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 734 毫秒
1.
2.
β-Difluoroalkylborons, featuring functionally important CF2 moiety and synthetically valuable boron group, have great synthetic potential while remaining synthetically challenging. Herein we report a hypervalent iodine-mediated oxidative gem-difluorination strategy to realize the construction of gem-difluorinated alkylborons via an unusual 1,2-hydrogen migration event, in which the (N-methyliminodiacetyl) boronate (BMIDA) motif is responsible for the high regio- and chemoselectivity. The protocol provides facile access to a broad range of β-difluoroalkylborons under rather mild conditions. The value of these products was demonstrated by further transformations of the boryl group into other valuable functional groups, providing a wide range of difluorine-containing molecules.

A hypervalent iodine-mediated gem-difluorination allows the facile synthesis of β-difluoroalkylborons. An unusual 1,2-hydrogen migration, triggered by boron substitution, is involved.

Organofluorine compounds have been widely applied in medicinal chemistry and materials science.1ad In particular, the gem-difluoro moiety featuring unique steric and electronic properties can act as a chemically inert isostere of a variety of polar functional groups.2ac Therefore, the construction of gem-difluoro-containing compounds has received considerable attention in recent years. Efficient methods including deoxyfluorination of carbonyl compounds,3a,b photoredox difluorination,4 radical difluorination,5 and cross-coupling reactions with suitable CF2 carriers6af are well developed. Alternatively, iodoarene-mediated oxidative difluorination reactions provide valuable access to these motifs by using simple alkenes as starting materials.7ai Previously, these reactions were generally associated with a 1,2-aryl or 1,2-alkyl migration (Scheme 1a).7af Recent developments also allowed the use of heteroatoms as migrating groups, thereby furnishing gem-difluoro compounds equipped with easily transformable functional groups (Scheme 1b). In this regard, Bi and coworkers reported an elegant 1,2-azide migrative gem-difluorination of α-vinyl azides, enabling the synthesis of a broad range of novel β-difluorinated alkyl azides.7g Jacobsen developed an iodoarene-catalyzed synthesis of gem-difluorinated aliphatic bromides featuring 1,2-bromo migration with high enantioselectivity.7h Almost at the same time, research work from our group demonstrated that not only bromo, but also chloro and iodo could serve as viable migrating groups.7iOpen in a separate windowScheme 1Hypervalent iodine-mediated β-difluoroalkylboron synthesis.We have been devoted to developing new methodologies for the assembly of boron-containing building blocks by using easily accessible and stable MIDA (N-methyliminodiacetyl) boronates8ac as starting materials.9ae Recently, we realized a hypervalent iodine-mediated oxidative difluorination of aryl-substituted alkenyl MIDA boronates.9d Depending on the substitution patterns, the reaction could lead to the synthesis of either α- or β-difluoroalkylborons via 1,2-aryl migration (Scheme 1c). Recently, with alkyl-substituted branched alkenyl MIDA boronates, Szabó and Himo observed an interesting bora-Wagner–Meerwein rearrangement, furnishing β-difluorinated alkylboronates with broader product diversity (Scheme 1d).10 While extending the scope of our previous work,9d we found that the use of linear alkyl-substituted alkenyl MIDA boronates also delivers β-difluoroalkylboron products. Intriguingly, instead of an alkyl- or boryl-migration, an unusual 1,2-hydrogen shift takes place. It should be noted that internal inactivated alkenes typically deliver the 1,2-difluorinated products, with no rearrangement taking place.11ad Herein, we disclose our detailed study of our second generation of β-difluoroalkylborons synthesis (Scheme 1e). The starting linear 1,2-disubstituted alkyl-substituted alkenyl MIDA boronates, unlike the branched ones,10 could be readily prepared via a two-step sequence consisting of hydroborylation of the terminal alkyne and a subsequent ligand exchange with N-methyliminodiacetic acid. This intriguing 1,2-H shift was found to be closely related to the boron substitution, probably driven thermodynamically by the formation of the β-carbon cation stabilized by a σ(C–B) bond via hyperconjugation.12adTo start, we employed benzyl-substituted alkenyl MIDA boronate 1a as a model substrate (9d the use of F sources such as CsF, AgF and Et3N·HF in association with PhI(OAc)2 (PIDA) as the oxidant and DCM as the solvent led to no reaction (entries 1 to 3). The use of Py·HF (20 equiv) successfully provided β-difluorinated alkylboronate 2a, derived from an unusual 1,2-hydrogen migration, in 39% yield (entry 4). By simply increasing the loading of Py·HF to 40 equivalents, a higher conversion and thus an improved yield of 61% was obtained (entry 5). No further improvement was observed by using a large excess of Py·HF (100 equiv) (entry 6). Other hypervalent iodine oxidants such as PhIO or PIFA were also effective but resulted in reduced yields (entries 7 and 8). A brief survey of other solvents revealed that the original DCM was the optimal one (entries 9 and 10).Optimization of reaction conditions
EntryF (equiv)OxidantSolventYield (%)
1CsF (2.0)PIDADCM0
2AgF (2.0)PIDADCM0
3Et3N·HF (40.0)PIDADCM0
4Py·HF (20.0)PIDADCM39
5 Py·HF (40.0) PIDA DCM 61
6Py·HF (100.0)PIDADCM55
7Py·HF (40.0)PIFADCM52
8Py·HF (40.0)PhIODCM26
9Py·HF (40.0)PIDADCE49
10Py·HF (40.0)PIDAToluene46
Open in a separate windowWith the optimized reaction conditions in hand, we set out to investigate the scope and limitation of this gem-difluorination reaction. The reaction of a series of E-type 1,2-disubstituted alkenyl MIDA boronates were first examined. As shown in Scheme 2, the reaction of substrates with primary alkyl (1b, 1e–g), secondary alkyl (1c, 1d), or benzyl (1h–k) groups proceeded efficiently to give the corresponding gem-difluorinated alkylboronates in moderate to good yields. Halides (1i–k, 1m) and cyano (1l) were well tolerated in this reaction. Of note, cyclic alkene 1n is also a viable substrate, affording an interesting gem-difluorinated cyclohexane product (2n).Open in a separate windowScheme 2Scope of 1,2-H migratory gem-difluorinations. a 4 h. b PIFA was used.To define the scope further, the substrates with Z configuration were also employed under the standard reaction conditions (eqn (1) and (2)). The same type of products were isolated with comparable efficiency, suggesting that the reaction outcome is independent of the substrate configuration and substrates with Z configuration also have a profound aptitude of 1,2-hydrogen migration. Nevertheless, the reaction of t-butyl substituted alkenyl MIDA boronate (1p) delivered a normal 1,2-difluorinated alkylboron product (eqn (3)). The 1,2-hydrogen migration was completely suppressed probably due to unfavorable steric perturbation. With an additional alkyl substituent introduced, a 1,2-alkyl migrated product was formed as expected (eqn (4)).1The gem-difluorination protocol was amenable to gram-scale synthesis of 2a (Scheme 3, 8 mmol scale of 1a, 1.24 g, 50%). To assess the synthetic utility of the resulting β-difluorinated alkylborons, transformations of the C–B bond were carried out (Scheme 3). Ligand exchange of 2a furnished the corresponding pinacol boronic ester 4 without difficulty, which could be ligated with electron-rich aromatics to obtain 5 and 6 in moderate yields. On the other hand, 2a could be oxidized with high efficiency to alcohol 7 using H2O2/NaOH. The hydroxyl group of 7 could then be converted to bromide 8 or triflate 9. Both serve as useful electrophiles that can undergo intermolecular SN2 substitution with diverse nitrogen- (10, 13), oxygen- (14), phosphorus- (11) and sulfur-centered (12) nucleophiles.Open in a separate windowScheme 3Product derivatizations. PMB = p-methoxyphenyl.To gain insight into the reaction mechanism, preliminary mechanistic studies were conducted. The reaction employing deuterated alkenyl MIDA boronate [D]-1a efficiently afforded difluorinated product [D]-2a in 72% isolated yield, clearly demonstrating that 1,2-H migration occurred (Scheme 4a). However, when the MIDA boronate moiety was replaced with a methyl group (15), no difluorinated product (derived from 1,2-migration) was detected at all, suggesting an indispensable role of boron for promoting the 1,2-migration event (Scheme 4b). Also, with a Bpin congener of 1a, the reaction led to large decomposition of the starting material, with no desired product being formed (Scheme 4b).Open in a separate windowScheme 4Mechanistic studies and proposals.Based on the literature precedent and these experiments, a possible reaction mechanism is proposed in Scheme 4c. With linear alkenyl MIDA boronates, the initial coordination of the double bond to an iodium ion triggered a regioselective fluoroiodination to deliver intermediate B. The regioselectivity could arise from an electron-donating inductive effect from boron due to its low electronegativity, consistent with previous observations.13a,b Thereafter, a 1,2-hydrogen shift, rather than the typical direct fluoride substitution of the C–I bond, provides carbon cation C. The formation of a hyperconjugatively stabilized cation is believed to be the driving force for this event.12ad The trapping of this cation finally forms the product.In conclusion, we demonstrated herein our second generation of β-difluoroalkylboron synthesis via oxidative difluorination of easily accessible linear 1,2-disubstituted alkenyl MIDA boronates. An unexpected 1,2-hydrogen migration was observed, which was found to be triggered by a MIDA boron substitution. Mild reaction conditions, moderate to good yields and excellent regioselectivity were achieved. The applications of these products allowed the facile preparation of a wide range of gem-difluorinated molecules by further transformations of the boryl group.  相似文献   

3.
A new catalytic asymmetric formal cross dehydrogenative coupling process for the construction of all-aryl quaternary stereocenters is disclosed, which provides access to rarely explored chiral tetraarylmethanes with excellent enantioselectivity. The suitable oxidation conditions and the hydrogen-bond-based organocatalysis have enabled efficient intermolecular C–C bond formation in an overwhelmingly crowded environment under mild conditions. para-Quinone methides bearing an ortho-directing group serve as the key intermediate. The precise loading of DDQ is critical to the high enantioselectivity. The chiral products have also been demonstrated as promising antiviral agents.

A one-pot oxidation of racemic triarylmethanes to form para-quinone methides followed by enantioselective construction of all-aryl quaternary stereocenters has been developed.

Cross dehydrogenative coupling (CDC) is a powerful tool to forge intermolecular C–C bonds from two C–H bonds without prefunctionalization.1 Specifically, the benzylic C–H bond is relatively prone to oxidation and thus it has evolved into a versatile arena for the implementation of this reaction, leading to efficient construction of various benzylic stereogenic centers. As a result, CDC has proved to be useful for the establishment of a wide range of 1,1-diaryl stereocenters (Scheme 1a).2 Recently, Liu and coworkers reported a elegant synthesis of enantioenriched triarylacetonitriles via in situ oxidation of α-diarylacetonitriles to para-quinone methides (p-QMs) followed by asymmetric nucleophilic addition with stereocontrol induced by a chiral phosphoric acid catalyst. This represents a rare example of formal CDC for the synthesis of 1,1,1-triarylalkanes (Scheme 1b).3 However, the establishment of tetraaryl-substituted carbon stereocenters by this approach remains unknown (Scheme 1c).Open in a separate windowScheme 1Catalytic asymmetric synthesis of chiral tetraarylmethanes.Distinct from the asymmetric synthesis of triaryl-substituted stereocenters,4 substantial steric hindrance in establishing tetraaryl-substituted quaternary stereocenters poses significant synthetic challenges.5–8 Indeed, even racemic or achiral syntheses of tetraarylmethanes have been an elusive topic of investigation in organic synthesis.6 In this context and in continuation of our effort in the studies of asymmetric reactions of para-quinone methides (p-QMs)9,10 as well as the synthesis of chiral tetraarylmethanes,8 we envisioned that suitable oxidation of racemic triarylmethane 1 is expected to generate triarylmethyl cation IM1 (Scheme 1c). With one aryl group as para-hydroxyphenyl, this cation could be stabilized in the form of p-QM IM2. Subsequent asymmetric nucleophilic addition by another electron-rich arene to the p-QM intermediate is expected to generate chiral tetraarylmethanes 2. The challenges associated with this one-pot process mainly include the compatibility problem between the oxidative condition and the catalytic asymmetric system in order to achieve both high efficiency and enantioselectivity.We commenced our study with racemic triarylmethane 1a as the model substrate. The initial study was directed to the search for a suitable oxidant to mildly generate the p-QM intermediate (11 At room temperature, the use of superstoichiometric amounts of Ag2O or benzoquinone was completely ineffective (entries 1 and 2). Similarly, the reaction did not proceed using oxygen as the oxidant in combination with catalyst Mn(acac)3 (entry 3). Subsequently, considerable efforts were devoted to screening many other oxidation systems, almost all of which were completely incapable for this oxidation (entries 4–8). However, eventually we were delighted to identify DDQ as the superior oxidant, leading to complete and clean conversion to the desired QM at room temperature (entry 9). In contrast, a combination of catalytic DDQ with 5 equivalents of MnO2 gave only 60% conversion (entry 10).Evaluation of oxidants
Entry[O]Conv. (%)
1Ag2O (5.0 equiv.)0
2Benzoquinone (1.5 equiv.)0
3Mn(acac)3 (10 mol%), O2 (1 atm)0
4KBr (1.2 equiv.), Oxone (1.2 equiv.)0
5K3Fe(CN)6 (1.5 equiv.)0
6AIBN (0.5 equiv.), TBHP (3.0 equiv.)0
7FeCl3 (10 mol%), TBHP (3.0 equiv.)0
8TEMPO (3.0 equiv.)0
9DDQ (1.0 equiv.)100
10DDQ (20 mol%), MnO2 (5.0 equiv.)60
Open in a separate windowWe next set out to evaluate the key C–C bond formation step (12,13 After oxidation, the nucleophile and catalyst were added to the reaction mixture. The reaction with catalyst (R)-A1 proceeded smoothly at room temperature to form the desired product 2a in 90% yield, but unfortunately in a racemic form (entry 1). Next, a range of chiral phosphoric acids were screened. To our delight, the BINOL-derived TRIP catalyst, (R)-A4, provided excellent enantioselectivity (93% ee, entry 4). However, those with H8BINOL- and SPINOL-derived catalysts (B and C) bearing the same 2,4,6-triisopropylphenyl substituents proved to be inferior. Finally, a slightly modified acid A5 was found to be the best (95% ee, entry 7). Decreasing the temperature to 0 °C improved the result (97% ee, entry 8). However, no further improvement was observed at a lower temperature. While DCM was comparable to DCE, other solvents (e.g., EtOAc and Et2O) significantly affected the enantioselectivity. Varying the concentration led to no improvement (entries 9–13). Finally, the catalyst loading could be reduced to 7.5 mol% without erosion in yield or enantioselectivity (entry 14). Notably, during the course of our study, the enantioselectivity was found to be sensitive to the amount of DDQ when it was used in excess. For example, with 1.5 equivalents of DDQ (entry 15), the enantioselectivity decreased to 51% ee. However, with 0.8 equivalents, the selectivity remained excellent, albeit with reduced yield. These results suggest that the excessive DDQ might be detrimental to stereocontrol. Unfortunately, this feature also prevented the two-step protocol from merging into one operation. The catalyst has to be added after complete consumption of DDQ to ensure high enantioselectivity (entry 17). Moreover, although the oxidation step was relatively fast (∼30 min) based on TLC analysis, keeping this mixture under stirring for an additional 4 h before adding the acid catalyst was critical to achieve high enantioselectivity, which is likely to ensure complete consumption of DDQ or precipitation of its reduced form DDQH2 from the solution (entry 18).Condition optimizationa
EntryCPATemp.Yield 2a (%)ee (%)
1(R)-A1rt900
2(R)-A2rt9547
3(R)-A3rt9249
4(R)-A4rt9693
5(R)-Brt9365
6(R)-Crt919
7(R)-A5rt9595
8(R)-A50 °C9597
Open in a separate windowaReaction conditions: 1a (0.025 mmol), 3a (0.05 mmol), catalyst (10 mol%), DCE (0.5 mL). Yield is based on analysis of the 1H NMR spectroscopy of the crude reaction mixture using CH2Br2 as an internal standard.
Change from the entry 8
9EtOAc as solvent>9541
10Et2O as solvent8870
11DCM as solvent>9593
12c = 0.1 M9695
13c = 0.025 M9593
147.5 mol% of (R)-A59597
151.5 equiv. of DDQ9451
160.8 equiv. of DDQ7796
17Mix all together at the beginning4762
181 h (not 5 h) for the first step9581
Open in a separate windowWith the optimized conditions (entry 14, Scheme 2). A wide range of diversely-substituted triarylmethanes participated in this process with good to excellent efficiency and enantioselectivity. In addition to OMe, other alkoxy groups (e.g., OBn and OAllyl, 2k–l), protected amine groups (e.g., sulfonamides, 2m–o), and even fluorine (2p–q) can serve as an effective directing group when they are present at the ortho position. Moreover, as shown in the case of 2f, the observed good enantioselectivity indicated that the directing ability of alkoxy and fluorine groups is remarkably different. The incorporation of a heterocycle, such as thiophene (2g), did not interfere with the reactivity or enantiocontrol. Some other pyrroles, including 2,4-dimethyl pyrrole (2x), were also good nucleophiles. 4,7-Dihydro-1H-indole also reacted smoothly to form the product 2v. Subsequent oxidation by DDQ could easily afford the indole-substituted tetraarylmethane 2weqn (1). Unfortunately, pyrroles with carbonyl substituents and other electron-rich arenes, such as indole, furan, 2-naphthol, and 1,3,5-trimethoxybenzene, were not reactive under the standard conditions (0 °C). At room temperature, indole could react to form the desired product 2y, but in only 21% ee, while the others remain unreactive.1Open in a separate windowScheme 2Reaction scope. Reaction scale: 1 (0.25 mmol), DDQ (0.25 mmol), DCE (5.0 mL), rt, 5 h; then 3 (0.50 mmol), (R)-A5 (18.8 μmmol), 0 °C, 3 h. Isolated yield is provided. The ee value was determined by chiral HPLC analysis. aRun at −20 °C for 12 h after catalyst addition. bRun at rt for 24 h after catalyst addition.The standard protocol could be scaled to 1.25 mmol without erosion in efficiency or enantiocontrol (Scheme 3). Moreover, the directing groups, such as the para-hydroxy group, could be easily converted or removed. For example, after triflation of the phenol unit in 2d, the triflate 3 could easily participate in coupling reactions to form the arylation, reduction, and allylation products 4–6. The high enantiopurity remained essentially intact.Open in a separate windowScheme 3Product transformations. [a] Tf2O, Et3N, DCM, 0 °C to rt; [b] PhB(OH)2, Pd(OAc)2, BrettPhos, K3PO4, tBuOH, 85 °C; [c] Et3SiH, Pd(OAc)2, dppp, DMF, 60 °C; [d] AllylBpin, Pd(OAc)2, BrettPhos, K3PO4, tBuOH, 85 °C.To understand the reaction mechanism, we carried out some control experiments. First, the intermediate QM, though unstable and easy to undergo addition, was obtained by careful isolation from the oxidation step in the presence of molecular sieves (Scheme 4a). Next, in the absence of DDQ, the standard reaction between QM and 2-methylpyrrole proceeded with high efficiency and excellent enantioselectivity (97% ee, Scheme 4b). However, with DDQ as an additive, the enantioselectivity decreased to 44% ee, which confirmed that it is detrimental to enantiocontrol.14 The methylated substrate 1a-Me was also examined. The desired tetraarylmethane 2a-Me was successfully formed, but in an almost racemic form (Scheme 4c). In this case, the corresponding oxonium cation served as an activated intermediate, rather than p-QM. This result indicated that the free hydroxyl group in the standard substrates is not necessary for DDQ oxidation, but the resulting p-QM intermediate is essential for excellent enantiocontrol.Open in a separate windowScheme 4Mechanistic study.Finally, the substrates bearing other ortho-substituents in place of the ortho-methoxyl group were examined. With ortho-methyl and ethyl groups (1r–s), low enantioselectivies were obtained in spite of excellent yields. In particular, the ethyl group has a similar size to the methoxyl group, but does not provide hydrogen bonding interactions. The dramatically low ee (17% ee) for this case provided strong evidence that steric hindrance is not key to the excellent asymmetric induction for 1a. Furthermore, substrate 1t (with ortho-OiPr) also provided a lower ee (72% ee) than 1a. These results suggested that it is the hydrogen bonding interaction with the ortho-directing group, not the steric or electronic effect, that leads to the excellent enantiocontrol in the standard protocol.8We also randomly selected a few of our products to test their potential antiviral activities in Rhabdomyosarcoma (RD) cells, which are commonly used to investigate enterovirus A71 (EV-A71) infections. Our compounds showed relatively high CC50 measured by MTT assay, indicating low cell toxicity (Fig. 1). Quantitation of viral genome RNA in the secreted virions showed potent inhibition of virus replication with IC50 ranging from 0.20 to 1.24 μM, indicating a high selectivity index (
CompoundCC50 (μM)IC50 (μM)Selectivity indexb
2k 29.30.20148.5
2u 33.20.24138.3
2r 28.21.2422.7
Open in a separate windowaCC50, 50% cytotoxic concentration measured by viability assay (without virus infection); IC50, the viral RNA copies were reduced by 50% compared with the control (without compound treatment) in the secreted virions.bA selectivity index (CC50/IC50) of >10 is considered to have good potential for drug development.Open in a separate windowFig. 1The antiviral effects examined by CPE assay and quantitation of viral RNA copies in the secreted virions. RD cells were treated with the indicated compounds and infected with EV-A71 at a MOI of 0.1, and the cell morphology was observed using a phase-contrast microscope 24 h post infection. The viral RNA genome copy number was determined by RT-qPCR.In conclusion, we have developed the first catalytic asymmetric formal cross dehydrogenative coupling for the efficient synthesis of enantioenriched chiral tetraarylmethanes, a family of challenging molecules to synthesize. Enabled by a one-pot oxidation and nucleophilic addition protocol, the intermolecular C–C bond was efficiently forged from two C–H bonds with high enantioselectivity under mild conditions, which benefitted from successful understanding and addressing the key compatibility issue between the DDQ oxidant and resulting DDQH2 with the catalytic asymmetric system. Finally, these new products have been demonstrated as promising antiviral agents.  相似文献   

4.
Manganese(i)-catalyzed access to 1,2-bisphosphine ligands     
Luo Ge  Syuzanna R. Harutyunyan 《Chemical science》2022,13(5):1307
Chiral bisphosphine ligands are of key importance in transition-metal-catalyzed asymmetric synthesis of optically active products. However, the transition metals typically used are scarce and expensive noble metals, while the synthetic routes to access chiral phosphine ligands are cumbersome and lengthy. To make homogeneous catalysis more sustainable, progress must be made on both fronts. Herein, we present the first catalytic asymmetric hydrophosphination of α,β-unsaturated phosphine oxides in the presence of a chiral complex of earth-abundant manganese(i). This catalytic system offers a short two-step, one-pot synthetic sequence to easily accessible and structurally tunable chiral 1,2-bisphosphines in high yields and enantiomeric excess. The resulting bidentate phosphine ligands were successfully used in asymmetric catalysis as part of earth-abundant metal based organometallic catalysts.

Chiral bisphosphine ligands are of key importance in transition-metal-catalyzed asymmetric synthesis of optically active products. Mn(i)-catalyzed hydrophosphination offers a two-step, one-pot synthetic sequence to access chiral 1,2-bisphosphines.

The vast majority of important catalytic transformations make use of very effective catalysts based on scarce, expensive and toxic noble transition metals and phosphine containing ligands that, especially when chiral, are often as expensive as the noble metals themselves due to their cumbersome synthetic accessibility.1 The past decade has witnessed significant progress towards the development of competitive catalysts that contain earth-abundant transition metals instead. These catalysts, however, still frequently rely on the use of chiral phosphine ligands. Bisphosphine ligands (Scheme 1A) for instance Pyrphos,2a Chiraphos,2b as well as Josiphos2c are among the most successful chiral ligands used in homogeneous catalysis. In recent years, bis(phosphine) monoxide compounds such as Bozphos,2d and Binap(o)2e have been shown to be powerful ligands in asymmetric catalysis as well. Unfortunately, the synthesis of these frequently and successfully used chiral phosphine-based ligands often requires stoichiometric amounts of chiral auxiliaries, enantiopure substrates, or separation by resolution to obtain them enantiomerically pure.1bfOpen in a separate windowScheme 1(A) Examples of phosphine ligands commonly used in homogeneous catalysis. (B) Catalytic asymmetric hydrophosphination of various Michael acceptors. (C) This work: Mn (i)-catalyzed access to chiral 1,2-bisphosphines.Catalytic asymmetric hydrophosphination is one of the most straightforward approaches for generating optically active P-chiral or C-chiral phosphines, from which chiral ligands can be derived.3 The potential of hydrophosphination reactions to access enantioenriched chiral phosphines catalytically was demonstrated for the first time by Glueck and coworkers in 2001 using a catalytic system based on Pt and the chiral bisphosphine ligand Me-DuPhos.4 Following the publication of this initial work, precious noble metal complexes such as chiral Pd or Pt catalysts have been widely used in the field of asymmetric hydrophosphination (Scheme 1B).5 Only few examples utilizing earth-abundant metals such as Ni,6 Cu7 and very recently Mn8 have been reported to date for catalytic asymmetric hydrophosphination. Apart from metal based catalytic systems, examples of asymmetric organocatalytic hydrophosphination reactions were also presented in the literature.9 So far, all successful methods that rely on the addition of phosphines to α,β-unsaturated conjugated systems provide chiral monophosphines.3 Interestingly, the only reported example of catalytic hydrophosphination that allows access to chiral 1,2-bisphosphine ligands utilizes a Michael acceptor with a P-containing electron-withdrawing group.7bWhile α,β-unsaturated phosphine oxides are bench stable and readily available Michael acceptors, their application is less common when compared to conventional carbonyl based Michael acceptors, which is in part due to their lower reactivity.10 Yin and co-workers found an elegant solution to this problem by transforming α,β-unsaturated phosphine oxides into phosphine sulphides. This allows a ‘soft–soft’ interaction to be established between the Cu(i) atom of the chiral Cu(i)-catalyst and the S atom of the phosphine sulphide, enabling catalytic hydrophosphination towards the synthesis of chiral bisphosphines.7b While successful in applying this strategy for catalytic synthesis of variety of chiral bisphosphines, nevertheless it requires 6-steps synthetic sequence starting from α,β-unsaturated phosphine oxides (Scheme 1C).7bHerein, we present a highly efficient, short and scalable catalytic protocol for the synthesis of chiral 1,2-bisphosphines from readily available, bench stable α,β-unsaturated phosphine oxides employing Mn(i)-catalyzed hydrophosphination as its core transformation (Scheme 1D).The last five years witnessed remarkable success of Mn(i)-complexes as catalysts for reductive transformations of carbonyl compounds including asymmetric variants.11–13 Next to these reports, we have recently demonstrated that such complexes are capable of catalytic H–P bond activation of diarylphosphines.8 Based on these findings we hypothesised that Mn(i)-complexes should be able to bring the phosphine oxide and the phosphine reagents into closer proximity thus allowing the hydrophosphination reaction to take place directly with α,β-unsaturated phosphine oxides. This approach would avoid the additional synthetic steps and purifications procedures necessitated by the installation and removal of the sulphur atom that are intrinsic to the method utilising phosphine sulphides.At the outset of this work, bench-stable α-substituted α,β-unsaturated phosphine oxide 1a was chosen as the model substrate in the reaction with HPPh2 (i)-complex, Mn(i)-L, developed by Clark and co-workers13a,d for hydrogenation and transfer hydrogenation of carbonyl compounds, was selected as the chiral catalyst. After extensive optimization, the reaction with 5 mol% t-PentOK, 2.5 mol% Mn(i)-L, 1.05 equiv. of HPPh2 in toluene at room temperature for 16 hours was found to be optimal. Under these conditions, the product 3aa was obtained with 96% isolated yield and over 99% ee (entry 1).Optimization of the reaction conditionsa
EntryDeviation standard conditionsConv.b (%)Eec (%)
1None>99 (96)d>99
2Without Mn(i)-L and t-PentOK0
3Without t-PentOK0
4Without Mn(i)-L99
5THF instead of toluene9996
61,4-Dioxane instead of toluene9897
7i-PrOH instead of toluene7595
8MeOH instead of toluene9052
9 t-BuOK instead of t-PentOK9997
10Barton''s base instead of t-PentOK9898
11 t-PentOK (2.5 mmol%)5699
12 t-PentOK (7.5 mmol%)9995
Open in a separate windowaGeneral conditions: 1a (0.1 mol), Mn(i) (2.5 mol%), t-PentOK (5 mol%), 2a (0.105 mol) in toluene (1.0 ml) at rt for 16 h.bDetermined by 1H NMR of reaction crude.cDetermined by HPLC on a chiral stationary phase.dIsolated yield.In the absence of both the base and the catalyst, as well as in the presence of only Mn(i)-L, no reaction occurs at room temperature (entries 2 and 3). In the presence of only the base (5 mol% of t-PentOK), however, 99% conversion towards the phosphine product 3aa was observed (entry 4).14The screening of various solvents (entries 5–8) revealed excellent yields and enantiomeric ratios when using any of the following solvents: toluene, THF, and 1,4-dioxane. Given that the stereocenter in this reaction is generated upon formal stereospecific protonation, it was surprising that only a slight decrease in enantiomeric purity of the final product was observed in protic solvents, such as i-PrOH. On the other hand, running the reaction in MeOH led to a significant decrease in both substrate conversion and product ee.As for the nature of the base we discovered that alkoxides and Barton''s base provide the best results regarding the product yield and enantiopurity. The optimal performance of the base in the Mn(i)-catalyzed reaction is achieved with between 1.5 and 2 equivalents of the base with respect to the catalyst. A higher or lower amount of the base results in lower enantioselectivity or lower yield, respectively (compare entries 1, 11 and 12).With the optimized conditions in hand, we moved to explore the scope of this methodology, first concentrating on the R2 substituent on the phosphine oxide. Various substitutions with aryl or alkyl groups led to excellent results in all cases (Scheme 2). Substrates with either an electron-donating group (3ba and 3ca) or an electron-withdrawing group (3da, 3ea, and 3fa) at the para-position of the phenyl ring led to the corresponding products with over 98% ee. The phenyl and ester functional groups at the para-position were also well tolerated, providing products 3ga and 3ha with high yields and enantiopurities. Similar results were obtained for substrates containing methyl- (3ia), chloro- (3ja) or methoxy- (3ka) substituents at the meta-position of the phenyl ring.Open in a separate windowScheme 2Product scope of Mn(i)-catalyzed asymmetric hydrophosphination of α,β-unsaturated phosphine oxidesa.aReaction conditions: 0.1 M of 1 in toluene, Mn(i)-L (2.5 mol%), t-PentOK (5 mol%), HP(Ar)2 (1.05 equiv) at rt. Isolated yields reported. For products 3aa and 3za the absolute configurations were determined by transforming them into the corresponding known compounds 6aa and 6da and for the remainder of the products by analogy (for details see ESI); b5 mol% Barton''s base used; c5 mol% Mn(i)-L,10 mol% t-PentOK used and reaction was carried out at rt for 72 h; d5 mol% Mn(i)-L,10 mol% t-PentOK used and reaction was carried out at rt for 5 days; e5 mol% Mn(i)-L,10 mol% t-PentOK used and reaction carried out at 60 °C; fthe reaction quenched with H2O2; gfor the absolute configuration of 3za, see the ESI.α,β-Unsaturated phosphine oxides containing a heteroaryl moiety, such as 2-naphthyl (3ma), 3-thienyl (3na), and 3-pyridinyl (3oa), were well applicable in our catalytic system. We were pleased to see that substrate 3pa, bearing a ferrocenyl substituent – an essential structural component for many successful chiral ligands – can also be hydrophosphinated with excellent results. Next, α-alkyl substituted substrates were evaluated. The enantioselectivities observed for substrates with linear (3qa) and branched aliphatic substituents (3ra and 3sa) were in line with the results obtained for their aromatic counterparts. Substrates bearing functional groups amenable to further transformations, namely hydroxyl- (3ta), cyano- (3ua) or chloro-substituents provided the corresponding phosphine products with equally good results. We then move to study the effect of varying the substituents at the phosphorus atom. Various unsaturated diaryl phosphine oxides are compatible with this catalytic system and afford the corresponding products 3wa, 3xa, and 3ya with excellent enantiomeric excess and high isolated yield.The relatively less reactive β-butyl-substituted α,β-unsaturated phosphine oxide is well tolerated as well, providing the corresponding enantioenriched oxide product 3za with 87% ee. On the other hand, no conversion to the product 3a′a was observed with β-phenyl-substituted α,β-unsaturated phosphine oxide. Interestingly, this catalytic system also supports α,β-unsaturated phosphonates, generating the corresponding final products (4a′a, 4b′a, 4c′a, and 4d′a) with enantiomeric excesses in the range of 89–95%. The catalytic protocol was also applied to a phosphinate substrate, allowing access to the product 4e′a with two chiral centers (dr 1 : 1) with high ee. Finally, screening of various phosphine reagents revealed some limitations of the protocol. Hydrophopshination with (p-Me-C6H4)2PH and (p-MeO-C6H4)2PH led to the corresponding products 5ab and 5ac with good yields and good to excellent enantioselectivities. However, no conversion was obtained with the sterically more demanding (o-Me-C6H4)2PH, (3,5-CF3-C6H3)2PH, nor with Cy2PH and (p-CF3-C6H4)2PH. Attempts to access P-chiral phosphine product via addition of racemic diarylphosphine to α, β-unsaturated phosphine oxides led to the racemic P-chiral phosphine 5a′h.To demonstrate the potential application of our catalytic protocol in chiral phosphine ligand synthesis, we performed a gram-scale reaction between 1b and 2a (Scheme 3A). To our delight, the catalyst loading could be decreased to 0.5 mol%, leading to the product 3ba without deterioration of the yield (91%) or the enantioselectivity (98%).Open in a separate windowScheme 3(A) Gram-scale Mn(i)-catalyzed reaction using 0.5 mol% Mn(i)-L. (B) One-pot synthesis of chiral 1,2-bisphosphine boranes. (C) Synthesis of chiral 1,2-bisphosphines. (D) Application of bisphosphine 7ca in Cu(i)-catalyzed hydrophosphination.Building on these results, we then developed a highly efficient one-pot method for the synthesis of four different chiral phosphine boranes (6aa–6da) (Scheme 3B) that yield the corresponding chiral 1,2-bisphosphine ligands (7aa–7da) in a single deprotection step (Scheme 3C). As is typical of any phosphines, the 1,2-bisphosphines 7 prepared in this study can easily oxidize during chromatographic purifications.7bTherefore, to minimise chromatographic purification, as well as to facilitate product separation, degassed water was used to wash the reaction mixture, followed by the removal of volatiles under high vacuum. The free ligands 7 were obtained in good yields and high purity. Importantly, the 1,2-bisphosphine 7aa is a known, efficient chiral ligand for Rh-catalyzed asymmetric hydrogenation of α-amino-α,β-unsaturated esters.7b We also examined our bisphosphine ligand 7ca in the Cu-catalyzed hydrophosphination of α,β-unsaturated phosphine oxide 1a (Scheme 3D), obtaining the desired product 3aa in good yield (90%) and high enantioselectivity (92%). Similarly, α,β-unsaturated carboxamide 8 was investigated,7c providing the corresponding product 9 in good yield (82%) and moderate ee (52%).From a mechanistic point of view, we wondered whether our base activated Mn-catalyst I is involved in the activation of the phosphine reagent 2avia ligand–metal cooperation, as proposed in our previous work on α,β-unsaturated nitriles,8 or whether it also plays a role in the activation of the phosphine oxide substrate 1. Preliminary NMR spectroscopic studies did not reveal any interaction between I and 1 (see ESI) leading us to hypothesise that the current transformation might follow a mechanistic path that primarily involves phosphine activation, as depicted in Scheme 4. Additional interaction between the NH and P Created by potrace 1.16, written by Peter Selinger 2001-2019 O moieties of the catalyst and phosphine oxide respectively is also possible and cannot be excluded at this stage. Detailed mechanistic studies are currently underway.Open in a separate windowScheme 4Hypothetical catalytic cycle.In summary, we have developed the first manganese(i) catalyzed enantioselective strategy for the hydrophosphination of α, β-unsaturated phosphine oxides. This methodology allows a high-yielding, catalytic, two-step sequence for the synthesis of enantiopure chiral 1,2-bisphosphine ligands, that were successfully applied in asymmetric catalysis. Since manganese is the third most abundant transition metal in the Earth''s crust, a general catalytic method to access chiral bisphosphine ligands using this metal is further step towards more sustainable homogeneous catalysis. Further work is currently underway in order to unravel the mechanism of this transformation.  相似文献   

5.
Allylic alcohol synthesis by Ni-catalyzed direct and selective coupling of alkynes and methanol     
Herong Chen  Zhijun Zhou  Wangqing Kong 《Chemical science》2021,12(27):9372
Methanol is an abundant and renewable chemical raw material, but its use as a C1 source in C–C bond coupling reactions still constitutes a big challenge, and the known methods are limited to the use of expensive and noble metal catalysts such as Ru, Rh and Ir. We herein report nickel-catalyzed direct coupling of alkynes and methanol, providing direct access to valuable allylic alcohols in good yields and excellent chemo- and regioselectivity. The approach features a broad substrate scope and high atom-, step- and redox-economy. Moreover, this method was successfully extended to the synthesis of [5,6]-bicyclic hemiacetals through a cascade cyclization reaction of alkynones and methanol.

Methanol is an abundant and renewable chemical raw material, but its use as a C1 source in C–C bond coupling reactions still constitutes a big challenge, and the known methods are limited to the use of expensive and noble metal catalysts such as Ru, Rh and Ir.

To address the sustainability issues in the production of new chemicals, the development of new catalytic processes that are free of by-products and using abundant renewable feedstocks is one of the most important challenges facing chemists today. The simplest alcohol, methanol, is very abundant, with a total annual production capacity of approximately 110 million metric tons per year,1 and is an important C1-feedstock in the chemical industry. Beller2 and Milstein3 made fundamental developments in catalytic dehydrogenation reactions of methanol.4 Krische and coworkers pioneered the study of Ir-catalyzed direct C–C coupling of methanol with reactive π-unsaturated reactants (1,3-dienes, 1,3-enynes and allenes).5 The groups of Glorius,6 Donohoe,7 Obora,8 Andersson9 and others10 demonstrated the direct methylation of ketones or amines using methanol. Despite these achievements, the catalytic C–C bond coupling reactions with methanol are still extremely rare and are limited to the use of precious and noble metal-catalysts such as Ru, Rh or Ir.11 The development and use of cheap and abundant metal catalysts for methanol activation is uphill and remains an important field that urgently needs to be developed.On the other hand, allylic alcohols are highly versatile building blocks in organic synthesis and the pharmaceutical industry, and much effort has been devoted to their synthesis. Among them, nickel-catalyzed reductive coupling of alkynes and aldehydes represents an effective and powerful method. However, this method generally requires the use of stoichiometric reducing reagents that are air-sensitive, metallic or pyrophoric (e.g. ZnR2, BEt3, and R3SiH, Scheme 1a).12 The direct cross-coupling of alcohols and alkynes to synthesize allylic alcohols without the use of any reductant or oxidant represents a significant advancement (Scheme 1b).13 However, this approach still poses many limitations that will require considerable effort to overcome. (1) Alkynes are limited to dialkyl alkynes, and poor regioselectivities were observed for unsymmetrical alkynes, which greatly limits the scope of application of the reaction. (2) Alcohols are restricted to active benzyl alcohols and higher alcohols. The direct cross-coupling of alkynes with methanol has not yet been reported.Open in a separate windowScheme 1Synthesis of allylic alcohols by Ni-catalyzed coupling reaction with alkynes.Although the alkyne–paraformaldehyde reductive coupling has been developed,14 the paraformaldehyde was itself prepared from synthesis gas (through methanol). Therefore, the development of a new strategy for the direct coupling of alkynes and methanol without the use of any reductant or oxidant is still of great value, but also extremely challenging: (1) alkynes are reactive and could rapidly dimerize to 1,3-dienes15 or cyclotrimerize to aromatic ring derivatives in the interaction with nickel.16 (2) Unsymmetric alkynes could result in a mixture of regioisomers that are difficult to separate. (3) The activation energy of methanol in the dehydrogenation process (ΔH = +84 kJ mol−1) is significantly higher than that of higher alcohols or even ethanol (ΔH = +68 kJ mol−1).17Herein we report the nickel-catalyzed direct and regioselective hydrohydroxymethylation of alkynes for the first time using methanol as a C1-feedstock, providing a broad and efficient approach for the synthesis of high added-value allylic alcohols in a high atom-, step- and redox-economic manner. In addition, a cascade cyclization reaction of alkynones and methanol has also been developed for the synthesis of [5,6]-bicyclic hemiacetals in good yields and excellent regio- and diastereoselectivity (Scheme 1c).In our initial experiments, we chose unsymmetrical internal alkyne 1a as a model substrate to optimize the reaction conditions (13a Even if the reaction temperature was increased to 100 °C, only a trace amount of allylic alcohol product 2a was observed (entry 2), which indicates that the use of methanol in the catalytic C–C coupling reactions is indeed a big challenge. Various N-heterocyclic carbene ligands (L2–L6) were investigated (entries 3–9). We found that the selectivity of allylic alcohol 2a is challenged by a number of side reactions, such as the hydrogenation (3a), dimerization (4a) and trimerization (5a) of alkyne 1a. L4 is the most effective, providing 2a with the highest yield (40%) and excellent regioselectivity (14/1), but an appreciable quantity of dimerization and trimerization by-products 4a and 5a was still obtained (entry 5). Krische14 reported that PCy3 could promote the reductive coupling of alkynes and paraformaldehyde, but we found that it is not effective for alkyne–methanol coupling (entry 8).Optimization of reaction conditionsa
EntryLigandAdditiveYieldb (2, %)Yieldb (3, %)Yieldb (4, %)Yieldb (5, %)
1c,d L1 No reaction
2c L1 66<220
3 L2 30 (14/1)e151320
4 L3 35<243
5 L4 40 (14/1)e92223
6 L5 30 (7/1)e61827
7 L6 No reaction
8Cy3PComplicated
9PPh3104<259
10 L4 A1 10<2<2<2
11 L4 A2 38 (14/1)e<2<2<2
12 L4 A3 60f (14/1)e6<2<2
13 L4 A4 No reaction
14 L4 A3 g54f (14/1)e867
Open in a separate windowaReactions conditions: 1a (0.2 mmol), Ni(COD)2 (10 mol%), ligand (20 mol%), tBuOK (12 mol%), additive (1 equiv.) in toluene (1 mL) and MeOH (3 mL) in a sealed tube at 100 °C.bDetermined by GC analysis using adamantane as the internal standard.cWithout tBuOK.dRoom temperature.eRegioselectivity (2a/2a′).fIsolated yield.g0.2 equivalent.Many examples have reported that olefins can affect the outcomes of transition metal-catalyzed cross coupling reactions through increased activity, stability, or selectivity.18 More recently, Montgomery et al.19 found that adding electron-deficient olefins to NHC–Ni(0) complexes can improve their catalytic performance. Inspired by this discovery, we examined various acrylates A1–A4 (entries 10–13). Excitingly, the addition of methyl methacrylate (A3) can indeed significantly improve the chemoselectivity of the reaction, providing the allylic alcohol 2a in 60% isolated yield and a more than 14/1 ratio of regioisomers (entry 12). The structure of acrylates has a great influence on the reaction outcome, indicating that they may act as additional ligands to coordinate with the nickel catalyst, thereby suppressing these undesired dimerization or cyclotrimerization side reactions. However, by-products formed by the reductive coupling of acrylates and alkynes have also been observed (see Section 3 in the ESI).20 It is worth mentioning that stoichiometric acrylate additives are not necessary. As shown in entry 14, even if 0.2 equivalent of A3 was used, 54% of the target product 2a can be obtained, thus showing the subtleties of our catalytic system.With the optimized reaction conditions in hand, we turned our attention to explore the substrate scope of alkynes (Scheme 2). We were pleased to find that various unsymmetrical aryl–alkyl alkynes were coupled with methanol to provide the corresponding allylic alcohols 2a–2q in moderate to good yields and high regioselectivities. Various functional groups, such as fluorine (2b), trifluoromethyl (2c), chlorine (2e), allyl (2f), bromine (2g), amine (2h–2j) and amide (2k and 2l) could all be well-tolerated. Heteroaromatic ring-substituted alkynes, such as 5-indole,21 2-dibenzothiophene and 2-dibenzofuran could also proceed smoothly to furnish allylic alcohols 2n–2p in 40–63% yield. It is worth mentioning that complex biologically active molecules such as estrone derivatives, could also be successfully incorporated into the desired product 2q in 62%, thus demonstrating the robustness and generality of this methodology for late-stage modification of complex biologically active molecules. Terminal alkynes were also found to be compatible with the reaction conditions, providing the corresponding products 2r–2s in moderate yields and excellent regioselectivity (>20/1). Symmetric diarylalkynes bearing electron-donating or electron-withdrawing groups were applicable to the reaction (2t–2y). Strikingly, both 1,2-di(furan-2-yl)ethyne and 1,2-di(thiophen-2-yl)ethyne were competent substrates and furnished the desired allylic alcohols 2x–2y in good yields.Open in a separate windowScheme 2Substrate scope of alkynes for the synthesis of allylic alcohols. Reactions were carried out with 1 (0.2 mmol), Ni(COD)2 (10 mol%), L4 (10 mol%), tBuOK (12 mol%), and methyl methacrylate (0.2 mmol) in toluene (1.0 mL) and MeOH (3.0 mL) in a sealed tube at 100 °C. Isolated yields are given. a The reaction was conducted with Ni(COD)2 (15 mol%), L4 (15 mol%), and tBuOK (18 mol%). b ((4-Bromophenyl)ethynyl)trimethylsilane was used. c The reaction was conducted with toluene (0.5 mL) and MeOH (1.5 mL).In addition, this transformation is not restricted to aryl-substituted alkynes. As shown in Scheme 2, oct-4-yne and cyclododecane were coupled with methanol to produce allylic alcohols 2z and 2aa in 87% and 54% yields, respectively. To further evaluate the influence of the electronic properties of the substituents on the regioselectivity, we tested the hydrohydroxymethylation reaction of unsymmetrical dialkyl-substituted alkynes bearing benzyloxy or dibenzylamino groups at the propargylic position. To our delight, the corresponding allylic alcohols 2ab–2ad were obtained in moderate yields, with remarkably high regioselectivity (>20/1). However, the regioselectivity of this reaction was decreased by the alkyne bearing a benzyloxy group at the homopropargylic position.To expand the potential synthetic applications of the transformation, we investigated the hydrohydroxymethylation of 1,3-enynes. The corresponding dienol 2af was obtained, which was selectively hydrohydroxymethylated on the alkyne but not on the alkene moiety. 1,6-Enyne was also compatible to give the corresponding allylic alcohol 2ag in 42% yield with >20/1 regioselectivity. This strategy can serve as a powerful supplement to the previous method reported by Krische et al.,22 in which alcohols were reacted with alkenes to obtain the corresponding homopropargylic alcohols.23Alkynone substrates were also tested, but the expected product was not detected due to their sensitivity to base. After slightly modifying the reaction conditions, we were pleased to find that various [5,6]-bicyclic hemiacetals 7 could be obtained in good yields with excellent regio- and diastereoselectivities through the cascade cyclization reaction of alkynones 6 with methanol (Scheme 3). We first explored the influence of the substituents (R1) at the terminus of the triple bond. A variety of para-substituted aromatic rings at the alkyne terminus could undergo tandem cyclization to provide the target hemiacetals 7b–7g in 54–78% yields. The structure of 7a was confirmed by an X-ray crystal diffraction study. The aryl groups with substituents at the meta and ortho position were also found to be compatible, leading to the corresponding products 7h–7j in 56–74% yields. Moreover, various (hetero)aryl rings such as naphthalene (7k), benzodioxan (7l), 3,4-dihydrobenzodioxine (7m), thiophene (7n), dibenzofuran (7o), dibenzothiophene (7p), indole (7q) and pyridine (7r) at the terminal of the triple bond could be successfully incorporated into the desired products in good yields. Strikingly, estrone was also compatible with this transformation to afford the desired product 7s in 65% yield. However, no desired product was observed when the methyl substituted alkynone substrate was used. We then investigated the influence of the substituents (R2) at the 2-position of the cyclopentane-1,3-diones. Ethyl, benzyl, and allyl were all well tolerated leading to the corresponding [5,6]-bicyclic hemiacetals 7t–7w in moderate yields.Open in a separate windowScheme 3Substrate scope of alkynones for the synthesis of [5, 6]-bicyclic hemiacetals. Reactions were carried out with 6 (0.2 mmol), Ni(COD)2 (15 mol%), IMes (15 mol%), LiF (10 mol%), and methyl methacrylate (0.2 mmol) in toluene (1.5 mL) and MeOH (0.5 mL) in a sealed tube at 40 °C. Isolated yields are given.To provide a deeper insight into the reaction mechanism, deuterium-labelling experiments were performed. 1a was reacted with CH3OD under our standard reaction conditions; however, no incorporation of deuterium was detected in product 2a (Scheme 4a), revealing that the hydroxyl of methanol is not the proton source. This result is different from the previous report by Zhou et al.,24 in which the Ni(0) catalyst underwent oxidative addition to the O–H bond of methanol to form methoxyl nickel hydride species and then migratory insertion into unsaturated bonds. Further investigation using CD3OD as solvent provided 2a-D in 41% yield, in which 99% of the deuterium was incorporated into the olefinic position, but the reaction rate is obviously slowed down (Scheme 4b). We also conducted the kinetic isotope effect (KIE) experiment. The intermolecular competition reaction between 1a and CD3OD or CH3OH under standard reaction conditions provided a KIE (kH/kD) value of 6.1 (Scheme 4c). Taken together, these results may indicate that the dehydrogenation of methanol to form the key formaldehyde intermediate is the rate-determining step of this transformation.Open in a separate windowScheme 4Deuterium-labelling experiments.On the basis of these experimental results and previous observations, a possible reaction mechanism is proposed in Scheme 5. The reaction is initiated by reducing alkyne to alkene and simultaneously oxidizing methanol to formaldehyde, as evidenced by the detection of catalytic amounts of alkene 3. Oxidative cyclization of acrylate-coordinated NHC–Ni(0) A 17 with alkyne and formaldehyde gives oxa-nickelacycle intermediate B. Subsequent protonation of nickelacycle species B with methanol affords the vinylnickel intermediate C, which can undergo β-H elimination to generate vinyl nickel hydride species D and formaldehyde.25 Reductive elimination of D will furnish allylic alcohol 2 and the catalytically active Ni(0) catalyst A. Further nucleophilic addition of the hydroxyl group to one of the ketone carbonyl groups will produce [5,6]-bicyclic hemiacetal 7. We speculate that the acrylate is used as an additional ligand, thereby inhibiting the alkyne dimerization to 1,3-dienes or cyclotrimerization to aromatic ring derivatives.Open in a separate windowScheme 5Proposed reaction mechanism.  相似文献   

6.
Redox-neutral manganese-catalyzed synthesis of 1-pyrrolines     
Tingting Feng  Canxiang Liu  Zhen Wu  Xinxin Wu  Chen Zhu 《Chemical science》2022,13(9):2669
This report describes a manganese-catalyzed radical [3 + 2] cyclization of cyclopropanols and oxime ethers, leading to valuable multi-functional 1-pyrrolines. In this redox-neutral process, the oxime ethers function as internal oxidants and H-donors. The reaction involves sequential rupture of C–C, C–H and N–O bonds and proceeds under mild conditions. This intermolecular protocol provides an efficient approach for the synthesis of structurally diverse 1-pyrrolines.

Described herein is a novel manganese-catalyzed radical [3 + 2] cyclization of cyclopropanols and oxime ethers, leading to valuable multi-functionalized 1-pyrrolines.

Pyrroline and its derivatives appear frequently as the core of the structure of natural products and biologically active molecules (Fig. 1A).1 Such compounds also serve as versatile feedstocks in various transformations, such as 1,3-dipolar cyclization, ring opening, reduction and oxidation, leading to diverse and valuable compounds.2–4 Over the past few decades, great effort has been devoted to the preparation of pyrrolines. This has resulted in several elegant approaches that rely on photoredox catalysis (Fig. 1B).5 The groups of Studer,5a,b Leonori,5c and Loh5df disclosed intramolecular addition of the intermediate iminyl radical to alkenes to construct pyrrolines. Generally, the synthetic value of a method can be further improved by using an intermolecular reaction pattern. For example, Alemán et al. recently reported a radical-polar cascade reaction involving the addition to ketimines of alkyl radicals formed in hydrogen atom transfer (HAT) reactions.5g That the existence of benzylic C–H bonds in the substrates is requisite for the HAT, compromises the substrate scope. Despite the appealing photochemical processes, development of new redox approaches to enrich the product diversity of pyrrolines, especially with inexpensive transition-metal catalysts, is still in demand.Open in a separate windowFig. 1(A) Importance of pyrrolines, and (B and C) synthetic approaches to pyrrolines.Prompted by extensive applications of cyclopropanols in synthesis6 and our achievements in manganese-catalyzed ring-opening reactions,7 we conceived a radical [3 + 2] cyclization using cyclopropanol as a C3 synthon and oxime ethers as a nitrogen source (Fig. 1C). Hypothetically, single-electron oxidation of cyclopropanol by Mnn generates the β-keto radical (I), which undergoes a radical [3 + 2] cascade reaction with an oxime ether to give the alkoxy radical species (II). Conversion of II to the intermediate (III), the pyrroline precursor, requires an extra H-donor to support a HAT process and an oxidant for recovery of Mnn to perpetuate the catalytic cycle. In this scenario, the strategic inclusion of oxime ether is crucial to the overall transformation. The oxime ether is not only an internal oxidant and H-donor, but should also be subject to in situ deprotection by cleaving the N–O bond during the reaction. The choice of a proper Mnn/Mnn−1 pair with suitable redox potentials is also vital to the catalytic cycle.Herein, we provide proof-of-principle studies for this hypothesis. The desired radical [3 + 2] cyclization of cyclopropanols and O-benzyl oxime ethers is accomplished with manganese catalysis. This redox-neutral process involves sequential rupture of C–C, C–H, and N–O bonds under mild conditions. The intermolecular protocol provides an ingenious approach to the synthesis of multi-functionalized 1-pyrrolines.With these considerations in mind, phenylcyclopropanol (1a) and oxime ether (2a) were initially chosen as model substrates to evaluate reaction parameters in the presence of manganese salt ( Created by potrace 1.16, written by Peter Selinger 2001-2019 N bond of 2a (entry 2). The optimization of organic solvents was then conducted (entries 3–8), and it was found that the use of fluorinated alcohols, such as trifluoroethanol (TFE) and hexafluoroisopropanol (HFIP) as solvents provided excellent yields (entries 7 and 8). Decreasing the amount of Mn(acac)3 to 1.2 equiv. gave a comparable yield (entry 9), but further reducing the amount compromised the yield (entry 10). Replacing Mn(acac)3 with Mn(OAc)3 or MnCl2 significantly decreased the reaction yield (entries 11 and 12). However, the use of Mn(acac)2 gave a similar yield to Mn(acac)3 (entries 13 vs. 9). The above results prompted us to think over the counteranion effect that the acetylacetone (acac) anion may be requisite to the reaction. Indeed, the synergistic use of stoichiometric MnCl2 and acetylacetone led to a good yield of the desired product (entry 14). More importantly, a comparable yield was obtained with only 0.2 equiv. of MnCl2 and added acetylacetone, realizing this reaction under a catalytic amount of Mn salts (entry 15). Given that the low solubility of the Mn salt may lead to poor efficiency, a reaction with 0.067 M concentration was carried out and gave a 89% yield (entry 16). Further reducing the amount of acetylacetone to 1.0 equiv. had no influence on the outcome of the reaction (entry 17), but the reaction efficiency slightly decreased when 0.6 equiv. of acetylacetone was used as the additive (entry 18). Use of a decreased amount (1.0 equiv.) of acetic acid led to the best yield (91%, entry 19), whereas the reaction in the presence of 0.5 equiv. acetic acid (entry 20) or without acetic acid (entry 21) also gave high yields. It is noted that acetic acid is not crucial to the reaction using MnCl2 as catalyst, as the reaction could generate cat. HCl in situ. The reaction with substoichiometric amount (0.6 equiv.) of acac gave a decreased but also good yield (entry 22). Reducing the catalytic loading of MnCl2 to 10 mol% slightly compromised the yield (entry 23).Optimization of the synthesis of 1-pyrrolines
EntryaMn salt (equiv.)Additive (equiv.)SolventYield (%)
1Mn(acac)3 (1.7)NoneCH3CN33
2bMn(acac)3 (1.7)NoneCH3CNTrace
3Mn(acac)3 (1.7)NoneDCM31
4Mn(acac)3 (1.7)NoneAcetone25
5Mn(acac)3 (1.7)NoneDMSOTrace
6Mn(acac)3 (1.7)NoneDMFTrace
7Mn(acac)3 (1.7)NoneTFE80
8Mn(acac)3 (1.7)NoneHFIP82
9Mn(acac)3 (1.2)NoneHFIP83
10Mn(acac)3 (0.9)NoneHFIP55
11Mn(OAc)3·2H2O (1.2)NoneHFIP36
12MnCl2 (1.2)NoneHFIPTrace
13Mn(acac)2 (1.2)NoneHFIP88
14MnCl2 (1.2)acac (3.6)HFIP80
15MnCl2 (0.2)acac (3.6)HFIP81
16cMnCl2 (0.2)acac (3.6)HFIP89
17cMnCl2 (0.2)acac (1.0)HFIP89
18cMnCl2 (0.2)acac (0.6)HFIP83
19c,dMnCl2 (0.2)acac (1.0)HFIP91
20c,eMnCl2 (0.2)acac (1.0)HFIP83
21c,bMnCl2 (0.2)acac (1.0)HFIP80
22c,dMnCl2 (0.2)acac (0.6)HFIP82
23c,dMnCl2 (0.1)acac (1.0)HFIP81
Open in a separate windowaReaction conditions: 1a (0.45 mmol), 2a (0.3 mmol), AcOH (2.0 equiv.), and Mn salt (as shown) in solvent (2.0 mL), at room temperature (rt) under N2, for 16 h.bWithout AcOH.c0.067 M reaction.d1.0 equiv. AcOH.e0.5 equiv. AcOH. acac = acetylacetone.With the optimized conditions in hand for the synthesis of 1-pyrrolines, the compatibility of various cyclopropanols was inspected (Scheme 1). Common functional groups on the phenyl ring, including halides (3b–3d), ester (3f), ether (3j), were compatible under the reaction conditions. Regardless of the presence of electron-withdrawing or -donating substituents at the para-position of this phenyl ring, the reactions readily proceeded with generally high yields (3b–3j). The cyclopropanol (1k) with an ortho-methyl substituent underwent a cyclization reaction with excellent yield, demonstrating that steric effects had little effect on product of the reaction (3k). By replacing the phenyl group with a naphthyl or thienyl group, the corresponding products (3l and 3m) were produced with slightly lower yields. When 2-substituted cyclopropanols were utilized, these reactions gave rise to a portfolio of trisubstituted 1-pyrrolines (3n–3u).The relative configuration of 3u was determined by comparison with a reported structure.8 Remarkably, this protocol provided a convenient method for the construction of an N-containing spiro skeleton (3t). The reaction with alkyl cyclopropanols could also furnish the desired products (3v–3x) smoothly and with good yields.Open in a separate windowScheme 1Scope of cyclopropanols. Reaction conditions: 1 (0.3 mmol), 2a (0.2 mmol), AcOH (0.2 mmol), MnCl2 (0.04 mmol), and acac (0.2 mmol) in HFIP (3.0 mL), at rt under N2. The d.r. values were determined by 1H NMR analysis with crude reaction mixture, and major isomers are shown with relative configurations. aThe reaction is scaled up for 10 times.Next, we studied the scope of oxime ethers (Scheme 2). Steric hindrance from the ester moiety in the oxime ethers appeared not to influence the reaction outcome. Oxime ethers bearing various esters, such as phenyl (3y), biphenyl (3z and 3ab), 2-naphthyl (3aa), 2,4-di-tert-butylphenyl (3ac and 3ad), and 2,6-dimethylphenyl (3ae) esters all reacted smoothly. In addition, the substrate with tert-butyl ester also readily underwent cyclization to afford the desired product 3af with excellent yield. Remarkably, the trifluoromethyl-substituted pyrroline (3ag) was afforded almost quantitatively from the corresponding ketoxime ether. However, if the trifluoromethyl group was replaced by a methyl or phenyl group, the reaction failed to give rise to the desired product (3ah or 3ai), and this might be attributed to poorer electrophilic nature of the methyl or phenyl substituted substrate.Open in a separate windowScheme 2Scope of oxime ethers. Reaction conditions: 1 (0.3 mmol), 2 (0.2 mmol), AcOH (0.2 mmol), MnCl2 (0.04 mmol), and acac (0.2 mmol) in HFIP (3.0 mL), at rt under N2. The d.r. values were determined by 1H NMR analysis with crude reaction mixture, and major isomers are shown with relative configurations.To illustrate the utility of this protocol, we carried out a set of synthetic applications using 1-pyrroline (3a) (Scheme 3). Upon treatment with acetyl chloride and pyridine at 42 °C, 1-pyrroline (3a) could be readily converted into the acyclic amino acid derivative (4). The reaction between 3a and LiAlH4 gave rise smoothly to the corresponding alcohol (5). In the presence of 2,3-dichloro-5,6-dicyano-1,4-benzoquin-4-one (DDQ) and triethylamine, the 2,5-disubstituted pyrrole (6) was obtained. Moreover, treatment of 3a with MeOTf and NaBH4 delivered the N-methyl proline derivative (7).9Open in a separate windowScheme 3Synthetic applications. Reaction conditions: (a) AcCl, pyridine, dry DCE, 42 °C, 63% yield; (b) LiAlH4, THF, reflux, 90% yield; (c) DDQ, Et3N, DCM, rt, 53% yield; (d) MeOTf, DCM, and then NaBH4, THF, 40% yield, cis : trans = 6.6 : 1.To probe the mechanistic pathways, we performed a radical trapping experiment in the presence of 2.0 equiv. of radical scavenger TEMPO. The radical trapping product (8) was detected by high-resolution mass spectrometry (HRMS) (Scheme 4A, top). In addition, the reaction was obviously suppressed when 1,1-diphenylethylene was added under standard condition (Scheme 4A, bottom). These results suggested that this process engaged in a radical pathway. Kinetic studies illustrated that the reaction immediately started with 20 mol% Mn(acac)2 but an approximate 15 min of induction period was appeared by using Mn(acac)3, which probably indicated that the reaction was initiated with Mn(ii) rather than Mn(iii), and the Mn(ii)/Mn(i) cycle might be involved in the transformation (Scheme 4B, for details see ESI).Open in a separate windowScheme 4(A and B) Mechanistic studies, and (C) proposed mechanism.On the basis of these results, a plausible mechanism for this radical process was proposed in Scheme 4C. Initially, the interaction between cyclopropanol (1a) and Mn(ii) salt gives rise to the alkoxy manganese species (I), which undergoes a ligand-to-metal charge transfer (LMCT) process, leading to the alkoxy radical (II).5f Subsequent ring-opening of the alkoxyl radical (II) provides the alkyl radical (III). The addition of intermediate (III) to the oxime ether, possibly activated by HFIP or HOAc, furnishes the N-centered radical (IV), which intramolecularly attacks the ketone to afford a new alkoxy radical (V).10 The subsequent 1,5-hydrogen atom transfer (HAT) process delivers the alkyl radical (VI) at the α-position adjacent to the O atom, thus driving N–O bond cleavage to generate the N-centered radical (VII),5b,11 and benzaldehyde which was detected by TLC. This radical intermediate (VII) undergoes a single electron transfer (SET) mediated by the reduced-state Mn(i) species, and protonation to yield the cyclic pyrrolidine (VIII). Finally, dehydration of this intermediate produces 1-pyrroline (3a).  相似文献   

7.
Dirhodium(ii)-catalysed cycloisomerization of azaenyne: rapid assembly of centrally and axially chiral isoindazole frameworks     
Shaotong Qiu  Xiang Gao  Shifa Zhu 《Chemical science》2021,12(41):13730
Described herein is a dirhodium(ii)-catalyzed asymmetric cycloisomerization reaction of azaenyne through a cap-tether synergistic modulation strategy, which represents the first catalytic asymmetric cycloisomerization of azaenyne. This reaction is highly challenging because of its inherent strong background reaction leading to racemate formation and the high capability of coordination of the nitrogen atom resulting in catalyst deactivation. Varieties of centrally chiral isoindazole derivatives could be prepared in up to 99 : 1 d.r., 99 : 1 er and 99% yield and diverse enantiomerically enriched atropisomers bearing two five-membered heteroaryls have been accessed by using an oxidative central-to-axial chirality transfer strategy. The tethered nitrogen atom incorporated into the starting materials enabled easy late-modifications of the centrally and axially chiral products via C–H functionalizations, which further demonstrated the appealing synthetic utilities of this powerful asymmetric cyclization.

Rh(ii)-catalyzed asymmetric cycloisomerization of azaenyne through a cap-tether synergistic modulation strategy was described. Diverse centrally and axially chiral isoindazoles were prepared and directed C–H late-stage modifications were developed.

Known as one of the most significant and reliable access methods to chiral heterocycles, asymmetric cycloisomerization of conjugated enyne has caught extensive attention and interest for its wide applications in synthetic route design and mechanistic investigation.1 Specifically, asymmetric cyclization of conjugated enynone (X = C, Z = O) has been successfully developed and applied to the rapid construction of various chiral furan-containing skeletons with high efficiency in an extremely operationally simple manner (Scheme 1a).2 However, compared to the fruitful research with enynone, it is surprising that the analogous asymmetric version of azaenyne (Z = N–R) still remains underdeveloped.3 In fact, no successful example of catalytic asymmetric cyclization of azaenyne has been reported in the literature despite the apparent significance of nitrogen-containing five-membered heterocycles in the synthetic and pharmaceutical community.4 In 2004, Haley and Herges reported a detailed experimental and theoretical study of the cyclization reaction of (2-ethynylphenyl)-phenyldiazene, which is a unique azaenyne.5 According to the DFT calculations, very close and low activation barriers for 5-exo-dig and 6-endo-dig cyclization pathways under catalyst-free conditions were found, which shed light on the inherent challenges of the asymmetric reaction of azaenyne (Scheme 1b). For instance, there was usually a regioselectivity issue (5-exo and 6-endo) in the cyclization reaction of azaenyne because of their close reaction barriers where the competitive 6-endo-dig cyclization3a,6 may lead to troublesome side-product formation. In addition, the low activation barrier deriving from the strong N-nucleophilicity of azaenyne may easily lead to self-cyclization which will cause severe background reactions to interfere with the asymmetric process. More troublingly, this transformation might suffer from catalyst deactivation arising from the high coordinating capability of the nitrogen atom in both starting materials and products, which might give more opportunities to the propagation of detrimental background reactions. In some cases, even a super-stoichiometric amount of transition metal has to be used to ensure effective conversion.3a,7 Therefore, although many nonchiral approaches have been reported,3,5 catalytic asymmetric cyclization of azaenyne still remains elusive due to the inherent obstacles aforementioned. With our continuous interest in alkyne chemistry,2a,8 herein we designed a cap-tether synergistic modulation strategy to tackle these challenges, envisioning that modulation of the tethered atom and protecting cap of nitrogen in the azaenyne would intrinsically perturb and alter the reactivity of the starting material, and therefore the azaenyne motif could be effectively harnessed as a promising synthon for asymmetric transformations (Scheme 1c). It should be noted that the obtained centrally chiral product produced from intramolecular C–H insertion of donor-type metal carbene9 might be potentially converted into the axially chiral molecule via a central-to-axial chirality conversion strategy.Open in a separate windowScheme 1Development of the asymmetric cyclization reaction of conjugated azaenyne.With this design in mind, different types of azaenynes bearing typical tethering atoms and capping groups were chosen to test our hypothesis and representative results are shown in Scheme 2. First, tBu-capping imine (X = C, R = tBu) was selected as a substrate to test our hypothesis.6a It was found that the imine exhibited low reactivity and the reaction temperature has to be elevated to 100 °C to initiate the transformation with or without catalyst. Unfortunately, the desired 5-exo-dig cyclization product was not detected, but isoquinoline from 6-endo-dig cyclization was obtained instead (Scheme 2a). To further regulate and control the regioselectivity and reactivity, triazene (X = N, R = N-piperidyl) was then investigated. Similarly, this substrate also showed low reactivity and it is still required to be heated at 100 °C for conversion. In the absence of a metal catalyst, an unexpected alkyne, deriving from the fragmentation of the triazene moiety, was produced in 41% yield. When 2 mol% Rh2(OPiv)4 was added as a catalyst, the side reaction could be efficiently suppressed and the reaction selectivity was apparently reversed. In this case, the target C–H insertion dihydrofuran was furnished as the major product in 30% yield but still accompanied by concomitant formation of 12% yield of undesired alkyne (Scheme 2b). The above investigations showed neither the imine nor triazene was an ideal substrate for the asymmetric reaction. Thus, we moved our attention to the diazene substrate (X = N, R = aryl). As demonstrated by Haley''s and Herges'' pioneering work, ortho-alkynyl diazene, compared with imine and triazene, was more unstable and tended to self-cyclization even at room temperature.5a As shown in Scheme 2c, the ortho-alkynyl diazene degrades and 5-exo-dig cyclization products could be observed even in DCE solvent without any catalyst at room temperature. When the phenyl capping group was installed in the substrate, the reaction furnished 10% yield of isoindazole derivative. The uncatalyzed self-cyclization reaction was obviously accelerated when an electron-rich capping group (4-MeO–C6H4–) was introduced, affording the corresponding product in 20% yield. Inspired by these findings, we assumed that installation of an electron deficient group on the capping phenyl would reduce the nucleophilicity of the nitrogen atom and thus the troublesome self-cyclization reaction might be effectively inhibited. To our delight, when a bromo-substituent was introduced onto the phenyl cap, the undesired self-cyclization was almost suppressed. When Rh2(OPiv)4 was added as a catalyst, the desired carbene-involved C–H insertion product was furnished in 90% yield at room temperature. Worthy of note was the total absence of any cinnoline formation from 6-endo-dig cyclization.3a,6b In short, the synthetic challenges associated with regioselectivity (5-exo-dig and 6-endo-dig), strong background reaction and catalyst deactivation could be successfully regulated and controlled via a tether-cap synergistic modulation strategy.Open in a separate windowScheme 2Typical substrate investigation.Encouraged by the above findings, ortho-alkynyl bromodiazene 1a was chosen as a model substrate and different types of chiral dirhodium catalysts10 were screened in DCE at room temperature for 48 h. As shown in EntryRh(ii)*SolventYieldb [%]erc1Rh2(R-DOSP)4DCE5629 : 712Rh2(5S-MEPY)4DCE1750 : 503Rh2(S-BTPCP)4DCE618 : 924Rh2(S-PTPA)4DCE9191 : 95Rh2(S-PTTL)4DCE8697 : 36Rh2(S-PTAD)4DCE9394 : 67Rh2(S-NTTL)4DCE9296 : 48Rh2(S-TCPTTL)4DCE9598 : 2 9 Rh 2 (S-TFPTTL) 4 DCE 98 d 98 : 210Rh2(S-TFPTTL)4DCM8898 : 211Rh2(S-TFPTTL)4Toluene9298 : 212Rh2(S-TFPTTL)4MeCN1692 : 813Rh2(S-TFPTTL)4 n-Hexane9698 : 214eRh2(S-TFPTTL)4DCE65f96 : 4 Open in a separate windowaUnless otherwise noted, reactions were performed at 0.1 M in DCE using 0.20 mmol substrate and catalyst (2 mol%) under a N2 atmosphere.bDetermined by 1H NMR spectroscopy.cThe er value of 2a was determined by HPLC using a chiral stationary phase.dIsolated yields.e1 mol% catalyst was used.f25% starting material was recovered.With the optimized reaction conditions in hand (Scheme 3, the catalytic process could be successfully applied to azaenynes 1 bearing different ether side chains. For example, in addition to 1a, various azaenyne derivatives containing benzylic ethers could be efficiently converted into the desired products 2b–i with excellent diastereoselectivities and enantioselectivities (>99 : 1 d.r., 97:3–99 : 1 er). The yields were typically higher than 90% for most substrates. Satisfyingly, the substrates with bulkier aryl groups were well-tolerated to afford the isoindazole products 2j–m in good yields with excellent diastereo- and enantiocontrol (>97 : 3 d.r., > 95 : 5 er). In addition to azaenynes with arylmethyl ether, this protocol was also successfully applied to substrates with allylic ether, propargyl ether and even aliphatic ether to furnish the cyclization products 2n–u in good yields with decent diastereo- and enantioselectivities (>93 : 7 d.r., > 90 : 10 er). In the cases of allylic and propargyl ether, only C–H insertion products (2n–p) were observed though cyclopropanation or cyclopropenation often took place competitively when using the allylic or propargyl substrate to trap the carbene intermediate.11 It was noted that the azaenynes with aliphatic ether, which represent challenging substrates2a in the asymmetric carbene transfer reactions, also showed good reactivities to afford the corresponding chiral dihydrobenzofurans (2q–u) with excellent diastereoselectivities (>93 : 7 d.r.) and enantioselectivities (>98 : 2 er). Interestingly, when phenyl and methoxyphenyl capping azaenynes, which potentially suffered from the undesired background reactions, were subjected to the standard conditions, chiral products (2v–w) could be obtained with high optical purity (>99 : 1 d.r., > 96 : 4 er) as well. These results might be attributed to the high catalytic activity of Rh2(S-TFPTTL)4 in the asymmetric cyclization process, which eventually led to complete suppression of the uncatalyzed self-cyclization. The scopes with respect to the group R1 on the fused phenyl ring were further investigated. Both electron-rich and -deficient substituents R1 were well accommodated, with the product yields ranging from 80% to 99%, enantiomeric ratios ranging from 95 : 5 to 97 : 3 and diastereomeric ratios higher than 99 : 1 (2x–z). In addition, azaenyne substituted with an alkyl side chain at the alkynyl carbon atom was also tested, giving tetrahydrofuran (2aa) with excellent diastereoselectivity (>99 : 1 d.r.), good enantioselectivity (90 : 10 er) and moderate yield (43%). In addition to the side chain of ether, this asymmetric protocol could even be extended to the more challenging nitrogen- and thio-tethered analogues, albeit with somewhat lower reactivities (46–65% yields) but good stereoselectivities (93 : 7 er and 84 : 16 d.r. for 2ab; 81 : 19 er and >99 : 1 d.r. for 2ac). Structures of the resulting products were confirmed by X-ray crystallographic analysis of their analogue 2h.Open in a separate windowScheme 3 aUnless otherwise noted, the reactions were performed under standard conditions for 48 h or monitored by TLC until the starting material disappeared. b5 mol% catalyst was used. cReactions were performed in n-hexane, using 2 mol% Rh2(S-TCPTTL)4 as the catalyst.The successful preparation of centrally chiral isoindazole through the asymmetric cyclization reaction prompted us to explore the further applications of this protocol. Axially chiral biaryl skeletons are undoubtedly regarded as one of the most prominent structural motifs for their ubiquity in natural products, pharmaceuticals and useful chiral ligands in asymmetric catalysis.12 Due to the lower rotational barrier, there are only limited examples of the enantioselective synthesis of axially chiral atropisomers featuring a five-membered ring, especially those bearing two pentatomic aromatics.13 Compared with the furan analogue, the extending cap in the isoindazole scaffold provides additional ortho steric hindrance making these molecules possible candidates for the preparation of five-five-membered biaryl atropisomers. Considering the unique chiral skeleton of dihydrofuranyl isoindazole 2, we began to explore their potential application in chiral atropisomer synthesis via a central-to-axial chirality transfer strategy. As shown in Scheme 4, oxidative aromatization of representative dihydrofuran candidate 2m furnished two configurationally unstable atropisomers, which might be attributed to their relatively low rotational barriers as five-membered atropisomers especially when the furan ring was incorporated (see ESI for details). Therefore, it was hypothesized that extending the fused phenyl to naphthyl might afford stable atropisomers by enhancing the ortho steric hindrance (Scheme 4b).Open in a separate windowScheme 4Investigation of central-to-axial chirality transfer.To our delight, as shown in Scheme 5, naphthyl-fused dihydrofurans 4 could be easily accessed through the above established dirhodium-catalyzed cyclization process and configurationally stable atropisomers 5 could be generated via further oxidative dehydrogenation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) as the oxidant (see ESI for the proposed mechanism). For example, asymmetric cyclization reactions proceeded smoothly to give the centrally chiral compounds 4 in good yields (54–99%) with excellent diastereoselectivities (92 : 8–99 : 1 d.r.) and enantioselectivities (95 : 5–99 : 1 er) under slightly modified reaction conditions. This reaction was compatible with a variety of arylmethyl side chains in azaenynes and well-accommodated with various functional groups (F, Cl, Br, OMe, and –CO2Me). Additionally, oxidative dehydrogenation of chiral candidates 4 with DDQ smoothly resulted in the formation of axially chiral atropisomers 5 in 90–99% yields with only slight loss of chiral integrity (90 : 10–97 : 3 er). An enantiomerically pure atropisomer could be obtained through a simple recrystallization procedure as exemplified by compound 5g. The structure and absolute configuration of isoindazole 4g and atropisomer 5g were confirmed by their single-crystal X-ray diffraction analysis.Open in a separate windowScheme 5 aConditions for cyclization of azaenyne: Rh2(S-TCPTTL)4 (2 mol%), n-hexane, rt for 48 h or monitored by TLC until the starting material disappeared; conditions for oxidative chirality transfer: DDQ (2 equiv.), DCE, −20 °C for 48 h or monitored by TLC until the starting material disappeared. b45 °C. cDDQ (5 equiv.). dRoom temperature. eAfter one recrystallization.With centrally and axial chiral molecules in hand, further transformations of these compounds were also explored. The tethered nitrogen atom in azaenynes not only showed a synergetic effect with the capping group on promoting asymmetric cyclization but also served as an innate directing group for late-stage modifications via C–H functionalization. As shown in Scheme 6, a variety of functional groups could be directly introduced onto the capping aromatic rings, allowing for rapid build-up of molecular complexity. For example, synthetically valuable alkenyl,14 allyl15 and alkynyl16 groups could be easily incorporated into the final structures, which had wide potential applications in organic synthesis (6a–c). Furthermore, C–H alkylation,17 amidation18 and selenylation19 were performed smoothly to afford the desired products 6d–g. It is noteworthy that unique chiral chelation backbones were constructed by amidation and selenylation of the isoindazole moiety (6e–g). In addition to centrally chiral compounds, axial chiral atropisomers 5 themselves could be efficiently converted to their functionalized scaffolds as well (6h–i) through a similar directed C–H functionalization process.Open in a separate windowScheme 6Late-stage modification of chiral isoindazoles. Reaction conditions: a4-octyne, [Rh(Cp*Cl2)]2, AgSbF6, Cu(OAc)2, DCE, 80 °C. bAllyl carbonate, [Rh(Cp*Cl2)]2, AgSbF6, PivOH, PhCl, 40 °C. cHypervalent iodine-alkyne, [Rh(Cp*Cl2)]2, Zn(OTf)2, DCE, 80 °C. dAlkene, [Rh(Cp*Cl2)]2, AgSbF6, AcOH, 1,4-dioxane, 50 °C. e3-Phenyl-1,4,2-dioxazol-5-one, [Cp*Co(MeCN)3](SbF6)2, DCE, 80 °C. fPhSeCl, [Rh(Cp*Cl2)]2, AgSbF6, THF, 60 °C.  相似文献   

8.
Three-component 1,2-carboamination of vinyl boronic esters via amidyl radical induced 1,2-migration     
Cai You  Armido Studer 《Chemical science》2021,12(47):15765
Three-component 1,2-carboamination of vinyl boronic esters with alkyl/aryl lithium reagents and N-chloro-carbamates/carboxamides is presented. Vinylboron ate complexes generated in situ from the boronic ester and an organo lithium reagent are shown to react with readily available N-chloro-carbamates/carboxamides to give valuable 1,2-aminoboronic esters. These cascades proceed in the absence of any catalyst upon simple visible light irradiation. Amidyl radicals add to the vinylboron ate complexes followed by oxidation and 1,2-alkyl/aryl migration from boron to carbon to give the corresponding carboamination products. These practical cascades show high functional group tolerance and accordingly exhibit broad substrate scope. Gram-scale reaction and diverse follow-up transformations convincingly demonstrate the synthetic potential of this method.

Three-component 1,2-carboamination of vinyl boronic esters with alkyl/aryl lithium reagents and N-chloro-carbamates/carboxamides is presented.

Alkenes are important and versatile building blocks in organic synthesis. 1,2-Difunctionalization of alkenes offers a highly valuable synthetic strategy to access 1,2-difunctionalized alkanes by sequentially forming two vicinal σ-bonds.1a–h Among these vicinal difunctionalizations, the 1,2-carboamination of alkenes, in which a C–N and a C–C bond are formed, provides an attractive route for the straightforward preparation of structurally diverse amine derivatives (Scheme 1a).2a–c Along these lines, transition-metal-catalyzed or radical 1,2-carboaminations of activated and unactivated alkenes have been reported.3a–p However, the 1,2-carboamination of vinylboron reagents, a privileged class of olefins,4a–h to form valuable 1,2-aminoboron compounds which can be readily used in diverse downstream functionalizations,5a–c,6a–d has been rarely investigated. To the best of our knowledge, there are only two reported examples, as shown in Schemes 1b and c. In 2013, Molander disclosed a Rh-catalyzed 1,2-aminoarylation of potassium vinyltrifluoroborate with benzhydroxamates via C–H activation (Scheme 1b).7 Thus, the 1,2-carboamination of vinylboron reagents is still underexplored but highly desirable.Open in a separate windowScheme 1Intermolecular 1,2-carboamination of alkenes.1,2-Alkyl/aryl migrations induced by β-addition to vinylboron ate complexes have been shown to be highly reliable for 1,2-difunctionalization of vinylboron reagents (Scheme 1c).4dh In 1967, Zweifel''s group developed 1,2-alkyl/aryl migrations of vinylboron ate complexes induced by an electrophilic halogenation.8 In 2016, the Morken group reported the electrophilic palladation-induced 1,2-alkyl/aryl migration of vinylboron ate complexes.9a–k Shortly thereafter, we,10a–c Aggarwal,11a–c and Renaud12 developed alkyl radical induced 1,2-alkyl/aryl migrations of vinylboron ate complexes. In these recent examples, the migration is induced by a C-based radical/electrophile, halogen and chalcogen electrophiles.13a,bIn contrast, N-reagent-induced migration of vinylboron ate complexes proceeding via β-amination is not well investigated. To our knowledge, as the only example the Aggarwal laboratory described the reaction of a vinylboron ate complex with an aryldiazonium salt as the electrophile, but the desired β-aminated rearrangement product was formed in only 9% NMR yield (Scheme 1c).13a No doubt, β-amino alkylboronic esters would be valuable intermediates in organic synthesis. Encouraged by our continuous work on amidyl radicals14a–i and 1,2-migrations of boron ate complexes,10a–c,15a–f we therefore decided to study the amidyl radical-induced carboamination of vinyl boronic esters for the preparation of 1,2-aminoboronic esters. N-chloroamides were chosen as N-radical precursors,16a–c as these N-chloro compounds can be easily prepared from the corresponding N–H analogues.17 Herein, we present a catalyst-free three-component 1,2-carboamination of vinyl boronic esters with N-chloroamides and readily available alkyl/aryl lithium reagents (Scheme 1d).We commenced our study by exploring the reaction of the vinylboron ate complex 2a with tert-butyl chloro(methyl)carbamate 3a applying photoredox catalysis. Complex 2a was generated in situ by addition of n-butyllithium to the boronic ester 1a in diethyl ether at 0 °C. After solvent removal, the photocatalyst fac-Ir(ppy)3 (1 mol%) and THF were added followed by the addition of 3a. Upon blue LED light irradiation, the mixture was stirred at room temperature for 16 hours. To our delight, the desired 1,2-aminoboronic ester 4a was obtained, albeit with low yield (26%, EntryPhotocatalystSolventT (°C)Yieldb (%)1 fac-Ir(ppy)3THFrt262 fac-Ir(ppy)3DMSOrt23 fac-Ir(ppy)3MeCNrt564Ru(bpy)3Cl2·6H2OMeCNrt695Na2Eosin YMeCNrt696cNa2Eosin YMeCNrt707cNoneMeCNrt458cNoneMeCN0789cNoneMeCN−2088 (85)10c,dNoneMeCN−202Open in a separate windowaReaction conditions: 1a (0.20 mmol), nBuLi (0.22 mmol), in Et2O (2 mL), 0 °C to rt, 1 h, under Ar. After vinylboron ate complex formation, solvent exchange to the selected solvent (2 mL) was performed.bGC yield using n-C14H30 as an internal standard; yield of isolated product is given in parentheses.c4 mL MeCN was used.dReaction carried out in the dark.With optimal conditions in hand, we then investigated the scope of this new 1,2-carboamination protocol keeping 2a as the N-radical acceptor (Scheme 2). This transformation turned out to be compatible with various primary amine reaction partners bearing carbamate (4a, 4b and 4d–4g) or acyl protecting groups (4c) (20–85%). Notably, N-chlorolactams can be used as N-radical precursors, as shown by the successful preparation of 4h (71%). Moreover, Boc-protected ammonia was also tolerated, delivering 4i in an acceptable yield (55%).Open in a separate windowScheme 21,2-Carboamination of 1a with various amidyl radical precursors. Reaction conditions: 1a (0.20 mmol, 1.0 equiv.), nBuLi (0.22 mmol, 1.1 equiv.), in Et2O (2 mL), 0 °C to rt, 1 h, under Ar; then [N]-Cl (0.24 mmol, 1.2 equiv.), −20 °C, 16 h, in MeCN (4 mL). Yields given correspond to yields of isolated products. aA solution of [N]-Cl (0.30 mmol, 1.5 equiv.) in MeCN (1 mL) was used. See the ESI for experimental details.We continued the studies by testing a range of vinylboron ate complexes (Scheme 3). To this end, various vinylboron ate complexes were generated by reacting the vinyl boronic ester 1a with methyllithium, n-hexyllithium, isopropyllithium and tert-butyllithium. For the n-alkyl-substituted vinylboron ate complexes, the 1,2-carboamination worked smoothly to afford 4j and 4k in good yields. However, the vinylboron ate complex derived from isopropyllithium addition provided the desired products in much lower yield (4l, 18% yield). When tert-butyllithium was employed, only a trace of the targeted product was detected (see ESI). As expected, cascades comprising a 1,2-aryl migration from boron to carbon worked well. Thus, by using PhLi for vinylboron ate complex formation, the 1,2-aminoboronic esters 4m–4o were obtained in 69–73% yields with the Boc (t-BuOCONClMe), ethoxycarbonyl-(EtOCONClMe) and methoxycarbonyl (Moc)-(MeOCONClMe) protected N-chloromethylamines (for the structures of 3, see ESI) as radical amination reagents. Keeping 3b as the N-donor, other aryllithiums bearing various functional groups at the para position of the aryl moiety, such as methoxy (4p), trimethylsilyl (4q), methyl (4r), phenyl (4s), trifluoromethoxy (4t), trifluoromethyl (4u), and halides (4v–4x) all reacted well in this transformation. Aryl groups bearing meta substituents are also tolerated, as documented by the preparation of 4y (81%). To our delight, a boron ate complex generated with a 3-pyridyl lithium reagent engaged in the cascade and the carboamination product 4z was isolated in high yield (82%).Open in a separate windowScheme 3Scope of vinylboron ate complexes. Reaction conditions: 1 (0.20 mmol, 1.0 equiv.), RMLi (0.22 mmol, 1.1 or 1.3 equiv.), Et2O or THF, under Ar; then [N]-Cl (0.30 mmol, 1.5 equiv.), −20 °C, 16 h, in MeCN. Yields given correspond to yields for isolated products. See the ESI for experimental details.The reason for the dramatic reduction in yield when α-branched alkyllithium or electron-rich aryllithium reagents were used might be that the corresponding vinylboron ate complexes could be oxidized by N-chloroamides via a single-electron oxidation process.18a–e Furthermore, the α-unsubstituted vinyl boronic ester and vinyl boronic ester bearing various α-substituents are suitable N-radical acceptors and the corresponding products 4aa–4ac were obtained in 48–70% yield.To gain insights into the mechanism of this 1,2-carboamination, a control experiment was conducted. The reaction could be nearly fully suppressed when the reaction was carried out in the presence of a typical radical scavenger (2,2-6,6-tetramethyl piperidine-N-oxyl, TEMPO), indicating a radical mechanism (Scheme 4a). Further, considering an ionic process, the N-chloroamides would react as Cl+-donors that would lead to Zweifel-type products, which were not observed under the applied conditions. The proposed mechanism is shown in Scheme 4b. As chloroamides have been recently proposed to undergo homolysis under visible light irradiation,19a,b we propose that initiation proceeds via homolytic N–Cl cleavage generating the electrophilic amidyl radical A, which then adds to the electron-rich vinylboron ate complex 2a to give the adduct boronate radical B. The radical anion B then undergoes single electron transfer (SET) oxidation with 3a in an electron-catalyzed process20a,b or chloride atom transfer with 3a to provide C or D along with the amidyl radical A, thereby sustaining the radical chain. Intermediates C or D can then react via a boronate 1,2-migration10c,11c,21 to eventually give the isolated product 4a.Open in a separate windowScheme 4Control experiment and proposed mechanism.To document the synthetic utility of the method, a larger-scale reaction and various follow-up transformations were conducted. Gram-scale reaction of 2a with 3a afforded the desired product 4a in good yield, demonstrating the practicality of this transformation (Scheme 5a). Oxidation of 4a with NaBO3 provided the β-amino alcohol 5 in 89% yield (Scheme 5b). The N-Boc homoallylic amine 6 was obtained by Zweifel-olefination with a commercially available vinyl Grignard reagent and elemental iodine in good yield (79%).22 Heteroarylation of the C–B bond in 4a was realized by oxidative coupling of 4a with 2-thienyl lithium to provide 7.23Open in a separate windowScheme 5Gram-scale reaction and follow-up chemistry.In summary, we have described an efficient method for the preparation of 1,2-aminoboronic esters from vinyl boronic esters via catalyst-free three-component radical 1,2-carboamination. Readily available N-chloro-carbamates/carboxamides, which are used as the N-radical precursors, react efficiently with in situ generated vinylboron ate complexes to afford the corresponding valuable 1,2-aminoboronic esters in good yields. The reaction features broad substrate scope and high functional group tolerance. The value of the introduced method was documented by Gram-scale reaction and successful follow-up transformations.  相似文献   

9.
Oxindole synthesis via polar–radical crossover of ketene-derived amide enolates in a formal [3 + 2] cycloaddition     
Niklas Radhoff  Armido Studer 《Chemical science》2022,13(13):3875
Herein we introduce a simple, efficient and transition-metal free method for the preparation of valuable and sterically hindered 3,3-disubstituted oxindoles via polar–radical crossover of ketene derived amide enolates. Various easily accessible N-alkyl and N-arylanilines are added to disubstituted ketenes and the resulting amide enolates undergo upon single electron transfer oxidation a homolytic aromatic substitution (HAS) to provide 3,3-disubstituted oxindoles in good to excellent yields. A variety of substituted anilines and a 3-amino pyridine engage in this oxidative formal [3 + 2] cycloaddition and cyclic ketenes provide spirooxindoles. Both substrates and reagents are readily available and tolerance to functional groups is broad.

Herein we introduce a simple, efficient and transition-metal free method for the preparation of valuable and sterically hindered 3,3-disubstituted oxindoles via polar–radical crossover of ketene derived amide enolates.

Oxindoles, in particular the 3,3-disubstituted congeners, are highly valuable substructures in medicinal chemistry. The oxindole core can be found in various biologically active compounds, that are for example used in the treatment of cancer or as antibacterial agents.1 In addition, the oxindole moiety also occurs in several complex natural products.2 The first oxindole synthesis was reported by Baeyer and Knop in 1866.3 That time, isatin was converted by sodium amalgam reduction to the corresponding oxindole. Since then, many methods for the preparation of 3,3-disubstituted oxindoles have been developed that proceed via functionalization of a pre-existing oxindole core.4 In addition, methods for the construction of 3,3-disubstituted oxindoles starting from acyclic precursors have also been introduced.5,6 Along these lines, transition metal-mediated reactions5,7 or homolytic aromatic substitutions (HAS)8–14 have found to be highly efficient for the construction of the oxindole core. Focusing on the latter approach, the intramolecular HAS proceeds via α-carbonyl radicals derived from radical addition to N-arylacrylamides,8 reduction of α-haloarylamides9 or oxidation of the corresponding enolates10–14 (Scheme 1a).Open in a separate windowScheme 1Selected strategies for the synthesis of oxindoles.In 2017, the group of Taylor developed a transition metal-free enolate oxidation-HAS-approach towards oxindoles at low temperature using elemental iodine as the oxidant and malonic acid derived N-aryl amides as substrates which are readily deprotonated.14The unique reactivity of ketenes15 has been explored extensively,16 especially in [2 + 2]-cycloadditions.17 Moreover, Staudinger,18 Lippman19 and Taylor20 showed that ketenes react with aryl nitrones in a tandem [3 + 2]-cycloaddition-[3,3]-sigmatropic-rearrangement cascade21 followed by hydrolysis to provide oxindoles (Scheme 1b). The use of chiral nitrones leads to chirality transfer and enantiomerically enriched oxindoles can be obtained via this approach.21,22 In contrast to the examples discussed in Scheme 1a, two σ-bonds are formed and the overall sequence can be regarded as a formal [3 + 2] cycloaddition. Despite good yields and high enantiomeric excess, nitrones have to be used as precursors and an aldehyde is formed as the byproduct diminishing reaction economy of these elegant cascades.To address these drawbacks, we decided to use the nucleophilic addition23–26 of deprotonated anilines to ketenes for the generation of the corresponding amide enolates that should then be oxidized in a single electron transfer process to α-amide radicals which can undergo a homolytic aromatic substitution providing direct access to sterically challenging 3,3-disubstituted oxindoles in a straightforward one-pot sequence (Scheme 1c). This polar–radical crossover reaction shows high atom economy and as the reaction with the nitrones can also be regarded as a formal [3 + 2] cycloaddition.We initiated the optimization study with N-methylaniline 1a and ethyl phenyl ketene 2a, which was prepared in an easy and scalable one-pot protocol starting from the corresponding carboxylic acid, as model substrates. Deprotonation of 1a with n-BuLi in THF and subsequent addition to the ketene 2a led to desired Li-enolate which was confirmed by protonation with water and isolation of the amide 4aa (56%). Pleasingly, addition of ferrocenium hexafluorophosphate (FcPF6, 2.2 equiv.) at room temperature to the Li-enolate afforded the desired oxindole 3aa in 29% yield (14 Light does not appear to play a crucial role in this transformation, as performing the reaction in the dark does not have a significant effect on the reaction outcome ( EntryBaseConc. (M)Oxidant (equiv.)Yield 3aa (%)b1 n-BuLi0.1FcPF6 (2.2)29 (18)c2d n-BuLi0.1CuCl2 (2.2)34c3 n-BuLi0.1I2 (2.2)41c4EtMgBr0.1I2 (2.2)44c5EtMgBr0.02I2 (2.2)78 6 EtMgBr 0.01 I 2 (2.2) 90 (82) c 7EtMgBr0.01I2 (1.2)258EtMgBr0.01NISe (2.2)399EtMgBr0.01I2 (1.2)f8010EtMgBr0.01I2 (2.2)g8211EtMgBr0.01I2 (2.2)h7412EtMgBr0.01I2 (2.2)i93Open in a separate windowaReactions (0.20 mmol) were conducted under argon atmosphere.b 1H NMR yield using 1,3,5-trimethoxybenzene as internal standard.cIsolated yield.dStep 1 and 2 were conducted at 0 °C.e N-Iodosuccinimide.fIodine addition at −78 °C, then slowly allowed to warm to room temperature.14gIn the dark.hIrradiation with blue LED (40 W, 467 nm, rt, 8 h).iRefluxing THF for step 3, reaction completed within 2 h.With the optimized reaction conditions in hand, we investigated the scope by first varying the R1-substituent at the N-atom using the ketene 2a as the reaction partner (Scheme 2). In general, increasing the steric bulk at the nitrogen leads to diminished yields of the targeted oxindoles. The lower yields go along with the formation of a larger amount of the corresponding α,β-unsaturated amide side product 5. Thus, as compared to the parent N-methyl derivative, all other N-alkyl derivatives were formed in lower yields (49%, 3ab; 35%, 3ac; 49%, 3ad). The N-benzyl protected oxindole 3af and the N-phenyl oxindole 3ae were isolated in 54% and 56% yield, respectively. Next, a diastereoselective oxindole synthesis was attempted using chiral anilines 1g and 1h. Surprisingly, despite the bulkiness of these nucleophiles containing styryl-type N-substituents, good yields were obtained for the oxindoles 3ag and 3ah (73–79%). Unfortunately, diastereocontrol was low in both cases (1.9 : 1 d.r. and 1.5 : 1 d.r.). Of note, addition of Mg-1g and Mg-1h to ketene 2a was rather slow under the standard reaction condition and a significant amount of unreacted aniline was recovered. That problem could be solved by prolonging the reaction time of both step 1 (deprotonation) and also step 2 (Mg-enolate formation).Open in a separate windowScheme 2Substrate scope – variation of substituents at the nitrogen. Reactions (0.20 mmol) were conducted under argon atmosphere. a For step 1 and 2 reaction time was 1 h.Next, the substrate scope was investigated by using different anilines in combination with the ketene 2a (Scheme 3). N-Methyl-p-toluidine 1i and N-methyl-p-haloanilines 1j–m could be successfully transformed to the corresponding oxindoles 3al–am in moderate to good yields (53–87%). Electron-withdrawing and also electron-donating substituents are tolerated and oxindoles derived from p-cyano- (3an, 92%), p-acetyl- (3ao, 30%), p-methoxycarbonyl- (3ap, 82%) and p-methoxy- (3aq, 71%) anilines were isolated in moderate to excellent yields documenting a high functional group tolerance of this reaction. The meta-methyl aniline afforded oxindole 3ar in 76% yield as a 1.8 : 1 mixture of the two regioisomers (only the major isomer drawn). For the pyridyl derivative 3as, a lower yield was obtained (39%), but reaction occurred with complete regiocontrol. Of note, ortho-methyl N-methylaniline provided the corresponding oxindole only in trace amounts (not shown).Open in a separate windowScheme 3Substrate Scope – variation of anilines and ketenes. Reactions (0.20 mmol) were conducted under argon atmosphere. a Isolated as an inseparable mixture (1 : 1.4) with the protonated enolate 4fa (56% combined yield).The ketene component was also varied using N-methylaniline 1a as the reaction partner. The transformation of methyl phenyl ketene 2b provided the oxindole 3ba in 58% yield. p-Bromophenyl ethyl ketene 2c and p-iodophenyl ethyl ketene 2d afforded the oxindoles 3ca and 3da in good yields (70% and 76%). For the ibuprofene-derived ketene 2e a lower yield was obtained (3ea, 40%) and the bulkier phenyl isopropyl congener 3fa was isolated in 27% yield as an inseparable mixture with the protonated enolate 4fa (56% combined yield). In the latter case, increasing the reaction time did neither lead to a higher yield of 3fa nor to a suppression of the formation of 4fa. The lower yield is likely caused by steric effects. Surprisingly, diphenyl ketene 2g delivered the targeted oxindole 3ga in acceptable 55% yield despite the steric demand of the two phenyl groups and the high stability of the corresponding α-amide radical. Spirocyclic oxindoles are of great interest due to their high pharmaceutical potential.27 We were pleased to find that our method also works for the preparation of such spiro compounds as documented by the successful synthesis of 3ha (26%).Mechanistically, we propose initial formation of the enolate A by nucleophilic attack of the deprotonated aniline to the ketene 2, which is then oxidized by elemental iodine to the α-amide radical B (pathway b). The radical nature of the transformation is supported by the fact that electronic effects on the arene show no influence on the efficiency of the cyclization, as would be shown by a conceivable polar aromatic substitution. Radical B readily cyclizes onto the aniline ring to generate the cyclohexadienyl radical D which is oxidatively rearomatized via cationic intermediate E to finally give the oxindole 3 (Scheme 4).10–14 Alternatively, enolate A can be iodinated with I2 to give the unstable iodide C which then undergoes C–I bond homolysis to generate the radical B (pathway a). Indeed, Taylor and coworkers14 observed under similar reaction conditions the decay of α-iodinated compounds of type Cvia C–I homolysis14,28 to give radicals of type B. Usually, we observed α,β-unsaturated amides analogous to 5aa as by-products. However, the corresponding protonated enolates were detected only in tiny amounts in most of these cases. This strongly suggests that those amides are not formed via disproportionation of radical B. HI-elimination seems more likely, pointing towards the presence of the iodinated species C and thus the contribution of pathway b to product formation. In addition, dimerization of radical B was also not observed.Open in a separate windowScheme 4Suggested mechanism.To further support pathway b, isolation of the iodinated intermediate C was attempted at low temperature. Upon addition of iodine (1.2 equiv.) to the preformed Mg-enolate A derived from aniline 1a and ketene 2a at −78 °C,14 TLC analysis showed a clean conversion to a single new compound, which was analyzed by rapid ESI-MS analysis and provided evidence for the formation of the iodinated intermediate C (Scheme 5). However, isolation of this highly unstable compound was not possible due to rapid HI-elimination to the amide 5aa. Note that oxindole formation worked well upon I2-addition at −78 °C and subsequent warming to room temperature (see Scheme 5). This is consistent with the observation from our optimization studies that irradiation with blue light does not contribute to the yield of oxindole 3aa (Open in a separate windowScheme 5Mechanistic experiments. (a) (1) EtMgBr (1.1 equiv.), rt, 30 min, (2) 2a (1.5 equiv.), −78 °C, 30 min, (3) I2 (1.2 equiv.), −78 °C, 15 min in THF (0.01 M). (b) Warm to room temperature in THF (0.01 M), 18 h. (c) NaI (1.2 equiv.) in acetone (0.77 M), rt, 18 h. (d) Irradiation with blue LED (40 W, 467 nm) in THF (0.01 M), rt, 8 h.  相似文献   

10.
Ligand design for Rh(iii)-catalyzed C–H activation: an unsymmetrical cyclopentadienyl group enables a regioselective synthesis of dihydroisoquinolones     
Todd K. Hyster  Derek M. Dalton  Tomislav Rovis 《Chemical science》2015,6(1):254-258
We report the regioselective synthesis of dihydroisoquinolones from aliphatic alkenes and O-pivaloyl benzhydroxamic acids mediated by a Rh(iii) precatalyst bearing sterically bulky substituents. While the prototypical Cp* ligand provides product with low selectivity, sterically bulky Cpt affords product with excellent regioselectivity for a range of benzhydroxamic acids and alkenes. Crystallographic evidence offers insight as to the source of the increased regioselectivity.C–H activation mediated processes have provided a unique retrosynthetic approach to access a variety of substituted heterocycles.1 One tactic that has received increased attention is the coupling of π-components with heteroatom containing molecules.2 A variety of transition metals are capable of catalyzing this type of transformation, providing access to dozens of heterocyclic motifs.13 A challenge for these methods is controlling the regioselectivity of migratory insertion across alkenes and alkynes after the metallacycle forming C–H activation (eqn 1).Steric and electronic effects are understood to control migratory insertion of unsymmetrical alkynes in Rh(iii) catalyzed isoquinolone syntheses (eqn 1). When the substituents are electronically similar, the larger group resides β- to Rh in the metallacycle to avoid unfavorable steric interactions (selectivity is generally >10 : 1).4 When the substituents are electronically different, the more electron-donating group prefers being α- to rhodium in the metallacycle, presumably to stabilize the electron poor metal.5,6 The type of C–H bond being activated also plays an important role in the regioselectivity of migratory insertion; aromatic substrates typically provide synthetically useful regioselectivities when electronically different alkynes are used (>10 : 1) but alkenyl C–H activation leads to products with lower regioselectivities, presumably due to minimal steric interactions during migratory insertion.7,8 We found that sterically bulky di-tert-butylcyclopentadienyl ligand (Cpt) enhances the regioselectivity of the alkyne migratory insertion event in these cases, delivering regioselectivities (>10 : 1) modestly above those achievable by Cp* ligated Rh complexes (<6 : 1). However, when the alkyne migratory insertion was poorly selective with RhCp* (<3 : 1), RhCpt complex was ineffective at providing synthetically useful levels of selectivity. Furthermore, the Cpt ligand was only effective with aryl substituted alkynes, presumably because of strong steric interactions between the ligand and alkyne in the insertion event. Migratory insertion of alkenes to access heterocycles using C–H activation chemistry is still relatively rare, with seminal studies by Glorius and Fagnou reporting the synthesis of dihydroisoquinolones.911 Similar to alkynes, alkenyl electron-donating groups favor the position adjacent to the metal in the metallacycle delivering high regioselectivity. In contrast to alkynes, aliphatic alkenes afford product with poor regioselectivity (2 : 1) (eqn 2).5h,12 We hypothesized competing steric and electronic effects cause the low regioselectivity, with steric effects favoring the formation of a 4-substituted product and electronics favoring the formation of a 3-substituted product.13 As a temporary solution to this problem, our group and others have employed tethering strategies to increase the regioselectivity of the migratory insertion event (eqn 3).14,15 Of course, regioselectivity controlled by the ligand on Rh would be the optimal solution to the selectivity problem (eqn 4).16 Consequently, we focused our attention toward developing an intermolecular variant of this reaction that would provide product with improved regioselectivity.As a model system, we explored the impact ligands have on the coupling of O-pivaloyl-benzhydroxamic acid 1a with 1-decene 2a to provide dihydroisoquinolones 3a and 3a′. When Cp* is used as a ligand, the desired products are isolated in excellent yield but poor selectivity (2.4 : 1 3a : 3a′) ( a

EntryCatalystYield (%)Regioselectivity
1[RhCp*Cl2]2 902.4 : 1
2 b [RhCpCF3Cl2]2 852.4 : 1
3 c [RhCpCl2]2 8212 : 1
4 d [RhCptCl2]2 9215 : 1
Open in a separate window aReaction conditions: 1a (.2 mmol), 1-decene (.2 mmol), precatalyst (1 mol%), CsOAc (200 mol%), MeOH (0.1 M). bCpCF3 = 1-trifluoromethyl-2-3,4,5-tetramethylcyclopentadienyl. cCp = 1,2-di-phenyl-3,4,5-trimethylcyclopentadienyl. dCpt = 1,3-di-t-butylcyclopentadienyl.To determine the effect that ligand electronics have on product regioselectivity, we employed an electron deficient 1-trifluoromethyl-2,3,4,5-tetramethylcyclopentadienyl ligand originally developed by Gassman (CpCF3)17 and found that this catalyst provides 3a and 3a′ products in good yield but without an increase in selectivity (2.4 : 1) (18,19 Since ligand electronics did not appear to affect product regioselectivity, we tested an electron rich, sterically bulky di-phenyl-tri-methyl Cp ligand (Cp) and were pleased to find a remarkable increase in selectivity from 2.4 : 1 to 12 : 1 (3a : 3a′). Pleased by this improvement, we tested the sterically bulky di-tert-butyl Cp ligand Cpt and were surprised to find that RhCpt provides the desired product in 91% yield with exquisite regioselectivity (15 : 1) ( a

EntryStarting materialYield b (%)Cp*Cpt
1X = CF3 (1b)501.5 : 119 : 1
2X = Cl (1c)762.2 : 119 : 1
3X = OMe (1d)701.9 : 116 : 1
4X = Ph (1e)751.7 : 114 : 1
5 951.9 : 115 : 1
6 842.5 : 119 : 1
7 881.8 : 119 : 1
Open in a separate window aReaction conditions: amide (.2 mmol), 1-decene (.2 mmol), precatalyst (1 mol%), CsOAc (200 mol%), MeOH (0.1 M). bIsolated yield of reaction using [RhCptCl2]2 as a precatalyst. meta-Substituents also provide exquisite levels of regioselectivity for alkene migratory insertion when Cpt is used (>15 : 1) ( a
Open in a separate window aReaction conditions: amide (.2 mmol), 1-decene (.2 mmol), precatalyst (1 mol%), CsOAc (200 mol%), MeOH (0.1 M). isolated yield of reaction using [RhCptCl2]2 as a precatalyst. b67% yield. c80% yield. d85% yield. e79% yield.We next explored the alkene tolerance of the method. Allyl benzene 2b furnishes a 1.6 : 1 ratio of dihydroisoquinolone with RhCp* ( a

EntryAlkeneYield b (%)Cp*Cpt
1 c 851.6 : 15.1 : 1
2 681.6 : 19.4 : 1
3 701.3 : 15.5 : 1
4 952.3 : 114 : 1
5 851.6 : 18 : 1
6 d 921.2 : 17.2 : 1
7 801.4 : 112 : 1
8 e 931 : 111 : 1
9 892 : 114 : 1
10 943 : 114 : 1
Open in a separate window aReaction conditions: 1a (.2 mmol), alkene (.2 mmol), precatalyst (1 mol%), CsOAc (200 mol%), MeOH (0.1 M). bIsolated yield of reaction using [RhCptCl2]2 as a precatalyst. cReaction conducted at 0 °C. dProducts isolated as a 1 : 1 ratio of diastereomers. eProduct isolated as a 2 : 1 ratio of diastereomers.While it is desirable to achieve high regioselectivity for a single regioisomer, it is even more attractive to use a ligand to access alternate regioisomers. Currently, the only example of Rh(iii)-catalyzed synthesis of 4-substituted dihydroisoquinolones is with potassium vinyltrifluoroborates where electronics are believed to control regioselectivity.20 We found that when vinylcyclohexane was submitted to a reaction with [RhCp*Cl2]2 as the precatalyst, the 3-substituted dihydroisoquinolone 4a was isolated in 90% yield with 11 : 1 regioselectivity (Fig. 1). However, when the same reaction was catalyzed by [RhCptCl2]2 the opposite isomer 4b was isolated in 75% yield and 10 : 1 (4b : 4a) regioselectivity. Given this unexpected discovery, we were interested in gleaning insight into how Cpt influences regioselectivity of alkene migratory insertion. A competition experiment between vinyl cyclohexane 2m and 1-decene 2a run to 10% conversion favored the formation of dihydroisoquinolone 3a in >19 : 1 ratio as determined by 1H NMR. This experiment suggests that enhanced steric interactions between the substrate and ligand slow the rate of migratory insertion.Open in a separate windowFig. 1Impact of ligand on reaction of vinyl cyclohexane.To investigate the steric differences between the RhCp* and RhCpt systems X-ray analysis was conducted on a 5-membered RhCpt metallacycle. While we were unable to obtain a 5-membered rhodacycle from our system, Jones and coworkers previously characterized 5-membered rhodacycle 5a from N-benzylidenemethanamine and [RhCp*Cl2]2.21 We found that a similar metallacycle 5b derived from [RhCptCl2]2 could be obtained in crystalline form under identical conditions and was evaluated by single crystal X-ray diffraction.A comparison of the bond lengths and angles reveals several notable differences between our Cpt rhodacycle and the Cp* rhodacycle reported by Jones (Fig. 2). The Rh–Cp centroid distance in 5b is 0.011 Å longer than 5a which is either the result of increased steric interactions, or an artifact of Cpt being a less electron-donating ligand. While there are subtle differences in many bond lengths and angles, the most striking difference is the angle C3–Rh–Cl, which is 98.03° in 5b while only 90.09° in 5a. The angle increase is likely the result of steric interactions caused by the tert-butyl moiety being situated directly over the Rh–Cl bond. As alkene exchange presumably occurs with Cl, we suggest that steric interactions between the t-butyl of the ligand and the alkene substituent affect both the alkene coordination and 1,2-insertion events.Open in a separate windowFig. 2X-Ray analysis.Based on the X-ray crystal structure and regioselectivity data, we propose the following model for regioselectivity of the 1,2-migratory insertion of alkenes, where steric contributions from the t-butyl groups influence both alkene coordination and insertion events to give high selectivity. With small alkyl alkenes, we propose that steric interactions from one t-butyl of Cpt disfavor alkene coordination (I) and subsequent insertion to give the β-substituted product 3a′ (Fig. 3). Coordination of the alkene with the steric bulk oriented away from the t-butyl group finds minimized steric interactions during coordination (II). Subsequent migratory insertion from II places the alkyl substituent α to Rh in the transition state, which we propose is able to stabilize a buildup of partial positive charge, making the α-substituted product 3a both sterically and electronically favored with Cpt. In the case of the Cp* ligand with small alkyl alkenes, neither steric nor electronic interactions dominate so low selectivity is observed.Open in a separate windowFig. 3Rationale for selectivity.However if the size of the alkene substituent is significantly increased, as in the case of vinyl cyclohexane, then Cpt favors the opposite regioisomer. While certainly a puzzling result, we propose that the selectivity can be explained by Cpt rotation such that the t-butyl groups both occupy the space above the metallacycle. Cpt rotation gears the O-piv toward the alkene coordination site disfavoring alkene coordination to this side (IV) favoring the α-substituted product 3a. At the same time, alkene coordination (III) with the cyclohexyl opposite the O-piv minimizes steric interactions enabling insertion of the large alkene and preferential formation of β-substituted product 3a′. While not conclusive, the observation that cyclohexyl alkene reacts significantly slower than n-octyl alkene suggests that migratory insertion of the cyclohexyl alkene proceeds through a higher energy and potentially highly ordered transition state, such as Cpt rotation.  相似文献   

11.
Ruthenium pincer complex-catalyzed heterocycle compatible alkoxycarbonylation of alkyl iodides: substrate keeps the catalyst active     
Han-Jun Ai  Yang Yuan  Xiao-Feng Wu 《Chemical science》2022,13(8):2481
The electron pair of the heteroatom in heterocycles will coordinate with metal catalysts and decrease or even inhibit their catalytic activity consequently. In this work, a pincer ruthenium-catalyzed heterocycle compatible alkoxycarbonylation of alkyl iodides has been developed. Benefitting from the pincer ligand, a variety of heterocycles, such as thiophenes, morpholine, unprotected indoles, pyrrole, pyridine, pyrimidine, furan, thiazole, pyrazole, benzothiadiazole, and triazole, are compatible here.

A pincer ruthenium-catalyzed heterocycle compatible alkoxycarbonylation of alkyl iodides has been developed.

Since the pioneering work on the catalytic alkoxycarbonylation of unactivated alkyl halides reported by Heck and Breslow in 1963,1 this transformation has attracted a great deal of interest due to its modularity and the direct employment of CO as a cheap and abundant C1 feedstock.2 However, compared with aryl halides, the development of alkoxycarbonylation of alkyl halides has been much more gradual.2,3 This situation is due to both the slow oxidative addition of C(sp3)–X bonds to the metal center and the easy β-hydride elimination of the alkyl-metal intermediate, particularly in the presence of carbon monoxide.4 Several catalytic systems for this process have been successfully developed in recent years (Scheme 1A), such as pure radical-based systems,5 palladium-based systems,6/palladium-based systems,7 rhodium-based systems,8 copper-based systems,9 and other metal carbonyl complex-based systems.10 Very recently, Neumann, Skrydstrup, and co-workers reported a nickel pincer-mediated alkoxycarbonylation for complete carbon isotope replacement, and this approach provided a procedure for generating carbon-labeled versions of potential simple carboxylate prodrug derivatives (Scheme 1B).11 Besides their advantages, in these cases the heterocycles, particularly those containing multiple N atoms or NH groups, are hardly compatible, which is considered as a remaining challenge. We attribute this to the Lewis-basic atoms in heterocyclic motifs being particularly detrimental to catalyst activity and potentially quenching the radical intermediates.12 Indeed, the development of heterocycle compatible catalytic systems remains an exciting task in the field of alkoxycarbonylation.Open in a separate windowScheme 1Approaches to alkoxycarbonylation of alkyl halides.On the other hand, heterocycles constitute important structural components of biologically active compounds and are ubiquitous in agrochemical and pharmaceutical industries.13 In a recent survey, 88% of small molecule drugs approved by the FDA between 2015 and June 2020 were found to contain at least one N-heterocycle.14 Specifically, heterocyclic subunits can modify the solubility, lipophilicity, polarity and hydrogen bonding ability of biologically active agents, thereby optimizing the corresponding ADME/Tox (absorption, distribution, metabolism, excretion, and toxicity) properties of drugs or drug candidates.15 Under this premise, the pursuit of new synthetic methods with good heterocycle compatibility is a worthwhile endeavor.Herein we report a heterocycle compatible catalytic system for alkoxycarbonylation of alkyl iodides. With a ruthenium pincer complex as the catalyst, the tight coordination of the pincer ligand can effectively prevent the ruthenium from deactivation by heterocycle coordination (Scheme 1C). To the best of our knowledge, this is the first example of a ruthenium pincer complex-catalyzed carbonylation reaction.16 This new catalytic system might lead to novel synthetic routes toward heterocyclic carbonyl-containing compounds.Pincer complexes of ruthenium are among the most effective catalysts for hydrogen transfer reactions between alcohols and unsaturated compounds.17 We initially used it to attempt the carbonylative coupling of acetophenone with iodobutane, as shown in eqn (1). Although we did not get the desired product I, the ester II could be obtained in 22% yield. By literature survey, we found there was no example showing that alkyl halides could be activated by ruthenium in previous reports on carbonylation reactions.3,16,18 We thus envisioned that the ruthenium pincer complex played a key role in this transformation.11,191With this discovery in mind, we started the investigation of this ruthenium-catalyzed alkoxycarbonylation of alkyl halides by examining the reaction of (3-iodopropyl)benzene (1) with isopropanol (2) at 100 °C under a CO atmosphere (10 bar) in the presence of a catalytic amount of various readily available ruthenium pincer complexes (). The improved yield of the desired product 3 was obtained when utilizing Milstein''s catalyst Ru-220 (21 were applied in the reaction; however, the selectivity obtained was unsatisfactory (eqn (2), when we removed isopropanol from the reaction, byproduct 2 which was produced by carbonylative homocoupling of the alkyl halide could be obtained in 71% yield.22 However, the reduced conversion and the absence of byproduct 3 implied that the alcohol plays more than a nucleophile role in this reaction. It is important to mention that the addition of water had no effect on the yield of byproduct 2. Concerning the effects from bases, organic bases, such as NEt3 and DBU, were tested, but no desired ester could be detected. Inorganic bases, including K2CO3 and K3PO4, were also tested, but very low yield of the ester was obtained. Notably, comparable yield of ester 3 can be obtained when LiOtBu was used as the base.2Optimization of the alkoxycarbonylation of 1a
Entry[Ru]Conv.b (%)Yieldb (%)
1Ru(acac)36015
2RuH(Cl)(CO)(PPh3)32111
3Ru-110024
4Ru-29341
5Ru-3534
6Ru-4495
7Ru-510032
8Ru-610038
9Ru-710081
10Ru-710063c
11Ru-710082d
12Ru-710072e
13Ru-710086d,f
Open in a separate windowaReaction conditions: 1 (0.2 mmol), 2 (0.6 mmol), [Ru] (5 mol%), Cs2CO3 (0.6 mmol), toluene (0.5 mL), CO (10 bar), 100 °C, 12 h.bDetermined by GC with hexadecane as the internal standard.cCO (1 bar), N2 (9 bar).d[Ru] (2.5 mol%) was used.e[Ru] (1 mol%) was used.f90 °C, average yield of two independent reactions.We next turned our attention to study the scope and the limitation of this transformation, as shown in Fig. 1. At the first stage, a variety of alcohols containing different functional groups and structural blocks were tested. In general, moderate to excellent yields were obtained under the standard conditions. For primary alcohols, the length of the carbon chain did not affect the good yield (4–7). The reaction tolerated the presence of ethers (8, 9), thioether (10), alkene (11), chlorine (12), trimethylsilyl (13), and amide (21). Benzyl alcohols and secondary alcohols were afterwards tested in this system and successfully transformed into the corresponding esters in good yields (14–17). With the further increase of the steric hindrance, tertiary alcohols hardly provided the desired products (18, 19). Phenol was also employed as the substrate in our attempt, and not surprisingly, phenyl 3-phenylpropyl ether (SN reaction product) was isolated as the main product (20).23 Interestingly, ethylene glycol could be converted to diester 22 in 83% yield, and no monocarbonylation product was detected, even though the alcohol was three equivalents. This suggests an interaction between the alcohol and the catalytic center, resulting in a higher rate of intramolecular reaction than intermolecular reaction. Subsequently, the excellent heterocycle compatibility of the method is nicely illustrated by the fact that thiophenes (23, 27), morpholine (24), unprotected indoles (25, 26), pyrrole (27), pyridine (28), pyrimidine (29), furan (30), thiazole (31), pyrazole (32), benzothiadiazole (33), and triazole (34) were perfectly tolerated under our protocol. The broad synthetic applicability of the reaction was also reflected in the successful alkoxycarbonylation of various primary iodides (35–44), secondary iodides (45–47), and even sterically hindered tertiary iodides (48–50).Open in a separate windowFig. 1Scope of Ru pincer complex-catalyzed alkoxycarbonylation. Reactions run with 0.2 mmol of alkyl iodide and 3 equiv. of alcohol. Yield of the isolated product. aTogether with a 68% yield of the SN reaction product (phenyl 3-phenylpropyl ether). bEthylene glycol (3 equiv.) was used. cReduced yield of the isolated product because of the volatility of the product.In particular, the secondary iodides generated the corresponding esters in near quantitative yields. We also evaluated a substrate containing the C(sp2)–I bond to probe the chemoselectivity of our process (41). No trace of arylate was detected in the crude mixture by GC-MS, hence illustrating the good chemoselectivity of this catalytic system and offering opportunities for further structure modification. While this new methodology allows for the formation of a wide range of heterocycle-containing esters, some limitations still remain in terms of substrate scope. Bromoalkanes and chloroalkanes cannot be successfully converted under these conditions, even with the addition of equivalent amounts of NaI.The alkoxycarbonylation could be applied to late-stage modification of a range of drugs and natural products, as shown in Fig. 2. trans-Sobrerol, a mucolytic, was successfully transformed, while the tertiary C–OH group was retained (51). A weak androgen, epiandrosterone, which is widely recognized to inhibit the pentose phosphate pathway and to decrease intracellular NADPH levels, provided 52 in 93% yield. Derivatives of estrone, cholesterol, and vitamin E also delivered the corresponding esters 53–55 in moderate to good yields. Common alcohol natural products, such as crotonyl alcohol, piperonyl alcohol, (−)-perillyl alcohol, (−)-borneol, (−)-menthol, and nerol, were tested as well and applicable to the reaction (56–61), which illustrated the utility of this method.Open in a separate windowFig. 2Modification of drugs and natural products. Reactions run with 0.2 mmol of alkyl iodide and 3 equiv. of alcohol. Yield of the isolated product.To gain more mechanistic insight into the reaction pathway, several experiments were conducted (Scheme 2). Under the standard conditions, the addition of TEMPO (radical capture agent) to the reaction led to the termination of the target reaction; meanwhile, the intermediate was captured (62) in 91% isolated yield (Scheme 2A, middle). In the control experiment, only limited conversion and no 62 was observed in the absence of the pincer catalyst (Scheme 2A, top), thus suggesting that the pincer/Ru activates the alkyl iodides to radicals. To ensure the radical pathway, we subsequently conducted radical inhibition experiments with BHT (butylated hydroxytoluene) as the radical inhibitor (Scheme 2B) and radical clock experiments (Scheme 2C). The model reaction was gradually suppressed with the addition of BHT. Furthermore, (iodomethyl)cyclopropane and 6-iodohex-1-ene under our optimized reaction conditions provided the corresponding ring-opening expansion product 64 and the cyclization product 65, respectively, with high selectivity.24Open in a separate windowScheme 2Mechanism studies.Based on the above results, we believe that the reaction involves a radical intermediate. In addition to this, as noted earlier, the alcohol appears to interact with the catalytic center and plays a role in promoting the activation of the alkyl halide. To probe this hypothesis, we removed the isopropanol from the reaction and utilized TEMPO to capture the radical intermediate (Scheme 2A, below). Compared with the reaction in the middle of Scheme 2A, the conversion and the yield of 62 significantly decreased in the absence of isopropanol. We explained that the (PNP)Ru(CO)X2 type complex is the catalyst resting state, and the alcohol may help it to return to the active state by hydrodehalogenation (Scheme 2D).25 Moreover, we could observe acetone during the optimization process, and when we subjected isopropanol alone to our optimized conditions, 57% yield of acetone could be detected,26 which suggests that (PNP)Ru(CO)HX can also undergo hydrodehalogenation to form (PNP)Ru(CO)H2.Based on the above results and previous reports,16–18 a plausible mechanism is proposed (Scheme 3). Initially, the active 16 electron ruthenium complex A will be formed under the assistance of the base. Through a SET process, alkyl iodide will be activated and a 17 electron ruthenium complex B will be formed together with the corresponding alkyl radical which will immediately react with B to give 18 electron ruthenium complex C. The acylruthenium complex D will be produced after a CO insertion step. The possibility that the acylruthenium complex D might also be produced from complex B and the in situ formed acyl radical cannot be excluded. After X ligand exchange, ruthenium complex E will be formed which will provide the final ester product after a reductive elimination step and regenerate the active ruthenium catalyst A to finish the catalyst cycle. Alternatively, the direct nucleophilic attack at the acyl carbonyl of complex D by alcohol to give the ester product and complex F is also possible. Then complex F will be transformed into complex A under the assistance of the base.Open in a separate windowScheme 3Proposed mechanism.  相似文献   

12.
Polycyclic heteroaromatics via hydrazine-catalyzed ring-closing carbonyl–olefin metathesis     
Eun Kee Cho  Phong K. Quach  Yunfei Zhang  Jae Hun Sim  Tristan H. Lambert 《Chemical science》2022,13(8):2418
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 15962 4853b 48274 48545 4864Open 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   

13.
An unusual formal migrative cycloaddition of aurone-derived azadienes: synthesis of benzofuran-fused nitrogen heterocycles     
Qiang Feng  An Wu  Xinhao Zhang  Lijuan Song  Jianwei Sun 《Chemical science》2021,12(22):7953
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   

14.
Assembly of multicyclic isoquinoline scaffolds from pyridines: formal total synthesis of fredericamycin A     
Fang-Xin Wang  Jia-Lei Yan  Zhixin Liu  Tingshun Zhu  Yingguo Liu  Shi-Chao Ren  Wen-Xin Lv  Zhichao Jin  Yonggui Robin Chi 《Chemical science》2021,12(30):10259
The construction of an isoquinoline skeleton typically starts with benzene derivatives as substrates with the assistance of acids or transition metals. Disclosed here is a concise approach to prepare isoquinoline analogues by starting with pyridines to react with β-ethoxy α,β-unsaturated carbonyl compounds under basic conditions. Multiple substitution patterns and a relatively large number of functional groups (including those sensitive to acidic conditions) can be tolerated in our method. In particular, our protocol allows for efficient access to tricyclic isoquinolines found in hundreds of natural products with interesting bioactivities. The efficiency and operational simplicity of introducing structural complexity into the isoquinoline frameworks can likely enable the collective synthesis of a large set of natural products. Here we show that fredericamycin A could be obtained via a short route by using our isoquinoline synthesis as a key step.

A concise approach for rapid assembly of multicyclic isoquinoline scaffolds from pyridines and β-ethoxy α,β-unsaturated carbonyl compounds was developed, which enabled the formal total synthesis of fredericamycin A.

Isoquinolines and their derivatives are common structural motifs in numerous natural products. Among them, the analogues of isoquinolines fused with rings from the benzene side such as 8-hydroxyisoquinolin-1[2H]-one (Fig. 1a) have been found in hundreds of natural products with interesting bioactivities.1 For example, fredericamycin A and the related family members, isolated from Streptomyces griseus, show both antimicrobial and anti-tumor activities.2 Ericamycin is a natural product isolated in the culture of Streptomyces varius n. sp. with anti-staphylococcal activities.3 Due to the widespread presence of isoquinolines in both natural and synthetic molecules, numerous approaches have been developed to assemble this class of scaffolds.4 The dominated strategies reported to date focus on forming the new pyridine ring of isoquinolines (Fig. 1b, left part). Classic methods include Bischler–Napieralski isoquinoline synthesis,4a,b Pictet–Gams isoquinoline synthesis,4a and Pomeranz–Fritsch reaction.4a These reactions, proven to be useful since as early as 1893,5 have their own merits and limitations. For instance, high reaction temperature (e.g. reflux in toluene) and strong acids are typically required and thus functional group tolerance can become challenging. On the other side, the introduction of structural complexities and substitution patterns is constrained as the substrates have to be pre-settled to favor the formation of pyridine moieties. Here we report a new approach to prepare isoquinoline scaffolds by constructing a new benzene ring (Fig. 1b, right part).6 Our method starts with pyridine derivatives as the substrates to react with readily available β-ethoxy α,β-unsaturated carbonyl compounds. The reaction cascade involves five main plausible mechanistic processes (Michael addition, Dieckmann condensation, elimination, aromatization and in situ methylation) to furnish isoquinoline-based products with medium to good yields. The tricyclic isoquinoline-containing products might serve as formal common starting points for rapid total synthesis of a large number of natural products, such as those exemplified in Fig. 1a. In the present study, we demonstrate that starting from the tricyclic isoquinoline adduct 6a prepared using our method, fredericamycin A can be synthesized in 8 steps (Fig. 1c). Our strategy for isoquinoline assembly offers complementary and in certain cases better solutions not readily provided by the classic methods. We expect our method to find impressive applications in concise modular synthesis of complex natural products and molecular libraries, especially those bearing isoquinoline units fused with additional cyclic structures.Open in a separate windowFig. 1Isoquinoline analogues and their synthesis.Our design and initial studies are illustrated in Scheme 1.7 We first used pyridine 1a to react with α-substituted cycloenones (2a–2d), in the hope of obtaining isoquinoline 3a as the target product (Scheme 1a). The use of 2a and 2b was inspired by studies from Tamura, in which α-Br in 1,4-naphthoquinone was used as a leaving group to form an aromatic ring.8 Unfortunately, no product was formed and most of the starting materials were recovered. When SPh (2c) or SOPh (2d) was incorporated at the α site of the cycloenone, side products 4a and 4b were isolated respectively in moderate yields. The Michael products 4a and 4b could not be further transformed into our desired cyclic product 3a under various conditions. We then studied the use of β-substituted cycloenones (2e–2g) to react with 1a (Scheme 1b). No reactions were observed when 2e or 2f was used. To our delight, when the halogen of 2e/2f was replaced with a methoxy unit (OCH3, substrate 2g), an encouraging amount of annulation product 3a was detected (10% yield). A side product 5a was also obtained (5% yield) in this initial study and it couldn''t be further transformed into the annulation product 3a under various alkaline conditions. It is noteworthy that, while β-alkoxy cycloenones (specifically, only β-alkoxy cyclohexenones) have been used in Staunton–Weinreb annulation9 to prepare fused aromatic compounds, no examples for those containing a heterocyclic aromatic ring were reported.10 Even for the construction of an aromatic ring without any heteroatom, low yields (mostly ranging from 0 to 30%) often occurred for this type of annulation starting with β-alkoxy cycloenones,9 which severely hampered its usage in Staunton–Weinreb annulation for the total synthesis of natural products. Our initial results showcased the possibility of direct assembly of isoquinoline scaffolds from β-methoxy cyclopentenone for the first time, though also in a low yield of 10%.Open in a separate windowScheme 1Proposed routes and initial studies for isoquinoline synthesis.With the initial results in hand, we performed additional condition optimization (11 The β-methoxy cyclopentenone 2g could also react to give 6a in a lower yield of 65% (entry 3). Other bases [such as triethylenediamine (DABCO), diazabicyclo[5.4.0]undec-7-ene (DBU), 4-dimethylaminopyridine (DMAP), lithium bis(trimethylsilyl)amide (LiHMDS) and potassium bis(trimethylsilyl)amide (KHMDS)] gave poorer results with yields ranging from 0 to 42% (entry 4). When THF was changed to other solvents, lower yields (<41%) were obtained (entry 5). Revising the ratio of 1a to 2h from 1 : 1.5 to 1.5 : 1 delivered 6a in 39% to 54% yields (entries 6–8). Lower reaction temperature (e.g. −78 °C) could not improve the outcome of this cascade transformation, but gave 23% yield of 6a together with 16% yield of recovered starting material 2h (entry 9). Long exposure to low temperature in step 1 could also lead to a considerable amount of the undesired elimination product 5a (ca. 29% yield), which was decomposed under the following methylation conditions (step 2). No product was observed in the absence of the methoxy group in 1a as it could stabilize the transition state via the formation of a metallate complex (entry 10).Screening of conditionsa
EntryVariation from standard conditionsYieldb (%)
1None72
2Without methylation14
3OCH3 instead of OEt in 2h65
4DABCO, DBU, DMAP, LiHMDS and KHMDS instead of LDA0–42
5Other solvents in step 1<41
6 1a : 2h = 1 : 139
7 1a : 2h = 1.5 : 154
8 1a : 2h = 1 : 1.542
9c−78 °C for step 123
10H instead of OCH3 in 1a0
Open in a separate windowaStandard conditions: 1a (0.2 mmol) and LDA (0.2 mmol) reacted in THF at −78 °C for 1 h; 2h (0.1 mmol) was added dropwise to the mixture before warming up to rt in 10 min. The reaction was quenched by the addition of saturated aqueous solution of NH4Cl after completion monitored by TLC. After the removal of solvents, the crude residue was treated directly with TBAB (0.2 eq.), NaOH (2.0 eq.) in water (1 mL), and Me2SO4 (4.0 eq.) in CH2Cl2 (1 mL).bIsolated yield.cRecovered starting material 2h: 16% yield.With the optimal reaction conditions in hand, we next examined the scope of the pyridine derivatives 1. As we can see from Scheme 2, substrates with the aliphatic substituents at C3 could afford the corresponding tricyclic isoquinoline products (6a and 6b) in acceptable yields. Besides, the incorporation of an aromatic ring at this site (6c–6j) also works well for this transformation, wherein electron-rich aromatic rings (6c–6g) could give higher yields than the corresponding electron-deficient ones (6h–6j). It should be noted that the relatively lower yield of 44% for 6h was partially due to the slow reaction rate as the recovered starting material was always detected in this transformation. When it comes to C4 substitution, the isoquinoline products with broad structural diversities such as alkyl (6k), alkenyl (6l–6n),12 alkynyl (6o), benzyl derivatives with different substituents on the phenyl ring (6p–6t), heteroaromatic ring (6u) and thioether (6v) could be obtained in 57–93% yields. Moreover, substrates bearing acid-hydrolyzable functionalities (6w) and with a relatively bulky secondary substituent (6x) also worked well under the optimized reaction conditions. Next, we examined the possibility of introducing a side chain at C5. To our delight, the substrate with an ethyl group instead of the methyl group on the aromatic ring reacted smoothly to deliver the corresponding isoquinoline 6y in 89% yield. Further study revealed that the exposure of the bicyclic substrate 5,6,7,8-tetrahydroisoquinoline derivative to the optimized reaction conditions could furnish the polycyclic product 6z in 92% yield. Finally, we relocated the nitrogen atom in the pyridine ring. The experimental results indicated that the substrate with nitrogen atom located at C3 can''t react to form the corresponding isoquinoline 6aa, possibly due to the mismatched dipole orientation. When the nitrogen atom was sited at the ortho-position of the methyl group in the aromatic ring, quinoline 6ab could not be detected either under the optimized reaction conditions. The control experiments showcased the decisive influence of the location of nitrogen atom in the aromatic ring on the reactivity of this cascade transformation.Open in a separate windowScheme 2Scope of pyridine derivatives.For the five-membered cycloenone derivatives 2 (Scheme 3), substrates with different substituents at the α′ position work well for this transformation (6ac–6ak),12 of which the incorporation of a quaternary carbon center (6aj) and a heteroatom (6ak) at this site was included. The introduction of an allyl group at the β′ position in cyclopentenone proved to be viable for this transformation, delivering 6al in 64% yield. More encouragingly, when the sterically hindered substrate with a quaternary carbon center located at the γ site was exposed to the optimized reaction conditions, the isoquinoline 6am was obtained in 65% yield. This is challenging, considering the fact that the reacting site is just adjacent to a sterically bulky all-carbon quaternary stereocenter. Bicyclic 3-ethoxy-1H-inden-1-one is also suitable for this cascade transformation, giving the tetracyclic 10H-indeno[1,2-g]isoquinolin-10-one derivative 6an in 89% yield. When it comes to six-membered cycloenone derivatives (6ao–6au), substrates with substituents at α′ and β′ positions all worked smoothly to provide the corresponding isoquinoline products in moderate to high yields. Notably, Kita reported a 5-step reaction sequence to get the tricyclic benzo[g]isoquinoline-derived product 6as starting from the 1a analogue in an overall yield of 22%.6b Using our developed method, 6as could be easily obtained in 53% yield from 1a. Unexpectedly, a side product 6av was isolated in moderate yield when it comes to the γ-substituted substrate. Further study revealed that cyclohept-2-en-1-one with a medium-sized ring (6aw), lactone (6ax), and lactam (6ay) all worked well for this annulation cascade, which significantly expanded the substrate scope of this powerful cascade transformation.Open in a separate windowScheme 3Scope of cycloenone derivatives and more.Finally, fredericamycin A was selected further as the target molecule to verify the flexibility of our method in the total synthesis of natural products, especially those containing 8-hydroxyisoquinolin-1[2H]-one units.13 Since its first isolation in 1981, fredericamycin A attracted much attention from the synthetic community due to its interesting chemical structure and significant anti-tumor activity.2,14,15 The synthetic route was inspired by the expeditious work from Bach.16a As shown in Scheme 4, we started our synthetic attempts with our developed multifold reaction sequence of pyridine 1a and β-ethoxy enone 2h, delivering the corresponding methyl ether 6a on a gram scale. To the best of our knowledge, this is the first example of isoquinoline synthesis directly starting from a pyridine derivative in a single step. The aromatic ketone 6a was subjected to a Mukaiyama aldol/pinacol rearrangement cascade with cyclobutene 7 to give spiro diketone 8 in 42% yield.7,16 After oxidation with DDQ, the pivotal synthon 9 was obtained in 62% yield.7 It should be noted that the addition of p-TsOH is necessary for this transformation as a sluggish reaction rate was detected in the absence of an acid. Meanwhile, a four-step access of phthalidyl chloride 10 was developed starting from a commercially available benzoic acid derivative.7,17 For the crucial Hauser–Kraus annulation18 between fragments 9 and 10, we found that the coupling product 11 was not stable and thus protected directly as the corresponding methyl ether. After extensive screening of reaction conditions,7 LiOtBu turned out to be the only efficient base for this annulation. Mechanistically, the intermolecular Michael addition of segments 9 and 10 was followed by successive transformations involving Dieckmann condensation of enolate V, extrusion of chloride anions from the diketone VI, and last aromatization of the advanced intermediate VII to afford the hexacyclic diphenol 11 with the full skeleton embedded in fredericamycin A. As far as we know, this is the first example of 3-halophthalide as the Hauser donor instead of the classic sulfonyl- or cyano-containing substrates in Hauser–Kraus annulation, as 3-halophthalide was previously reported not suitable for this annulation.18aIn situ methylation of the newly formed phenol hydroxyls delivered Kita''s intermediate 12 in 51% yield in 2 steps. A further 4-step sequence ensured the accomplishment of fredericamycin A.19 The overall synthetic route clearly showcased the power of ingenious introduction of multifold reaction cascades to realize the best performance from the point of step economy.Open in a separate windowScheme 4Formal synthesis of fredericamycin A.  相似文献   

15.
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   

16.
3-Aminooxetanes: versatile 1,3-amphoteric molecules for intermolecular annulation reactions     
Zengwei Lai  Renwei Zhang  Qiang Feng  Jianwei Sun 《Chemical science》2020,11(36):9945
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   

17.
Catalytic asymmetric hydrogenation of (Z)-α-dehydroamido boronate esters: direct route to alkyl-substituted α-amidoboronic esters     
Yazhou Lou  Jun Wang  Gelin Gong  Fanfu Guan  Jiaxiang Lu  Jialin Wen  Xumu Zhang 《Chemical science》2020,11(3):851
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   

18.
Regioselective B(3,4)–H arylation of o-carboranes by weak amide coordination at room temperature     
Yu-Feng Liang  Long Yang  Becky Bongsuiru Jei  Rositha Kuniyil  Lutz Ackermann 《Chemical science》2020,11(39):10764
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   

19.
Aluminum-catalyzed tunable halodefluorination of trifluoromethyl- and difluoroalkyl-substituted olefins     
Zhong Liu  Xian-Shuang Tu  Le-Tao Guo  Xiao-Chen Wang 《Chemical science》2020,11(42):11548
Herein, we report unprecedented aluminum-catalyzed halodefluorination reactions of trifluoromethyl- and difluoroalkyl-substituted olefins with bromo- or chlorotrimethylsilane. The interesting feature of these reactions is that one, two, or three fluorine atoms can be selectively replaced with bromine or chlorine atoms by modification of the reaction conditions. The generated products can undergo a variety of subsequent transformations, thus constituting a valuable stock of building blocks for installing fluorine-containing olefin motifs in other molecules.

Aluminum-catalyzed halodefluorination reactions of fluoroalkyl-substituted olefins are developed. The reactions can selectively deliver mono-, di-, or trisubstituted products.

Combined with the use of fluorine-18 for positron emission tomography, the discovery that incorporating fluorine atoms into drug molecules can improve their bioavailability, metabolic stability, and target specificity has driven the rapid development of new methods for generating C–F bonds and forming bond connections with fluorine-containing structural motifs over the past decade.1 However, synthesis of compounds bearing fluorovinyl (F–C Created by potrace 1.16, written by Peter Selinger 2001-2019 C) and gem-difluoroallyl (F2C–C Created by potrace 1.16, written by Peter Selinger 2001-2019 C) groups remains a challenge, despite the presence of these structural motifs in numerous drugs, such as tezacitabine,2 seletracetam,3 and tafluprost4 (Scheme 1a). We envisioned that synthesis of fluorovinyls containing an allylic bromine atom (F–C Created by potrace 1.16, written by Peter Selinger 2001-2019 C–C–Br) would facilitate the preparation of such compounds because the bromine atom would serve as a handle for a wide variety of substitution and cross-coupling reactions. The existing methods for their preparation generally rely on reactions of fluorovinyls containing an allylic hydroxyl group or gem-difluorinated vinyloxiranes with brominating reagents.5 Direct methods for their synthesis from readily accessible substrates are lacking.Open in a separate windowScheme 1Synthesis of fluorovinyls via Lewis acid activation of trifluoromethylalkenes.Elegant work from the groups of Maruoka,6 Oshima,7 Ozerov,8 Müller,9 Stephan,10 Oestreich,11 Chen,12 and Young13 on C–F bond activation reactions has proven that Lewis acid-promoted abstraction of fluoride from alkyl fluorides is a powerful tool for generating carbocations that can be trapped by nucleophiles. When trifluoromethylalkenes were studied as substrates, Ichikawa et al. reported that aryldefluorination of trifluoromethylalkenes can be accomplished with a stoichiometric amount of EtAlCl2via fluoride abstraction and subsequent Friedel–Crafts reactions between the resulting allylic carbocation and arenes (Scheme 1b).14 In addition, Braun and Kemnitz and colleagues carried out hydrodefluorination reactions of trifluoromethylalkenes with hydrosilanes catalyzed by Lewis acidic nanoscopic aluminum chlorofluoride (Scheme 1b).15 In light of these reports and our experiences in developing Lewis acid-catalyzed reactions,16 we speculated that 3,3-difluoroallyl bromides (F2C Created by potrace 1.16, written by Peter Selinger 2001-2019 C–C–Br) could be directly prepared from trifluoromethylalkenes and a suitable bromide source via Lewis acid activation of the C–F bonds and subsequent nucleophilic attack of the bromide anion at the distal olefinic carbon of the resulting allylic carbocation, a process that has no precedent in the literature.Herein, we report our discovery that by using an aluminum-based Lewis acid catalyst and bromotrimethylsilane (TMSBr) or chlorotrimethylsilane (TMSCl) as a halide source, we were able to achieve the proposed C–F bond activation/substitution reaction (Scheme 1c). Furthermore, simply by adjusting the stoichiometry of the reactants and the reaction temperature, we could selectively obtain mono-, di-, or trisubstituted products. Mechanistic studies indicated the multi-substitution reaction was achieved by thermally promoted 1,3-halogen migration of the initially formed product, followed by further halodefluorination. Notably, the previously reported defluorination reactions of trifluoromethylalkenes, either Lewis acid-catalyzed14,15 or promoted via other methods,17–19 usually provide monosubstitution products; that is, our finding that we could selectively generate multiply substituted products is also unprecedented.To test various reaction conditions, we chose α-aryl-substituted trifluoromethylalkene 1a as a model substrate (Table 1). TMSBr was selected as the bromide source because we expected the generated silyl cation to be an excellent scavenger for the displaced fluoride anion. We began by evaluating several Lewis acid catalysts and found that no reaction occurred when 1a was treated with B(C6F5)3, Zn(OTf)2, Sc(OTf)3, Al(OTf)3, or ZrCl4 (5 mol%) and 3 equiv. of TMSBr in DCE at 80 °C for 24 h (entries 1–5). However, we were encouraged to find that AlCl3 would catalyze the proposed bromodefluorination reaction, giving monobrominated product 2a and dibrominated product 3a (ref. 20) in 17% and 2% yields, respectively (entry 6). Investigation of additional aluminum-based Lewis acids showed that AlEtCl2 and Al(C6F5)3(tol)0.5 (ref. 21) had higher activities: AlEtCl2 gave 2a and 3a in 5% and 38% yields, respectively (entry 7), and Al(C6F5)3(tol)0.5 gave 16% and 32% yields, respectively (entry 8). Because Al(C6F5)3(tol)0.5 is a solid and therefore easier to store and handle than AlEtCl2 (a liquid), we chose Al(C6F5)3(tol)0.5 for further investigation. Changing the solvent to toluene inhibited the formation of 3a, but failed to improve the yield of 2a (entry 9). Coordinative solvents (acetonitrile and dioxane) shut down the reaction entirely (entries 10 and 11). When the reaction temperature was increased to 120 °C, 2a and 3a were obtained in 13% and 68% yields, respectively (entry 13). Gratifyingly, when 4 equiv. of TMSBr relative to 1a was used, 3a was generated as the sole reaction product in 90% yield (Z/E = 55 : 45, entry 14). Next, we tried using TMSBr as the limiting reagent to determine whether we could obtain the monobrominated product (2a) as the major product. Indeed, when 3 equiv. of 1a was treated with 1 equiv. of TMSBr at 80 °C, 2a was obtained as the sole product, although the yield was only 30% (entry 15). Further screening of reaction conditions revealed that using 9.0 mol% of Al(C6F5)3(tol)0.5 and running the reaction at 60 °C for 48 h (entry 16) gave the highest yield of 2a (76%; the yield of 3a was 8%).Optimization of reaction conditionsa
EntryLewis acid 1a/TMSBr T (°C)SolventYieldb2a (%)Yieldb3a (%)
1B(C6F5)31 : 380DCEn.d.n.d.
2Zn(OTf)21 : 380DCEn.d.n.d.
3Sc(OTf)31 : 380DCEn.d.n.d.
4Al(OTf)31 : 380DCEn.d.n.d.
5ZrCl41 : 380DCETracen.d.
6AlCl31 : 380DCE172
7AlEtCl21 : 380DCE538
8cAl(C6F5)3(tol)0.51 : 380DCE1632
9cAl(C6F5)3(tol)0.51 : 380Toluene16n.d.
10cAl(C6F5)3(tol)0.51 : 380CH3CNn.d.n.d.
11cAl(C6F5)3(tol)0.51 : 380Dioxanen.d.n.d.
12cAl(C6F5)3(tol)0.51 : 3100DCE2548
13cAl(C6F5)3(tol)0.51 : 3120DCE1368
14cAl(C6F5)3(tol)0.51 : 4120DCEn.d.90d
15cAl(C6F5)3(tol)0.53 : 180DCE30n.d.
16eAl(C6F5)3(tol)0.53 : 160DCE768
Open in a separate windowaUnless otherwise specified, reactions were performed with 0.1 mmol of 1a and 5 mol% of a Lewis acid in 1 mL of solvent for 24 h under N2.bYields were determined by 1H NMR using CH2Br2 as the internal standard; the 2a/3a ratios were determined by 19F NMR; n.d. = not detected.c4.5 mol% Al(C6F5)3(tol)0.5 was used as catalyst.dThe Z/E ratio was 55 : 45.eThe reaction was carried out with 9.0 mol% of Al(C6F5)3(tol)0.5 for 48 h.With the optimal conditions in hand, we first explored the scope of the monosubstitution reaction by testing various trifluoromethyl- and difluoroalkyl-substituted olefins 1 (Table 2, left column). From 1a, monobrominated product 2a could be isolated in pure form in 64% yield by means of preparative HPLC. When the α-phenyl ring bore an ortho-phenyl substituent, the reaction still afforded 2b in 58% yield despite the increased steric bulk around the reaction site. When the α substituent was changed to a 9-phenanthryl group (1c), monobrominated product 2c was isolated in 75% yield. We also tested other halogenating reagents with 1c: TMSI gave iodinated product 2c-I in 51% yield, whereas TMSCl was poorly reactive, giving a <10% yield of product. Furthermore, substrates with 1-naphthyl (2d), 4-dibenzothiophenyl (2e), and 4-dibenzofuranyl (2f) moieties at the α position were all suitable. Interestingly, even the reaction of conjugated diene 1g was feasible, giving brominated diene 2g in 67% isolated yield. In addition, a series of α-alkyl-substituted trifluoromethylalkenes gave the desired products (2h–2k) in moderate yields. Difluoroalkyl-substituted alkenes were also reactive; specifically, benzene-fused methylenecycloalkanes 1l–1n gave the corresponding products (2l–2n) in 59–88% yields. Finally, acyclic substrate 1o afforded (E)-2o as the predominant isomer (E/Z > 10 : 1) in 50% yield.Scope of the mono and disubstitution reactiona
Open in a separate windowaCondition A: reactions were performed with 0.6 mmol of 1, 0.2 mmol of TMSBr, and 9.0 mol% of Al(C6F5)3(tol)0.5 in 1.5 mL of DCE at 60 °C for 48 h; condition B: reactions were performed with 0.2 mmol of 1, 0.8 mmol of TMSBr, and 4.5 mol% of Al(C6F5)3(tol)0.5 in 1.5 mL of DCE at 120 °C for 24 h; isolated yields are reported.bThe reaction was performed at 80 °C.cTMSI was used instead of TMSBr.dThe reaction was carried out with 13.5 mol% of Al(C6F5)3(tol)0.5.e4 equiv. of 1 was used.fThe reaction was performed with 5 equiv. of TMSBr.Next the scope of the disubstitution reaction was investigated (Table 2, right column). Trifluoromethyl-substituted alkenes bearing electron-donating or electron-withdrawing groups on the α-aryl ring were reactive, affording the corresponding products (3a and 3p–3r) in 69–85% yields with Z/E ratios of approximately 1 : 1. 1-Naphthyl (3d), 4-dibenzothiophenyl (3e), 4-dibenzofuranyl (3f), and aliphatic (3h–3j, 3s, and 3t) substituents at the α position were well tolerated. Interestingly, even alkynyl-substituted trifluoromethylalkenes afforded the desired disubstituted products (3u and 3v) in good yields. In addition, difluoroalkyl-substituted alkenes 1n and 1o gave completely defluorinated products 3n and 3o in 46% and 60% yields, respectively. Notably, under these conditions, the monobrominated products either did not form or formed in only trace amounts, as indicated by GC-MS or NMR spectroscopy. Moreover, the E and Z isomers of dibrominated products were found interconvertible under the reaction conditions (for details, see the ESI) so the Z/E ratios of products might be the result of the thermodynamic equilibrium.It is also worth mentioning that some substrates shown in Table 2 were not compatible either with the monosubstitution reaction or with the disubstitution reaction. For example, substrates bearing coordinative functional groups, such as methoxy, carbonyl, sulfonyl and alkyne (1p, 1q, 1r, 1t, 1u, and 1v), gave very low yields (<20%) for monosubstitution, perhaps because the relatively low reaction temperature (60 °C) was not sufficient to break the coordination of these functional groups to the Lewis acid catalyst. Furthermore, Al(C6F5)3(tol)0.5 is probably a precatalyst because Al(C6F5)3(tol)0.5 rapidly decomposes in DCE to give a mixture of unidentified aluminum species21b that are active for the halodefluorination reaction (for details, see the ESI).We performed several control experiments to explore the reaction mechanism. When substrate 1a was treated with mesitylene in the presence of 1 equiv. of Al(C6F5)3(tol)0.5, Friedel–Crafts allylation of the aromatic ring generated product 4 in 96% yield (Scheme 2a).22 This result demonstrates that the aluminum Lewis acid could abstract fluoride from the trifluoromethylalkene to generate an allylic carbocation. Furthermore, when 2a was subjected to the conditions used for the disubstitution reaction, 3a was isolated in 65% yield (Scheme 2b), indicating that the dibrominated products were generated via monobrominated intermediates. However, subjecting nonbrominated 5 to the same conditions did not result in substitution of the vinylic fluorine atom by the bromine atom (6, Scheme 2c), which excludes the vinylic nucleophilic substitution (SNV) mechanism23 for the conversion from 2a to 3a. We thus suspected that the allylic bromine atom in 2a was involved in this conversion. Indeed, when 2a was heated at 120 °C in toluene for 12 h, 1,3-migration of the bromine atom gave bromodifluoromethylalkene 7 in 83% NMR yield (Scheme 2d).24 And, treatment of 7 with TMSBr in the presence of the catalyst at 120 °C gave 3a in 77% yield (Scheme 2e). Taken together, these results indicate that dibrominated products were generated via isomerization of the monobrominated product to form bromodifluoromethylalkenes, which then underwent a second bromodefluorination reaction. In addition, silylium Et3Si[B(C6F5)4]25 was found incapable of catalyzing the bromodefluorination reaction (Scheme 2f). This result suggests that the Lewis acidic aluminum is probably a catalyst, rather than an initiator, and TMS+ from TMSBr abstracts the fluoride from the aluminum–fluoride adduct to regenerate the active catalyst.Open in a separate windowScheme 2Control experiments.These results led us to wonder whether all three fluorine atoms of a trifluoromethylalkene could be replaced with bromine atoms via a 1,3-bromo migration reaction of the dibrominated product to give a dibromofluoromethylalkene, which would then undergo bromodefluorination. After screening various reaction conditions, we discovered that tribrominated products could be obtained by using a large excess (e.g., 10 equiv.) of TMSBr and extending the reaction time; however, in all cases, substantial amounts of the dibrominated products were always produced as well (see Table S1 in the ESI), which made separation of the product difficult. However, we were delighted to find that when TMSCl was used in large excess (7 equiv.) and the reaction temperature was 120 °C, trichlorinated compounds were the major or only products (Table 3). However, these conditions were suitable only for substrates bearing α-aryl substituents. The moderate to low yields of these reactions were due mainly to decomposition of the starting materials rather than to the formation of mono- or dichlorinated byproducts.Scope of trisubstitution reactiona
Open in a separate windowaUnless otherwise specified, reactions were performed with 0.2 mmol of 1, 1.4 mmol of TMSCl, and 9.0 mol% of Al(C6F5)3(tol)0.5 in 1.5 mL of DCE at 120 °C for 24 h; isolated yields are reported.As mentioned above, bromine atoms are among the most useful substituents for introducing other functional groups. To explore the utility of the above-described reactions, we carried out some transformations of the products (Scheme 3). For example, treatment of monobrominated product 2d with estrone under basic conditions delivered phenoxy-substituted product 9 in 65% yield via an SN2′ reaction. Additionally, azide and an indole were also suitable nucleophiles for SN2′ reactions, giving 10 and 11 in 50% and 79% isolated yields, respectively. Furthermore, a Suzuki coupling reaction of 2d with an arylboronic acid delivered coupling product 12 in 61% yield, and treatment of 2d with hexaldehyde gave alcohol 13 (63% yield) via an indium-mediated gem-difluoroallylation reaction.5b Reaction of dibrominated product 3a with an allyl Grignard reagent selectively replaced the allylic bromide to give compound 14. Subsequent electrophilic fluorination of 14 with Selectfluor in the presence of MeOH afforded α-CF2Br-substituted ether 15 in 41% yield. In addition, 14 could undergo a Pd-catalyzed intramolecular Heck reaction to generate fluoro-substituted cyclopentene 16 in 54% yield. Notably, both (Z)- and (E)-14 underwent these last two transformations to give a single product, thus eliminating the need to separate the isomers.Open in a separate windowScheme 3Transformations of products 2d and 3a.  相似文献   

20.
Electrophilic fluoroalkylthiolation induced diastereoselective and stereospecific 1,2-metalate migration of alkenylboronate complexes     
Feng Shen  Long Lu  Qilong Shen 《Chemical science》2020,11(30):8020
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   

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