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 31 毫秒
1.
General photoactivation of electron donor–acceptor (EDA) complexes between arylsulfonium salts and 1,4-diazabicyclo[2.2.2]octane with visible light or natural sunlight was discovered. This practical and efficient mode enables the production of aryl radicals under mild conditions, providing an unrealized opportunity for two-step para-selective C–H functionalization of complex arenes. The novel mode for generating aryl radicals via an EDA complex was well supported by UV-vis absorbance measurements, nuclear magnetic resonance titration experiments, and density functional theory (DFT) calculations. The method was applied to the regio- and stereo-selective arylation of various N-heterocycles under mild conditions, yielding an assembly of challengingly linked heteroaryl–(hetero)aryl products. Remarkably, the meaningful couplings of bioactive molecules with structurally complex drugs or agricultural pharmaceuticals were achieved to display favorable in vitro antitumor activities, which will be of great value in academia or industry.

General photoactivation of EDA complexes between arylsulfonium salts and 1,4-diazabicyclo[2.2.2]octane was discovered. This practical mode enables the generation of aryl radicals for C–H functionalization of arenes.  相似文献   

2.
Development of efficient and stereoselective synthesis of prostaglandins (PGs) is of utmost importance, owing to their valuable medicinal applications and unique chemical structures. We report here a unified synthesis of PGs cloprostenol, bimatoprost, PGF, fluprostenol, and travoprost from the readily available dichloro-containing bicyclic ketone 6a guided by biocatalytic retrosynthesis, in 11–12 steps with 3.8–8.4% overall yields. An unprecedented Baeyer–Villiger monooxygenase (BVMO)-catalyzed stereoselective oxidation of 6a (99% ee), and a ketoreductase (KRED)-catalyzed diastereoselective reduction of enones 12 (87 : 13 to 99 : 1 dr) were utilized in combination for the first time to set the critical stereochemical configurations under mild conditions. Another key transformation was the copper(ii)-catalyzed regioselective p-phenylbenzoylation of the secondary alcohol of diol 10 (9.3 : 1 rr). This study not only provides an alternative route to the highly stereoselective synthesis of PGs, but also showcases the usefulness and great potential of biocatalysis in construction of complex molecules.

We report a unified chemoenzymatic asymmetric synthesis of five prostaglandins, featuring two enzymatic redox transformations and a copper(ii)-catalyzed regioselective p-phenylbenzoylation.  相似文献   

3.
A highly general and straightforward approach to access chiral bis(indolyl)methanes (BIMs) bearing quaternary stereocenters has been realized via enantioconvergent dehydrative nucleophilic substitution. A broad range of 3,3′-, 3,2′- and 3,1′-BIMs were obtained under mild conditions with excellent efficiency and enantioselectivity (80 examples, up to 98% yield and >99 : 1 er). By utilizing racemic 3-indolyl tertiary alcohols as precursors of alkyl electrophiles and indoles as C–H nucleophiles, this organocatalytic strategy avoids pre-activation of substrates and produces water as the only by-product. Mechanistic studies suggest a formal SN1-type pathway enabled by chiral phosphoric acid catalysis. The practicability of the obtained enantioenriched BIMs was further demonstrated by versatile transformation and high antimicrobial activities (3al, MIC: 1 μg mL−1).

A highly general and straightforward approach to access chiral bis(indolyl)methanes (BIMs) bearing quaternary stereocenters has been realized via enantioconvergent dehydrative nucleophilic substitution.  相似文献   

4.
The transformations of N-phenacyl(p-nitrophenacyl, benzyl)-2,5-dimethyl-4-nitrophenyl (nitrobenzyl, benzyl)pyridinium bromides under the influence of potassium carbonate solution were studied. Stable pyridinium ylids were obtained in the case of -phenylpyridines that contain a nitro group in the benzene ring and in the case of -benzylpyridines with an N-nitrosubstituted phenacyl group. The conclusion that electron-acceptor substituents have a stabilizing effect on the stability of the ylids was confirmed. Under these conditions -nitrobenzyl derivatives are converted to substituted 1,4-dihydropyridines. The positions at which deprotonation of the starting quaternary pyridinium salts occurs and the formation of 1,4- and 1,2-dihydropyridines were established by PMR spectroscopy. The corresponding pyridinium salts were converted to a new group of indolizines containing a p-nitrophenyl (p-nitrobenzyl) substituent in the 2 or 7 position by the Chichibabin method. It was established that substituted dihydropyridines are converted to indolizines; ideas that confirm the scheme of the previously proposed mechanism for the formation of indolizines through a step involving ylids are expressed.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 3, pp. 384–389, March, 1979.  相似文献   

5.
4-Aryl-1,2,6-trimethyl-3,5-diacetyl-1,4-dihydropyridines and the corresponding pyridinium salts, which upon reduction with NaBH4 form 4-aryl-1,2,6-trimethyl-3,5-diacetyl-1,2-dihydropyridines, were synthesized.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 4, pp. 508–513, April, 1983.  相似文献   

6.
A direct Pd(ii)-catalyzed kinetic resolution of heteroaryl-enabled sulfoximines through an ortho-C–H alkenylation/arylation of arenes has been developed. The coordination of the sulfoximine pyridyl-motif and the chiral amino acid MPAA ligand to the Pd(ii)-catalyst controls the enantio-discriminating C(aryl)–H activation. This method provides access to a wide range of enantiomerically enriched unreacted aryl-pyridyl-sulfoximine precursors and C(aryl)–H alkenylation/arylation products in good yields with high enantioselectivity (up to >99% ee), and selectivity factor up to >200. The coordination preference of the directing group, ligand effect, geometry constraints, and the transient six-membered concerted-metalation–deprotonation species dictate the stereoselectivity; DFT studies validate this hypothesis.

A Pd/MPAA catalysed KR of heteroaryl substituted sulfoximines through C–H alkenylation and arylation (up to >99% ee) is developed. In-depth DFT studies uncover the salient features.  相似文献   

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

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

Known as one of the most significant and reliable access methods to chiral heterocycles, asymmetric cycloisomerization of conjugated enyne has caught extensive attention and interest for its wide applications in synthetic route design and mechanistic investigation.1 Specifically, asymmetric cyclization of conjugated enynone (X = C, Z = O) has been successfully developed and applied to the rapid construction of various chiral furan-containing skeletons with high efficiency in an extremely operationally simple manner (Scheme 1a).2 However, compared to the fruitful research with enynone, it is surprising that the analogous asymmetric version of azaenyne (Z = N–R) still remains underdeveloped.3 In fact, no successful example of catalytic asymmetric cyclization of azaenyne has been reported in the literature despite the apparent significance of nitrogen-containing five-membered heterocycles in the synthetic and pharmaceutical community.4 In 2004, Haley and Herges reported a detailed experimental and theoretical study of the cyclization reaction of (2-ethynylphenyl)-phenyldiazene, which is a unique azaenyne.5 According to the DFT calculations, very close and low activation barriers for 5-exo-dig and 6-endo-dig cyclization pathways under catalyst-free conditions were found, which shed light on the inherent challenges of the asymmetric reaction of azaenyne (Scheme 1b). For instance, there was usually a regioselectivity issue (5-exo and 6-endo) in the cyclization reaction of azaenyne because of their close reaction barriers where the competitive 6-endo-dig cyclization3a,6 may lead to troublesome side-product formation. In addition, the low activation barrier deriving from the strong N-nucleophilicity of azaenyne may easily lead to self-cyclization which will cause severe background reactions to interfere with the asymmetric process. More troublingly, this transformation might suffer from catalyst deactivation arising from the high coordinating capability of the nitrogen atom in both starting materials and products, which might give more opportunities to the propagation of detrimental background reactions. In some cases, even a super-stoichiometric amount of transition metal has to be used to ensure effective conversion.3a,7 Therefore, although many nonchiral approaches have been reported,3,5 catalytic asymmetric cyclization of azaenyne still remains elusive due to the inherent obstacles aforementioned. With our continuous interest in alkyne chemistry,2a,8 herein we designed a cap-tether synergistic modulation strategy to tackle these challenges, envisioning that modulation of the tethered atom and protecting cap of nitrogen in the azaenyne would intrinsically perturb and alter the reactivity of the starting material, and therefore the azaenyne motif could be effectively harnessed as a promising synthon for asymmetric transformations (Scheme 1c). It should be noted that the obtained centrally chiral product produced from intramolecular C–H insertion of donor-type metal carbene9 might be potentially converted into the axially chiral molecule via a central-to-axial chirality conversion strategy.Open in a separate windowScheme 1Development of the asymmetric cyclization reaction of conjugated azaenyne.With this design in mind, different types of azaenynes bearing typical tethering atoms and capping groups were chosen to test our hypothesis and representative results are shown in Scheme 2. First, tBu-capping imine (X = C, R = tBu) was selected as a substrate to test our hypothesis.6a It was found that the imine exhibited low reactivity and the reaction temperature has to be elevated to 100 °C to initiate the transformation with or without catalyst. Unfortunately, the desired 5-exo-dig cyclization product was not detected, but isoquinoline from 6-endo-dig cyclization was obtained instead (Scheme 2a). To further regulate and control the regioselectivity and reactivity, triazene (X = N, R = N-piperidyl) was then investigated. Similarly, this substrate also showed low reactivity and it is still required to be heated at 100 °C for conversion. In the absence of a metal catalyst, an unexpected alkyne, deriving from the fragmentation of the triazene moiety, was produced in 41% yield. When 2 mol% Rh2(OPiv)4 was added as a catalyst, the side reaction could be efficiently suppressed and the reaction selectivity was apparently reversed. In this case, the target C–H insertion dihydrofuran was furnished as the major product in 30% yield but still accompanied by concomitant formation of 12% yield of undesired alkyne (Scheme 2b). The above investigations showed neither the imine nor triazene was an ideal substrate for the asymmetric reaction. Thus, we moved our attention to the diazene substrate (X = N, R = aryl). As demonstrated by Haley''s and Herges'' pioneering work, ortho-alkynyl diazene, compared with imine and triazene, was more unstable and tended to self-cyclization even at room temperature.5a As shown in Scheme 2c, the ortho-alkynyl diazene degrades and 5-exo-dig cyclization products could be observed even in DCE solvent without any catalyst at room temperature. When the phenyl capping group was installed in the substrate, the reaction furnished 10% yield of isoindazole derivative. The uncatalyzed self-cyclization reaction was obviously accelerated when an electron-rich capping group (4-MeO–C6H4–) was introduced, affording the corresponding product in 20% yield. Inspired by these findings, we assumed that installation of an electron deficient group on the capping phenyl would reduce the nucleophilicity of the nitrogen atom and thus the troublesome self-cyclization reaction might be effectively inhibited. To our delight, when a bromo-substituent was introduced onto the phenyl cap, the undesired self-cyclization was almost suppressed. When Rh2(OPiv)4 was added as a catalyst, the desired carbene-involved C–H insertion product was furnished in 90% yield at room temperature. Worthy of note was the total absence of any cinnoline formation from 6-endo-dig cyclization.3a,6b In short, the synthetic challenges associated with regioselectivity (5-exo-dig and 6-endo-dig), strong background reaction and catalyst deactivation could be successfully regulated and controlled via a tether-cap synergistic modulation strategy.Open in a separate windowScheme 2Typical substrate investigation.Encouraged by the above findings, ortho-alkynyl bromodiazene 1a was chosen as a model substrate and different types of chiral dirhodium catalysts10 were screened in DCE at room temperature for 48 h. As shown in
EntryRh(ii)*SolventYieldb [%]erc
1Rh2(R-DOSP)4DCE5629 : 71
2Rh2(5S-MEPY)4DCE1750 : 50
3Rh2(S-BTPCP)4DCE618 : 92
4Rh2(S-PTPA)4DCE9191 : 9
5Rh2(S-PTTL)4DCE8697 : 3
6Rh2(S-PTAD)4DCE9394 : 6
7Rh2(S-NTTL)4DCE9296 : 4
8Rh2(S-TCPTTL)4DCE9598 : 2
9 Rh 2 (S-TFPTTL) 4 DCE 98 d 98 : 2
10Rh2(S-TFPTTL)4DCM8898 : 2
11Rh2(S-TFPTTL)4Toluene9298 : 2
12Rh2(S-TFPTTL)4MeCN1692 : 8
13Rh2(S-TFPTTL)4 n-Hexane9698 : 2
14eRh2(S-TFPTTL)4DCE65f96 : 4
Open in a separate windowaUnless otherwise noted, reactions were performed at 0.1 M in DCE using 0.20 mmol substrate and catalyst (2 mol%) under a N2 atmosphere.bDetermined by 1H NMR spectroscopy.cThe er value of 2a was determined by HPLC using a chiral stationary phase.dIsolated yields.e1 mol% catalyst was used.f25% starting material was recovered.With the optimized reaction conditions in hand (Scheme 3, the catalytic process could be successfully applied to azaenynes 1 bearing different ether side chains. For example, in addition to 1a, various azaenyne derivatives containing benzylic ethers could be efficiently converted into the desired products 2b–i with excellent diastereoselectivities and enantioselectivities (>99 : 1 d.r., 97:3–99 : 1 er). The yields were typically higher than 90% for most substrates. Satisfyingly, the substrates with bulkier aryl groups were well-tolerated to afford the isoindazole products 2j–m in good yields with excellent diastereo- and enantiocontrol (>97 : 3 d.r., > 95 : 5 er). In addition to azaenynes with arylmethyl ether, this protocol was also successfully applied to substrates with allylic ether, propargyl ether and even aliphatic ether to furnish the cyclization products 2n–u in good yields with decent diastereo- and enantioselectivities (>93 : 7 d.r., > 90 : 10 er). In the cases of allylic and propargyl ether, only C–H insertion products (2n–p) were observed though cyclopropanation or cyclopropenation often took place competitively when using the allylic or propargyl substrate to trap the carbene intermediate.11 It was noted that the azaenynes with aliphatic ether, which represent challenging substrates2a in the asymmetric carbene transfer reactions, also showed good reactivities to afford the corresponding chiral dihydrobenzofurans (2q–u) with excellent diastereoselectivities (>93 : 7 d.r.) and enantioselectivities (>98 : 2 er). Interestingly, when phenyl and methoxyphenyl capping azaenynes, which potentially suffered from the undesired background reactions, were subjected to the standard conditions, chiral products (2v–w) could be obtained with high optical purity (>99 : 1 d.r., > 96 : 4 er) as well. These results might be attributed to the high catalytic activity of Rh2(S-TFPTTL)4 in the asymmetric cyclization process, which eventually led to complete suppression of the uncatalyzed self-cyclization. The scopes with respect to the group R1 on the fused phenyl ring were further investigated. Both electron-rich and -deficient substituents R1 were well accommodated, with the product yields ranging from 80% to 99%, enantiomeric ratios ranging from 95 : 5 to 97 : 3 and diastereomeric ratios higher than 99 : 1 (2x–z). In addition, azaenyne substituted with an alkyl side chain at the alkynyl carbon atom was also tested, giving tetrahydrofuran (2aa) with excellent diastereoselectivity (>99 : 1 d.r.), good enantioselectivity (90 : 10 er) and moderate yield (43%). In addition to the side chain of ether, this asymmetric protocol could even be extended to the more challenging nitrogen- and thio-tethered analogues, albeit with somewhat lower reactivities (46–65% yields) but good stereoselectivities (93 : 7 er and 84 : 16 d.r. for 2ab; 81 : 19 er and >99 : 1 d.r. for 2ac). Structures of the resulting products were confirmed by X-ray crystallographic analysis of their analogue 2h.Open in a separate windowScheme 3 aUnless otherwise noted, the reactions were performed under standard conditions for 48 h or monitored by TLC until the starting material disappeared. b5 mol% catalyst was used. cReactions were performed in n-hexane, using 2 mol% Rh2(S-TCPTTL)4 as the catalyst.The successful preparation of centrally chiral isoindazole through the asymmetric cyclization reaction prompted us to explore the further applications of this protocol. Axially chiral biaryl skeletons are undoubtedly regarded as one of the most prominent structural motifs for their ubiquity in natural products, pharmaceuticals and useful chiral ligands in asymmetric catalysis.12 Due to the lower rotational barrier, there are only limited examples of the enantioselective synthesis of axially chiral atropisomers featuring a five-membered ring, especially those bearing two pentatomic aromatics.13 Compared with the furan analogue, the extending cap in the isoindazole scaffold provides additional ortho steric hindrance making these molecules possible candidates for the preparation of five-five-membered biaryl atropisomers. Considering the unique chiral skeleton of dihydrofuranyl isoindazole 2, we began to explore their potential application in chiral atropisomer synthesis via a central-to-axial chirality transfer strategy. As shown in Scheme 4, oxidative aromatization of representative dihydrofuran candidate 2m furnished two configurationally unstable atropisomers, which might be attributed to their relatively low rotational barriers as five-membered atropisomers especially when the furan ring was incorporated (see ESI for details). Therefore, it was hypothesized that extending the fused phenyl to naphthyl might afford stable atropisomers by enhancing the ortho steric hindrance (Scheme 4b).Open in a separate windowScheme 4Investigation of central-to-axial chirality transfer.To our delight, as shown in Scheme 5, naphthyl-fused dihydrofurans 4 could be easily accessed through the above established dirhodium-catalyzed cyclization process and configurationally stable atropisomers 5 could be generated via further oxidative dehydrogenation with 2,3-dichloro-5,6-dicyano-1,4-benzoquinone (DDQ) as the oxidant (see ESI for the proposed mechanism). For example, asymmetric cyclization reactions proceeded smoothly to give the centrally chiral compounds 4 in good yields (54–99%) with excellent diastereoselectivities (92 : 8–99 : 1 d.r.) and enantioselectivities (95 : 5–99 : 1 er) under slightly modified reaction conditions. This reaction was compatible with a variety of arylmethyl side chains in azaenynes and well-accommodated with various functional groups (F, Cl, Br, OMe, and –CO2Me). Additionally, oxidative dehydrogenation of chiral candidates 4 with DDQ smoothly resulted in the formation of axially chiral atropisomers 5 in 90–99% yields with only slight loss of chiral integrity (90 : 10–97 : 3 er). An enantiomerically pure atropisomer could be obtained through a simple recrystallization procedure as exemplified by compound 5g. The structure and absolute configuration of isoindazole 4g and atropisomer 5g were confirmed by their single-crystal X-ray diffraction analysis.Open in a separate windowScheme 5 aConditions for cyclization of azaenyne: Rh2(S-TCPTTL)4 (2 mol%), n-hexane, rt for 48 h or monitored by TLC until the starting material disappeared; conditions for oxidative chirality transfer: DDQ (2 equiv.), DCE, −20 °C for 48 h or monitored by TLC until the starting material disappeared. b45 °C. cDDQ (5 equiv.). dRoom temperature. eAfter one recrystallization.With centrally and axial chiral molecules in hand, further transformations of these compounds were also explored. The tethered nitrogen atom in azaenynes not only showed a synergetic effect with the capping group on promoting asymmetric cyclization but also served as an innate directing group for late-stage modifications via C–H functionalization. As shown in Scheme 6, a variety of functional groups could be directly introduced onto the capping aromatic rings, allowing for rapid build-up of molecular complexity. For example, synthetically valuable alkenyl,14 allyl15 and alkynyl16 groups could be easily incorporated into the final structures, which had wide potential applications in organic synthesis (6a–c). Furthermore, C–H alkylation,17 amidation18 and selenylation19 were performed smoothly to afford the desired products 6d–g. It is noteworthy that unique chiral chelation backbones were constructed by amidation and selenylation of the isoindazole moiety (6e–g). In addition to centrally chiral compounds, axial chiral atropisomers 5 themselves could be efficiently converted to their functionalized scaffolds as well (6h–i) through a similar directed C–H functionalization process.Open in a separate windowScheme 6Late-stage modification of chiral isoindazoles. Reaction conditions: a4-octyne, [Rh(Cp*Cl2)]2, AgSbF6, Cu(OAc)2, DCE, 80 °C. bAllyl carbonate, [Rh(Cp*Cl2)]2, AgSbF6, PivOH, PhCl, 40 °C. cHypervalent iodine-alkyne, [Rh(Cp*Cl2)]2, Zn(OTf)2, DCE, 80 °C. dAlkene, [Rh(Cp*Cl2)]2, AgSbF6, AcOH, 1,4-dioxane, 50 °C. e3-Phenyl-1,4,2-dioxazol-5-one, [Cp*Co(MeCN)3](SbF6)2, DCE, 80 °C. fPhSeCl, [Rh(Cp*Cl2)]2, AgSbF6, THF, 60 °C.  相似文献   

8.
Sulfur stereogenic centers in transition-metal-catalyzed asymmetric C–H functionalization: generation and utilization     
Wentan Liu  Jie Ke  Chuan He 《Chemical science》2021,12(33):10972
Transition-metal-catalyzed enantioselective C–H functionalization has emerged as a powerful tool for the synthesis of enantioenriched compounds in chemical and pharmaceutical industries. Sulfur-based functionalities are ubiquitous in many of the biologically active compounds, medicinal agents, functional materials, chiral auxiliaries and ligands. This perspective highlights recent advances in sulfur functional group enabled transition-metal-catalyzed enantioselective C–H functionalization for the construction of sulfur stereogenic centers, as well as the utilization of chiral sulfoxides to realize stereoselective C–H functionalization.

This perspective highlights sulfur functional groups enabled enantioselective C–H functionalization for the construction of sulfur stereogenic centers, and the utilization of chiral sulfoxide to realize stereoselective C–H functionalization.  相似文献   

9.
Pyridylphosphonium salts as alternatives to cyanopyridines in radical–radical coupling reactions     
Jacob W. Greenwood  Benjamin T. Boyle  Andrew McNally 《Chemical science》2021,12(31):10538
Radical couplings of cyanopyridine radical anions represent a valuable technology for functionalizing pyridines, which are prevalent throughout pharmaceuticals, agrochemicals, and materials. Installing the cyano group, which facilitates the necessary radical anion formation and stabilization, is challenging and limits the use of this chemistry to simple cyanopyridines. We discovered that pyridylphosphonium salts, installed directly and regioselectively from C–H precursors, are useful alternatives to cyanopyridines in radical–radical coupling reactions, expanding the scope of this reaction manifold to complex pyridines. Methods for both alkylation and amination of pyridines mediated by photoredox catalysis are described. Additionally, we demonstrate late-stage functionalization of pharmaceuticals, highlighting an advantage of pyridylphosphonium salts over cyanopyridines.

Cyanopyridines form dearomatized radical anions upon single-electron reduction and participate in photoredox coupling reactions. Pyridylphosphonium salts replicate that reactivity with a broader scope and increase the utility of these processes.

Modern photoredox catalysis and electrochemistry have enabled new synthetic methods that proceed via open-shell intermediates.1 Under this regime, pyridine functionalization strategies have been developed where 4-cyanopyridines undergo single-electron reduction to form dearomatized radical species that couple with other stabilized radicals (Scheme 1A).2 The cyano group is critical for efficient reactivity via pyridyl radical anions; alternatives such as 4-halopyridines more readily undergo elimination to pyridyl radicals after single-electron reduction resulting in a distinct set of coupling processes.3 We aimed to show that pyridylphosphonium salts could replicate the reactivity of cyanopyridines and allow a broader set of inputs into dearomatized pyridyl radical coupling reactions.4Open in a separate windowScheme 1Expansion of radical coupling reactions to complex pyridines.Cyanopyridines have facilitated pyridine alkylation, allylation, and alkenylation reactions providing access to valuable building blocks for medicinal and agrochemical programs.5 The cyano group is essential for these methods, but a problem arises when applying this chemistry to complex pyridines, such as those found in pharmaceutical and agrochemical candidates. These structures are often devoid of pre-installed functional groups, and it is often challenging to install a cyano group from C–H precursors regioselectively.6 We envisioned pyridylphosphonium salts, regioselectively constructed from the C–H bonds of a diverse set of pyridines, could serve as alternatives to cyanopyridines.7 Herein, we report couplings between alkyl BF3K salts and preliminary studies of carboxylic acids and amines with pyridylphosphonium salts, including late-stage functionalization of complex pyridine-containing pharmaceuticals using this strategy.Recently, we reported a radical coupling reaction between a boryl-stabilized cyanopyridyl radical and a boryl-stabilized pyridylphosphonium radical.7a The intermediate radicals arose via an unusual inner-sphere process that would be difficult to extend to other coupling reactions. A significant advance would be to show that pyridylphosphonium salts could function more generally as radical anion precursors and mimic the reactivity of cyano-pyridines. In particular, showing their viability in photoredox and electrochemical processes would translate to numerous synthetic transformations. To demonstrate this principle, we envisioned a redox-neutral alkylation reaction (Scheme 1B) via a radical coupling between radical zwitterion I, formed through single-electron reduction of a pyridylphosphonium salt (Eredp/2 = −1.51 V vs. SCE) and benzyl radical II, resulting from single-electron oxidation of a BF3K salt (Ered = +1.10 V vs. SCE for a primary benzylic salt).8 Loss of triphenylphosphine from dearomatized intermediate (III) would furnish the alkylated pyridine product. Notably, the redox events could invert, where the photocatalyst oxidizes the BF3K salt first and reduces the pyridylphosphonium salt second, broadening the scope of amenable photocatalysts.We began our investigation by examining a series of photocatalysts for the coupling reaction of phosphonium salt 1a, formed with complete regioselectivity for the 4-position from 2-phenylpyridine, and benzylic BF3K salt 2a under irradiation from a 455 nm Kessil light (Scheme 1B are potentially interchangeable.1b The Adachi-type photocatalyst 3DPAFIPN improved the yield to 77% with a further increase to 82% after increasing the reaction concentration (entries 3 and 4). Adding 2,6-lutidine, previously shown as an effective additive for photoredox cross-coupling reactions of BF3K salts by the Molander group,9 had no impact on the yield of 2-phenylpyridine salt 1a (entry 5) and the [Ir(ppy)2(dtbbpy)]PF6 catalyst was marginally less efficient under the same conditions (entry 6). We observed that 2,6-lutidine did substantially improve the yield when isomeric 3-Ph salt 1b was employed (entries 7 and 8); without 2,6-lutidine, the crude 1H NMR indicates significant amounts of decomposition occurred, including 3-phenylpyridine, and the 4- vs. 2-position product ratio was 3 : 1. This outcome suggests that protiodephosphination and non-selective Minisci-type pathways can occur under these conditions. With 2,6-lutidine, the crude reaction pathway is cleaner, and the 4- vs. 2-position ratio improved to 8 : 1. At this point, we have not established the role of 2,6-lutidine, although it is conceivable that it reacts with BF3 produced as the reaction progresses. In 2-substituted systems, steric hindrance around the pyridine N-atom of the salt would deter BF3-coordination, whereas, in 3-substituted systems, such as salt 1b, coordination is more likely and may have a deleterious effect on the reaction (vide infra). Given the structural variation of pyridines that we anticipated applying to this process and how those structures could impact boron speciation during the reaction, we elected to use 2,6-lutidine as an additive in all subsequent reactions.10Optimization of pyridine alkylation, photocatalyst data and effect of BF3·OEt2 as an additivea
Open in a separate windowaConditions: 1a (1.0 equiv.), 2a (2.0 equiv.), photocatalyst (2 mol%), additive (3.0 equiv.), rt.bYields determined by 1H NMR analysis using 1,3,5-trimethoxybenzene as internal standard.cIsolated yield on 0.50 mmol scale.dIsolated yield on 2.00 mmol scale.e3 : 1 4- vs. 2-regioisomeric ratio determined from the crude 1H NMR.f8 : 1 4- vs. 2-regioisomeric ratio determined from the crude 1H NMR.gUsed 365 nm LEDs instead of 455 nm Kessil light for 89 h.hAll redox potentials reported vs. SCE and all values compiled from previous literature reports.1iCounterion omitted in structure for simplicity.We conducted a series of further experiments to explore the effect of light and photocatalyst type on the reaction ().11 Furthermore, a photocatalyst with a redox potential window misaligned with the redox events in Scheme 1B, [Mes-Acr]BF4, is also competent (entry 11). An energy transfer mechanism was considered based on entry 9, but the low triplet state energies for [Mes-Acr]BF4 make this pathway unlikely (12–14Employing the optimized conditions, we investigated the scope of pyridylphosphonium salts in this coupling process ( Open in a separate windowaIsolated yields of single regioisomers. Conditions: 1 (1.0 equiv.), 2a (2.0 equiv.), 3DPAFIPN (2 mol%), 2,6-lutidine (3.0 equiv.), 1,4-dioxane (0.3 M), rt.b11 : 1 crude regioisomeric ratio. Isolated as a single regioisomer. Grey circle denotes the site of alkylation for the minor regioisomer.cWith 1 equiv. TfOH.Next, we converted a series of drug-like fragments and pharmaceuticals into phosphonium salts in this alkylation reaction. These examples represent the most significant advantage of this chemistry as installing a cyano group would be challenging from the C–H bond and limits the ability to make analog compounds. In addition, these structures contain multiple reactive sites and functional groups that could interfere with the coupling process. Nevertheless, we synthesized benzylated fragments 3n–3r without difficulty. Notably, other heterocycles are compatible, such as thiazoles and protected piperidines and pyrrolidines. The pyridine-pyrimidine biaryl 3p is particularly interesting as the phosphonium salt formed site-selectively on the pyrimidine ring, and the photoredox coupling proceeded in good yield on this heterocycle. Lastly, we demonstrated coupling with four FDA-approved pharmaceuticals and an agrochemical that illustrates functional group tolerance for protonated tertiary amines, amides, aryl halides, benzyl ethers, and sulfones (3s–3w). These examples validate this tactic for late-stage functionalization of complex pyridines.Scheme 2A shows the scope of the BF3K salts in the photoredox alkylation reaction. Secondary benzylic salts with electron-withdrawing and electron-donating groups are suitable coupling partners (3x–3z). In the case of 3y, we added a 1.2 : 1 mixture of benzylic and homobenzylic BF3K salts but only observed the benzylated product, presumably because the primary isomer is more difficult to oxidize. Secondary naphthyl and primary benzylic BF3K salts are proficient, resulting in 3aa and 3ab. The reaction also tolerates α-amino BF3K salts as evidenced by heterobenzylic amine derivative 3ac. At this stage, non-stabilized radicals were not successful in this process.Open in a separate windowScheme 2Scope of radical coupling partners. aIsolated yields of single regioisomers. Conditions: 1a (1.0 equiv.), 2 (2.0 equiv.), 3DPAFIPN (2 mol%), 2,6-lutidine (3.0 equiv.), 1,4-dioxane (0.3 M), rt. bBF3K starting material is 1.2 : 1 mixture of regioisomers (benzylic : primary). c>20 : 1 regioisomeric ratio and 5.7 : 1 mono : bis alkylated product in crude 1H NMR spectrum. Isolated as single monoalkylated regioisomer.Finally, we investigated whether pyridylphosphonium salts are competent with other radical precursors. In Scheme 2B, we obtained a preliminary result (unoptimized) of coupling with a carboxylic acid. These abundant compounds would improve the scope of radical coupling partners, and further studies are currently underway in our laboratory. In addition, Wu recently reported a method for photoredox catalyzed amination using cyanopyridines as coupling partners, and we attempted to replicate this transformation using pyridylphosphonium salts (Scheme 2C).15 Applying salt 1a to the reaction protocol with N-methyl aniline resulted in diaryl amine 4.16 Similarly, using N,O-dimethylhydroxylamine as a coupling partner, followed by in situ cleavage of the N–O bond, formed aniline 5 in reasonable yield. Consistent with the results in 相似文献   

10.
Total synthesis of (−)-funebrine via Au-catalyzed regio- and stereoselective γ-butyrolactonization of allenylsilane     
Takuya OkadaKazuhiko Sakaguchi  Tetsuro ShinadaYasufumi Ohfune 《Tetrahedron letters》2011,52(44):5744-5746
The stereoselective total synthesis of (−)-funebrine from 2-butyn-1-ol was described. The crucial steps in the synthesis involved the stereoselective enolate Claisen rearrangement of the (S)-α-acyloxy-α-alkynylsilane 8, the Au-catalyzed regio- and stereoselective lactonization of the allenylsilane 7, and the Paal-Knorr pyrrole condensation using an unsymmetrical 1,4-diketone 4b.  相似文献   

11.
Atropoenantioselective palladaelectro-catalyzed anilide C–H olefinations viable with natural sunlight as sustainable power source     
Johanna Frey  Xiaoyan Hou  Lutz Ackermann 《Chemical science》2022,13(9):2729
Enantioselective electrocatalyzed transformations represent a major challenge. We herein achieved atropoenantioselective pallada-electrocatalyzed C–H olefinations and C–H allylations with high efficacy and enantioselectivity under exceedingly mild reaction conditions. With (S)-5-oxoproline as the chiral ligand, activated and non-activated olefins were suitable substrates for the electro-C–H activations. Dual catalysis was devised in terms of electro-C–H olefination, along with catalytic hydrogenation. Challenging enantiomerically-enriched chiral anilide scaffolds were thereby obtained with high levels of enantio-control in the absence of toxic and cost-intensive silver salts. The resource-economy of the transformation was even improved by directly employing renewable solar energy.

Asymmetric pallada-electrocatalyzed C–H activation of achiral anilides were accomplished by catalyst control with high levels of enantioselectivity. Dual catalysis was devised, while photovoltaic cells could be used to empower the electrocatalysis.  相似文献   

12.
Combining palladium and ammonium halide catalysts for Morita–Baylis–Hillman carbonates of methyl vinyl ketone: from 1,4-carbodipoles to ion pairs     
Yang Yang  Bo Zhu  Lei Zhu  Ying Jiang  Chun-Ling Guo  Jing Gu  Qin Ouyang  Wei Du  Ying-Chun Chen 《Chemical science》2021,12(34):11399
Here we report that Morita–Baylis–Hillman carbonates from diverse aldehydes and methyl vinyl ketones can be directly utilised as palladium-trimethylenemethane 1,4-carbodipole-type precursors, and both reactivity and enantioselectivity are finely regulated by adding a chiral ammonium halide as the ion-pair catalyst. The newly assembled intermediates, proposed to contain an electronically neutral π-allylpalladium halide complex and a reactive compact ion pair, efficiently undergo asymmetric [4 + 2] annulations with diverse activated alkenes or isatins, generally with high regio-, diastereo- and enantio-selectivity, and even switchable regiodivergent or diastereodivergent annulations can be well realised by tuning the substrate or catalyst assemblies. An array of control experiments, including UV/Vis absorption study and density functional theory calculations, are conducted to rationalise this new double activation mode combining a palladium complex and an ammonium halide as an ion-pair catalyst.

A double activation catalytic system combining a palladium complex and an ammonium halide was developed to promote the asymmetric [4 + 2] annulations of Morita–Baylis–Hillman carbonates of methyl vinyl ketone.  相似文献   

13.
Introduction of heteroatom-based substituents into 1,4-dihydropyridines by means of a halogen-mediated, oxidative protocol: diamination, sulfonylation, sulfinylation, bis-sulfanylation, and halo-phosphonylation processes     
Lavilla R  Kumar R  Coll O  Masdeu C  Spada A  Bosch J  Espinosa E  Molins E 《Chemistry (Weinheim an der Bergstrasse, Germany)》2000,6(10):1763-1772
The natural tendency of 1,4-dihydropyridines to undergo "biomimetic" oxidation to afford pyridinium salts can be switched off and, through the use of reagents that interact electrophilically with the enamine moiety present in the heterocyclic system, it is possible to promote alternative oxidations. In this way, efficient regio- and stereocontrolled synthetic methods have been developed that lead to diversely substituted di- and tetrahydropyridines. These include iodoazidation, diamination, bis-sulfonamidation, sulfonylation, sulfinylation, thiocyanation, sulfanylation, bis-sulfanylation, and halo-phosphonylation processes.  相似文献   

14.
Nickel/Brønsted acid dual-catalyzed regio- and enantioselective hydrophosphinylation of 1,3-dienes: access to chiral allylic phosphine oxides     
Jiao Long  Yuqiang Li  Weining Zhao  Guoyin Yin 《Chemical science》2022,13(5):1390
While chiral allylic organophosphorus compounds are widely utilized in asymmetric catalysis and for accessing bioactive molecules, their synthetic methods are still very limited. We report the development of asymmetric nickel/Brønsted acid dual-catalyzed hydrophosphinylation of 1,3-dienes with phosphine oxides. This reaction is characterized by an inexpensive chiral catalyst, broad substrate scope, and high regio- and enantioselectivity. This study allows the construction of chiral allylic phosphine oxides in a highly economic and efficient manner. Preliminary mechanistic investigations suggest that the 1,3-diene insertion into the chiral Ni–H species is a highly regioselective process and the formation of the chiral C–P bond is an irreversible step.

Asymmetric hydrophosphinylation of 1,3-dienes with phosphine oxides using an inexpensive chiral catalyst has been demonstrated, providing access to chiral allylic phosphine oxides with broad substrate scope and high regio- and enantioselectivity.  相似文献   

15.
Solvent directed chemically divergent synthesis of β-lactams and α-amino acid derivatives with chiral isothiourea     
Dong-Sheng Ji  Hui Liang  Kai-Xuan Yang  Zhi-Tao Feng  Yong-Chun Luo  Guo-Qiang Xu  Yucheng Gu  Peng-Fei Xu 《Chemical science》2022,13(6):1801
A protocol for the chemically divergent synthesis of β-lactams and α-amino acid derivatives with isothiourea (ITU) catalysis by switching solvents was developed. The stereospecific Mannich reaction occurring between imine and C(1)-ammonium enolate generated zwitterionic intermediates, which underwent intramolecular lactamization and afforded β-lactam derivatives when DCM and CH3CN were used as solvents. However, when EtOH was used as the solvent, the intermediates underwent an intermolecular esterification reaction, and α-amino acid derivatives were produced. Detailed mechanistic experiments were conducted to prove that these two kinds of products came from the same intermediates. Furthermore, chemically diversified transformations of β-lactam and α-amino acid derivatives were achieved.

A protocol for the solvent directed chemically divergent synthesis of β-lactam and α-amino acid derivatives with chiral isothiourea was reported.  相似文献   

16.
Light-driven reduction of aromatic olefins in aqueous media catalysed by aminopyridine cobalt complexes     
Carla Casadevall  David Pascual  Jordi Aragn  Arnau Call  Alicia Casitas  Irene Casademont-Reig  Julio Lloret-Fillol 《Chemical science》2022,13(15):4270
A catalytic system based on earth-abundant elements that efficiently hydrogenates aryl olefins using visible light as the driving-force and H2O as the sole hydrogen atom source is reported. The catalytic system involves a robust and well-defined aminopyridine cobalt complex and a heteroleptic Cu photoredox catalyst. The system shows the reduction of styrene in aqueous media with a remarkable selectivity (>20 000) versus water reduction (WR). Reactivity and mechanistic studies support the formation of a [Co–H] intermediate, which reacts with the olefin via a hydrogen atom transfer (HAT). Synthetically useful deuterium-labelled compounds can be straightforwardly obtained by replacing H2O with D2O. Moreover, the dual photocatalytic system and the photocatalytic conditions can be rationally designed to tune the selectivity for aryl olefin vs. aryl ketone reduction; not only by changing the structural and electronic properties of the cobalt catalysts, but also by modifying the reduction properties of the photoredox catalyst.

A dual catalytic system based on earth-abundant elements reduces aryl olefins to alkanes in aqueous media under visible light. Mechanistic studies allow for rational tunning of the system for the selective reduction of aryl olefins vs ketones and vice versa.  相似文献   

17.
A novel approach to functionalized (E)-1,4-diaryl-1-butenes by Heck reaction and their applications for the construction of dibenzylbutyrolactone lignan skeletons by radical cyclization     
Rekha Singh 《Tetrahedron letters》2005,46(28):4719-4722
A highly regio- and stereoselective [Pd(allyl)Cl]2 catalyzed Heck reaction of aryl iodides and electronically neutral terminal olefins generated in situ by fluoride induced protiodesilylation of alkenylsilanol derivatives under mild conditions has been developed. The products, viz. terminally substituted styrenes and (E)-1,4-diaryl-1-butenes were obtained in very good yields. The dibenzylbutyrolactone lignan skeletons have been prepared employing two regio- and stereoselective Bu3SnH-mediated radical cyclization routes.  相似文献   

18.
Regio- and stereoselectivity of the reactions of pyridinium and picolinium salts and ylids with nitriles containing a nucleofugic group     
A. M. Shestopalov  V. P. Litvinov  Yu. A. Sharanin  I. A. Aitov  L. A. Rodinovskaya 《Russian Chemical Bulletin》1991,40(4):835-838
The reaction of pyridinium, 2-picolinium, and 3-picolinium ylids with tetracyanoethylene, ethoxymethylenemalononitrile, and (dimethylthio)methyl-enecyanamide in the presence of organic bases proceeds regio- and stereo-selectively to give substituted pyridinium 1,4-ylids, 2-(2-propen-1-ylidene)-1,2-dihydropyridines, and 5-(1-pyridino)pyrimidin-4-olates, respectively. The reaction of 1-phenacylpyridinium bromide and tetracyanoethylene leads to 1-phenacylpyridinium tricyanoethenolate.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 4, pp. 939–942 April, 1990.  相似文献   

19.
Catalytic enantioselective synthesis of benzocyclobutenols and cyclobutanols via a sequential reduction/C–H functionalization     
Jun Chen  Zhan Shi  Chunyu Li  Ping Lu 《Chemical science》2021,12(31):10598
We report here a sequential enantioselective reduction/C–H functionalization to install contiguous stereogenic carbon centers of benzocyclobutenols and cyclobutanols. This strategy features a practical enantioselective reduction of a ketone and a diastereospecific iridium-catalyzed C–H silylation. Further transformations have been explored, including controllable regioselective ring-opening reactions. In addition, this strategy has been utilized for the synthesis of three natural products, phyllostoxin (proposed structure), grandisol and fragranol.

We report here a sequential enantioselective reduction/C–H functionalization to install contiguous stereogenic carbon centers of benzocyclobutenols and cyclobutanols.

Molecules with inherent ring strain have gained considerable interest in the synthetic community.1 Among them, four-membered ring molecules have been recognized as powerful building blocks in organic synthesis.2 Driven by ring strain releasing, the reactions of carbon–carbon bond cleavage have been extensively studied in recent years.3 Meanwhile, cyclobutane motifs represent important structural units in natural product and bioactive molecules as well (Scheme 1).4 Therefore, a general and robust method to constitute four-membered ring derivatives is of great value, especially in an enantiomerically pure form.5Open in a separate windowScheme 1Representative cyclobutane-containing bioactive molecules.[2 + 2]-Cycloaddition6 and the skeleton rearrangement reaction7 are two primary methods to prepare chiral cyclobutane derivatives. Recently, the precision modification of four-membered ring skeletons to access enantioenriched cyclobutane derivatives has attracted emerging attention. Several strategies have been developed, including allylic alkylation,8 α-functionalization,9 conjugate addition10 and C–H functionalization11 of prochiral or racemic cyclobutane derivatives (Scheme 2a).12 However, the enantioselective synthesis of chiral benzocyclobutene derivatives is still underdeveloped.13 Although two efficient palladium-catalyzed C–H activation strategies have been developed by Baudoin14 and Martin15 groups via similar intermediate five-membered palladacycles, no enantioenriched benzocyclobutene derivative has been prepared by employing the above two methods. In 2017, Kawabata reported an elegant example of asymmetric intermolecular α-arylation of enantioenriched amino acid derivatives to afford benzocyclobutenones with tetrasubstituted carbon via memory of chirality (Scheme 2b).16 In 2018, Zhang reported an iridium-catalyzed asymmetric hydrogenation of α-alkylidene benzocyclobutenones in good enantioselectivities (3 examples, 83–88% ee).12c To the best of our knowledge, there is no report on enantioselective synthesis of benzocyclobutene derivatives with all-carbon quaternary centers.Open in a separate windowScheme 2Asymmetric synthesis of cyclobutanes and their derivatives. (a) Enantioselective functionalization of four-membered ring substrates. (b) Synthesis of chiral benzocyclobutenone via memory of chirality. (c) This work: sequential enantioselective reduction/C–H functionalization.In line with our continued interest in precision modification of four-membered ring skeletons,9d,10c,12a we initiated our studies on the synthesis of chiral benzocyclobutenes via enantioselective functionalization of highly strained benzocyclobutenones. It is well known that benzocyclobutene derivatives are labile to undergo a ring-opening reaction to release their inherent ring strains.17 Therefore, it is a challenging task to modify the benzocyclobutenone and preserve the four-membered ring skeleton at the same time. We envisioned that a carbonyl group directed C–H functionalization18 of the gem-dimethyl group could furnish enantioenriched α-quaternary benzocyclobutenones (Scheme 2c). This could be viewed as an alternative approach to achieve the alkylation of benzocyclobutenone, which was otherwise directly inaccessible using enolate chemistry through the unstable anti-aromatic intermediate.19 In addition, a highly regioselective C–H activation would be required to functionalize the methyl group instead of the aryl ring. Here we report our work on sequential enantioselective reduction and intramolecular C–H silylation to provide enantioenriched benzocyclobutenols and cyclobutanols with all-carbon quaternary centers. The excellent diastereoselectivity and regioselectivity of silylation were attributed to rigid structural organization of the 4/5 fused ring. Furthermore, this strategy has been utilized to accomplish the total synthesis of natural products phyllostoxin (proposed structure), grandisol and fragranol.We commenced our studies with enantioselective reduction of readily prepared dimethylbenzocyclobutenone 1a (Scheme 3).15,20 Surprisingly, enantioselective reduction of the carbonyl group of cyclobutanone derivatives received little attention. The first reduction of parent benzocyclobutenone was studied in 1996 by Kündig using chlorodiisopinocamphenylborane21 or chiral oxazaborolidines (CBS reduction),22 and only moderate enantioselectivity (44–68% ee) was obtained.23 Although copper-catalyzed asymmetric hydrosilylation of benzocyclobutenone 1a using CuCl/(R)-BINAP gave the benzocyclobutenol ent-2a in 88% ee, optimization of ligands gave no further improvement (Scheme 3a, see Tables S1–S4 for details).24 Gladly, excellent enantioselective reduction could be achieved in 94% yield and 97% ee under Noyori''s asymmetric transfer hydrogenation conditions (Scheme 3b, conditions A, RuCl[(S,S)-Tsdpen](p-cymene)).25 The product 2a showed remarkable stability and no ring-opening byproduct 2a′ was observed. The reduction of parent benzocyclobutenone was examined under conditions A, and benzocyclobutenol was obtained in 90% yield and 81% ee. Apparently, the steric influence imposed by the α-dimethyl group enhanced the enantioselectivity of the reduction. Similarly, the CBS reduction ((S)-B–Me) of benzocyclobutenone 1a gave better results compared with parent benzocyclobutenone, affording the product 2a in 86% yield and 92% ee (Scheme 3c).Open in a separate windowScheme 3Enantioselective reduction of benzocyclobutenone 1a. (a) Copper hydride reduction. (b) Ru-catalyzed asymmetric transfer hydrogenation. (c) CBS reduction.We then examined the substrate scope of the reduction reaction (26 was chosen to improve the yield and enantioselectivity. Besides, benzocyclobutenol 2g with nitro substitution could be obtained in 96% yield and 93% ee. Treatment of pyrrolidinyl substituted benzocycobutenone 1h with catalyst (S,S)-Ts-DENEB afforded desired product 2h in 49% yield and 89% ee, together with ring-opening product 2h′ (18%).Enantioselective reduction of benzocyclobutenonesa
Open in a separate windowaConditions A: 1a (0.5–2.0 mmol), RuCl[(S,S)-Tsdpen](p-cymene) (1–2 mol%), HCOOH/Et3N (5/2), rt. All results are corrected to the (S)-catalyst. The ee values were determined by HPLC analysis; see the ESI for more details.b(S,S)-Ts-DENEB (1–2 mol%) was used, rt or 60 °C.3,3-Disubstituted cyclobutanones were also explored (l-selectride gave cis-4i as a single product in 99% yield and 96% ee. The reaction of 3j gave similar results, and enantioenriched cyclobutanols cis-4j could be furnished in 78% yield and 97% ee from ent-trans-4j (98% ee) following the above oxidation–reduction procedure. The absolute configurations of 2a, ent-2j and trans-4i were unambiguously determined by single-crystal X-ray diffraction analysis of their corresponding nitrobenzoate derivative.27Enantioselective reduction of cyclobutanones 3a
Open in a separate windowaConditions B: 3a (1.0–5.0 mmol), (S)-B–Me (10 mol%), BH3·Me2S (0.6 equiv.), THF, rt.b(S)-B–Me (20 mol%), BH3·Me2S (1.0 equiv.).c(−)-Ipc2BCl (1.2 equiv.), THF, −20 °C. (−)-Ipc2BCl = (−)-diisopinocampheylchloroborane.Inspired by powerful and reliable directed C–H silylation chemistry pioneered by Hartwig,28 we envisioned that the transition-metal catalyzed intramolecular C–H silylations of the above alcohols would provide a single diastereomer owing to rigid structural organization. The challenges here are the control of regioselectivity in the cyclization step and inhibition of the ring-opening pathway. Benzocyclobutenol 2a was chosen as a model substrate to study this intramolecular C–H silylation. The transition-metal catalyst system and alkene acceptors were screened (Scheme 4, see Tables S5–S9 for details). Acceptor norbornene (nbe) derivative A gave the optimal yield in the cyclization step (63% NMR yield), and other phenanthroline ligands gave inferior results. The reaction showed remarkable regio- and diastereoselectivity; no silylation of the arene was detected.With optimal intramolecular silylation conditions in hand, sequential hydroxysilylation/C–H silylation/phenyllithium addition reaction of 2a provided desired product 5a in 56% overall yield without any obvious erosion of enantiomeric purity ( Open in a separate windowaConditions C: i. 2a (0.5 mmol), [Ir(COD)OMe]2 (0.05 mol%), Et2SiH2 (1.2 equiv.), THF, 30 °C; ii. [Ir(COD)Cl]2 (2.5 mol%), Me4Phen (6 mol%), A (1.0 equiv.), THF, 100 °C; iii. PhLi, THF, −78 °C; see the ESI for more details.biii. KHCO3 (2.5 equiv.), H2O2 (10 equiv.), THF/MeOH (1 : 1), 50 °C.Open in a separate windowScheme 4Optimization of intramolecular C–H silylation of benzocyclobutenol 1a.Cyclobutanols were examined under optimal conditions as well (27 The diols cis-6h′ and trans-6h′ could be achieved upon treatment of cyclization products with H2O2 instead of phenyllithium. In addition, bicyclic substrates 4i, 4j smoothly furnished the corresponding enantioenriched products cis-6i, 6j and trans-6i, 6j with four contiguous carbon centers in good yields.Stereospecific C–H functionalization of cyclobutanols 4a
Open in a separate windowaReaction conditions: 4 (0.5 mmol), Ru(PPh3)3Cl (0.2 mol%), Et2SiH2 (1.5 equiv.), THF, 35 °C; ii. [Ir(COD)Cl]2 (2.5 mol%), Me4Phen (6 mol%), A (1.0 equiv.), THF, 100 °C; iii. PhLi, THF, −78 °C; see the ESI for more details.bii. [Ir(COD)Cl]2 (5 mol%), Me4Phen (12 mol%).ciii. KHCO3 (2.5 equiv.), KF (2.5 equiv.), H2O2 (10 equiv.), THF/MeOH (1 : 1), 50 °C.dent-cis-4i (70% ee) was used.eent-trans-4i (97% ee) was used.At this point, we conducted further transformations to explore the utilities of the chiral benzocyclobutene derivatives (Scheme 5). The oxidation of benzocyclobutenol 5a afforded benzocyclobutenone 7 smoothly using Dess–Martin periodinane. This product could be viewed as the result of the alkylation of α-substituted benzocyclobutenone via elusive enolate intermediate I.Open in a separate windowScheme 5Further transformations of benzocyclobutenol 5a.Subsequent Tamao–Fleming oxidation29 with a concomitant cyclobutanone oxidation provided alcohol 8 in 57% yield, albeit with partial loss of enantiopurity. Furthermore, the regioselective Bayer–Villiger oxidation of 7 was achieved using MMPP,30 giving phthalide 9 in 63% yield and 97% ee. Poor regioselectivity was observed when parent benzocyclobutenone was treated with a base.31 In contrast, exposure of 7 to sodium methoxide afforded phenylacetic acid derivative 10 as a single product in 94% yield and 97% ee via proximal bond cleavage.Phyllostoxin (11) was isolated from fungal pathogen Phyllosticta cirsii, and it could represent a potential natural herbicide (Scheme 6).32 The structure was proposed to contain chiral α-quaternary benzocyclobutenone moiety. We envisioned that our strategy would provide a straightforward way to assemble the quaternary center of benzocyclobutenone, thereby confirming the proposed structure and determining the absolute configuration. Our synthesis commenced with enantioselective transfer hydrogenation of substrate 1o. Enantioenriched benzocyclobutenol 2o could be obtained in 93% yield and 99% ee using catalyst (R,R)-Ts-DENEB. Standard procedure, including hydrosilylation/C–H silylation/oxidation, provided diol 5o′ in 89% overall yield and 99% ee. Various oxidation conditions were examined to oxidize diol 5o′, including Swern oxidation, Dess–Martin periodinane and PCC; unfortunately, the reaction only gave messy mixtures. Thus we turned to selective protection of the diol. Selective benzoylation could be achieved via three-step manipulation, giving primary alcohol 12 in 82% overall yield. Swern oxidation and nucleophilic addition of EtMgBr, followed by global deprotection, provided triol 13 in 54% yield over 3 steps. Of mention, benzoyl migration was observed in the EtMgBr addition step. Finally, selective acylation of the phenol and subsequent oxidation furnished benzocyclobutenone 11 in 39% overall yield. However, the optical rotation and NMR spectral data did not match those reported for the natural product.Open in a separate windowScheme 6Total synthesis of the proposed structure of phyllostoxin. Conditions: [Ir(COD)OMe]2, Et2SiH2, THF, rt; ii. [Ir(COD)Cl]2, Me4Phen, A, THF, 100 °C; iii. KHCO3, H2O2, THF/MeOH (1 : 1), 50 °C.The monoterpene grandisol (14) was known as a main component of the sex pheromone of the cotton boll weevil, Anthonomous grandis Boheman, and other insects.33,34 The diastereomer fragranol (15) was isolated in many essential oil aerial parts of plant species such as Achillea fragrantissima, A. falcata and Geranium tuberosum.33 Surprisingly, in comparison to grandisol, there is only one report on enantioselective synthesis of fragranol yet.35 We postulated that our strategy would enable a divergent synthesis of these two diastereomers, starting from an optical resolution of cyclobutanone 3k (Scheme 7). As expected, the CBS reduction of 1x provided cyclobutanols cis-4k and trans-4k (90% yield, 1 : 1.1 dr, 90–99% ee). Subsequent C–H functionalization and oxidation gave diastereomers cis-6k′ and trans-6k′ in good yield. And both diastereomers could be easily separated by column chromatography. Debenzylation, selective silylation of the primary alcohol and Barton–McCombie deoxygenation provided cyclobutanes 17 and 20 uneventfully. Starting from cyclobutane 17, deprotection and subsequent oxidation afforded lactone 18 in 56% overall yield, which led to formal total synthesis of (−)-grandisol 14. Starting from cyclobutane 20, regioselective dehydration with Martin sulfurane and removal of the TBS group furnished alkene 21 in 70% overall yield. Finally, (−)-fragranol 15 was obtained in three additional steps, which included oxidation to an aldehyde, olefination/hydrolysis and reduction.Open in a separate windowScheme 7Divergent synthesis of grandisol and fragranol. Conditions: Ru(PPh3)3Cl, Et2SiH2, THF, 35 °C; ii. [Ir(COD)Cl]2, Me4Phen, A, THF, 100 °C; iii. KHCO3, H2O2, THF/MeOH (1 : 1), 50 °C.  相似文献   

20.
Bromination of 2,6-dimethyl-3,5-dimethoxycarbonyl-4-(2′-difluoromethoxyphenyl)-1,4-dihydropyridine (foridone)     
I. P. Skarstin'sh  V. V. Kastron  G. Ya. Dubur  I. B. Mazheika  É. É. Liepin'sh 《Chemistry of Heterocyclic Compounds》1989,25(7):791-795
The action of mild brominating agents (pyridinium bromide-perbromide, dioxan dibromide N-bromosuccinimide, and N-bromoacetamide) on 2,6-dimethyl-3,5-dimethoxy-carbonyl-4-(2-difluoromethoxyphenyl)-1,4-dihydropyridine (Foridone) has been studied. Depending on reaction conditions, furo-, difuro-, and dibromomethyl-1,4-dihydropyridines are obtained together with certain oxidized forms.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 7, pp. 948–952, July, 1989.  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