NMR spectra of the crude reaction mixture using CH2Br2 as an internal standard.We next examined the scope of this CO2-fixation process. Unfortunately, at a larger scale (0.5 mmol), the same condition (entry 11, Table 2) could not lead to complete conversion within 12 h. Therefore, further optimization aiming to accelerate the reaction was performed. Indeed, a higher concentration (1.0 M) resulted in a higher rate without affecting the yield. As shown in Scheme 5, a wide variety of 3-aminooxetanes were smoothly converted to the corresponding oxazolidinones in high yields. Both electron-donating and electron-withdrawing substituents on the N-benzyl group did not affect the efficiency. Heterocycle-based N-benzyl or N-allylic substituents are all suitable substrates. However, for regular alkyl substituents, such as homobenzyl (7h) or n-butyl (7j), the stronger base catalyst TBD was needed to achieve good efficiency. Furthermore, this reaction can tolerate steric hindrance in the 3-position of the oxetane (7k), where a quaternary carbon center could be incorporated. However, increasing the size of the N-substituent, such as the secondary alkyl groups in 7i and 7l, did influence the reactivity, thus requiring a higher temperature (100 °C). This process exhibited good compatibility with diverse functional groups, such as ethers, pyridines, aryl halides, olefins, silyl-protected alcohols, and phthalimides. Finally, this protocol is also capable of generating various oxazolidinones embedded in a different structural context, such as chiral oxazaolidinone 7l, bis(oxazolidinone) 7m, and polyheterocycle-fused oxazolidinone 7o.Open in a separate windowScheme 5Formal [3 + 2] annulation between 3-aminooxetanes and CO2. aReaction scale: 1 (0.5 mmol), CO2 (1 atm), DBU (10 mol%), MeCN (0.5 mL). Isolated yield. bRun with TBD as the catalyst. cRun with DMF as solvent at 100 °C.In summary, 3-aminooxetanes have been systematically demonstrated, for the first time, as versatile 1,3-amphoteric molecules. They are a new addition to the limited family of amphoteric molecules. Though previously unappreciated, these molecules exhibited various advantages over the related known 1,3-amphotric molecules (e.g., α-aziridine aldehydes and amino isocyanates), including easy access and extraordinary stability. The perfect position of the nucleophilic nitrogen together with the orthogonal electrophilic carbon allowed them to participate in a diverse range of intermolecular formal [3 + 2] annulations with polarized π-systems, leading to rapid access to various valuable nitrogen heterocycles. Different types of polarized double bonds, from reactive (thio)isocyanates to inert CO2, all participated efficiently in these highly selective annulations with or without a suitable catalyst. Furthermore, the involvement of more functional groups in such amphoteric reactivity allowed manifold activation modes, thereby greatly enriching the reactivity of the already versatile oxetane unit to a new dimension. These reactions, proceeding in an intermolecular convergent manner from readily available substrates, provide expedient access to various valuable nitrogen heterocycles, thus being complementary to those traditional methods that either required multiple steps or less available substrates. More studies on the 1,3-amphoteric reactivity of 3-oxetanes, particularly those with other partners as well as their asymmetric variants, are ongoing in our laboratory.  相似文献   

13.
The use of hydrazine-catalyzed ring-closing carbonyl–olefin metathesis (RCCOM) to synthesize polycyclic heteroaromatic (PHA) compounds is described. In particular, substrates bearing Lewis basic functionalities such as pyridine rings and amines, which strongly inhibit acid catalyzed RCCOM reactions, are shown to be compatible with this reaction. Using 5 mol% catalyst loadings, a variety of PHA structures can be synthesized from biaryl alkenyl aldehydes, which themselves are readily prepared by cross-coupling.

Hydrazine catalysis enables the ring-closing carbonyl–olefin metathesis (RCCOM) to form polycyclic heteroaromatics, especially those with basic functionality.

Polycyclic heteroaromatic (PHA) structures comprise the core framework of many valuable compounds with a diverse range of applications (Fig. 1A).1 For example, polycyclic azines (e.g. quinolines) are embedded in many alkaloid natural products, including diplamine2 and eupolauramine3 to name just a few. These types of structures are also of interest for their biological activity, such as with the inhibitor of the Src-SH3 protein–protein interaction shown in Fig. 1A.4 Many nitrogenous PHAs are also useful as ligands for transition metal catalysis, as exemplified by the widely used ligand 1,10-phenanthroline.5 Meanwhile, chalcogenoarenes6 such as dinaphthofuran7 and benzodithiophene8 have attracted high interest for both their medicinal properties9 and especially for their potential use as organic light-emitting diodes (OLEDs), organic photovoltaics (OPVs), and organic field-effect transistors (OFETs).10 These and numerous other examples have inspired the development of a wide variety of strategies to construct PHAs.1,11–14 Although these approaches are as varied as the structures they target, the wide range of molecular configurations within PHA chemical space and the challenges inherent in exerting control over heteroatom position and global structure make novel syntheses of these structures a topic of continuing interest.Open in a separate windowFig. 1(A) Examples of PHAs. (B) RCCOM strategy for PHA synthesis. (C) Lewis base inhibition for Lewis acid vs. hydrazine catalyzed RCCOM. (D) Hydrazine-catalyzed RCCOM for PHA synthesis.One potentially advantageous strategy for PHA synthesis is the use of ring-closing carbonyl–olefin metathesis15 (RCCOM) to forge one of the PHA rings, starting from a suitably disposed alkenyl aldehyde precursor 2 that can be easily assembled by cross-coupling (Fig. 1B). In related work, the application of RCCOM to form polycyclic aromatic hydrocarbons (PAHs) was reported by Schindler in 2017.16 In this case, 5 mol% FeCl3 catalyzed the metathesis of substrates to form phenanthrenes and related compounds in high yields at room temperature. This method was highly attractive for its efficiency, its use of an earth-abundant metal catalyst, and the production of benign acetone as the only by-product. Nevertheless, one obvious drawback to the use of Lewis acid activation is that the presence of any functionality that is significantly more Lewis basic than the carbonyl group can be expected to strongly inhibit these reactions (Fig. 1C). Such a limitation thus renders this method incompatible with a wide swath of complex molecules, especially PHAs comprised of azine rings. This logic argues for a mechanistically orthogonal RCCOM approach that allows for the synthesis of PHA products with a broader range of ring systems and functional groups.We have developed an alternative approach to catalytic carbonyl–olefin metathesis that makes use of the condensation of 1,2-dialkylhydrazines 5 with aldehydes to form hydrazonium ions 6 as the key catalyst–substrate association step.17–19 This interaction has a much broader chemoorthogonality profile than Lewis acid–base interactions and should thus be much less prone to substrate inhibition than acid-catalyzed approaches. In this Communication, we demonstrate that hydrazine-catalyzed RCCOM enables the rapid assembly of PHAs bearing basic functionality (Fig. 1D).For our optimization studies, we chose biaryl pyridine aldehyde 7 as the substrate (20 salt 11 was also productive (entry 2), albeit somewhat less so. Notably, iron(iii) chloride generated no conversion at either ambient or elevated temperatures (entries 3 and 4). Trifluoroacetic acid (TFA) was similarly ineffective (entry 5). Meanwhile, a screen of various solvents revealed that, while the transformation could occur in a range of media (entries 6–9), THF was optimal. Finally, by raising the temperature to 90 °C (entry 10) or 100 °C (entry 11), up to 96% NMR yield (85% isolated yield) of adduct 8 could be obtained in the same time period.Optimization studiesa
EntryCatalystSolventTemp. (°C)8 yield (%)
110THF8067
211THF8053
3FeCl3DCErt0
4FeCl3DCE800
5TFATHF800b
610i-PrOH8031
710CH3CN8028
810EtOAc8026
910Toluene8024
1010THF9087
1110THF10096c
Open in a separate windowaConditions: substrate 8 (0.2 mmol) and 5 mol% catalyst in 0.4 mL of solvent (0.5 M) in a 5 mL sealed tube were heated to the temperature indicated for 15 h. Yields were determined by 1H NMR using CH2Br2 as an internal standard.b2 equiv. of TFA was used.c85% isolated yield.Using the optimized conditions, we explored the synthesis of various PHAs (Fig. 2). In addition to benzo[h]isoquinoline (8), products 12 and 13 with fluorine substitution at various positions could be generated in good yields. Similarly, benzoisoquinolines 14 and 15 bearing electron-donating methoxy groups and the dioxole-fused product 16 were also accessed efficiently. Furthermore, a phenolic ether product 17 with a potentially acid-labile N-Boc group was generated in modest yield. We found that an even more electron-donating dimethylamino group was also compatible with this chemistry, allowing for the production of 18 in 68% yield. On the other hand, adduct 19 bearing a strongly electron-withdrawing trifluoromethyl group was isolated in only modest yield. The naphtho-fused isoquinoline 20 could be generated as well; however, 20 mol% catalyst was required to realize a 35% yield. The thiophene-fused product 21 was furnished in much better yield, also with the higher catalyst loading. Although not a heterocyclic system, we found that the reaction to form phenanthrene (22) was well-behaved, providing that compound in 83% yield. In addition, an amino-substituted phenanthrene 23 was also formed in good yield. Other thiophene-containing PAHs such as 24–26 were produced efficiently. On the other hand, adduct 27 was generated only in low yield. Naphthofuran (28), which is known to have antitumor and oestrogenic properties,21 was synthesized in good yield. Finally, pharmaceutically important structures such as benzocarbazole2229 and naphthoimidazole2330 could be accessed in moderate yields with increased catalyst loading.Open in a separate windowFig. 2Substrate scope studies for hydrazine 1-catalyzed RCCOM synthesis of polycyclic heteroaromatics. a Conditions: substrate and catalyst 1·(TFA)2 (5 mol%) in THF (0.5 M) were heated to 100 °C in a 5 mL sealed tube for 15 h. Yields were determined on purified products. b 20 mol% catalyst.We also examined the scope of the olefin substitution pattern (
EntrySubstrateTime (h)Yield (%)
1 1596
2 485
3b 4827
4 4854
5 4864
Open in a separate windowaConditions: 5 mol% 10 in THF (0.5 M) in a 5 mL sealed tube were heated to the temperature indicated for 15–48 h. Conversions and yields were determined by 1H NMR using CH2Br2 as an internal standard.bMixture of E/Z (2 : 1) isomers.The vinyl substrate 31 led to very little desired product (entry 2), while the propenyl substrate 32 (2 : 1 mixture of E and Z isomers) was somewhat improved but still low-yielding (entry 3). Finally, styrenyl substrates 33 and 34 (entries 4 and 5) led to improved yields relative to 31 and 32, with the cis isomer 34 being slightly more efficient (entry 5).In order to better understand the facile nature of this RCCOM reaction, we conducted DFT calculations for each step of the proposed reaction pathway (Fig. 3A). Condensation of the substrate 7 with [2.2.1]-hydrazinium 10 to afford the hydrazonium Z-35 was found to be exergonic by −13 kcal mol−1. Isomerization of Z-35 to E-35 comes at a cost of ∼3 kcal mol−1, but the total activation energy for cycloaddition (cf.36), taking into account this isomerization, was still relatively modest at only +21.0 kcal mol−1 with an overall exergonicity of −11.1 kcal mol−1. The energetic change for proton transfer in the conversion of cycloadduct 37a to the cycloreversion precursor 37b was negligible (+1.2 kcal mol−1). Interestingly, including the proton migration step, the cumulative energy barrier for cycloreversion 38 was found to be only +21.7 kcal mol−1, nearly the same as for the cycloaddition. Undoubtedly, the formation of an aromatic ring greatly facilitates this step relative to other types of substrates. Unsurprisingly, the cycloreversion to produce benzoisoquinoline 8 along with hydrazonium 39 was calculated to be strongly exergonic. Finally, the hydrolysis of 39 to regenerate hydrazinium catalyst 10 (and acetone) required an energy input approximately equal to that gained from the condensation with the substrate to form 35.Open in a separate windowFig. 3(A) Computational study of hydrazine 10-catalyzed RCCOM of biaryl aldehyde 7. Calculations were performed at the PCM(THF)-M06-2X/6-311+G(d,p)//6-31G(d) level of theory.24,25 All energies are given in units of kcal mol−1. (B) 1H NMR spectroscopy of the RCCOM reaction of 7 catalyzed by 10 at 60 °C in THF-d8 with mesitylene as internal standard for 5 hours. (C) Plot of the data showing conversion vs. time. SM = starting material 7; CA = cycloadduct 37; Prd = product 8.Given the low activation energy barriers of both the cycloaddition and cycloreversion steps, we reasoned it should be possible for the reaction to proceed at a relatively low temperature. In fact, we observed 82% conversion of biaryl aldehyde 7 to cycloadduct 37 (72%) and benzoisoquinoline 8 (10%) at 40 °C over 6 hours. Attempts to isolate the cycloadduct 37 resulted in complete conversion to 8 during column chromatography. Meanwhile, at 60 °C over approximately 4 hours, 95% of the starting material 7, via the intermediate cycloadduct 37, was converted to benzoisoquinoline product 8 (Fig. 3B and C). The rate of consumption of the cycloadduct was consistent with first-order behavior, and upon fitting, revealed the rate constant for cycloreversion as kCR = 2.14 × 10−4 s−1, with a half-life of 54 minutes. These observations corroborate the computational results, in particular showing that the cycloreversion step is quite facile with these types of substrates compared to other hydrazine-catalyzed COM reactions we have investigated17 and that cycloaddition and cycloreversion have energetically similar activation energies.In conclusion, the development of catalytic carbonyl–olefin metathesis reactions has opened new possibilities for the rapid construction of complex molecules. The current work demonstrates this strategy as a means to rapidly access polycyclic heteroaromatics, which often require lengthy sequences that can be complicated by the presence of basic functionality. The ability of the hydrazine catalysis platform to accommodate such functional groups provides a novel approach to polycyclic heteroaromatic synthesis and greatly expands the landscape of structures accessible by RCCOM.  相似文献   

14.
Correction: Expanding medicinal chemistry into 3D space: metallofragments as 3D scaffolds for fragment-based drug discovery     
Christine N. Morrison  Kathleen E. Prosser  Ryjul W. Stokes  Anna Cordes  Nils Metzler-Nolte  Seth M. Cohen 《Chemical science》2022,13(32):9450
Correction for ‘Expanding medicinal chemistry into 3D space: metallofragments as 3D scaffolds for fragment-based drug discovery’ by Christine N. Morrison et al., Chem. Sci., 2020, 11, 1216–1225, https://doi.org/10.1039/C9SC05586J.

The authors regret that in the original article, inhibitory values reported for some metallofragments were incorrect. Unfortunately, DMSO stock solutions of reportedly active ferrocene-based metallofragments were found to decompose in the presence of light, which resulted in inaccurate inhibition values. The authors maintain that the core conclusions of the paper are accurate and the utility of three-dimensional metal complexes for fragment-based drug discovery has merit.In the original article, ‘class A’ metallofragments are comprised of ferrocene derivatives (Fig. 1). Some of these ferrocene fragments (specifically those containing carbonyl groups) are reported as broadly inhibiting several protein targets. It was noted in our original report that the ferrocene scaffold was likely promiscuous due to its lipophilicity and potential redox activity, but that it might still serve as a useful metallofragment for fragment-based drug discovery (FBDD) campaigns. However, re-evaluation of these compounds against the influenza endonuclease (PAN) failed to reproduce our original inhibition results for the class A metallofragments using freshly prepared stocks, indicating a problem with the materials used in the original study.Open in a separate windowFig. 1Chemical structures of class A metallofragments.Several compounds from class A were originally reported as having near complete (100%) inhibition against PAN endonuclease at an inhibitor concentration of 200 μM (and2).2). However, when re-evaluated under identical conditions, using freshly prepared DMSO stock solutions, inhibition was only observed with one fragment of this class (A22, Fig. 1), with the previously reported highly active fragments (A4, A7–A21, CompoundA1A2A3A4A5A7A8A9A10A11Reported12 ± 6<1<145 ± 148 ± 7103 ± 5103 ± 453 ± 546 ± 790 ± 5Corrected3 ± 10n.d.18 ± 36 ± 321 ± 59 ± 310 ± 54 ± 216 ± 410 ± 7Open in a separate windowan.d. = not determined.
CompoundA12A14A15A16A17A18A19A20A21A22
Reported66 ± 526 ± 655 ± 719 ± 8100 ± 4107 ± 632 ± 880 ± 410 ± 1688 ± 9
Corrected9 ± 410 ± 518 ± 115 ± 65 ± 3<111 ± 9<1< 193 ± 1
Open in a separate windowReported and re-evaluated percent inhibition values of representative metallofragments against PAN endonuclease at 200 μM inhibitor concentration. Each compound was tested in triplicate from either two or three independent experimentsa
CompoundA1B1C1D1E1F1G1
Reported12 ± 64 ± 670 ± 2320 ± 1118 ± 982 ± 516 ± 6
Re-evaluated<519 ± 875 ± 1114 ± 9<510 ± 14<5
Open in a separate windowan.d. = not determined.
CompoundH1I1J1K1L1M1DPBA
Reported31 ± 626 ± 725 ± 699 ± 312 ± 426 ± 4n.d.
Re-evaluated25 ± 9<541 ± 683 ± 330 ± 854 ± 597 ± 1
Open in a separate windowIn the original article, one representative member of each metallofragment class was assessed for stability by NMR. Compound A1 (ferrocene) proved stable in DMSO and class A metallofragments were stored as DMSO stocks at −80 °C, but were not consistently protected from light. As noted above, many of the derivatives in class A contain a ferrocenyl carbonyl motif. It has been previously reported that ferrocenyl ketones can undergo photoaquation (λ > 280 nm) in wet DMSO to produce a monocyclopentadienyliron cation, the anionic ligand, and free cyclopentadiene.1 Suspecting issues with photostability, we dissolved several of the ferrocenyl fragments in DMSO-d6, exposed them to ambient room light (fluorescent light bulb), and monitored stability by NMR. Indeed, photoinstability was confirmed by the observance of free cyclopentadienyl peaks appearing in the 1H NMR spectrum (Fig. 2). It should also be noted that while the fresh stock of A22 retained significant inhibition against PAN, it also exhibits sensitivity to light in DMSO.Open in a separate windowFig. 2Compound A7 in DMSO-d6 (left) and after exposure to ambient light for 24 h (right) demonstrating the photoinstability of this compound.Based on these findings, the authors regret that the inhibitory data associated with class A metallofragments are incorrect, likely because of photodecomposition of these ferrocene derivatives. To confirm if other classes of metallofragments were correctly reported, a representative member of each class was evaluated against PAN endonuclease at an inhibitor concentration of 200 μM using freshly prepared DMSO stocks. Each compound was tested in triplicate in two or three independent experiments, with the addition of 2,4-dioxo-4-phenylbutanoic acid (DPBA) as a positive control.2 Fortunately, these experiments largely reproduced our original findings. Although several fragments showed slightly greater activity upon re-evaluation (J1, L1, M1, Fig. 3), only one fragment initially identified as a hit (>50% inhibition) failed to show activity when re-examined (F1, Fig. 3). Other than compound F1, all selected compounds designated as ‘hits’ (>50% inhibition) retained a high level of inhibitory activity upon re-evaluation. Taken together, the authors believe the inaccuracies stemming from photostability issues are limited to class A compounds; however, these inaccuracies would include all other inhibition data reported for class A compounds, including assay data against other enzyme targets, IC50 values, and thermal shift assay (TSA) binding data. Furthermore, the hit rate against each target is likely lower than reported, with PAN having an adjusted hit rate of ∼28% (20/71).Open in a separate windowFig. 3Chemical structures of representative metallofragments from each class re-examined for inhibition activity against PAN endonuclease.The authors maintain that three-dimensional metallofragments represent a useful new line of inquiry for FBDD and our ongoing studies seek to further test this hypothesis. The core message of our original study – the ability of metallofragments to be useful scaffolds for FBDD that occupy hard-to-access three-dimensional chemical space – remains unchanged. However, as demonstrated by our error, the authors acknowledge that metallofragments may pose unique challenges that must be carefully considered and controlled for when using them in FBDD campaigns.The authors would like to take this opportunity to thank the readers who alerted them to the concerns regarding the inhibitory activities and allowed them to reinvestigate. Both the authors and the Royal Society of Chemistry appreciate their support.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

15.
Reductive radical-initiated 1,2-C migration assisted by an azidyl group     
Xueying Zhang  Zhansong Zhang  Jin-Na Song  Zikun Wang 《Chemical science》2020,11(30):7921
We report here a novel reductive radical-polar crossover reaction that is a reductive radical-initiated 1,2-C migration of 2-azido allyl alcohols enabled by an azidyl group. The reaction tolerates diverse migrating groups, such as alkyl, alkenyl, and aryl groups, allowing access to n+1 ring expansion of small to large rings. The possibility of directly using propargyl alcohols in one-pot is also described. Mechanistic studies indicated that an azidyl group is a good leaving group and provides a driving force for the 1,2-C migration.

We report here a novel reductive radical-polar crossover reaction that is a reductive radical-initiated 1,2-C migration of 2-azido allyl alcohols enabled by an azidyl group.

Since the groups of Ryu and Sonoda described the reductive radical-polar crossover (RRPCO) concept in the 1990s,1 it has attracted considerable attention in modern organic synthesis.2 By using this concept, a variety of complex molecules could be assembled in a fast step-economic fashion which is not possible using either radical or polar chemistry alone. However, only two RRPCO reaction modes are known to date: nucleophilic addition and nucleophilic substitution (Fig. 1A). The first RRPCO reaction is the nucleophilic addition of organometallic species, which is generated in situ from the reduction of a strong reducing metal with a carbon-centered radical intermediate and cations (E+ = H+, I+, Br+, path 1).3 However, the necessity for a large amount of harmful and strong reducing metals has greatly limited the scope and functional group tolerance of the reaction. Recently, photoredox catalysis has not only successfully overcome the shortcomings of using toxic strong reducing metals in the RRPCO reaction,4 but also enabled the development of several new RRPCO reaction types, including the nucleophilic addition with carbonyl compounds or carbon dioxide (path 2),5 the cyclization of alkyl halides/tosylates (path 3),6 and β-fluorine elimination (path 4).7 Although the RRPCO reaction has been greatly advanced by photoredox catalysis, it is still in its infancy, and the development of a novel RRPCO reaction is of great importance.Open in a separate windowFig. 1(A) Reductive radical-polar crossover reactions; (B) this work: reductive radical-initiated 1,2-C migration assisted by an azidyl group.Herein, we wish to report a new type of reductive radical-polar crossover cascade reaction that is the reductive radical-initiated 1,2-C migration under metal-free conditions (Fig. 1B). The development of this approach is not only to further expand the application of the RRPCO reaction, but also to solve the problems associated with the oxidative radical-initiated 1,2-C migration, such as the necessity for an oxidant and/or transition metal for the oxidative termination of the radicals, and also required sufficient ring strain to avoid the generation of epoxy byproducts.8 To realize this reaction, a driving force is needed to drive the 1,2-C migration after reductive termination, to avoid the otherwise inevitable protonation of the generated anion.9 Inspired by the leaving group-induced semipinacol rearrangement,10 we envisaged that 2-azidoallyl alcohols11 might be the ideal substrates for the reductive radical-initiated 1,2-C migration because these compounds contain both an allylic alcohol motif, which is vital for the radical-initiated 1,2-C migration, and an azidyl group, a good leaving group,12 which may facilitate the 1,2-C migration after the reductive termination of the radicals.With the optimal conditions established (ESI, Table S1), we then explored the scope of this radical-initiated 1,2-migration. As shown in Table 1, a series of naphthenic allylic alcohols could undergo n+1 ring expansion with minimal impact on the product yield (Table 1, 3aa–aq). Notably, only the alkyl groups were migrated when using benzonaphthenic allylic alcohols in the reaction. These results might be attributed to the aryl group possessing greater steric resistance. The structure of 3an was further verified by single-crystal diffraction. Interestingly, the vinyl azide derived from a pharmaceutical ethisterone was also a viable substrate, affording the migration product 3aq in 57% yield, which highlighted the applicability of this strategy in the late-stage modification of pharmaceuticals. Moreover, the acyclic allylic alcohol with an alkyl chain also successfully delivered the migration product 3ar in 64% yield.Substrate scope of 2-azidoallyl alcoholsab
Open in a separate windowaStandard reaction conditions: 1 (0.5 mmol), TMSN3 (2.0 mmol), 2a (3.0 mmol) in H2O (0.7 mL) and DMSO (1.4 mL) at 50 °C in air for 48 h.bIsolated yields.Next, we extend the reaction scope to a range of aryl allylic alcohols. In comparison with alkyl allylic alcohols, aryl allylic alcohols gave the migration products in higher yields. The structure of 3ba was unambiguously confirmed by X-ray single crystal diffraction (CCDC 1897779). As demonstrated by the arene scope (Table 1, 3ba–bl), a variety of aryl allylic alcohols, including electron-withdrawing phenyl, electron-donating phenyl, polysubstituted phenyl, and fused rings, afforded the corresponding products in moderate to high yields (67–89%). Unsurprisingly, the substrates containing electron-donating groups afforded higher yields than those containing electron-withdrawing groups.Phenols and their derivatives are important structural constituents of numerous pharmaceuticals, agrochemicals, polymers, and natural products.13 The most common method for synthesising phenols is the hydroxylation of aryl halides.14 However, the method usually requires transition metals and harsh reaction conditions. Interestingly, by using the current strategy, inexpensive and abundant cyclopentadiene moieties can also be easily converted into phenols (Table 1, 3ca–cc) in moderate to good yield. Thus, this strategy provides metal-free and mild conditions for accessing phenols.Next, we investigated the migration capabilities of different groups (Table 2). When using a substrate that contains two different alkyl groups (1da), the product with the less sterically hindered alkyl group is obtained in a higher migration ratio. A comparison of aryl groups and alkyl groups in the same allylic alcohols showed that the migration of aryl groups was more facile, and the migration ratio ranged from 1 : 4 to 1 : 1.3 (3db–dd). The results of the migration ratio of different aryl groups (3de–dh) revealed that aryl moieties with electron-donating groups possessed higher migration ratios than aryl moieties with electron-withdrawing groups.Investigation of the migration efficiency
Entry 1 R1R2Yielda (%)
3d 3d′
1 1da Me t-Bu1542
2 1db MeC6H55326
3 1dc Me4-MeOC6H55614
4 1dd Me4-CF3C6H54232
5 1de C6H54-MeC6H54240
6 1df C6H54-MeOC6H54639
7 1dg C6H54-ClC6H54144
8 1dh C6H54-CF3C6H53648
Open in a separate windowaIsolated yields.After the evaluation of the scope of our allylic alcohols, we turned our attention to sulfonyl radical precursors (Table 3). We carried out the reaction of various sodium sulfinates with allylic alcohol 1ba under standard conditions. Pleasingly, the sodium sulfinates with straight chain alkyl (3ea), cyclic alkyl (3eb), and aryl (3ec–ef) groups were all suitable for this radical-initiated 1,2-carbon migration, and afforded corresponding products in 71–91% yield.Substrate scope of sodium sulfinatesa
Open in a separate windowaIsolated yields.In this work, the 2-azidoallyl alcohols substrates were derived from propargylic alcohols through a silver-catalyzed hydroazidation of alkynes.15 Consequently, we hypothesized that the radical-initiated 1,2-carbon migration could be directly achieved from propargylic alcohols in a one pot process. With a slight modification of the reaction conditions, we realized the one-pot preparation of the desired products from propargylic alcohols (Table 4). Propargylic alcohols containing cyclic alkyl (3ag and 3ah), heterocyclic alkyl (3ak and 3al), acyclic alkyl (3ar), and aryl (3ba) groups all gave the desired migration products, although the yields were slightly lower than those from the reactions of the 2-azidoallyl alcohols. It should be noted that the ring expansion products could be directly generated from a bioactive compound, ethisterone (3aq). Performing such a reaction in a single step could greatly reduce the cost of pharmaceutical modification. The fused phenol (3cd) could also be obtained in moderate yield via the one-step reaction. In addition, the migration order of the different substituted groups (3db) was nearly identical to that observed in vinyl azide-based protocol. Furthermore, alkyl sodium sulfinates (3ea) were also well tolerated.Substrate scope of propargyl alcoholsa,b
Open in a separate windowaStandard reaction conditions: 4 (0.5 mmol), TMSN3 (2.0 mmol), 2 (3.0 mmol), Ag2CO3 (0.05 mmol) in H2O (0.7 mL) and DMSO (1.4 mL) at 50 °C in air for 48 h.bIsolated yields.To gain more insight into the mechanism of radical-initiated 1,2-carbon migration, we conducted various experiments to confirm the presence or absence of radical and carbanion intermediates (Scheme 1). When the reaction of 1ba was performed in the presence of TEMPO (6.0 equiv.), the reaction was suppressed under the standard conditions (Scheme 1, eqn (1)), supporting the involvement of a radical intermediate. To prove the formation of a carbanion intermediate, we carried out two deuterium labeling experiments (Scheme 1, eqn (2) and (3)). The resulting products [d]-3ba and MA-1 contain the deuterium atom α in the carbonyl group, confirming the formation of a carbanion intermediate. To identify the key intermediate of the 1,2-migration, we prepared a potential intermediate M1 and subjected it to the standard conditions (Scheme 1, eqn (4)). But, the product 3ba was not observed and almost all of the M1 was recovered, which indicates that M1 is not a key intermediate. However, the product 3ba was obtained in a yield of 41% while M2 was subjected to the standard conditions (eqn (5)). If the hydroxyl group in the 2-azidoallyl alcohols was protected (M3), the reaction would not give the corresponding migration product (3ga), but generate product 5 with a yield of 51% (eqn (6)).11c These results proved that the reaction involved a 1,3-H migration process thereby enabling an oxygen anion intermediate IV (other mechanistic studies are discussed in ESI Fig. S1).Open in a separate windowScheme 1Mechanistic investigations.Based on the above experimental results and relevant literature, a possible reaction pathway was proposed as shown in Fig. 2. First, TolSO2TMS (I) is generated by the anion exchange of TolSO2Na with TMSN3. Such intermediates are known to be somewhat unstable,16 as similar to the analogous compounds, such as TolSO2I,17 and TMSTePh18 and thus undergo homolysis. Therefore, we anticipated that TolSO2TMS (I) should also yield sulfonyl and trimethylsilyl radicals.19 Then the 2-azidoallyl alcohol 1ba is readily attacked by the sulfonyl radical, leading to carbon-centered radical II. Subsequently, the carbon-centered radical II undergoes single electron transfer by the oxidation of sulfinate to the sulfonyl radical yielding the carbanion III.20 A 1,3-H shift of carbanion III affords the intermediate IV21 which rapidly undergoes 1,2-migration with the assistance of the azidyl leaving group, generating the desired product. It is worth noting that the present work is a novel radical reaction mode for vinyl azides compared to the existing reports that involve N–N bond breaking in the presence of radicals. Moreover, the development of this strategy is of great significance for the application of vinyl azides in the reconstruction of C–C bonds.Open in a separate windowFig. 2Proposed mechanism.On the other hand, the coupling of sulfonyl radicals produces intermediate V.22 The azidyl anion that is generated in the reaction is more prone to attack intermediate V to afford tosyl azide.23 Subsequently, tosyl azide is reduced to p-toluenesulfonamide by the trimethylsilyl radical.24 The sideproducts tosyl azide and p-toluenesulfonamide were isolated by column chromatography, and the associated TMSOH and TMS2O have been detected by GC-MS.25  相似文献   

16.
Direct synthesis of pentasubstituted pyrroles and hexasubstituted pyrrolines from propargyl sulfonylamides and allenamides     
Changqing Ye  Yihang Jiao  Mong-Feng Chiou  Yajun Li  Hongli Bao 《Chemical science》2021,12(26):9162
Multisubstituted pyrroles are important fragments that appear in many bioactive small molecule scaffolds. Efficient synthesis of multisubstituted pyrroles with different substituents from easily accessible starting materials is challenging. Herein, we describe a metal-free method for the preparation of pentasubstituted pyrroles and hexasubstituted pyrrolines with different substituents and a free amino group by a base-promoted cascade addition–cyclization of propargylamides or allenamides with trimethylsilyl cyanide. This method would complement previous methods and support expansion of the toolbox for the synthesis of valuable, but previously inaccessible, highly substituted pyrroles and pyrrolines. Mechanistic studies to elucidate the reaction pathway have been conducted.

This method is a toolbox for the synthesis of valuable, but previously inaccessible, highly substituted pyrroles and pyrrolines.

Pyrroles are molecules of great interest in a variety of compounds including pharmaceuticals, natural products and other materials. Pyrrole fragments for example are key motifs in bioactive natural molecules, forming the subunit of heme, chlorophyll and bile pigments, and are also found in many clinical drugs, including those in Fig. 1a.1 Although many classical methods of pyrrole synthesis, including the Paal–Knorr condensation,2 the Knorr reaction,3 the Hantzsch reaction,4 transition metal-catalyzed reactions,5 and multicomponent coupling reactions,6 have been developed over many years, the efficient synthesis of multisubstituted pyrroles is still challenging. In condensation syntheses of pyrroles, the major problems lie in the extended syntheses of complex precursors and limited substitution patterns that are allowed. Multicomponent reactions are superior when building pyrrole core structures with more substituents. Among these, the [2+2+1] cycloaddition reaction of alkynes and primary amines is attractive because of the readily available alkyne and amine substrates and the ability to construct fully substituted pyrroles.7 However, with the exception of some rare examples,8 most [2+2+1] cycloaddition reactions afford pyrroles with two or more identical substituents. The synthesis of multisubstituted pyrroles with all different substituents from simple starting materials therefore remains a major challenge.Open in a separate windowFig. 1Previous reports and this work on propargylamides transformation.Easily accessible propargylamides are classical, privileged building blocks broadly utilized for the synthesis of a large variety of heterocyclic molecules such as pyrroles, pyridines, thiazoles, oxazoles and other relevant organic frameworks.9 For example, Looper10et al. reported the synthesis of 2-aminoimidazoles from propargyl cyanamides and Eycken11 reported a method starting from propargyl guanidines which undergo a 5-exo-dig heterocyclization as shown in Fig. 1b. Subsequently, Wan12et al. revealed the cyclization of N-alkenyl propargyl sulfonamides into pyrroles via sulfonyl migration. Inspired by these transformations and multi-substituted pyrrole synthesis, we report herein a direct synthesis of pentasubstituted pyrroles and hexasubstituted pyrrolines with all different substituents from propargyl sulfonylamides and allenamides.Previously, Zhu,13 Ji14 and Qiu13b,15 reported efficient syntheses of 2-aminopyrroles from isocyanides. Ye16 and Huang17 independently developed gold-catalyzed syntheses of 2-amino-pentasubstituted pyrroles with ynamides. Despite the many advantages of these methods, they all afford protected amines, rather than free amines. The deprotection of these amines may cause problems in further transformations of the products. Our method delivers pyrroles with an unprotected free amino group and are often complementary to the previously well-developed classical methods.Initially, the cyclization reaction of N-(1,3-diphenylprop-2-yn-1-yl)-N-ethylbenzenesulfonamide (1a) with trimethylsilyl cyanide (TMSCN) was carried out with Ni(PPh3)2Cl2 as a catalyst, a base (Cs2CO3) and DMF as a solvent. Different metal catalysts, such as Ni(PPh3)2Cl2, Pd(OAc)2, Cu(OAc)2, and Co(OAc)2 provided the desired product with similar yields ( EntryCat.BaseSolventYield1Ni(PPh3)2Cl2Cs2CO3DMF67%2Pd(OAc)2Cs2CO3DMF65%3Cu(OAc)2Cs2CO3DMF65%4Co(OAc)2Cs2CO3DMF63%5Cs2CO3DMF65%6KFDMFTrace7K3PO4DMFTrace8K2CO3DMF48%9KOHDMF52%10KOtBuDMF46%11Et3NDMFTrace12Cs2CO3CH3CN18%13Cs2CO3DME23%14Cs2CO3TolueneTrace15Cs2CO3DCETrace16Cs2CO3DioxaneTraceOpen in a separate windowaReaction conditions: 1a (0.1 mmol, 1 equiv.), TMSCN (0.3 mmol, 3 equiv.), cat. (0 or 10 mol%), base (0.3 mmol, 3 equiv.) and solvent (1 mL), at 80 °C for 10 h; isolated yield.With the optimal reaction conditions in hand, we investigated the scope of this reaction. As shown in Fig. 2, the transformation tolerates a broad variety of substituted propargylamides (1). The R1 group could be an aryl group containing either electron-donating groups or electron-withdrawing groups, and the corresponding products (2b–2h) were obtained in yields of 62–80%. The substituent R1 could also be an alkyl group such as 1-hexyl in which case the reaction provided the corresponding pyrrole (2i) in 53% yield. Exploration of the R2 substituent was also conducted. Electron-rich and electron-deficient substituents in the aromatic ring of R2 gave the desired products (2j–2o) with yields of 70–81%. The product bearing a furyl group (2p) can be produced in 61% yield. However, when R2 group is an aliphatic group, the reaction failed to provide the desired product. Substituent groups R3, such as benzyl (2q) or 3,4-dimethoxyphenylethyl (2r) were also compatible in the reaction, providing the corresponding products in moderate yields. Significantly, this method has the potential to produce core structures (for example 2s) similar to that in Atorvastatin. Interestingly, when alkynyl substituted isoquinolines (1t–1v) were used as the substrates, the reactions smoothly afforded fused pyrrolo[2,1-α]isoquinoline derivatives (2t–2v), members of a class of compounds that are found widely in marine alkaloids and exhibit anticancer and antiviral activity.18Open in a separate windowFig. 2Substrate scope of propargylamides. Reaction conditions: 1 (0.20 mmol, 1 equiv.), TMSCN (0.60 mmol, 3 equiv.), Cs2CO3 (0.60 mmol, 3 equiv.) and DMF (2 mL), at 80 °C for 10 h; isolated yield.Allenes are key intermediates in the synthesis of many complex molecules.19 As a subtype of allenes, allenamines are also useful as reaction intermediates.20 Although the transformation of allenamides to multisubstituted pyrroles has not been previously recorded, this reaction probably goes through the allenamide intermediates which can be derived from propargyl sulfonamides under basic conditions. To verify this hypothesis, the trisubstituted allenamide (3) was synthesized and subjected to the standard reaction conditions. A pyrrole (2a) was isolated in 82% yield from this reaction (Fig. 3). This result confirmed our assumption and raised a new question: is it possible to build hexasubstituted pyrrolines from tetrasubstituted allenamides? A range of tetrasubstituted allenamides21 was tested under the standard reaction conditions, and the hexasubstituted pyrrolines were obtained as is shown in Fig. 4. The R1 group could be an aryl substituent or an alkyl chain, and the corresponding products (5a–5e) were obtained with good yields. Various aryl groups with either electron-donating groups or electron-withdrawing groups in the aromatic ring of R2 provided the desired products (5f–5k) in 62–83% yields. In addition, the difluoromethyl group can also be replaced by a phenyl group, and the reaction provided the corresponding product 5l in 82% yield. It is worth noting that these pyrroline products are not easily accessible from other methods.Open in a separate windowFig. 3Synthesis of substituted pyrroles from allenes.Open in a separate windowFig. 4Substrate scope of tetrasubstituted allenamides. Reaction conditions: 4 (0.10 mmol, 1 equiv.), TMSCN (0.30 mmol, 3 equiv.), K2CO3 (0.30 mmol, 3 equiv.) and DMF (1 mL), at 80 °C for 10 h, isolated yield.Some synthetic applications of this method are shown in Fig. 5. The amide is a naturally occurring and ubiquitous functional group. When using benzoyl chloride to protect the free amino group of the fully-substituted pyrrole (2a), a bis-dibenzoyl amide (6) was obtained in the presence of a base, triethylamine while the monobenzoyl protected amide (7) was obtained in the presence of pyridine as the base (Fig. 5a). This method also provides a straightforward approach to pyrrole fused lactam structures (Fig. 5b). For examples, a five-membered lactam and a six-membered lactam were generated separately in a one pot reaction, directly from, (8 and 10), respectively. Taking advantage of this method, an analogue of the drug Atorvastatin was synthesized in 5 steps (Fig. 5c), demonstrating the synthetic value of the reaction.Open in a separate windowFig. 5Synthetic applications.Mechanistic experiments were performed (Fig. 6) to explore the mechanism of the reaction. When 3 equivalents of TEMPO were added, the reaction was not inhibited and the desired product (2a) was formed in 62% yield (Fig. 6a). This result suggested that the reaction might not involve a radical process. To probe the reaction further, a kinetic study was conducted (Fig. 6b). According to this study, the propargylamide (1a) was completely converted to an allenamide (3a) in 10 min under the standard conditions. The multi-substituted pyrrole (2a) was then gradually produced from the intermediate allenamide and no other reaction intermediates were observed or identified. On the other hand, DFT calculations of substrates 3b and 4a were carried out at the B3LYP-D3(SMD)/Def2-TZVP//B3LYP-D3/Def2-SVP level of theory to identify the natural bond orbital (NBO) charges on the carbons of the allene moieties. NBO charges on the internal carbon in both 3b and 4a are 0.11 and 0.18, respectively (Fig. 6c) indicating that the nucleophilic addition of cyanide anion onto the internal carbon should be reasonable as opposed to its addition onto the terminal carbon. Pathways of the cyano addition to 3b were also calculated (Fig. 6d). The transition state of cyano addition on the internal carbon (TS1), is indeed much lower than addition on the terminal carbon (TS2). The intermediate of internal carbon addition int1, is more stable than int2, implying that the internal carbon addition pathway is not only kinetically but also thermodynamically favoured.Open in a separate windowFig. 6Mechanistic studies and proposed mechanism.Based on the results of these mechanistic studies, a plausible reaction mechanism for the synthesis of pentasubstituted pyrroles and hexasubstituted pyrrolines is proposed and is shown in Fig. 6e. First, under basic conditions, the propargylamide isomerizes to an intermediate allenamide (A), which can be attacked nucleophilically by the cyanide anion to afford an intermediate imine (B) with release of the sulfonyl group. Then, the second cyanide anion attacks the imine to form an intermediate (C), which can undergo cyclization and protonation to afford the fully substituted pyrrole (2). Similarly, the hexasubstituted pyrroline product (5) can be obtained from double nucleophilic attack of the intermediate (A) by the cyanide ion.  相似文献   

17.
Palladium/GF-Phos-catalyzed asymmetric carbenylative amination to access chiral pyrrolidines and piperidines     
Yue Sun  Chun Ma  Zhiming Li  Junliang Zhang 《Chemical science》2022,13(37):11150
The cross-coupling of N-tosylhydrazones has emerged as a powerful method for the construction of structurally diverse molecules, but the development of catalytic enantioselective versions still poses considerable challenges and only very limited examples have been reported. We herein report an asymmetric palladium/GF-Phos-catalyzed carbenylative amination reaction of N-tosylhydrazones and (E)-vinyl iodides pendent with amine, which allows facile access to a range of chiral pyrrolidines and piperidines in good yields (45–93%) with up to 96.5 : 3.5 er. Moreover, mild conditions, general substrate scope, scaled-up preparation, as well as the efficient synthesis of natural product (−)-norruspoline are practical features of this method.

An efficient asymmetric palladium/GF-Phos-catalyzed carbenylative amination reaction to access structurally diverse chiral pyrrolidines and piperidines in good yields with high chemo-, regio- and enantioselectivities has been developed.

N-tosylhydrazones, readily prepared from aldehydes or ketones, served as a safe source of carbene precursors and have attracted much attention of chemists.1N-tosylhydrazone-mediated applications have been continuously developed, such as cyclopropanation or cyclopropenation, X–H insertion, ylide formation, cycloaddition, aza-Wacker-type cyclization, asymmetric allylic substitution, etc.2 Among them, transition-metal-catalyzed cross-coupling is one of the powerful protocols for C–X or C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond formation in organic synthesis involving versatile intermediates, of which in situ generation of diazo compounds and carbene migratory insertion are considered key steps.3–5 Over the past decades, considerable progress has been made in the asymmetric cross-coupling reactions of N-tosylhydrazones with various coupling partners, including cyclobutanols, terminal alkynes, silacyclobutanes and so on.4 Relatively, only a few examples focus on the cross-coupling reactions of aryl halides with N-tosylhydrazones involving benzyl metal intermediates [Scheme 1A, eqn. (a)].6 For example, Gu,6a Wu,6b Lassaletta6c and coworkers have developed a palladium-catalyzed asymmetric synthesis of axial chiral compounds from aryl bromides and N-tosylhydrazones, ending with β-H elimination. Very recently, we realized palladium/GF-Phos catalyzed asymmetric three component cross-coupling reactions of aryl halides, N-tosylhydrazones, with terminal alkynes.6f In contrast, much less progress has been made in N-tosylhydrazone-based carbenylative insertions from vinyl halides, which would generate a π-allylic metal intermediate followed by nucleophile attack, providing a unique approach for building C–X bonds, especially for N-heterocyclic compounds [Scheme 1A, eqn. (b)].7N-heterocycles are important structural motifs for the development of various types of valuable chemicals and materials.8 Importantly, optically active 2-substituted pyrrolidine and piperidine derivatives are privileged scaffolds in many natural products and pharmaceuticals with a wide range of biological activities,9 as well as the backbone of organocatalysts in asymmetric catalysis (Fig. 1).10Open in a separate windowScheme 1Asymmetric transition-metal-catalyzed carbenylative cross-coupling reactions.Open in a separate windowFig. 1Selected natural products and pharmaceuticals containing chiral 2-substituted pyrrolidine and piperidine units.Notably, Van Vranken and coworkers reported an elegant palladium-catalyzed carbenylative amination reaction of N-tosylhydrazones and (E)-vinyl iodides pendent with amine, providing facile access to pyrrolidine and piperidine ring systems that are common to alkaloid natural products (Scheme 1B).11 Unfortunately, only up to 58.5 : 41.5 er was obtained after they made a lot of efforts to screen a series of chiral phosphine ligands, indicating that this asymmetric reaction indeed poses considerable challenges in addition to competitive side reactions such as the dimerization of vinyl iodides,12 the formation of diene via the palladatropic rearrangement/β-H elimination or allene via β-H elimination from Csp2,13 and the π-allylpalladium intermediate trapped by the byproduct sulfinic acid salt.14 Given the significance of chiral pyrrolidines and piperidines as core structures in alkaloid natural products, the development of an asymmetric version of this elegant carbenylative amination reaction is highly desirable. In recent years, our group has developed a series of chiral sulfinamide phosphine ligands (so-called Sadphos), which showed unique potential in asymmetric transition-metal catalysis,6f,15 so we wondered whether Sadphos could address this challenging asymmetric carbenylative amination reaction (Scheme 1C).Initially, our study began with (E)-vinyl iodide 1a and N-tosylhydrazone 2a in the presence of Pd2(dba)3, t-BuOLi, Et3N, and triethylbenzylammonium chloride (TEBAC) in THF at 30 °C. A series of commercially available chiral ligands were first screened (Fig. 2). Only (R, R)-DIOP (L1), (R)-DTBM-SegPhos (L3) and (R)-MOP (L4) provided the desired product 3aa with poor enantioselectivity and other ligands such as (R, R)-Ph-BPE (L2), (R, S)-Josiphos (L5) and (S, S)-iPr-FOXAP (L6) showed low reactivity. We next turned to systematically investigate Sadphos, such as Wei-Phos,16 Xiao-Phos,15d,17 Ming-Phos,15a,18 Xu-Phos,15b,19 Xiang-Phos20 and PC-Phos15c,21 (Fig. 2). To our delight, PC1 delivered 3aa in 32% yield and 85.5 : 14.5 er. Inspired by this result, we further screened PC2–PC5 which vary in the substituent of phenyl, but unfortunately none of them showed better results. Surprisingly, the reactivity of this reaction could be greatly improved with our recently developed GF-Phos GF1, delivering 71% yield. When steric hindered tert-butyl groups were introduced on the phenyl group (GF2), the product 3aa was obtained in 77% yield with 91.5 : 8.5 er. After screening different palladium catalysts and solvents (Open in a separate windowFig. 2Screened chiral ligands.Optimization of reaction conditionsa
Entry[Pd]BaseSolventYieldb (%) er c
1Pd2(dba)3Et3NTHF7791.5 : 8.5
2Pd(acac)2Et3NTHF8986.5 : 13.5
3Pd(OAc)2Et3NTHF8288 : 15
4PdBr2Et3NTHF7888 : 12
5Pd2(dba)3·CHCl3Et3NTHF7592 : 8
6Pd2(dba)3·CHCl3Et3NToluene2392.5 : 7.5
7Pd2(dba)3·CHCl3Et3NDMF9080 : 20
8Pd2(dba)3·CHCl3Et3NMTBE2893 : 7
9Pd2(dba)3·CHCl3Et3N1,4-Dioxane3888.5 : 11.5
10Pd2(dba)3·CHCl3Et3N2-Me-THF8993 : 7
11dPd2(dba)3·CHCl3Et3N2-Me-THF2694.5 : 5.5
12Pd2(dba)3·CHCl3DABCO2-Me-THF7694 : 6
13Pd2(dba)3·CHCl3Cs2CO32-Me-THF9392.5 : 7.5
14Pd2(dba)3·CHCl3KOH2-Me-THF8993 : 7
15Pd2(dba)3·CHCl3None2-Me-THF8393 : 7
16ePd2(dba)3·CHCl3None2-Me-THF6988 : 12
17fPd2(dba)3·CHCl3None2-Me-THF8194.5 : 5.5
Open in a separate windowaReaction conditions: 1a (0.1 mmol), 2a (0.16 mmol), [Pd] (5 mol%), GF2 (15 mol%), t-BuOLi (2.2 equiv.), TEBAC (1.0 equiv.), base (2.0 equiv.) in 0.1 M solvent at 30 °C for 12 h.bDetermined by GC analysis with n-tetradecane as an internal standard.cThe er value was determined by chiral HPLC.d15 °C for 12 h.eWithout TEBAC.f15 mol% Ag2CO3. THF = tetrahydrofuran. MTBE = tert-butyl methyl ether. DMF = N,N-dimethylformamide. DCE = 1,2-dichloroethane. DMSO = dimethyl sulfoxide.We also found that, besides t-BuOLi, there was little effect on the yield or enantioselectivity by changing another base. The study was therefore continued without it ().The scope of the carbenylative amination reaction was then studied using the optimized reaction conditions (22 Multisubstituted phenyl and naphthyl groups were also well-tolerated (3am, 3an, 3ap–3as). It is note-worthy that the 2,4,6-trimethylphenyl-substituted substrate delivered 3ao in 57% yield with 7/1 E/Z selectivity, probably due in part to the steric hindrance. Moreover, N-tosylhydrazones containing heterocycles reacted smoothly to furnish the expected products 3at–3aw. Besides diverse substituted N-tosylhydrazones 2, various kinds of vinyl iodide derivatives 1 with functional groups such as halides, methyl, tert-butyl, methoxy and 1-naphthyl at different positions on the phenyl ring also worked well and afforded 3ba–3ja in good yields. Surprisingly, when the protective group on the nitrogen atom was replaced by a p-toluenesulfonyl or p-nitrophenylsulfonyl group, the corresponding cyclic products 3ka, 3lx, and 3ly were successfully produced in high yields and enantioselectivities.Scope for enantioselective formation of pyrrolidinesa
Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2 (0.48 mmol), Pd2(dba)3·CHCl3 (2.5 mol%). GF2 (15 mol%), t-BuOLi (2.2 equiv.), TEBAC (1.0 equiv.), Ag2CO3 (15 mol%) in 0.1 M 2-MeTHF at 30 °C for 6 h.b1.8 mmol scale, 24 h.c2.0 mmol scale, 20 h.Subsequently, we further turned our efforts to the synthesis of piperidine derivatives. As shown in Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2 (0.48 mmol), Pd2(dba)3·CHCl3 (2.5 mol%), GF2 (15 mol%). t-BuOLi (2.2 equiv.), TEBAC (1.0 equiv.), Ag2CO3 (15 mol%) in 0.1 M 2-MeTHF at 30 °C for 6 h.b12 h.To evaluate the synthetic utility of this asymmetric carbenylative amination reaction, we carried out a gram–scale reaction under standard conditions, providing the product 3aj in 85% yield with 95.5 : 4.5 er (Scheme 2a). Of note, a 2-step deprotection of 3lx with p-toluenethiol/K2CO3 and HCl (1 M) enabled the synthesis of natural product (−)-norruspoline in 51% overall yield. Additionally, replacing the protecting group of 3ly with the Boc group afforded 6 in 67% yield without the loss of enantioselectivity and it has been previously shown that 6 is a synthetic intermediate for the preparation of natural product (−)-indolizidine 201 (Scheme 2b).23 A linear relationship was demonstrated by a nonlinear effect study on the ee value of GF2 and product 3aa, which implied that the catalytically active structure contains only a single chiral ligand. (please find more details in the ESI).Open in a separate windowScheme 2Gram-scale synthesis and synthetic applications.Based on our study and previous work,24 a catalytic cycle pathway to rationalize the synthesis of chiral pyrrolidines is illustrated in Scheme 3. First, the oxidative addition of vinyl iodide 1a to a Pd0/GF-Phos complex would generate vinyl PdII species A. In the presence of a base, N-tosylhydrazone 2ain situ generated a diazo intermediate and formed palladium carbene B with vinyl PdII species A, followed by migratory insertion to generate the π-allylpalladium intermediate C, as displayed in path a. Alternatively, the reaction proceeds in a palladium carbene/oxidative addition sequence as in path b. Next, the nucleophilic attack of the nitrogen atom on π-allylpalladium delivered product 3aa and regenerated the Pd0 complex, thus completing the entire catalytic cycle. In light of the structure of the chiral ligand GF2 and the absolute configuration of product (S)-3, a chirality induction model for stereochemical induction was proposed (Fig. 3).Open in a separate windowScheme 3Proposed catalytic cycle.Open in a separate windowFig. 3Proposed chirality induction model.In conclusion, we have developed a palladium/GF-Phos catalyzed asymmetric carbenylative amination of (E)-vinyl iodides with N-tosylhydrazones via a carbene migratory insertion/Tsuji-Trost sequence to build C–N/C–C more efficiently. This catalytic system exhibits general functional group tolerance and enables rapid access to a variety of chiral 2-substituted pyrrolidines and piperidines in moderate to good yields with high chemo-, regio-, enantioselectivities under mild conditions. Our approach can be applied to the direct synthesis of significant natural product (−)-norruspoline and provides an alternative route for the formal synthesis of (−)-indolizidine 201.  相似文献   

18.
Rh(iii)-catalyzed tandem annulative redox-neutral arylation/amidation of aromatic tethered alkenes     
Chao Chen  Chen Shi  Yaxi Yang  Bing Zhou 《Chemical science》2020,11(44):12124
Transition-metal-catalyzed directed C–H functionalization has emerged as a powerful and straightforward tool to construct C–C bonds and C–N bonds. Among these processes, the intramolecular annulative alkene hydroarylation reaction has received much attention because this intramolecular annulation can produce more complex and high value-added structural motifs found in numerous natural products and bioactive molecules. Despite remarkable progress, these annulative protocols developed to date remain limited to hydroarylation and functionalization of one side of alkenes, thus largely limiting the structural diversity and complexity. Herein, we developed a rhodium(iii)-catalyzed tandem annulative arylation/amidation reaction of aromatic tethered alkenes to deliver a variety of 2,3-dihydro-3-benzofuranmethanamine derivatives bearing an all-carbon quaternary stereo center by employing 3-substituted 1,4,2-dioxazol-5-ones as an amidating reagent to capture the transient C(sp3)–Rh intermediate. Notably, by simply changing the directing group, a second, unsymmetrical ortho C–H amidation/annulation can be achieved to provide tricyclic dihydrofuro[3,2-f]quinazolinones in good yields.

A rhodium(iii)-catalyzed tandem annulative arylation/amidation reaction of aromatic tethered alkenes was developed to deliver a variety of 2,3-dihydro-3-benzofuranmethanamine derivatives.

Transition-metal-catalyzed C–H functionalization for the direct conversion of C–H bonds to C–C bonds and C–N bonds has evolved into a widespread and effective strategy for fine chemical production.1 Among these processes, the hydroarylation of C–C double bonds via a C–H addition has been well-established and become an effective strategy to access synthetically useful structural motifs.1 Recently, this alkene hydroarylation reaction has received much attention in an intramolecular fashion2 (Scheme 1a) because this intramolecular annulation can produce more complex and high value-added structural motifs found in numerous natural products and bioactive molecules (Fig. 1).3 Despite remarkable progress in this area, most of the annulative protocols developed to date remain limited to one-component intramolecular alkene hydroarylation and functionalization of one side of alkenes. More challenging C–H arylation of intramolecular alkenes followed by a tandem coupling with a different coupling partner have unfortunately proven elusive thus far, thus largely limiting the structural diversity and complexity.Open in a separate windowFig. 1Representative bioactive 2,3-dihydrobenzofurans.Open in a separate windowScheme 1Transition-metal-catalyzed C–H functionalization to construct the C–C and C–N bonds.On the other hand, nitrogen-containing molecules have gained great attention due to their widespread presence in natural products and widespread use in pharmaceutical science.4 During the last two decades, transition-metal-catalyzed direct C(sp2)–H amination/amidation assisted by chelating directing group is a well-established strategy.5 Recently, several examples of C(sp3)–H amination/amidation have also been reported for the efficient installation of C–N bonds.6,7 Mechanistically, the reaction is initiated by a chelation-assisted C–H metalation to form a C(sp3)–M species, which is then coupled with amination reagents to construct the C–N bonds.In this context, we wondered if a catalytic annulative C–H arylation of a O-bearing olefin-tethered arenes might be possible, thus leading to a C(sp3)–M intermediate, which upon capture with a amidation reagent to construct a new C–N bond and provide bioactive 2,3-dihydro-3-benzofuranmethanamine derivatives. Inherently, the tandem annulative 1,2-arylation/amidation of alkenes has several challenges. First, the resulting C(alkyl)–M intermediate is liable to undergo protonation to provide the alkene hydroarylation products.1,2 Moreover, a potential competing β-H elimination of the resulting C(alkyl)–M intermediate also required to be suppressed. In addition, compared with the C(sp2)–M species, the resulting C(alkyl)–M species is relatively unstable and also has a low reactivity.To address these challenges and with our continuing interest in the Rh(iii)-catalyzed C–H functionalization,8 we introduced a Weinreb amide as a directing group and 3-substituted 1,4,2-dioxazol-5-ones as the amide sources9 to trigger a new tandem annulative 1,2-arylation/amidation of alkenes via a Rh(iii)-catalyzed C–H activation,10 providing a variety of synthetically challenging 2,3-dihydro-3-benzofuranmethanamine derivatives bearing an all-carbon quaternary stereo center (Scheme 1b). More importantly, through simply changing the directing group, a second, unsymmetrical ortho C–H amidation/annulation could be realized to provide tricyclic dihydrofuro[3,2-f]quinazolinone derivatives. This protocol provides a good complement to previously reported carboamination reactions.11To begin our studies, Weinreb amide 1a was reacted with methyl dioxazolone 3a in the presence of various catalyst and AgSbF6 at 70 °C in DCE (Table 1, entries 1–4). The use of [Cp*RhCl2]2 as the catalyst was found to be crucial to give the desired tandem annulative product 4a, with other catalysts, such as [Ru(p-cymene)Cl2]2, [Cp*IrCl2]2, and Cp*Co(CO)I2, resulting in no desired product. Attempt to increase or lower the reaction temperature led to a slightly low yield (entries 5 and 6). Interestingly, when employing a NH–OMe amide 2a as the substrate and using 3 equivalent of 3a, a second, unsymmetrical ortho C–H amidation/annulation was achieved to provide the tricyclic dihydrofuro[3,2-f]quinazolinone 5a in 50% yield (entry 7). A screen of additives (entries 8–10) identified LiOAc as the optimal additive, affording the desired product 5a in 93% yield (entry 10). The Rh(iii) catalyst was found to be crucial for this tandem annulative arylation/amidation reaction, with no reactivity in its absence (entries 11 and 12).Optimization of reaction conditionsa
EntryXCatalyst (5 mol%)Additive (20 mol%)Solvent T (°C)Yield of 4a (%)Yield of 5a (%)
1Me[Cp*RhCl2]2DCE70890
2Me[Ru(p-cymene)Cl2]2DCE7000
3Me[Cp*IrCl2]2DCE7000
4MeCp*Co(CO)I2DCE7000
5Me[Cp*RhCl2]2DCE90690
6Me[Cp*RhCl2]2DCE50700
7bH[Cp*RhCl2]2DCE70050
8bH[Cp*RhCl2]2Cu(OAc)2DCE70087
9bH[Cp*RhCl2]2KOAcDCE70086
10bH[Cp*RhCl2]2LiOAcDCE70093
11bHLiOAcDCE7000
12MeDCE7000
Open in a separate windowaConditions: 1a (0.1 mmol), 3a (0.12 mmol), catalyst (5 mol%), AgSbF6 (20 mol%) and additive (20 mol%) in DCE (1 mL) for 12 h. Yield isolated by column chromatography.bConditions: 2a (0.1 mmol), 3a (0.3 mmol), catalyst (5 mol%), AgSbF6 (20 mol%), additive (20 mol%) in DCE (1 mL) for 12 h. Yield isolated by column chromatography.Having determined the optimal reaction conditions, we sought to evaluate the substrate scope (Scheme 2). First, the amidation reagents were explored and 1,4,2-dioxazol-5-ones substituted with primary alkyl (4a and 4e), secondary alky (4b, 4c and 4f), tertiary alkyl (4d) and aryl group (4g–j) all coupled smoothly with 1a, providing the 2,3-dihydro-3-benzofuranmethanamines 4a–4j in good yields. The structure of 4f was unambiguously confirmed by an X-ray crystallographic analysis (CCDC 2015893). The scope with regards to the arene moiety was then examined. The substrates 1 containing either electron-donating or electron-withdrawing substituents at different positions on the arene ring were well tolerated and provide the desired products 4k–t in good yields. We were pleased that 2-naphthalenecarboxamide effectively underwent this tandem annulative 1,2-arylation/amidation reaction, affording the desired product 4r in good yield. Notably, the various substituted allyl groups such as ethyl, cyclopentyl, phenyl, and phenoxymethyl groups were found to be compatible with the reaction conditions (4u–x). In addition, 3-N-tethered and 3-S-tethered substrates failed to give the desired tandem annulative products 4z and 4za.Open in a separate windowScheme 2Substrate scope of tandem annulative arylation/amidation reaction of aromatic tethered alkenes. Conditions: 1 (0.1 mmol), 3 (0.12 mmol), [Cp*RhCl2]2 (5 mol%), AgSbF6 (20 mol%) in DCE (1 mL) at 70 °C for 12 h. Yield isolated by column chromatography.Next, we proceeded to explore the scope of this unsymmetrical twofold C–H functionalization reaction (Scheme 3). Under the optimal reaction conditions, amidating reagents bearing alkyl or aryl groups are fully tolerated, affording the tricyclic dihydrofuro[3,2-f]quinazolinones 5a–5h in good yields. The structure of 5e was unambiguously confirmed by an X-ray crystallographic analysis (CCDC 2014245). Electronic and steric modification of the aryl group was also tolerated. Both electron-deficient (5j, 5m–o) and electron-rich (5i, 5k, 5l, 5q and 5r) substrates gave the corresponding tricyclic systems in good yields. Meta and para substitutions of a methyl group were also tolerated and delivered the products 5i and 5k, indicating a high tolerance for steric hindrance. Interestingly, when 2-naphthalenecarboxamide was used, a third C–H amidation of naphthalene ring took place, affording the product 5s in 65% yield. Notably, the current method effectively resulted in the ethyl-, cyclopentyl, phenyl, and phenoxymethyl-substituted products 5t–w bearing an all-carbon quaternary stereo center in good yield, respectively.Open in a separate windowScheme 3Substrate scope of unsymmetrical twofold C–H functionalization reaction. Conditions: 2 (0.1 mmol), 3 (0.30 mmol), [Cp*RhCl2]2 (5 mol%), AgSbF6 (20 mol%), LiOAc (20 mol%) and DCE (1 mL) at 70 °C for 12 h. Yield isolated by column chromatography.To check the practicability of this protocol, this two procedures could be readily scaled up with comparable efficiency in the presence of 2.5 mol% of Rh(iii) catalyst on a 2.0 mmol scale (eqn (1) and (2)). The product 5a could be readily converted into potential useful intermediates, such as amines 6a and free amino quinazolinone analog 6b, respectively (eqn (3)).To gain insight into the reaction mechanism, hydrogen/deuterium (H/D) exchange were carried out. A H/D exchange at the ortho-position of the amide group in the re-isolated 1a and 2a was observed in the absence or presence of 3a, indicative of the reversibility of the ortho C–H activation (eqn (4)–(6)). Treatment of 2a with 1 equivalent of 3d at room temperature for 1 h delivered the 6d as the sole product, indicating that the intramolecular tandem annulative 1,2-arylation/amidation of alkenes is faster than ortho C–H amidation/annulation (eqn (7)). In addition, the use of 3 equivalent of 3d at 70 °C for 2 h provided 6e as the main product and subsequent treatment of 6e under the standard conditions gave 5d in 60% yield (eqn (8)), indicating that the second ortho C–H amidation occurs first, followed by an intramolecular dehydration to give the desired quinazolinone product. Finally, treatment of substrate 7 with 3a under the standard reaction conditions did not give any product 8, ruling out the possibility of the insertion of a nitrene to double bond (eqn (9)).Based on above-mentioned experimental results, a plausible reaction pathway is proposed in Scheme 4. [Cp*RhCl2]2 precursor reacts with AgSbF6 to form an active cationic Rh(iii) species, which undergoes a C–H bond activation to form cyclometalated complex Int-A. Coordination of the tethered olefin and a subsequent migratory insertion affords the intermediate Int-B, which undergoes an oxidative addition into the N–O bond of 3a, followed by a CO2 extrusion, to provide the Rh(v) nitrenoid species Int-C. Reductive elimination occurs to deliver the intermediate Int-D which then is protonated to release product 4a or Int-E and regenerate the catalyst. Int-E can undergo a second ortho C–H activation to give Int-F, which can be oxidized by 3a again to afford the Rh(v) nitrenoid species Int-G, with a CO2 extrusion. Subsequent reductive elimination and protonation give the Int-I which undergoes an intramolecular dehydration to deliver the product 5a.Open in a separate windowScheme 4Proposed reaction mechanism.  相似文献   

19.
Intermolecular oxidative amination of unactivated alkenes by dual photoredox and copper catalysis     
Xiangli Yi  Xile Hu 《Chemical science》2021,12(5):1901
  相似文献   

20.
Benzannulation of isobenzopyryliums with electron-rich alkynes: a modular access to β-functionalized naphthalenes     
An Wu  Hui Qian  Wanxiang Zhao  Jianwei Sun 《Chemical science》2020,11(30):7957
Described here is a modular strategy for the rapid synthesis of β-functionalized electron-rich naphthalenes, a family of valuable molecules lacking general access previously. Our approach employs an intermolecular benzannulation of in situ generated isobenzopyrylium ions with various electron-rich alkynes, which were not well utilized for this type of reaction before. These reactions not only feature a broad scope, complete regioselectivity, and mild conditions, but also exhibit unusual product divergence depending on the substrate substitution pattern. This divergence allows further expansion of the product diversity. Control experiments provided preliminary insights into the reaction mechanism.

A substituent-controlled divergent benzannulation provides rapid access to various β-functionalized naphthalenes from electron-rich alkynes.

Functionalized naphthalenes are important structural motifs widely present in bioactive natural products, pharmaceuticals and functional materials.1,2 They also serve as valuable intermediates in organic synthesis.3 Among them, naphthalenes bearing an electron-donating group (EDG) at the β-position (e.g., β-naphthols, β-naphthylamines, and β-naphthyl thioethers) are particularly versatile (Fig. 1).2,3 For example, BINOL, which is derived from β-naphthol, is an extensively utilized synthetic precursor toward a wide range of privileged chiral ligands and catalysts.3 While various approaches have been developed for the synthesis of naphthalenes, efficient and selective de novo strategies toward these β-functionalized ones still remain in high demand. Moreover, to the best of our knowledge, a general and modular approach for rapid access to all these electron-rich β-functionalized naphthalenes remains unknown.4Open in a separate windowFig. 1Useful β-naphthol, β-naphthylamine, and β-naphthyl thioether units.Isobenzopyrylium ions are readily accessible and versatile intermediates in organic synthesis (Scheme 1).5 They are known to participate in cycloaddition reactions with diverse carbon–carbon double bonds or triple bonds for the synthesis of naphthalenes.5,6 While extensive progress has been achieved in this topic, challenges still remain to be addressed. For example, the majority of these benzannulations with alkynes have to be executed at high temperature, except those intramolecular cases or with stoichiometric activators. Moreover, electron-rich alkynes have not been well explored as reaction partners for the benzannulations with isobenzopyrylium ions. Nevertheless, such reactions would provide expedient access to the valuable β-naphthol and β-napthylamine derivatives with diverse substitution patterns (Scheme 1). In this context, here we report our effort in achieving a general and modular strategy toward these electron-rich naphthalenes with high efficiency and regioselectivity under mild conditions. We also observed interesting selectivity divergence controlled by the substituent of the isobenzopyrylium substrates, giving rise to different types of naphthalene products (e.g., 2-hydroxy-1-naphthaldehydes, Scheme 1).Open in a separate windowScheme 1Benzannulation of isobenzopyryliums with electron-rich alkynes.We began our study with isochromene 1a as the isobenzopyrylium precursor. Siloxy alkyne 2a was first employed as the model electron-rich alkyne in view of the extraordinary versatility of this type of alkyne in various cyclization reactions, including benzannulation.7–9 Various Brønsted and Lewis acids were evaluated as catalysts for this reaction. Unfortunately, most of them did not show obvious catalytic activity toward the formation of a naphthalene product (Table 1). These reactions either had little conversion or resulted in a mixture of undesired products. Nevertheless, further screening led to the identification of AgOTf and BF3·OEt2 as capable catalysts (entries 5–6), with the latter being superior, leading to the formation of naphthalene 3a as the major product (75% yield, entry 6). Increasing the catalyst loading to 20 mol% further improved the product yield to 85% (with an isolated yield of 81%, entry 7). Further increasing the catalyst loading proved to be not beneficial (entry 8). Notably, the use of substrate 1a′ bearing an ethoxy leaving group also provided an equally good result. Other solvents, such as toluene and MeCN, were also evaluated, but all gave lower product yields. Finally, it is worth noting that the mild conditions used here are in sharp contrast to the typical high temperature required for the previously known catalytic intermolecular benzannulations involving isobenzopyryliums and alkynes.5dEvaluation of reaction conditions
EntryCatalystYielda (%)
1MeSO3H<5b
2HNTf2<5c
3Sc(OTf)3<5c
4TiCl4<5b
5AgOTf30c
6BF3·OEt275
7 BF 3 ·OEt 2 (20 mol%) 85 (81) d
8BF3·OEt2 (1.0 equiv.)83
9BF3·OEt2 (20 mol%), with 1a′(75)d
Open in a separate windowaReaction scale: 1a (0.05 mmol), 2a (0.06 mmol), catalyst (10 mol%), and DCM (0.5 mL). Yield is based on the analysis of the 1H NMR spectrum of the crude product using CH2Br2 as the internal standard.bUnreacted substrates account for the major remainder of the mass balance.cAn unidentifiable mixture accounts for the major remainder of the mass balance.dYield in parentheses is isolated yield.With the optimized conditions, the generality of this process was examined (Scheme 2). A range of isochromenes 1 and siloxy alkynes 2 participated in this benzannulation to form the desired silylated β-naphthols 3 with moderate to good efficiency. Notably, these products were all generated as a single isomer. This is particularly noteworthy when compared with the cases using phthalazines, which gave a mixture of regioisomers if the phthalazine was unsymmetrically substituted.8g The excellent regioselectivity observed in our reaction is likely attributed to the significant polarization of both cycloaddition partners. While alkyl-substituted siloxy alkynes reacted efficiently, unfortunately aryl-substituted ones led to low yield.Open in a separate windowScheme 2Reaction scope. Reaction scale: 1 (0.5 mmol), 2 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 24 h.During the above scope study, we found that the reaction of 3-unsubstituted isochromene 1g and siloxy alkyne 2a did not form the desired product 3a. Instead, 2-hydroxy-1-naphthaldehyde 4a was formed as the major product (64% yield, eqn (1)). Careful analysis of this product structure indicated that the original leaving part in 1a (in the case of 3a) was not completely cleaved from the molecule. Instead, only the C–O bond cleavage took place, which led to a dangling aldehyde group. Another important observation is that the TIPS group was lost in the naphthalene product. It was proposed that the methoxide leaving group in 1a might help remove this silyl group and assist the above C–O bond cleavage. In contrast, in the formation of 3a, this methoxide group was a part of the whole leaving unit during rearomatization and thus unavailable for desilylation (vide infra).In fact, such 2-hydroxy-1-naphthaldehyde products, resembling the highly versatile salicylaldehyde family, are indeed a useful substructure of bioactive molecules, such as gossypol (Fig. 1).2b They can also serve as key intermediates toward functional materials, such as molecular sensors (e.g., AHN, Fig. 1).2e In view of these important applications, considerable efforts were next devoted to further improving the reaction efficiency (see the ESI for details). Finally, we found that the reaction yield of 4a could be improved to 85% with 2.0 equivalents of BF3·OEt2 and one equivalent of 2,4,6-collidine as the additive. We believe that the role of collidine is to help reversibly stabilize the isobenzopyrylium intermediate and prevent its decomposition during the reaction progress.9b1The above protocol for the synthesis of 2-hydroxy-1-naphthaldehydes is general for a range of 3-substituted isobenzopyryliums (Scheme 3). The reaction efficiency was not affected by electron-donating or electron-withdrawing groups. Various siloxy alkynes were also suitable reaction partners, including aryl- and alkyl-substituted ones. Again, these products were all formed as a single regioisomer. While the majority of these reactions were highly chemoselective toward the formation of aldehydes 4, it is worth noting that tBu- and TBS-substituted alkynes resulted in a mixture of 3 and 4, with the former being major. This is likely due to the substantial steric hindrance in close proximity to the silyl group, whose cleavage was obstructed and thus the driving force for the subsequent C–O bond cleavage was weakened, thereby altering the pathway to preferentially form 3 (vide infra). Finally, the structures of 4a and 4b were unambiguously confirmed by X-ray crystallography.Open in a separate windowScheme 3Scope for the synthesis of 2-hydroxy-1-naphthaldehydes 4. Reaction scale: 1 (0.5 mmol), 2 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 16 h. b Run for 19 h.The divergent reaction patterns observed with siloxy alkynes prompted us to explore other electron-rich alkynes. Thioalkyne 5a was next employed for the reaction with 1a. Indeed, direct extension with BF3·OEt2 as the catalyst successfully resulted in regioselective formation of β-naphthyl thioethers 6a in about 70% yield (Table 2, entries 1 and 2). Further screening of other catalysts was performed to further improve the reaction efficiency (see the ESI for details). While Brønsted acids led to a significant drop in the yield, we were pleased to find that the Lewis acid AgNTf2 exhibited excellent catalytic activity, furnishing 6a in 89% yield (entry 6).Condition optimizationa
EntryCatalystYieldb (%)
1BF3·OEt270
2BF3·OEt2 (20 mol%)72 (68)b
3HNTf220
4MeSO3H7
5AgOTf80
6 AgNTf 2 89 (83) b
Open in a separate windowaReaction scale: 1a (0.05 mmol), 5a (0.06 mmol), catalyst (10 mol%), and DCM (0.5 mL). The yields are based on NMR analysis of the crude product using CH2Br2 as the internal standard.bIsolated yield.With the above conditions, we carried out a brief examination of the reaction scope (Scheme 4). Interestingly, a similar selectivity divergence was also observed with thioalkynes. Indeed, under identical conditions, the 3-butyl-substituted and 3-unsubstituted isobenzopyryliums all reacted efficiently, with the former leading to only 2-naphthyl thioethers 6, but the latter preferentially to 2-sulfenyl-1-naphthaldehydes 7. It is worth noting that only one regioisomer was observed in all these products.Open in a separate windowScheme 4Scope for the synthesis of 2-naphthyl thioethers. Reaction scale: 1 (0.3 mmol), 5 (0.36 mmol), 12 h, and isolated yield. a Run for 37 h. b Run for 19 h. The unreacted alkyne accounted for the major mass balance.Finally, we were curious about the reactivity of ynamides in such benzannulations.10 To our delight, in the presence of 20 mol% of BF3·OEt2, the reaction between isochromene 1a and different ynamides 8 successfully afforded the desired 2-naphthylamine products 9 with excellent regioselectivity and good efficiency (Scheme 5). Different electron-withdrawing groups on the ynamides were all compatible with this protocol. However, to our surprise, the use of 3-unsubstituted isochromene 1g did not lead to the expected naphthaldehyde 10 at all. Instead, the same product 9a was obtained as the only product, even if two equivalents of BF3·OEt2 were used together with collidine as the additive (conditions in Scheme 3). This result is in dramatic contrast to the cases of siloxy alkynes and thioalkynes.Open in a separate windowScheme 5Scope of ynamides. Reaction scale: 1 (0.5 mmol), 8 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 30 h. b Yield in parentheses is obtained with BF3·OEt2 (2.0 equiv.), 2,4,6-collidine (1.0 equiv.), 0 °C, and DCM.Next, further studies were performed to help understand the unusual selectivity divergence observed with siloxy alkynes and thioalkynes. It is obvious that the substituent at the 3-position of the isochromene substrates has a crucial impact on the product distribution. To further probe the possible steric and electronic effects, we incorporated various other substituents, such as different aryl groups and the bulky tBu group (Table 3). We found that these reactions afforded products 3a and 4′ with variable ratios. With the tBu group, almost the same product distribution as the nBu group was observed, exclusively forming 3a (Table 3, entry 2). However, aryl substitution reversed this ratio, giving product 4′ as the major product (entries 3–5). More importantly, the preference toward 4′ is more pronounced with electron-deficient aryl groups. p-Fluorophenyl substitution led to almost only 4′. These results suggested that the product distribution appeared to be more related to the electronic effect than the steric effect.Influence of the 3-substituents in isochromenesa
Open in a separate windowaReaction scale: 1 (0.05 mmol), 2a (0.06 mmol), and 12 h. Yield is based on the analysis of the NMR yield of the crude mixture using CH2Br2 as an internal standard.bRun with 10 mol% of BF3·OEt2.Possible mechanisms are depicted in Scheme 6 using a siloxy alkyne as an example. The reaction begins with Lewis acid activation on the acetal motif to form the isobenzopyrylium intermediate I. The subsequent cycloaddition with the alkyne triple bond forms the key bicyclic oxonium intermediate II, which has a resonance form II′. This step might also be stepwise by forming one C–C bond first via ketenium III. In either case, the regioselectivity is precisely controlled by matching the polarity of the two partners, which explains the exclusive formation of only one regioisomer in all the cases. Depending on the R substituent, oxonium II can proceed via three possible pathways. In path a (R = alkyl), the methoxide attacks the oxonium carbon to form intermediate IV, which then undergoes retro-[4 + 2] cycloaddition with concomitant rearomatization to form naphthalene 3, together with the release of ester RCO2Me. However, if the oxonium carbon is unsubstituted (R = H), this intermediate is relatively unstable due to less stabilization of the positive charge (also viewed as a secondary cation in II′, versus tertiary carbocation if R ≠ H). Thus, the silyl enol ether motif tends to push the electron toward this unstable oxonium (a “push–pull” scenario) to cleave the bridging C–O bond. This step, likely assisted by the methoxide attack on the silyl unit, leads to 1,3-dicarbonyl intermediate V. The subsequent tautomerization/aromatization leads to the observed 2-hydroxy-1-naphthaldehyde 4.6b As an exception, if the siloxy alkyne is bulky (e.g., R′ = tBu or TBS), the reaction preferentially gives product 3 (Scheme 3, 4l–m). We believe that the increased steric repulsion by bulky R′ significantly disfavors the approach of methoxide to the neighboring silyl group in IIb. Instead, the methoxide prefers to add to the oxonium carbon, which allows path a to operate and forms 3 as the major product. Another exceptional case is the use of ynamides, which exclusively give products 3 even if R = H. Although actual rationalization would require more sophisticated mechanistic study, we currently reason that the enamide unit in the IIb analogue might not be electron-rich enough to exert sufficient driving force to cleave the C–O bond. This can also be viewed from the standpoint of the limited stabilization of the resulting iminium ion next to an electron-withdrawing group if this C–O bond is cleaved. Consequently, the methoxide prefers to attack the oxonium carbon to favor path a. Finally, with 3-aryl-substituted isochromenes (R = Ar), we believe that the oxonium intermediate IIc is well stabilized by aryl resonance. Thus, the effective delocalization of the positive charge makes this oxonium carbon less electrophilic for nucleophilic attack by methoxide. Instead, the methoxide serves as a base to deprotonate the bridgehead hydrogen, which triggers rearomatization and C–O cleavage to form product 4′. The more electron-deficient aryl group makes this bridgehead hydrogen more acidic, thus further favoring this pathway, which explains the trend in entries 3–5, Table 3.Open in a separate windowScheme 6Proposed mechanism.To further substantiate the above rationale, we carried out some control experiments. First of all, to confirm the fate of the leaving part in the isochromene substrates when β-naphthols 3 were formed, we used a large homobenzyl group in isochromene 1p (eqn (2)). After its reaction with 2a, we were able to isolate ester 11, which is consistent with path a of the proposed mechanism. Next, we also suspected that products 3 and 4 might interconvert to each other depending on the reaction conditions. To probe this possibility, a mixture of 3a and methyl formate was treated with BF3·OEt2 and 2,4,6-collidine, the standard conditions toward products 4 (eqn (3)). However, product 4a was not observed at all, indicating that 3 is not an intermediate toward 4. Similarly, subjecting 4a and MeOH (or MeONa) to BF3·OEt2 did not lead to the deformylation product β-naphthol 3a′, suggesting that 4 is unlikely an intermediate in the formation of 3 (eqn (4)). These observations are consistent with the proposed mechanism, in which the product distribution is kinetically controlled by the barrier in each case and these paths are likely irreversible, but not thermodynamically controlled by product stability.234In summary, we have developed a modular strategy for the rapid synthesis of valuable β-functionalized electron-rich naphthalenes, specifically, β-naphthol, β-naphthylamine, and β-naphthyl thioether derivatives. It also represents the first systematic study of the benzannulations of versatile isobenzopyryliums with general electron-rich alkynes. With suitable choice of the isochromene substrates and the Lewis acid catalysts, different types of electron-rich alkynes, such as siloxy alkynes, ynamides, and thioalkynes, participated in the intermolecular cycloaddition reactions under mild conditions with high efficiency and complete regioselectivity. Moreover, depending on the substitution pattern of the isochromene substrates, unusual divergence toward different naphthalene products was observed, thus allowing further diversification of the naphthalene products. Control experiments provided preliminary insights into the intriguing mechanism. Further detailed investigations toward a better understanding are ongoing.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号