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 62 毫秒
1.
The asymmetric total synthesis of (+)-xestoquinone and (+)-adociaquinones A and B was achieved in 6–7 steps using an easily accessible meso-cyclohexadienone derivative. The [6,6]-bicyclic decalin B–C ring and the all-carbon quaternary stereocenter at C-6 were prepared via a desymmetric intramolecular Michael reaction with up to 97% ee. The naphthalene diol D–E ring was constructed through a sequence of Ti(Oi-Pr)4-promoted photoenolization/Diels–Alder, dehydration, and aromatization reactions. This asymmetric strategy provides a scalable route to prepare target molecules and their derivatives for further biological studies.

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

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

2.
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   

3.
The cross-coupling of N-tosylhydrazones has emerged as a powerful method for the construction of structurally diverse molecules, but the development of catalytic enantioselective versions still poses considerable challenges and only very limited examples have been reported. We herein report an asymmetric palladium/GF-Phos-catalyzed carbenylative amination reaction of N-tosylhydrazones and (E)-vinyl iodides pendent with amine, which allows facile access to a range of chiral pyrrolidines and piperidines in good yields (45–93%) with up to 96.5 : 3.5 er. Moreover, mild conditions, general substrate scope, scaled-up preparation, as well as the efficient synthesis of natural product (−)-norruspoline are practical features of this method.

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

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

4.
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   

5.
Reductive radical-initiated 1,2-C migration assisted by an azidyl group     
Xueying Zhang  Zhansong Zhang  Jin-Na Song  Zikun Wang 《Chemical science》2020,11(30):7921
We report here a novel reductive radical-polar crossover reaction that is a reductive radical-initiated 1,2-C migration of 2-azido allyl alcohols enabled by an azidyl group. The reaction tolerates diverse migrating groups, such as alkyl, alkenyl, and aryl groups, allowing access to n+1 ring expansion of small to large rings. The possibility of directly using propargyl alcohols in one-pot is also described. Mechanistic studies indicated that an azidyl group is a good leaving group and provides a driving force for the 1,2-C migration.

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

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

6.
Correction: Polariton chemistry: controlling molecular dynamics with optical cavities     
Raphael F. Ribeiro  Luis A. Martínez-Martínez  Matthew Du  Jorge Campos-Gonzalez-Angulo  Joel Yuen-Zhou 《Chemical science》2020,11(45):12371
  相似文献   

7.
Enantioselective α-amination enabled by a BINAM-derived phase-transfer catalyst     
H. M. Nelson  J. S. Patel  H. P. Shunatona  F. D. Toste 《Chemical science》2015,6(1):170-173
  相似文献   

8.
Development of an active site titration reagent for α-amylases     
Ryan P. Sweeney  Phillip M. Danby  Andreas Geissner  Ryan Karimi  Jesper Brask  Stephen G. Withers 《Chemical science》2021,12(2):683
α-Amylases are among the most widely used classes of enzymes in industry and considerable effort has gone into optimising their activities. Efforts to find better amylase mutants, such as through high-throughput screening, would be greatly aided by access to precise and robust active site titrating agents for quantitation of active mutants in crude cell lysates. While active site titration reagents designed for retaining β-glycosidases quantify these enzymes down to nanomolar levels, convenient titrants for α-glycosidases are not available. We designed such a reagent by incorporating a highly reactive fluorogenic leaving group onto unsaturated cyclitol ethers, which have been recently shown to act as slow substrates for retaining glycosidases that operate via a covalent ‘glycosyl’-enzyme intermediate. By appending this warhead onto the appropriate oligosaccharide, we developed efficient active site titration reagents for α-amylases that effect quantitation down to low nanomolar levels.

α-Amylases are among the most widely used classes of enzymes in industry and considerable effort has gone into optimising their activities.

Amylases are among the most common classes of enzymes employed in industrial settings, being used in detergents, bread, beer, biofuel, and many other sectors. Accordingly, α-amylases account for 25% of the world''s multi-billion dollar enzyme market.1,2 α-Amylases are endo-acting enzymes that cleave starch into malto-oligosaccharides, which are further degraded by exo-acting α-glucosidases, glucoamylases, β-amylases and α-glucan phosphorylases and lyases. They are found in CAZy GH families 13, 57, 119 and 126, with the vast majority in the large GH13 family.3 GH13 enzymes adopt a (β/α)8 fold with three highly conserved active site carboxylic acids.4–6 They employ a classical double-displacement mechanism7 in which one of the glutamic acids provides acid catalytic assistance to the leaving group departure while an aspartate attacks the anomeric centre, forming a covalent glycosyl enzyme intermediate. In a second step, water attacks the anomeric centre with base assistance from the glutamate residue (Fig. 1A and B).Open in a separate windowFig. 1Koshland mechanism of retaining β- and α-glycosidases (A & B). The same mechanism has been observed for the hydrolysis of “β”-valienols (C), and for “α”-valienols (D).Given their industrial importance, a huge amount of attention has been given to the discovery and improvement of α-amylases to attain optimal performance for particular applications. These approaches typically require high-throughput analysis of large numbers of gene products or mutants thereof.8–10 Identification of the best candidates then ideally requires high-throughput assay coupled with a method for determining the enzyme concentration in each sample. This can be a challenging task in the absence of purification, as would be the case for truly high-throughput approaches. The “gold standard” method to quantify active enzyme concentration is active site titration.11 Active site titrants react stoichiometrically with their target enzymes and release one equivalent of a quantifiable agent, which is typically either a chromophore or fluorophore. For enzymes that operate via a covalent intermediate, such as retaining glycosidases, the active site titrants are usually chromogenic or fluorogenic substrates that form this intermediate with a rate constant (kon) that is much greater than that for its hydrolysis (koff) – ideally with koff approaching zero.Our lab has previously developed active site titration reagents for several retaining β-glycosidases12,13 and neuraminidases.14,15 By replacing the substituent on the position adjacent to the anomeric centre of the sugar (the hydroxyl at C-2 for many monosaccharides) with a fluorine atom, both the formation and the hydrolysis of the glycosyl-enzyme intermediate are slowed, largely through inductive destabilisation of the transition state. Further incorporation of a reactive fluorogenic leaving group generates a reagent that, upon covalently inactivating the glycosidase, releases a stoichiometric and quantifiable amount of fluorophore. The fluorogenic response is then measured to determine the amount of active glycosidase that is present in solution.Unfortunately, this same strategy does not work for retaining α-glycosidases. In those cases, koff remains greater than kon, likely due to the inherently greater reactivity of the β-glycosyl-enzyme intermediate,16,17 and the compounds are simply substrates with low turnover numbers. By use of 2,2-dihalosugars with yet more reactive leaving groups, this problem could be solved in some cases, but their synthesis is challenging, and inactivation rates were low, or non-existent in some cases.18,19 Alternative approaches were called for.Recently, a new class of glycosidase substrates was reported in which the sugar moiety is replaced by an equivalently hydroxylated cyclohexene.20–23 Hydrolysis of these enol ethers likely occurs via an allylic cation of almost identical reactivity to that of the equivalent oxocarbenium ion. Glycosidases cleave these substrates via the classical Koshland mechanism7 (Fig. 1C and D), but considerably more slowly than their natural substrates. However, incorporation of a good leaving group will accelerate, relatively, the first step such that, in some cases, they act as mechanism-based inactivators making them candidates for development of an active site titrant for α-amylases.Since α-amylases are endo-acting enzymes that do not usually cleave monosaccharide glycosides, an ‘extended” oligosaccharide version containing a total of 2 or 3 sugar/pseudosugar moieties would be needed. Substrates longer than this would be prone to internal glycoside cleavage. Since 2-chloro-4-nitrophenyl maltotrioside (CNP-G3) functions as a substrate for most amylases, we focused on addition of a maltosyl unit to a valienol moiety containing a 6,8-difluorocoumarin (F2MU) leaving group at its “anomeric centre”. The low pKa of this coumarin, 4.7,14 results in a greater reactivity of the reagent and also ensures the coumarin will be deprotonated and thus fluorescent, upon release at neutral pH.Synthesis of partially protected alcohol 2 from gluconolactone 1via literature methods24 was followed by attachment of F2MU via a Mitsunobu reaction and subsequent removal of the protecting groups under acidic conditions, generating known pseudo-glycoside 3.23 To check this concept before we synthesized the longer version, we tested compound 3 as a titrant of a simple α-glucosidase and found that it did indeed titrate the enzyme (Fig. S5). Since elongation of this pseudosugar via classical organic synthesis would require substantial protecting group chemistry, we elected instead to employ an enzymatic coupling strategy using the GH13 cyclodextrin transglycosidase, CGTase. This enzyme can use glycosyl fluorides, such as α-maltosyl fluoride, to effect glycosyl transfer onto suitable acceptors. However, a significant competing reaction would involve self-condensation of glycosyl fluorides ultimately forming cyclodextrins. To avoid this problem, we employed a maltosyl fluoride donor (4), in which the 4′-hydroxyl had been capped with a methyl group.25,26 Incorporation of 4′-methoxy groups does not alter the reaction with α-amylases, as this site in the normal substrate is occupied by additional sugar residues. Thus CGTase-catalysed glycosylation between known glycosyl fluoride 4 and pseudo-glycoside 3, gave the pseudo-trisaccharide 5 in 64% isolated yield (Scheme 1).Open in a separate windowScheme 1Synthesis of titration reagent 5.With this reagent in hand, we proceeded to screen its ability to inactivate a small panel of α-amylases. As shown in Fig. 2, time-dependent inactivation was observed for all enzymes tested, with the most industrially relevant enzymes, Effusibacillus pohliae amylase (EPA) and Aspergillus oryzae amylase (AOA), being inactivated the fastest.Open in a separate windowFig. 2Time-dependent inactivation of a small panel of amylases, showing remaining % activity versus time. Red box with X: AOA (91 nM); blue square: EPA (66.7 nM); purple cross: PPA (500 nM); green triangle: HPA (125 nM). AOA = A. oryzae amylase; EPA = E. pohliae amylase; HPA = human pancreatic amylase; PPA = porcine pancreatic amylase.Kinetic parameters for inactivation were then determined by directly monitoring the release of F2MU by UV-Vis (Table 1). To determine kon and koff (Scheme 2), we monitored chromophore (F2MU) release by absorbance at 370 or 380 nm (dependent on the concentration of 5 in the measurements of each enzyme). After mixing 5 with each enzyme individually, a burst phase followed by a steady-state phase was observed. For each enzyme, this was then repeated with varying concentrations of 5. Initial rates of F2MU release versus concentration of 5 were fit to a Michaelis–Menten equation to provide kon. The rate constant of cyclitol release, koff, was determined by measuring rates of the steady-state region at a saturating concentration (5× Ki). We found that several amylases: Effusibacillus pohliae amylase (EPA), Aspergillus oryzae amylase (AOA), Rhizomucor pusillus amylase (RPA) and porcine pancreatic amylase (PPA), inactivated quickly (highest kon, lowest koff, and greatest kon/Ki), and are therefore ideal candidates for titration with compound 5. Human pancreatic amylase (HPA), on the other hand, while inactivating rapidly, binds the reagent relatively poorly.Open in a separate windowScheme 2Kinetic parameters for the hydrolysis of 5 by several amylases (at 25 °C for EPA, AOA, and RPA and 30 °C for human pancreatic amylase [HPA] and porcine pancreatic amylase [PPA])
Enzyme: K i (μM) k on (min−1) k on/Ki (min−1 nM−1) k off (min−1)
HPA30400.200.07n.d.
PPA2280.482.10.024
EPA45.01.7028.90.002
AOA73.00.344.60.001
RPA1600.251.60.004
Open in a separate windowConfirmation that the inactivation observed was a result of stoichiometric covalent derivatisation of the enzyme was obtained for three representative enzymes by monitoring the enzyme molecular weights before and after inactivation, by electrospray ionisation mass spectrometry. As shown in Table 2, Fig. 3 and in ESI Fig. S1, the mass of each enzyme after the reaction was increased by the expected 496 mass units relative to that of the unlabelled enzyme. Complete reaction is evident by the absence of any unlabelled enzyme peak after addition of 5 (Fig. 3 and S1). It should also be noted, that all glycoforms of AOA underwent complete reaction (Fig. 3).Open in a separate windowFig. 3MS-plot of AOA before (left side, blue) and after (right side, green) addition of 5 showing the expected increase of 496 mass units.Intact-MS results confirming the stoichiometric addition of pseudo-trisaccharide (+496 m/z) to each amylase
Enzyme:Enzyme (M + H+)Enzyme + 5 (M + H+)
HPA56 066.756 563.0
EPA75 190.675 686.0
AOA53 863.354 358.6
Open in a separate windowHaving demonstrated that these reagents function as effective time-dependent inactivators, we then evaluated their utility as active site titration agents. Initially, we tested the ability of compound 5 to titrate ∼100 nM EPA (Fig. 4A) and observed a classical burst of fluorescence followed by a steady-state turnover phase. The active enzyme concentration can be quantified from such plots by extrapolating the steady-state portion back to the y-intercept (t = 0) and fitting the burst to eqn (1), as described previously.11Burst = [E]0 × (kon/kon + koff)21Open in a separate windowFig. 4(A) Titration of EPA with 100 μM of 5. The release of F2MU was monitored fluorimetrically (λEx = 353 nm, λEm = 451 nm). The steady-state signal was extrapolated back to the y-axis to get the corresponding burst amplitude and fit to eqn (1) to give enzyme concentration. (B) Titration of a set of serial dilutions of EPA. Red small box with X: 100 nM; yellow triangle: 50 nM; green (small) square: 25 nM; blue-green cross: 12.5 nM; blue (large) square: 6.3 nM; pink star: 3.1 nM. Burst responses show a detectable response down to 3 nM. (C) Plot of dilution factor versus the calculated [Novamyl], based on the burst response in Fig. 4B with the colours corresponding to the concentrations indicated in Fig. 4B.The values we obtained in this way are shown in Table 3 and compared with the concentrations claimed by the manufacturers: agreement was excellent for EPA and AOA. To explore a case in which total protein concentration was likely to be different from active concentration, we titrated a commercial porcine pancreatic amylase (PPA) that was purchased from Sigma-Aldrich several years prior to this set of experiments. The concentration of the active enzyme was found to be almost 4-fold lower than the original stated value. This discrepancy is likely due to the degradation of the enzyme over the long storage period. Indeed, such discrepancies in active enzyme concentration and total protein concentration are exactly what the titration agent is designed to detect.Concentrations of several amylases determined by titration with reagent 5, compared with manufacturers'' listed concentrationa
Enzyme:[Manufacturers] (nM)[Titration] (nM)
AOA9193
EPA100116
PPA610164
Yeast α-glucosidaseb88007500
Open in a separate windowaDetermined by declared activity measurements.bSubstrate 3 was used for titration of this enzyme. The manufacturer''s concentration was determined by measuring the enzyme activity with PNP-Glc (Vmax) and using their reported units of activity vs. PNP-Glc.Finally, the sensitivity and linearity of the response of our titration reagent 5 were tested using a crude, commercial batch of EPA (5 mg mL−1; 66.7 μM). After initial dilution of this batch of EPA to 100 nM, a series of 2× dilutions gave a reliable and detectable fluorogenic response down to enzyme concentrations as low as 3 nM (Fig. 4B and C).  相似文献   

9.
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   

10.
Rh(iii)-catalyzed tandem annulative redox-neutral arylation/amidation of aromatic tethered alkenes     
Chao Chen  Chen Shi  Yaxi Yang  Bing Zhou 《Chemical science》2020,11(44):12124
Transition-metal-catalyzed directed C–H functionalization has emerged as a powerful and straightforward tool to construct C–C bonds and C–N bonds. Among these processes, the intramolecular annulative alkene hydroarylation reaction has received much attention because this intramolecular annulation can produce more complex and high value-added structural motifs found in numerous natural products and bioactive molecules. Despite remarkable progress, these annulative protocols developed to date remain limited to hydroarylation and functionalization of one side of alkenes, thus largely limiting the structural diversity and complexity. Herein, we developed a rhodium(iii)-catalyzed tandem annulative arylation/amidation reaction of aromatic tethered alkenes to deliver a variety of 2,3-dihydro-3-benzofuranmethanamine derivatives bearing an all-carbon quaternary stereo center by employing 3-substituted 1,4,2-dioxazol-5-ones as an amidating reagent to capture the transient C(sp3)–Rh intermediate. Notably, by simply changing the directing group, a second, unsymmetrical ortho C–H amidation/annulation can be achieved to provide tricyclic dihydrofuro[3,2-f]quinazolinones in good yields.

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

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

11.
Ligand-promoted palladium-catalyzed β-methylene C–H arylation of primary aldehydes     
Ke Yang  Zhi Li  Chong Liu  Yunjian Li  Qingyue Hu  Mazen Elsaid  Bijin Li  Jayabrata Das  Yanfeng Dang  Debabrata Maiti  Haibo Ge 《Chemical science》2022,13(20):5938
The transient directing group (TDG) strategy allowed long awaited access to the direct β-C(sp3)–H functionalization of unmasked aliphatic aldehydes via palladium catalysis. However, the current techniques are restricted to terminal methyl functionalization, limiting their structural scopes and applicability. Herein, we report the development of a direct Pd-catalyzed methylene β-C–H arylation of linear unmasked aldehydes by using 3-amino-3-methylbutanoic acid as a TDG and 2-pyridone as an external ligand. Density functional theory calculations provided insights into the reaction mechanism and shed light on the roles of the external and transient directing ligands in the catalytic transformation.

Aliphatic aldehydes are among the most common structural units in organic and medicinal chemistry research. Direct C–H functionalization has enabled efficient and site-selective derivatization of aliphatic aldehydes.

Simple aliphatic functional groups enrich the skeletal backbones of many natural products, pharmaceuticals, and other industrial materials, influencing the utility and applications of these substances and dictating their reactivity and synthetic modification pathways. Aliphatic aldehydes are some of the most ubiquitous structural units in organic materials.1 Their relevance in nature and industry alike, combined with their reactivity and synthetic versatility, attracted much attention from the synthetic organic and medicinal chemistry communities over the years (Fig. 1).2 Efficient means to the functionalization of these molecules have always been highly sought after.Open in a separate windowFig. 1Select aliphatic aldehyde-containing medicines and biologically active molecules.Traditionally, scientists have utilized the high reactivity of the aldehyde moiety in derivatizing a variety of functional groups by the means of red-ox and nucleophilic addition reactions. The resourceful moiety was also notoriously used to install functional groups at the α-position via condensation and substitution pathways.3 Although β-functionalization is just as robust, it has generally been more restrictive as it often requires the use of α,β-unsaturated aldehydes.4,5 Hence, transition metal catalysis emerged as a powerful tool to access β-functionalization in saturated aldehydes.6 Most original examples of metal-catalyzed β-C–H functionalization of aliphatic aldehydes required the masking of aldehydes into better metal coordinating units since free unmasked aldehydes could not form stable intermediates with metals like palladium on their own.7 Although the masking of the aldehyde moiety into an oxime, for example, enabled the formation of stable 5-membered palladacycles, affording β-functionalized products, this system requires the installation of the directing group prior to the functionalization, as well as the subsequent unmasking upon the reaction completion, compromising the step economy and atom efficiency of the overall process.8 Besides, some masking and unmasking protocols might not be compatible with select substrates, especially ones rich in functional groups. As a result, the development of a one-step direct approach to the β-C–H functionalization of free aliphatic aldehydes was a demanding target for synthetic chemists.α-Amino acids have been demonstrated as effective transient directing groups (TDGs) in the remote functionalization of o-alkyl benzaldehydes and aliphatic ketones by the Yu group in 2016.9 Shortly after, our group disclosed the first report on the direct β-C–H arylation of aliphatic aldehydes using 3-aminopropanoic acid or 3-amino-3-methylbutanoic acid as a TDG.10 The TDG was found to play a similar role to that of the oxime directing group by binding to the substrate via reversible imine formation, upon which, it assists in the assembly of a stable palladacycle, effectively functionalizing the β-position.11 Since the binding of the TDG is reversible and temporary, it is automatically removed upon functionalization, yielding an efficient and step-economic transformation. This work was succeeded by many other reports that expanded the reaction and the TDG scopes.12–14 However, this system suffers from a significant restriction that demanded resolution; only substitution of methyl C–H bonds of linear aldehydes was made possible via this approach (Scheme 1a–e). The steric limitations caused by incorporating additional groups at the β-carbon proved to compromise the formation of the palladacycle intermediate, rendering the subsequent functionalization a difficult task.12Open in a separate windowScheme 1Pd-catalyzed β-C–H bond functionalization of aliphatic aldehydes enabled by transient directing groups.Encouraged by the recent surge in use of 2-pyridone ligands to stabilize palladacycle intermediates,15,16 we have successfully developed the first example of TDG-enabled Pd-catalyzed methylene β-C–H arylation in primary aldehydes via the assistance of 2-pyridones as external ligands (Scheme 1f). The incorporation of 2-pyridones proved to lower the activation energy of the C–H bond cleavage, promoting the formation of the intermediate palladacycles even in the presence of relatively bulky β-substituents.17 This key advancement significantly broadens the structural scopes and applications of this process and promises future asymmetric possibilities, perhaps via the use of a chiral TDG or external ligand or both. Notably, a closely related work from Yu''s group was published at almost the same time.18We commenced our investigation of the reaction parameters by employing n-pentanal (1a) as an unbiased linear aldehyde and 4-iodoanisole (2a) in the presence of catalytic Pd(OAc)2 and stoichiometric AgTFA, alongside 3-amino-3-methylbutanoic acid (TDG1) and 3-(trifluoromethyl)-5-nitropyridin-2-ol (L1) at 100 °C (ii) sources proved Pd(OAc)2 to be the optimal catalyst, while Pd(TFA)2, PdCl2 and PdBr2 provided only moderate yields (entries 10–12). Notably, a significantly lower yield was observed in the absence of the 2-pyridone ligand, and no desired product was isolated altogether in the absence of the TDG (entries 13 and 14). The incorporation of 15 mol% Pd catalyst was deemed necessary after only 55% yield of 3a was obtained when 10 mol% loading of Pd(OAc)2 was instead used (entry 15).Optimization of reaction conditionsa
EntryPd sourceL (mol%)TDG1 (mol%)Solvent (v/v, mL)Yield (%)
1Pd(OAc)2L1 (30)TDG1 (40)HFIP30
2Pd(OAc)2L1 (30)TDG1 (40)AcOH<5
3Pd(OAc)2L1 (30)TDG1 (40)HFIP/AcOH (1 : 1)28
4Pd(OAc)2L1 (30)TDG1 (40)HFIP/AcOH (9 : 1)47
5Pd(OAc)2L1 (30)TDG1 (40)HFIP/AcOH (1 : 9)<5
6Pd(OAc)2L1 (30)TDG1 (60)HFIP/AcOH (9 : 1)50
7Pd(OAc)2L1 (30)TDG1 (80)HFIP/AcOH (9 : 1)25
8Pd(OAc)2L1 (60)TDG1 (60)HFIP/AcOH (9 : 1)70(68)b
9Pd(OAc)2L1 (75)TDG1 (60)HFIP/AcOH (9 : 1)51
10Pd(TFA)2L1 (60)TDG1 (60)HFIP/AcOH (9 : 1)60
11PdCl2L1 (60)TDG1 (60)HFIP/AcOH (9 : 1)52
12PdBr2L1 (60)TDG1 (60)HFIP/AcOH (9 : 1)54
13Pd(OAc)2TDG1 (60)HFIP/AcOH (9 : 1)9
14Pd(OAc)2L1 (60)HFIP/AcOH (9 : 1)0
15cPd(OAc)2L1 (60)TDG1 (60)HFIP/AcOH (9 : 1)55
Open in a separate windowaReaction conditions: 1a (0.2 mmol), 2a (0.4 mmol), Pd source (15 mol%), AgTFA (0.3 mmol), L1, TDG1, solvent (2.0 mL), 100 °C, 12 h. Yields are based on 1a, determined by 1H-NMR using dibromomethane as an internal standard.bIsolated yield.cPd(OAc)2 (10 mol%).To advance our optimization of the reaction conditions, a variety of 2-pyridones and TDGs were tested (Scheme 2). Originally, pyridine-2(1H)-one (L2) was examined as the external ligand, but it only yielded the product (3a) in 7% NMR yield. Similarly, other mono- and di-substituted 2-pyridone ligands (L3–L10) also produced low yields, fixating L1 as the optimal external ligand. Next, various α- and β-amino acids (TDG1–10) were evaluated, yet TDG1 persisted as the optimal transient directing group. These amino acid screening results also suggest that a [5,6]-bicyclic palladium species is likely the key intermediate in this protocol since only β-amino acids were found to provide appreciable yields, whereas α-amino acids failed to yield more than trace amounts of the product. The supremacy of TDG1 when compared to other β-amino acids is presumably due to the Thorpe–Ingold effect that perhaps helps facilitate the C–H bond cleavage and stabilize the [5,6]-bicyclic intermediate further.Open in a separate windowScheme 2Optimization of 2-pyridone ligands and transient directing groups.With the optimized reaction conditions in hand, substrate scope study of primary aliphatic aldehydes was subsequently carried out (Scheme 3). A variety of linear primary aliphatic aldehydes bearing different chain lengths provided the corresponding products 3a–e in good yields. Notably, relatively sterically hindered methylene C–H bonds were also functionalized effectively (3f and 3g). Additionally, 4-phenylbutanal gave rise to the desired product 3h in a highly site-selective manner, suggesting that functionalization of the methylene β-C–H bond is predominantly favored over the more labile benzylic C–H bond. It is noteworthy that the amide group was also well-tolerated and the desired product 3j was isolated in 60% yield. As expected, with n-propanal as the substrate, β-mono- (3k1) and β,β-disubstituted products (3k2) were isolated in 22% and 21% yields respectively. However, in the absence of the key external 2-pyridone ligand, β-monosubstituted product (3k1) was obtained exclusively, albeit with a low yield, indicating preference for functionalizing the β-C(sp3)–H bond of the methyl group over the benzylic methylene group.Open in a separate windowScheme 3Scope of primary aliphatic aldehydes. Reaction conditions: 1 (0.2 mmol), 2a (0.4 mmol), Pd(OAc)2 (15 mol%), AgTFA (0.3 mmol), L1 (60 mol%), TDG1 (60 mol%), HFIP (1.8 mL), HOAc (0.2 mL), 100 °C, 12 h. Isolated yields. aL1 (60 mol%) was absent and yields are given in parentheses.Next, substrate scope study on aryl iodides was surveyed (Scheme 4). Iodobenzenes bearing either an electron-donating or electron-withdrawing group at the para-, meta-, or ortho-position were all found compatible with our catalytic system (3l–3ah). Surprisingly, ortho-methyl- and fluoro-substituted aryl iodides afforded the products in only trace amounts. However, aryl iodide with ortho-methoxy group provided the desired product 3ac in a moderate yield. Notably, a distinctive electronic effect pattern was not observed in the process. It should be mentioned that arylated products bearing halogen, ester, and cyano groups could be readily converted to other molecules, which significantly improves the synthetic applicability of the process. Delightfully, aryl iodide-containing natural products like ketoprofen, fenchol and menthol were proven compatible, supplying the corresponding products in moderate yields. Unfortunately, (hetero)aryl iodides including 2-iodopyridine, 3-iodopyridine, 4-iodopyridine and 4-iodo-2-chloropyridine failed to produce the corresponding products. Although our protocol provides a novel and direct pathway to construct β-arylated primary aliphatic aldehydes, the yields of most examples are modest. The leading reasons for this compromise are the following: (1) aliphatic aldehydes are easily decomposed or oxidized to acids; (2) some of the prepared β-arylated aldehyde products may be further transformed into the corresponding α,β-unsaturated aldehydes.Open in a separate windowScheme 4Scope of aryl iodides. Reaction conditions: 1a (0.2 mmol), 2 (0.4 mmol), Pd(OAc)2 (15 mol%), AgTFA (0.3 mmol), L1 (60 mol%), TDG1 (60 mol%), HFIP (1.8 mL), HOAc (0.2 mL), 100 °C, 12 h. Isolated yields.Density functional theory (DFT) calculations were performed to help investigate the reaction mechanism and to elucidate the role of the ligand in improving the reactivity (Fig. 2). The condensation of the aliphatic aldehyde 1a with the TDG to form imine-1a was found thermodynamically neutral (ΔG° = −0.1 kcal mol−1). As a result, it was permissible to use imine-1a directly in the calculations. According to the calculations results, the precatalyst [Pd(OAc)2]3, a trimeric complex, initially experiences dissociation and ligand metathesis with imine-1a to generate the Pd(ii) intermediate IM1, which is thermodynamically favorable by 21.9 kcal mol−1. Consequently, the deprotonated imine-1a couples to the bidentate ligand to form the stable six-membered chelate complex IM1. Therefore, IM1 is indeed the catalyst resting state and serves as the zero point to the energy profile. We have identified two competitive pathways for the Pd(ii)-catalyzed C–H activation starting from IM1, one of which incorporates L1 and another which does not. On the one hand, an acetate ligand substitutes one imine-1a chelator in IM1 to facilitate the subsequent C–H activation leading to IM2, which undergoes C(sp3)–H activation through concerted metalation-deprotonation (CMD) viaTS1 (ΔG = 37.4 kcal mol−1). However, this kinetic barrier is thought to be too high to account for the catalytic activity at 100 °C. On the other hand, the chelate imine-1a could be replaced by two N-coordinated ligands (L1) leading to the Pd(ii) complex IM3. This process is endergonic by 6.4 kcal mol−1. To allow the ensuing C–H activation, IM3 dissociates one ligand (L1) producing the active species IM4, which undergoes TS2 to cleave the β-C(sp3)–H bond and form the [5,6]-bicyclic Pd(ii) intermediate IM5. Although this step features an energy barrier of 31.2 kcal mol−1, it is thought to be feasible under the experimental conditions (100 °C). Possessing similar coordination ability to that of pyridine, the ligand (L1) effectively stabilizes the Pd(ii) center in the C–H activation process, indicating that this step most likely involves a manageable kinetic barrier. This result explicates the origin of the ligand-enabled reactivity (TS2vs.TS1). Additionally, we considered the γ-C(sp3)–H activation pathway viaTS2′ which was found to have a barrier of 35.5 kcal mol−1. The higher energy barrier of TS2′ compared to that of TS2 is attributed to its larger ring strain in the [6,6]-bicyclic Pd(ii) transition state, which reveals the motive for the site-selectivity. Reverting back to the supposed pathway, upon the formation of the bicyclic intermediate IM5, it undergoes ligand/substrate replacement to afford intermediate IM6, at which the Ar–I coordinates to the Pd(ii) center to enable oxidative addition viaTS3 (ΔG = 27.4 kcal mol−1) leading to the five-coordinate Pd(iv) complex IM7. Undergoing direct C–C reductive elimination in IM7 would entail a barrier of 29.6 kcal mol−1 (TS4). Alternatively, iodine abstraction by the silver(i) salt in IM7 is thermodynamically favorable and irreversible, yielding the Pd(iv) intermediate IM8 coordinated to a TFA ligand. Subsequently, C–C reductive coupling viaTS5 generates the Pd(ii) complex IM9 and concludes the arylation process. This step was found both kinetically facile (6.1 kcal mol−1) and thermodynamically favorable (30.7 kcal mol−1). Finally, IM9 reacts with imine-1avia metathesis to regenerate the palladium catalyst IM1 and release imine-3a in a highly exergonic step (21.0 kcal mol−1). Ultimately, imine-3a undergoes hydrolysis to yield the aldehyde product 3a and to release the TDG.Open in a separate windowFig. 2Free energy profiles for the ligand-promoted Pd(ii)-catalyzed site-selective C–H activation and C–C bond formation, alongside the optimized structures of the C–H activation transition states TS1 and TS2 (selected bond distances are labelled in Å). Energies are relative to the complex IM1 and are mass-balanced.  相似文献   

12.
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   

13.
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   

14.
C(sp3)–C(sp3) coupling of non-activated alkyl-iodides with electron-deficient alkenes via visible-light/silane-mediated alkyl-radical formation     
Sanesh Mistry  Roopender Kumar  Andrew Lister  Matthew J. Gaunt 《Chemical science》2022,13(44):13241
Here, we present a remarkably mild and general initiation protocol for alkyl-radical generation from non-activated alkyl-iodides. An interaction between a silane and an alkyl iodide is excited by irradiation with visible light to trigger carbon–iodide bond homolysis and form the alkyl radical. We show how this method can be developed into an operationally simple and general Giese addition reaction that can tolerate a range of sensitive functionalities not normally explored in established approaches to this strategically important transformation. The new method requires no photocatalyst or other additives and uses only commerical tris(trimethylsilyl)silane and visible light to effectively combine a broad range of alkyl halides with activated alkenes to form C(sp3)–C(sp3) bonds embedded within complex frameworks.

Here, we present a remarkably mild and general initiation protocol for alkyl-radical generation from non-activated alkyl-iodides.

The efficient and straightforward construction of C(sp3)–C(sp3) bonds is a crucial process in organic synthesis. Over the past 80 years, the polar conjugate addition reaction has become a powerful method to forge a variety of C(sp3)–C(sp3) bonds.1 Alongside two-electron nucleophiles, alkyl-radicals – neutral yet nucleophilic species – have emerged as alternatives to organometallic reagents for additions to electron deficient alkenes.2 Since the 1960s, a variety of methods have been reported for the formation of alkyl-radicals; early examples include the decomposition of in situ generated organomercurial hydrides, the fragmentation of xanthate or Barton esters, or the UV-mediated homolysis of alkyl halides, amongst many others.3 Although these strategies tolerate a broad range of functionalities, the initiation processes can be complicated by the need for aggressive reaction conditions and frequently require toxic reagents such as tributyltin hydride, with notable exceptions.4,5The emergence of photoredox catalysis has obviated many of the potential drawbacks to the generation and use of alkyl-radicals. The exploitation of the multifaceted reactivity of visible light excited transition metal or organic-photocatalysts, whose properties can be tuned through modification of the ligand, metal and/or scaffold, facilitates optimization of the single electron transfer event towards alkyl-radical generation from a wide range of functionalized alkyl groups.6 In addition, the reactivity of electron donor–acceptor (EDA) complexes has also provided a straightforward means to form alkyl-radicals from a variety of precursors.7 As such, a plethora of methods have been developed for the generation of C(sp3)-centred radicals from a variety of commercially available native functionalities, which dramatically expand the scope of alkyl-radical chemistry. In this context, the single electron reduction of non-activated alkyl halides provides a useful means to generate alkyl radicals.8 As an example, Leonori and co-workers recently developed a method wherein halogen atom abstraction pathways were leveraged using radical species forged through photocatalyst-mediated oxidation event leading to a general alkyl-radical generation.9 Related to the current study, Jørgensen and co-workers published a visible-light mediated reduction of alkyl halides under very mild conditions. Accordingly, there remains a need for further innovation towards orthogonal, general and benign methods of alkyl-radical generation that tolerate a broad range of functionalities, thereby enabling the construction of a greater variety of C(sp3)–C(sp3) bonds.10Recently, we reported a general reaction to form tertiary alkylamines via the addition of alkyl-radicals (generated from non-activated alkyl-iodides) to in situ-generated all-alkyl iminium ions.11a This carbonyl alkylative amination (CAA) reaction was promoted by the action of blue LEDs and tris(trimethylsilyl)silane ((Me3Si)3Si–H). No photoredox catalyst is required. We believe that the alkyl-radical formation step, devoid of traditional initiating reagents, proceeds through the visible-light excitation of a transient ternary EDA complex, which stimulates homolysis of the carbon–iodide bond that would be otherwise stable under such irradiation conditions (Fig. 1B). The presence of an enamine was important to the initiation pathway, as revealed by an absorption band in the UV/vis spectrum of its mixture with an alkyl-iodide and (Me3Si)3Si–H.11a Gouverneur and co-workers have also reported an elegant example of visible-light mediated addition of more functionalized alkyl halides, such as iodofluoromethane, to electron deficient alkenes.12 They proposed that light mediated homolytic cleavage of iodofluoromethane was responsible for radical initiation prior to a classical chain process.Open in a separate windowFig. 1(A) Selected visible-light mediated methods for the generation of alkyl-radicals; (B) previous work – a method for tertiary amine formation exploiting a visible-light activation of a ternary EDA complex to promote alkyl-radical formation. (C) Previous work from Gouverneur & Gaunt labs on radical fluoromethylation. (D) This work – alkyl-radical formation promoted solely by visible light and tris-trimethylsilyl silane demonstrated through a remarkably practical and straightforward Giese reaction.Gouverneur et al. also showed methyl iodide was only efficient as a radical source under these conditions when an organic photocatalyst was present and the reaction of other simple non-activated alkyl iodides was only demonstrated in the presence of iodofluoromethane, which was presumably responsible for the initiation pathway (vide supra). Our prior work in this area also identified iodofluoromethane as a visible-light activated source of fluoromethyl radical and its addition to iminium ions and electron deficient alkenes (Fig. 1C).11b Taken together, these works reveal that the use of visible light and (Me3Si)3Si–H to initiate radical formation from non-activated alkyl halides has not been achieved in an unbiased transformation without the requirement of an initiation process via of the reaction components or a photocatalyst. Accordingly, we questioned whether a pathway mediated by visible-light and (Me3Si)3Si–H alone might facilitate alternative modes of radical initiation from non-activated alkyl halides, and therefore enable the general coupling of unbiased alkyl fragments with a wider range of acceptors under practical, straightforward reaction conditions.Herein, we report the successful realization of this idea through the development of a remarkably straightforward visible-light mediated method for alkyl-radical generation from non-activated alkyl iodides using only non-toxic tris(trimethylsilyl)silane as a reagent (Fig. 1D). While we are not certain of the precise pathway for the radical initiation, it seems likely that excitation of a species resulting from the interaction of tris(trimethylsilyl)silane and the alkyl iodide, leading to carbon–iodide bond homolysis. The utility of this activation mode is demonstrated through a broad and chemoselective Giese addition to electron deficient alkenes and is notable by its tolerance to a range of synthetically valuable functionalities in both alkyl iodide and alkene components. In comparison to other methods for Giese-addition,2,3,8,9,12 the conditions are mild and do not require expensive catalysts or cocktails of additives.Our studies were stimulated from an observation arising from the development of the visible light mediated carbonyl alkylative amination (shown in Fig. 1B). High yields of the tertiary amine product, arising from the union of alkyl-radical, aldehyde and secondary amine were maintained when using a 455 nm long-pass filter, which discounted UV-mediated carbon–iodide bond homolysis as the initiation pathway for alkyl-radical formation.11a To explore the formation of an alkyl-radical independently from the enamine component, the reaction conditions were simplified to comprise a representative alkyl halide and (Me3Si)3Si–H, which allowed us to first assess any impact solvent might have on the radical forming process. As shown in 13 However, 47% of 5 was still obtained after visible-light irradiation of a reaction mixture from which air had been rigorously excluded (entry 10), suggesting an alternative initiation pathway excluding oxygen could also operate.14 A reaction at 80 °C in the absence of light showed no conversion to 5. This data shows the nature of the solvent is not relevant for the initiation step and suggests a straightforward radical initiation process that results from visible-light excitation of an intermediate arising from an interaction between the alkyl halide and (Me3Si)3Si–H.Effect of different parameters on radical initiationa
EntrySolventDeviation in conditionsYield 5 (%)
1CH2Cl233
2THF68
3MeOH85
4EtOH55
5C6H1284
6PhH41
7PhMe34
8EtOH16 h86
9EtOH16 h, 455 nm filter82
10EtOH16 h, degassed47
11EtOH80 °C, dark0
Open in a separate windowaYields of 5 were calculated by 1H NMR using 1,1,2,2-tetrachloroethane as internal standard.With the operationally simple and mild reaction conditions for the homolysis of non-activated alkyl halides, we next focussed on benchmarking the process against existing transformations: namely the Giese addition reaction of alkyl-radicals to electron deficient alkenes. Therefore, using acrylamide 2a (as a representative alkene acceptor), 3.0 equivalents of iso-propyl iodide 1a (as a representative non-activated alkyl halide) and 1.5 equivalents of (Me3Si)3Si–H in MeOH at 0.1 M, we were pleased to find visible light irradiation of this reaction mixture led to the formation of alkylamide 3a in 59% assay yield ( EntrySolvent(Me3Si)3Si–HAlkyl-iodideConc.Yield 3aa (%)1MeOH1.5 equiv.3.0 equiv.0.1 M592MeOH2.0 equiv.3.0 equiv.0.2 M663EtOH2.0 equiv.3.0 equiv.0.2 M794EtOH2.0 equiv.2.0 equiv.0.2 M775EtOH2.0 equiv.1.5 equiv.0.2 M706EtOH1.5 equiv.1.5 equiv.0.2 M47Open in a separate windowaYields of 3a were calculated by 1H NMR using 1,1,2,2-tetrachloroethane as internal standard.We next turned attention to evaluating the scope of the reaction ( Open in a separate windowNext, the scope of the reaction in the alkyl halide component was investigated. Carbon-centred radicals formed through the visible-light/TTMS initiation pathway from simple tertiary alkyl-iodides (1-adamantyl to 3t and tert-butyl to 3u) underwent Giese addition with acrylamide 2a in excellent yields. A series of alkyl-iodides based on heterocyclic scaffolds were found to undergo smooth radical formation and Giese addition, providing products that can be further elaborated and could be utilized in the construction of pharmaceutically relevant molecules (3v–3ab). Finally, a selection of primary alkyl-iodides (3ac–ae) were found to be compatible with the radical activation mode and generated the linear alkyl products in good yields.We had recognized a report by the Merck discovery group who showed that a visible-light mediated Ir-catalyzed Giese addition using alkyl bromides also utilized (Me3Si)3Si–H as a reagent to propagate alkyl-radical formation (Fig. 2A).8b A distinct difference of our system is the departure from any requirement for a photocatalyst, which not only reduces the cost of the transformation but could offer an alternative protocol for Giese addition when redox sensitive groups appear in one of the coupling partners.Open in a separate windowFig. 2(A) Photocatalytic Giese addition with alkyl bromides mediated by (Me3Si)3Si–H; (B) catalyst-free visible-light/(Me3Si)3Si–H mediated Giese addition.Several simple mechanistic experiments were conducted to supplement those shown in Fig. 3A), confirming the intermediacy of an alkyl radical. This result suggests a pathway involving β-scission of the cyclopropylmethyl radical (to int-I) prior to addition to the alkene (to int-II), which is followed by 5-exo trig cyclization (to int-III) and HAT of the resulting methyl radical with (Me3Si)3Si–H to form cyclopentane 7. A reaction conducted using d5-EtOD showed no incorporation of deuterium in the product, thereby eliminating solvent participation in the radical interception step (Fig. 3B), as expected from the observations outlined in Open in a separate windowFig. 3(A) Evidence for alkyl-radical formation via radical trap experiment; (B) experiments to demonstrate HAT is not from solvent and the essential role of the silane.Despite the strength of our experimental observations, conclusive evidence that unravels the visible-light mediated radical initiation pathway between (Me3Si)3Si–H and the non-activated alkyl-iodide remains elusive. 1H NMR, UV-vis or IR spectroscopy titration studies aimed at identifying an interaction between a variety of alkyl-halides and (Me3Si)3Si–H could not be obtained (see ESI for details). However, visible light must be exciting a transient intermediate comprising the silane and alkyl-iodide as the predominant pathway here because the process so efficiently homolyzes the carbon–iodide bond15 to yield alkyl-radicals. Consequently, we hypothesized that such an interaction may not be observable due to its transient nature and short lifetime or would be present in minute quantities beyond the detection limits of these spectroscopic methods. One possibility for the radical initiation is the excitation of a halogen bonded intermediate between the alkyl iodide and silane – whereby a coulombic attraction between a region of electron deficiency around the polarizable halogen atom (σ-hole)16,17 and the electron rich hydridic component of the silane – could result in weakening of the carbon–iodine bond, subsequently absorbing low energy visible light to stimulate homolysis (Fig. 1D). We do, however, acknowledge that an as yet undetermined pathway could be responsible for the radical initiation pathway. Despite the uncertainty over the mode of radical initiation, our understanding of the mechanism responsible for this catalyst-free Giese addition is detailed in Fig. 4. Visible-light and silane-mediated carbon–iodide bond homolysis, generates an alkyl-radical to initiate the process. Addition of the alkyl-radical to the alkene acceptor generates a new electrophilic radical, which undergoes HAT with (Me3Si)3Si–H to form the product. The resulting (Me3Si)3Si radical now undergoes halogen atom transfer (XAT) with a new molecule of the alkyl-iodide, thereby propagating the radical chain.Open in a separate windowFig. 4Current proposed mechanism for metal-free Giese addition.In summary, we have developed an operationally straightforward method for alkyl-radical generation from non-activated alkyl-iodides and applied this protocol through the development of mild Giese addition. Through the action of visible light and (Me3Si)3Si–H, radical initiation from an alkyl-iodide is achieved under mild reaction conditions. No photocatalysts are required for this process. The range of alkyl-iodides and alkenes is very broad, and the reaction tolerates many sensitive functional groups. Considering current demand for saturated scaffolds in the drug discovery programs, we believe this strategy offers a potentially powerful method through which to combine two readily available classes of building blocks into complex molecules of biological and pharmaceutical interest. While the pathway of radical initation remains unclear, it is clear that the activation mode afforded by the combination of (Me3Si)3Si–H, alkyl iodide and visible-light provides a mild and general means for forming open shell alkyl intermediates.11,12 Subsequently, it is likely that this method will be useful to practitioners of synthetic chemistry in both academic and industrial settings.  相似文献   

15.
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   

16.
Direct electrochemical hydrodefluorination of trifluoromethylketones enabled by non-protic conditions     
John R. Box  Alexander P. Atkins  Alastair J. J. Lennox 《Chemical science》2021,12(30):10252
CF2H groups are unique due to the combination of their lipophilic and hydrogen bonding properties. The strength of H-bonding is determined by the group to which it is appended. Several functional groups have been explored in this context including O, S, SO and SO2 to tune the intermolecular interaction. Difluoromethyl ketones are under-studied in this context, without a broadly accessible method for their preparation. Herein, we describe the development of an electrochemical hydrodefluorination of readily accessible trifluoromethylketones. The single-step reaction at deeply reductive potentials is uniquely amenable to challenging electron-rich substrates and reductively sensitive functionality. Key to this success is the use of non-protic conditions enabled by an ammonium salt that serves as a reductively stable, masked proton source. Analysis of their H-bonding has revealed difluoromethyl ketones to be potentially highly useful dual H-bond donor/acceptor moieties.

The electrochemical hydrodefluorination of trifluoromethylketones under non-protic conditions make this single-step reaction at deeply reductive potentials uniquely amenable to challenging electron-rich substrates and reductively sensitive functionalities.

The difluoromethyl group (CF2H) has attracted significant recent attention in medicinal chemistry,1,2 which complements the well-documented importance and growing use of fluorine in small molecule pharmaceuticals.3–6 The CF2H group is an H-bond donor7,8 that is also lipophilic,9,10 a unique combination that positions it as an increasingly valuable tool within drug-discovery.11 CF2H has been used as a bioisostere of OH and SH in serine and cystine moieties, respectively, as well as NH2 groups, where greater lipophilicity and rigidity provide advantages to pharmacokinetics and potency.12–14The hydrogen-bond acidity of CF2H groups is exceptionally dependent on the atom or group to which it is appended (Fig. 1A).1,2 The H-bond acidity of alkyl-CF2H groups is half that of O–CF2H and even a quarter of SO2–CF2H groups.1 This mode of control allows the H-bonding strength and, therefore its function, to be finely tuned. While much research has focused on the synthesis, behaviour and use of XCF2H groups, where X = O, S, SO, SO2, Ar, it is surprising that the corresponding carbonyl containing moiety (X = CO) has remained relatively elusive in these contexts. Not only would difluoromethyl ketones (DFMK) be expected to provide a relatively strong H-bond, but the carbonyl unit provides a complementary, yet proximal mode of intermolecular interaction (Fig. 1B). Indeed, the dual action of neighbouring H-bond donor and acceptor functionalities provides the fundamental basis for many biological systems, including in the secondary structure assembly mechanisms for proteins and DNA/RNA nucleobase pairing, as well as in enzyme/substrate complexes. Indeed, the DFMK functionality has demonstrated important utility in biological applications, including anti-malarial and -coronaviral properties.15 Finally, the carbonyl provides a useful synthetic handle for further derivatization.Open in a separate windowFig. 1H-Bonding in DFMKs and their synthesis via hydrodefluorination.While some progress has been made on the synthesis of DFMKs,16 there still remains a need for a general and more broadly accessible route to their preparation. Current strategies for DFMK preparation require multi-step processes, expensive reagents, installation of activating groups, or are inherently low yielding.15a,16–25 The hydrodefluorination of trifluoromethyl ketones (1) potentially represents the most accessible strategy, as the starting materials are most readily prepared through a high-yielding trifluoroacetylation of C–H or C–X bonds.26–29 In 2001, Prakash demonstrated the viability of this approach using 2 equivalents of magnesium metal as stoichiometric reductant to drive the defluorination, with a second hydrolysis step (HCl (3–5 M) or fluoride, overnight stirring) to reveal the product.30 The scope in this 2-step process (6 substrates) reflects the limitations of using a reductant, such as Mg, that has a fixed reduction potential, as well as incompatibilities arising from Mg/halide exchange with aryl halides. Similar limitations with the use of electron-rich substrates were revealed in related contributions from Uneyama.31In order to access more electron-rich and reductively challenging substrates, such as those containing medicinally relevant heterocycles, we postulated that electrochemical reduction could be employed (Fig. 1C). Electrosynthesis is becoming an increasingly valuable enabling technology and has seen a recent resurgence due to the precise control, unique selectivity, and the potential scalability and sustainability benefits that it offers.32–36 This strategy would avoid the undesirable use of stoichiometric metals and the ‘deep-reduction’ potentials required are readily accessed by simply selecting the applied potential. Pioneering early work from Uneyama on the cathodic formation of silylenol ether intermediate 2, suggested this approach could be viable.37,38 The fundamental challenge in designing a practical, single-step process under highly reducing potentials (<−2.0 V vs. Fc/Fc+), is to avoid the reduction of the proton source, which would otherwise compete to generate H2 gas and leave the starting material untouched. Uneyama does not demonstrate hydrodefluorination, presumably due to this problem. Additional challenges posed by ‘deep-reduction’ include a lack of tolerance for reduction-sensitive functionality (alkene, C–X bonds etc.), low mass balance due to substrate decomposition and the undesirable use of sacrificial metal anodes.39 Solving these problems should provide generally applicable, safe and scalable conditions for the hydrodefluorination of readily accessible trifluoromethyl ketones (1).Given the electron-rich nature of indoles, their ubiquity in bioactive compounds, and their ease of functionalisation, we chose indole 1a as the model substrate for optimisation. The highly reductive potentials required will render it a challenging substrate, which should lead to more general conditions suitable for other important substrate classes. Indeed, when we applied the Mg conditions of Prakash to this substrate, no silyl enol ether intermediate (2a) was observed, nor product 3a, and the starting material remained completely untouched ( EntryConditions different from aboveReductantProton source 1a a/%(2a) 3aa/%1 Mg 0, THF, no electricity (Prakash conditions for3)Mg0—100(0) n/a2bUndivided cell, TBAPF6Sacrificial Mg anode—100(0) n/a3bPb:C (cath:an), 0 oC, 30 mA (Uneyama conditions for2)TBABr (4 eq.)—33(32) 04b—TBABr (2 eq.)(a) Acetic acid; (b) oxalic acid.51; 1000; 05b—TBABr (2 eq.)Dimethylurea8206b—TBABr (2 eq.)TEAPF6 (4 eq.)49457TMSCl (0 eq.)TBABr (2 eq.)TEAPF6 (4 eq.)8308bTMSCl (6 eq.)TBABr (2 eq.)TEAPF6 (4 eq.)49499c TMSCl (3 + 3 eq.) TBABr (2 eq.) TEAPF 6 (4 eq.) 0 97 10cEntry 9, but Pt:Gr (cath:An)TBABr (2 eq.)TEAPF6 (4 eq.)09411cEntry 9, but Ni:Pt (cath:An)TBABr (2 eq.)TEAPF6 (4 eq.)08312cEntry 9, but Stainless Steel:Pt (cath:An)TBABr (2 eq.)TEAPF6 (4 eq.)08513cEntry 9, but Gr:Pt (cath:An)TBABr (2 eq.)TEAPF6 (4 eq.)018Open in a separate windowa 19F NMR yields.bTMSCl only added to cathodic chamber.cTMSCl added to both cathodic and anodic chambers.The electrochemical conditions of Uneyama for preparing silylenol ethers (2) were applied to our indole 1a (entry 3). Unsurprisingly, no hydrodefluorinated product was observed, however intermediate 2a was formed in a 32% yield. In an effort to improve this yield we explored several solvents, reductants, additives and electrode materials, all of which were conducted in a divided cell at constant current and ambient temperature.40 In addition, as we were keen to develop a single-step protocol, by avoiding the second hydrolysis step that can readily form homo-coupled aldol side products,38 we surveyed a range of added proton sources for in situ delivery of 3a. The addition of carboxylic acids, such as acetic or oxalic acid (entry 4), gave no desired product, as the competing reduction of protons to H2 gas dominated. Dimethylurea was recently used as a proton source in an electrochemical ‘deep-reduction’,41 but it returned no trace of intermediate 2a or product 3a (entry 5). We hypothesized that increasing the conductivity of the system, with additional tetraalkylammonium salts (from 2 to 4 eq.), the formation of intermediate 2a may be facilitated by avoiding large cell potentials. While this change did facilitate a lower cell potential, we discovered these salts behaved as reductively stable yet competent masked proton donors: 4 eq. NEt4PF6 gave 45% yield of product 3a, with no sign of intermediate 2a (entry 6). The detection of triethylamine in solution suggests donation through a Hoffmann elimination.42 With the exception of NMe4+, other tetraalkylammonium salts were also competent proton donors (NEt4+ > NBu4+ > NPr4+).A critical improvement to the yield was observed when the use of the radical anion trapping agent, TMSCl, was optimised. With no TMSCl, 3a was not observed (entry 7), and a loading of 6 equivalents saw little improvement over 3 equivalents (entry 8 vs. 6). Experiments hitherto described were conducted with TMSCl added only to the cathodic chamber (entries 2–8). Only when the 6 equivalents was split between both chambers was a drastic improvement observed (entry 9), giving an optimised yield of 97%. Notably, the increase in conversion still occurred with only 2 F, implying that a lower steady-state concentration may be important in the cathode chamber. To test this hypothesis, TMSCl was slowly added to the catholyte by syringe-pump addition over the course of the reaction, which gave a similar yield of 94%.40 Although intermediate 2a is transient and was never observed, the importance of TMSCl to trap and stabilise reduced 1a was revealed by DFT (B3LYP/6-311+g(d)) calculations,40 which suggested a thermodynamically highly challenging reaction in its absence.The oxidation of bromide to tribromide occurs on the anode, which is an ideal counter-electrode process: not only is bromide an inexpensive and metal-free sacrificial reductant, but as the produced Br3 is anionic, it does not rapidly migrate to the cathodic chamber, preventing unwanted side reactions.43 The generated Br3 can even be used in follow-up bromination reactions.44 An increase in the applied cell potential during the reaction signifies the consumption of Br, and the oxidation of Br3 to Br2 (Fig. 2).45 Despite needing 3 equivalents of Br to form 2 equivalents of Br3 after 2 F, the loading of Br could be reduced to 2 equivalents without affecting yield. No over-reduction of 3a to the monofluoromethyl ketone was observed, which is significant considering the small difference in reduction potentials.40 This emphasises the importance of a flat chronopotentiometry trace that is achieved with Br oxidation. Other reductants were found to be sub-optimal, including diisopropylamine and oxalic acid.40Open in a separate windowFig. 2Reaction of 1a to 3a with 3 different Br concentrations.A graphite anode performed equally well as platinum for the counter electrode reaction (entry 10). Only marginally reduced yields were observed with nickel and stainless-steel cathodes (entries 11 and 12), however, a drastic decrease in the yield was observed with a graphite cathode (entry 13), possibly due to substrate grafting.39We proceeded to explore the substrate scope with our optimized conditions, Fig. 3. As expected, our electrochemical conditions were suitable for the hydrodefluorination of electron-poor acetophenone derivatives (1b, 1c). However, unlike with the use of Mg,30 substrates containing electron donating substituents are now well tolerated (1d–k). In addition, no hydrodebromination was observed for 1b, highlighting the selectivity and orthogonality granted by the use of our Mg-free, non-protic conditions. A selection of extended π-systems was tolerated, producing pyridyl 3l, biphenyl 3m, benzothiophene 3n, primary amine 3o, and pyrimidines 3p and 3q and in moderate to excellent yields. Chromoionophore dye 1r and stilbene 1s and were transformed in excellent yield, demonstrating tolerance to reductively sensitive alkenes, which would otherwise hydrogenate under protic electrochemical conditions.46 Anthracenyl 1t and naphthyl substrates 1u and 1v all transformed efficiently in good to excellent yields, the latter of which underwent direct double hydrodefluorination. 4.5% over-reduction was observed in the double defluorination product, 3v, which was the only instance where this side-product was observed in greater than 1% quantities.40 The good mass-balance and faradaic efficiency is notable considering the delocalization of charge around extended π-systems increases the likelihood of grafting.47Open in a separate windowFig. 3Isolated yields of DFMKs tested under the reaction conditions at 0.5 mmol scale. NMR yields in parentheses. aReaction run at 10 mA; breaction run in IKA Divided ProSyn: quantitative yield based on RSM; c5 mmol scale, Ni foil:Gr (cath:an); disolated as the corresponding ketone following purification on silica.49The model indole substrate 1a gave an excellent yield of DFMK at 0.5 mmol scale, which gave equally high yields when scaled up 10-fold (5 mmol), thereby demonstrating the robustness and practicality of the technique. We were also able to successfully prepare 3a in a commercially available divided cell set-up.40 Alternative groups on nitrogen, including Boc, perfluoropyridyl and benzyl (3w–y), as well as the free indole 3z, were well tolerated and gave moderate to good yields of 3. Tosyl and acetyl groups on nitrogen were less well tolerated.40 As with the acetophenones, indoles with electron donating (1aa) and withdrawing (1ab) groups proceeded to product. Methoxy demethylation of 3aa should lead to the corresponding phenol,48 which is difficult to prepare using other methodologies due to competing side-reactions. Halide substitution also successfully yielded DFMKs (3ac–ag). The inclusion of the aryl-iodide functionality is especially notable due to its facile reduction; when a silver cathode was used to convert 1ag, hydrodeiodination was observed, but which was absent under our non-protic conditions with a Pt cathode. Increased steric bulk around the reacting center in thiophenyl and phenyl-substituted substrates 1ah and 1ai had no negative influence and gave good yields of product.Heterocyclic trifluoromethylketones were successfully hydrodefluorinated under the standard conditions, including indole 3aj, carbazole 3ak, pyrrole 3al, pyridine 3am, and pyrazoles 3an and 3ao, the latter of which leads to a compound with anti-malarial activity.15a Alkyl trifluoromethylketones are more difficult to reduce compared to aromatic trifluoromethylketones, and are therefore challenging substrates to hydrodefluorinate, and impossible to convert using other methods. Nevertheless, oleyl 1ap, cyclohexyl 1aq and ethyl 1ar substrates were all amenable to the conditions, although the smaller alkyl products were cumbersome to isolate due to their volatility. The non-protic optimized conditions ensured no loss of mass-balance at these enhanced reduction potentials (|Ecell| = ca. 3.4–3.7 V for alkyl substrates vs. ca. 2.3–2.7 V for acetophenones and indoles). Finally, we tested the conditions on trifluoroacetamide 1as, thioester 1at and imines 1au and 1av. For each of these, the corresponding product was returned in moderate to good yields. Despite some complications in their isolation, these results are notable considering their difference in structure and lack of precedent. Unsuccessful substrates included a nitro-substituted indole, which was insoluble in the reaction medium, and hydrated TFMKs.40We tested a variety of substrates with the Mg-mediated conditions reported by Prakash to gauge the level of complementary between the methods.30 While acetophenone derivatives 1k and 1am were amenable to reduction with Mg, bromide substitution in 1b was unsurprisingly not tolerated with Grignard formation dominating. Indoles – 1a, 1ai, pyrazole – 1an, alkyl – 1aq, 1ar and anilide – 1as based trifluoromethylketones were untouched by Mg in all cases, with starting materials recovered only.To explore the value of the DFMK moiety in synthesis, we derivatized it in a variety of ways, Fig. 4. Resubjecting the product 3a to our non-protic hydrodefluorination conditions led to monofluorinated product 4, providing an alternative to the use of electrophilic fluorine sources.50 Reduction of the ketone in 3ae to the methyl ether and alcohol successfully gave products, 5 and 6, respectively. The dithiane of 3a, which is a useful synthetic intermediate, was formed in excellent yield (7). A Corey–Chaykovsky methenylation gave epoxide 8 in good yield. A Horner–Wadsworth–Emmons reaction transformed the carbonyl to give alkene 9. Nucleophilic attack of the ketone was demonstrated with a trifluoromethylation reaction to give highly fluorinated alcohol 10. Orthogonal reactivity was also demonstrated with a Suzuki–Miyaura cross-coupling that gave biaryl 11. Interestingly, deuterium was not exchanged into 3a when stirred in a mixture of D2O and MeCN, providing evidence for a less favourable enolization.Open in a separate windowFig. 4[A] Derivatization of DFMKs. X = H (3a) for 4, 7, and 8, X = Br (3ae) for others; [B] H-bond strength (A-value) correlated to σm Hammett parameter; [C] intermolecular H-bond revealed in X-ray crystal structure of 3ae; [D] DFT calculated (B3LYP/6-311+g(d)) relative energies of conformers with rotation around HC–CO bond. Brown arrows indicate direction of dipole.The H-bond strength (A-value) was measured for a series of phenyl substituted X–CF2H derivatives using the NMR method from Abraham, Fig. 4B.51–53 These experiments confirmed the sensitivity of the H-bonding ability to the identity of X. DFMK 3g and sulfoxide–CF2H were found to be comparable H-bond donors, which were only marginally less than the sulfone–CF2H. The H-bond strength correlated best with the σm parameter, reflecting the strong influence of inductive effects. Multiple regression analysis showed that any contribution of σp was statistically insignificant (P value = 0.33).Analysis of the X-ray crystal structure of 3ae, showed an inter-molecular H-bond between the CF2H and a carbonyl from a neighbouring molecule (Fig. 4C). DFT was used to calculate the relative conformer energy with rotation about the (O)C–CF2H dihedral bond (Fig. 4D). The lowest energy conformer eclipsed the H with the carbonyl, implying the possibility of an energy lowering intra-molecular H-bond. However, analysis of the other derivatives in the set (C(O)CH3, C(O)CFH2 and C(O)CF3) revealed that the alignment of dipoles was the dominant effect (brown arrows, Fig. 4D).40 The absence of an unusually low or even negative A-value also provides evidence against an intramolecular H-bond.51 Interestingly, in the solid-state structure (Fig. 4C), the highest energy conformer (with dipoles aligned) is adopted, highlighting the stronger propensity of this moiety to engage in H-bonding interactions.In conclusion, we have developed a mono-selective hydrodefluorination to access a broad scope of DFMKs, enabled by non-protic electrochemical conditions at deeply reducing potentials. These moieties have been studied and diversified and reveal themselves to be potentially useful dual H-bond donor/acceptor moieties. This is especially interesting considering the structurally related trifluoromethylketones are known reversible protease inhibitors;54,55 thus, the additional H-bonding moiety could enhance interaction within enzymatic active sites.15  相似文献   

17.
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   

18.
Ruthenium-catalyzed formal sp3 C–H activation of allylsilanes/esters with olefins: efficient access to functionalized 1,3-dienes     
Dattatraya H. Dethe  Nagabhushana C. Beeralingappa  Saikat Das  Appasaheb K. Nirpal 《Chemical science》2021,12(12):4367
Ru-catalysed oxidative coupling of allylsilanes and allyl esters with activated olefins has been developed via isomerization followed by C(allyl)–H activation providing efficient access to stereodefined 1,3-dienes in excellent yields. Mild reaction conditions, less expensive catalysts, and excellent regio- and diastereoselectivity ensure universality of the reaction. In addition, the unique power of this reaction was illustrated by performing the Diels–Alder reaction, and enantioselective synthesis of highly functionalized cyclohexenone and piperidine and finally synthetic utility was further demonstrated by the efficient synthesis of norpyrenophorin, an antifungal agent.

Ru-catalysed oxidative coupling of allylsilanes and allyl esters with activated olefins has been developed via isomerization followed by C(allyl)–H activation providing efficient access to stereodefined 1,3-dienes in excellent yields.

1,3-Dienes not only are widespread structural motifs in biologically pertinent molecules but also feature as a foundation for a broad range of chemical transformations.1–14 Indeed, these conjugated dienes serve as substrates in many fundamental synthetic methodologies such as cycloaddition, metathesis, ene reactions, oxidoreduction, or reductive aldolization. It is well-understood that the geometry of olefins often influences the stereochemical outcome and the reactivity of reactions involving 1,3-dienes.15 Hence, a plethora of synthetic methods have been developed for the stereoselective construction of substituted 1,3-dienes.16–24 The past decade has witnessed a huge advancement in the field of metal-catalyzed C–H activation/functionalization.25–27 Although, a significant amount of work in the field of C(alkyl)–H and C(aryl)–H activation has been reported; C(alkenyl)–H activation has not been explored conspicuously, probably due to the complications caused by competitive reactivity of the alkene moiety, which can make chemoselectivity a significant challenge. Over the past few years, several different palladium-based protocols have been developed for C(alkenyl)–H functionalization, but the reactions are generally limited to employing conjugated alkenes, such as styrenes,28–31 acrylates/acrylamides,32–36 enamides,37 and enol esters/ethers.38,39 To date, only a few reports have appeared in the literature for expanding this reactivity towards non-conjugated olefins, which can be exemplified by camphene dimerization,40 and carboxylate-directed C(alkenyl)–H alkenylation of 1,4-cyclohexadienes.41 In 2009, Trost et al. reported a ruthenium-catalyzed stereoselective alkene–alkyne coupling method for the synthesis of 1,3-dienes.42 The same group also reported alkene–alkyne coupling for the stereoselective synthesis of trisubstituted ene carbamates.43 A palladium catalyzed chelation control method for the synthesis of dienes via alkenyl sp2 C–H bond functionalization was described by Loh et al.44 Recently, Engle and coworkers reported an elegant approach for synthesis of highly substituted 1,3-dienes from two different alkenes using an 8-aminoquinoline directed, palladium(ii)-mediated C(alkenyl)–H activation strategy.45 Allyl and vinyl silanes are known as indispensable nucleophiles in synthetic chemistry.46 Alder ene reactions of allyl silanes with alkynes are reported for the synthesis of 1,4-dienes.47 Innumerable methods are known for the preparation of both allyl and vinyl silanes48–52 but limitations are associated with many of the current protocols, which impedes the synthesis of unsaturated organosilanes in an efficient manner. Silicon-functionalized building blocks are used as coupling partners in the Hiyama reaction53 and are easily converted into iodo-functionalized derivatives (precursor for the Suzuki cross-coupling reaction), but there is little attention given for the synthesis of functionalized vinyl silanes. Herein, we report a general approach for the stereoselective synthesis of trisubstituted 1,3-dienes by the Ru-catalyzed C(sp3)–H functionalization reaction of allylsilanes (Scheme 1).Open in a separate windowScheme 1Highly stereoselective construction of 1,3-dienes.In 1993, Trost and coworkers reported an elegant method for highly chemoselective ruthenium-catalyzed redox isomerization of allyl alcohols without affecting the primary and secondary alcohols and isolated double bonds.54,55 Inspired by the potential of ruthenium for such isomerization of double bonds in allyl alcohols, we sought to identify a ruthenium-based catalytic system that can promote isomerization of olefins in allylsilanes followed by in situ oxidative coupling with an activated olefin to form substituted 1,3-dienes. We initiated our studies by choosing trimethylallylsilane 1a and acrylate 2a by using a commercially available [RuCl2(p-cymene)]2 catalyst in the presence of AgSbF6 as an additive and co-oxidant Cu(OAc)2 in 1,2-DCE at 100 °C. Interestingly, it resulted into direct formation of (2E,4Z)-1,3-diene 3aa as a single isomer in 55% yield. It is likely that this reaction occurs by C(allyl)–H activation of the π-allyl ruthenium complex followed by oxidative coupling with the acrylate and leaving the silyl group intact (Table 1). π-Allyl ruthenium complex formation may be highly favorable due to the α-silyl effect which stabilizes the carbanion forming in situ in the reaction.56 Next, the regioselective C–H insertion of vinyl silanes could be controlled by stabilization of the carbon–metal (C–M) bond in the α-position to silicon. This stability arises due to the overlapping of the filled carbon–metal orbital with the d orbitals on silicon or the antibonding orbitals of the methyl–silicon (Me–Si) bond.57 The stereochemistry of the diene was established by 1D and 2D spectroscopic analysis of the compound 3aa. To quantify the C–H activation mediated coupling efficiency, an extensive optimization study was conducted (allylsilanes followed by in situ oxidative coupling with an activated olefin to form substituted 1,3-dienes). The change of solvents from 1,2-DCE to t-AmOH, DMF, dioxane, THF or MeCN did not give any satisfactory result, rather a very sluggish reaction rate or decomposition of starting materials was observed in each case (entry 2–6).Optimization of reaction conditionsa
EntryAdditive (20 mol%)Oxidant (2 equiv.)SolventYieldb (%)
1AgSbF6Cu(OAc)2DCE55
2AgSbF6Cu(OAc)2t-AmOH10
3AgSbF6Cu(OAc)2DMF0
4AgSbF6Cu(OAc)2Dioxane8
5AgSbF6Cu(OAc)2THF21
6AgSbF6Cu(OAc)2MeCN0
7cAgSbF6Cu(OAc)2DCE35
8dAgSbF6Cu(OAc)2DCE82
9eAgSbF6Cu(OAc)2DCE45
10dAg2CO3Cu(OAc)2DCE0
11dAgOAcCu(OAc)2DCE20
12dAgSbF6DCE0
Open in a separate windowaReaction conditions: 1a (0.24 mmol), 2a (0.2 mmol), [Ru(p-cymene)Cl2]2 (5 mol%), additive (20 mol%) and oxidant (2 equiv.) at 100 °C in a specific solvent (2.0 mL), under argon, for 16 h.bIsolated yields are of product 3aa.cThe reaction was performed at 120 °C.dThe reaction was performed at 80 °C.eThe reaction was performed at 60 °C. t-AmOH – tertiary amyl alcohol, DMF – N,N-dimethylformamide, DCE – 1,2-dichloroethane.The increase of temperature from 100 °C to 120 °C resulted in the formation of diene in lower yield (entry 7). To our delight, it was found that a substantial enhancement in the yield (82%) was observed when the reaction was performed at 80 °C (entry 8). In particular, this was found to be the best reaction condition since further lowering of the temperature led to noteworthy attenuation of the reaction rate and yield (entry 9). Interestingly, the reaction was not efficient, when AgSbF6 was replaced with other additives, such as Ag2CO3 and AgOAc. It was also observed that, co-oxidant Cu(OAc)2 is necessary for the success of this reaction (entry 12).With these optimized conditions in hand, various allyl sources and acrylates have been tested (Table 2). It was found that a variety of acrylates 2 bearing alkyl and sterically crowded cyclic substituents successfully underwent the coupling reaction with allyl silane 1a to afford corresponding silyl substituted (2E,4Z)-1,3-dienes in good yields (3aa–3af). Similarly, dimethyl benzylallylsilane 1b reacted smoothly with acrylates such as methyl, isobutyl and n-butyl to generate desired dienes 3ba, 3bb and 3bc in 83%, 85% and 82% yield respectively. Interestingly, sterically crowded, tert-butyldimethyl allylsilane 1c showed its reactivity towards the coupling reaction with n-butyl acrylate to provide required diene 3cb in 80% yield. It is worth mentioning that allylsilanes 1a and 1b also exhibited their coupling reactivity with phenyl vinyl sulfone and successfully generated corresponding 1,3-dienes 3ag and 3bg in 78% and 76% yield respectively. When tert-butyldiphenylallylsilane 1d was subjected to the coupling reaction with methyl acrylate 2a, end–end coupling product 3da was isolated in 68% yield. This may be attributed to the steric crowding offered by bulky groups on silicon which prevents allyl to vinyl isomerization.Substrate scope for oxidative coupling of allylsilanes with acrylates and vinyl sulfonesa
Open in a separate windowaReaction conditions: 1 (0.24 mmol), 2 (0.2 mmol), [Ru(p-cymene)Cl2]2 (5 mol%), AgSbF6 (20 mol%) and Cu(OAc)2·H2O (2 equiv.) at 80 °C in 1,2-dichloroethane (2.0 mL), under argon, 16 h.bIsolated yields are of product 3. TMS – trimethylsilyl, TBDMS – tertiarybutyldimethyl silyl.To extend the substrate scope of the reaction, we next examined the scope of allylesters by employing 2a as the coupling partner. First, we carried out the coupling reaction between allyl ester derivative 4a and methyl acrylate 2a under standard conditions. To our delight, a single isomer of acetate substituted (2E,4Z)-1,3-diene 5aa was isolated with a good yield (75%) (Table 3). This result may be extremely unusual due to the weak thermodynamic driving force for the double bond migration of allyl esters and tendency of many metal catalysts to insert themselves into the C(allyl)–O bond to form a stable carboxylate complex.58 Even for unsubstituted allyl esters very few reports of double bond migrations exist.59–62 It is worth mentioning that unlike the Tsuji–Trost reaction,63–65 the C(allyl)–O bond doesn''t break to form the π-allyl palladium complex as an electrophile, instead it forms a nucleophilic π-allylruthenium complex (umpolung reactivity) keeping the acetate group intact, which further reacts with an electrophile. The stereochemistry of the diene was established by 1D and 2D spectroscopic analysis of the compound 5ga and also by comparison of spectroscopic data with those of an authentic compound.66 Next we turned our attention to expand the scope of the coupling reaction between various acrylates and allyl esters. It was found that a variety of allyl esters bearing alkyl substituents on the carbonyl carbon could provide moderate to good yields of the corresponding stereodefined (2E,4Z)-1,3,4-trisubstituted 1,3-dienes successfully. As can be seen from Table 2, alkyl substituents (4b–4d) had little influence on the yields (65–75%). Gratifyingly, we noticed that the presence of a bulky substituent in 4 also showed its viability towards the coupling reaction, albeit with modest yields (5ea & 5fa). Also, various acrylate derivatives reacted smoothly to generate the 1,3-dienes in excellent yield. A simple allyl acetate 4g reacted with a series of different acrylates 2 to afford the desired products in good yields.Substrate scope for oxidative coupling of various allyl esters with different acrylates and vinyl sulfonesa
Open in a separate windowaReaction conditions: 4 (0.24 mmol), 2 (0.2 mmol), [Ru(p-cymene)Cl2]2 (5 mol%), AgSbF6 (20 mol%) and Cu(OAc)2·H2O (2 equiv.) at 80 °C in 1,2-dichloroethane (2.0 mL), under argon, 16 h.bIsolated yields are of product 5.Several acrylates such as methyl-, ethyl-, n-butyl-, isobutyl-, n-heptyl-, cyclohexylmethyl-, benzyl-, etc. were tested and good to very good yields of the products were obtained. Also, gram scale synthesis of 5gh (1.35 g) by the reaction of acetate 4g with 2h gave identical results in terms of yield (69%) and diastereoselectivity, indicating the robustness and practicality of this method. Markedly, a C2-symmetric diacrylate (2e) also reacted with allyl acetate to form a mono-coupled product 5ge, though in a somewhat lower yield. In contrast to the allyl esters, the coupling was not affected by the steric bulk of the acrylate substituents as depicted in Table 3. Even the borneol derivative 2j and menthol derivative 2l, which can offer considerable steric hindrance, were found to be equally effective in the formation of 5gj and 5gl in very good yields. A somewhat reduced yield of the product 5gm was observed while using phenyl acrylate (2m) perhaps due to competitive reactive sites. Interestingly, the versatility of this methodology was not restricted only to acrylates, since phenyl vinyl sulfone was also found to be equally efficient for oxidative C–H functionalization with different allyl esters and a successful C–C coupling reaction was observed in each case with moderate yield and excellent diastereoselectivity.Interestingly treatment of allylsilanes under standard reaction conditions in the absence of an acrylate coupling partner led to isomerization of various allylsilanes to afford corresponding vinylsilanes 6b–6e in excellent yields (Scheme 2a). When allylsilane 1d was subjected to isomerization in the presence of CD3CO2D, a significant amount of deuterium scrambling at the α-position (>20%) as well as at the methyl group (>45%) was observed in corresponding vinylsilane, indicating that the isomerization step is reversible and the rate determining step (Scheme 2b). It is also observed that when vinylsilane 6b was made to react with methyl acrylate 2a under standard conditions, it successfully underwent highly regioselective C–H activation and afforded coupling product 3b′a in 80% yield (Scheme 2c). This result confirms that the coupling reaction proceeds via vinyl silane intermediate 6.Open in a separate windowScheme 2Isomerization of allylsilanes and deuterium study.It is delightful to mention that diene 3aa successfully underwent the Diels–Alder reaction with N-phenyl maleimide 7 in toluene at 80 °C, to afford single isomer 8 in 70% yield which ensures the pragmatism of the method (Scheme 3). The unique power of this ruthenium-catalyzed C–H functionalization strategy is illustrated by the late-stage diversification of the diene 5gh, to a very reactive Michael acceptor 9 (conventional route for preparation of 9 requires in situ oxidation of α-hydroxyketones using 10 equiv. MnO2 followed by the Wittig reaction, which generates a superstoichiometric amount of phosphine waste)67,68via selective hydrolysis of the acetate group, which is useful in the synthesis of ester-thiol 10,69 cyclohexenone 11 and polysubstituted piperidine 12 (ref. 70) (Scheme 4). Thus the Micheal acceptor 9 on reaction with thiophenol generated compound 10 in excellent yield and high regioselectivity. On the other hand compound 9 on reaction with heptanal in the presence of Hayashi–Jørgensen''s catalyst afforded the Michael adduct 13 in 72% yield and excellent diastereoselectivity. Keto-aldehyde 13 was converted to highly substituted cyclohexenone 11 and piperidine 12.Open in a separate windowScheme 3Application to the Diels–Alder reaction.Open in a separate windowScheme 4Application to the organocatalytic Michael addition reaction.The potential of this Ru-catalysed reaction was further demonstrated by norpyrenophorin synthesis.71–74 Norpyrenophorin 14 is a synthetic 16-membered lactone which has essentially the same physiological activity as the natural fungicide pyrenophorin 15 and the antibiotic vermiculin 16.73 A brief retrosynthetic analysis revealed that the dimeric macrocycle 14 could be dissected into monomer 17 which could be easily accessed from oxidative coupling of 2a with 18 using the C–H activation reaction (Scheme 5). Ruthenium catalysed oxidative coupling of symmetric allylester 18 with 2a generated the key intermediate 19 in 32% yield. Selective hydrolysis of acetyl enolate 19 was accomplished by the treatment with K2CO3 in methanol to provide 20 in 70% yield. In accordance with some previously reported studies, the active ketone functionality of 20 was protected as ketal by treatment with ethylene glycol in refluxing benzene to afford substrate 21. Selective hydrolysis of acetate was achieved using Bu2SnO to generate alcohol 22 and finally, aluminium–selenium adduct mediated72 ring closing lactonization followed by deketalization ensured the completion of synthesis of 14 in 23% yield (two steps) (Scheme 6). A similar type of dimerization reaction could be envisioned to synthesize the natural products pyrenophorin 15 and vermiculin 16.Open in a separate windowScheme 5Retrosynthetic analysis of norpyrenophorin.Open in a separate windowScheme 6Synthesis of norpyrenophorin.Based on the above result and previous report, a plausible mechanism for this oxidative coupling reaction is depicted in Scheme 7. The catalytic cycle is initiated by substrate 4g coordination to in situ generated reactive cationic ruthenium complex [Ru(OAc)L]+ A, followed by weakly coordinating ester group directed C–H activation of allyl ester to give a π-allyl ruthenium intermediate C, which again would undergo isomerization to produce intermediate D. In the case of allyl silanes, an α-silyl effect might play an important role for the isomerisation of allylsilanes to vinylsilanes via the silylated allyl anion.56 Regioselective C–H activation of in situ generated vinyl acetate would give intermediate E. Induction of stability to the carbon–metal bond by the silyl group favours regioselective C–H insertion in the case of vinyl silanes.57 Coordination followed by 1,4-addition of vinyl ruthenium species to the activated olefins (acrylate, 2a) would generate intermediate G, which would further undergo β-hydride elimination to provide a single isomer of 1,3-diene H and intermediate I could undergo reductive elimination followed by reoxidation of in situ forming Ru(0) species in the presence of Cu(OAc)2 to regenerate the reactive ruthenium(ii) complex A for the next catalytic cycle.Open in a separate windowScheme 7Plausible reaction mechanism.  相似文献   

19.
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   

20.
Difluorination of α-(bromomethyl)styrenes via I(I)/I(III) catalysis: facile access to electrophilic linchpins for drug discovery     
Joel Hfliger  Keith Livingstone  Constantin G. Daniliuc  Ryan Gilmour 《Chemical science》2021,12(17):6148
Simple α-(bromomethyl)styrenes can be processed to a variety of 1,1-difluorinated electrophilic building blocks via I(I)/I(III) catalysis. This inexpensive main group catalysis strategy employs p-TolI as an effective organocatalyst when combined with Selectfluor® and simple amine·HF complexes. Modulating Brønsted acidity enables simultaneous geminal and vicinal difluorination to occur, thereby providing a platform to generate multiply fluorinated scaffolds for further downstream derivatization. The method facilitates access to a tetrafluorinated API candidate for the treatment of amyotrophic lateral sclerosis. Preliminary validation of an enantioselective process is disclosed to access α-phenyl-β-difluoro-γ-bromo/chloro esters.

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

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

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