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 78 毫秒
1.
A protocol for stereoselective C-radical addition to a chiral glyoxylate-derived N-sulfinyl imine was developed through visible light-promoted photoredox catalysis, providing a convenient method for the synthesis of unnatural α-amino acids. The developed protocol allows the use of ubiquitous carboxylic acids as radical precursors without prior derivatization. The protocol utilizes near-stoichiometric amounts of the imine and the acid radical precursor in combination with a catalytic amount of an organic acridinium-based photocatalyst. Alternative mechanisms for the developed transformation are discussed and corroborated by experimental and computational studies.

A protocol for stereoselective C-radical addition to a chiral glyoxylate-derived N-sulfinyl imine was developed through visible light-promoted photoredox catalysis, providing a convenient method for the synthesis of unnatural α-amino acids.  相似文献   

2.
Catalytic enantioselective [2,3]-rearrangements of in situ generated ammonium ylides from glycine pyrazoleamides and allyl bromides were achieved by employing a chiral N,N′-dioxide/MgII complex as the catalyst. This protocol provided a facile and efficient synthesis route to a series of anti-α-amino acid derivatives in good yields with high stereoselectivities. Moreover, a possible catalytic cycle was proposed to illustrate the reaction process and the origin of stereoselectivity.

The Lewis acid catalyzed asymmetric [2,3]-rearrangement of quaternary ammonium ylides formed in situ from glycine pyrazoleamides and allyl bromides.  相似文献   

3.
In this paper, we describe a small library of easy-to-prepare chiral (cyclopentadienone)iron pre-catalysts for enantioselective CO and CN hydrogenations. Starting from readily accessible achiral materials, six chiral (cyclopentadienone)iron complexes (1a-f) possessing a stereogenic plane were synthesized in racemic form. Based on the screening of pre-catalysts (±)-1a-f in the hydrogenation of ketones and ketimines, we selected two complexes (1a and 1d) for resolution by semipreparative enantioselective HPLC. The absolute configuration of the separated enantiomers of 1a and 1d was assigned by XRD analysis (1a) and by comparison between experimental and DFT-calculated ECD and ORD spectra (1d). The enantiopure pre-catalysts (S)-1a and (R)-1d were tested in the asymmetric hydrogenation of several ketones and ketimines and showed good activity and modest enantioselectivity, the e.e. values ranging from very low to moderate (54%).  相似文献   

4.
Stilbene derivatives have long been known to undergo “acid-catalyzed” ZE isomerization, where a strong mineral acid at high concentration is practically necessary. Such severe reaction conditions often cause undesired by-reactions and limit their potential application. Herein, we present a trace mild acid-catalyzed ZE isomerization found with stilbene derivatives fused with a norbornene moiety. By-reactions, such as the migration of the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C double bond and electrophilic addition reactions, were completely inhibited because of the ring strain caused by the fused norbornene component. Direct photolysis of the E isomers at selected wavelengths led to the EZ photoisomerization of these stilbene derivatives and thus constituted a unique class of molecular switches orthogonally controllable by light and acid. The catalytic amount of acid could be readily removed, and the ZE isomerization could be controlled by turning on/off the irradiation of a photoacid, which allowed repeated isomerization in a non-invasive manner. Moreover, the Z isomer produced by photoisomerization could spontaneously self-recover to the E isomer in the presence of a catalytic amount of acid. The kinetics of ZE isomerization were adjustable by manipulating catalytic factors and, therefore, unprecedented molecular photoswitches with adjustable self-recovery were realized.

Quantitative ZE isomerization was catalyzed by trace mild acids to offer molecular switches orthogonally controllable by acid and light.  相似文献   

5.
《中国化学快报》2023,34(1):107569
We report herein an I2/PhI(OAc)2 catalytic system for the pragmatic construction of CN bonds through CH/NH oxidative coupling protocol. Divergent pyrrolo[2,3-b]indoles were efficiently prepared via I2-catalyzed intramolecular C–H amination reactions from (E/Z)-2-indolylenamines under metal-free conditions. Various functional groups are tolerated under mild reaction conditions and the resulting pyrrolo[2,3-b]indoles were obtained with mostly good to excellent yields. It was interesting to observe that both the (E)- and (Z)-isomers of the starting materials were efficiently transformed into the targeted product. The I+-mediated catalytic cycle was proposed based on mechanistic studies for this reaction.  相似文献   

6.
Stereoselective Mannich reactions of aldehydes with ketimines provide chiral β-amino aldehydes that bear an α-tert-amine moiety. However, the structural variation of the ketimines is limited due to the formation of inseparable E/Z isomers, low reactivity, and other synthetic difficulties. In this study, a highly diastereodivergent synthesis of hitherto difficult-to-access β-amino aldehydes that bear a chiral α-tert-amine moiety was achieved using the amine-catalyzed Mannich reactions of aldehydes with less-activated Z-ketimines that bear both alkyl and alkynyl groups.

Stereoselective Mannich reactions of aldehydes with ketimines provide chiral β-amino aldehydes that bear an α-tert-amine moiety.  相似文献   

7.
Cα-Tetrasubstituted α-amino acids are ubiquitous and unique structural units in bioactive natural products and pharmaceutical compounds. The asymmetric synthesis of these molecules has attracted a lot of attention, but a more efficient method is still greatly desired. Here we describe the first sequential four-step acylation reaction for the efficient synthesis of chiral Cα-tetrasubstituted α-amino acid derivatives from simple N-acylated amino acids via an auto-tandem catalysis using a single nucleophilic catalyst. The synthetic efficiency is improved via a direct enantioselective C-acylation; the methodology affords the corresponding Cα-tetrasubstituted α-amino acid derivatives with excellent enantioselectivities (up to 99% ee). This step-economic, one-pot, and auto-tandem strategy provides facile access to important chiral building blocks, such as peptides, serines, and oxazolines, which are often used in medicinal and synthetic chemistry.

The first four-step sequential reaction for the synthesis of Cα-tetrasubstituted chiral α-amino acid derivatives via auto-tandem catalysis has been developed.  相似文献   

8.
A range of chemical reactions occurring on the surfaces of metal nanoparticles exhibit enhanced rates under plasmonic excitation. It is not straightforward to distinguish between photochemical and photothermal effect using Arrhenius fitting of the reaction rates alone.

In the recently published article: “Thermal effects – an alternative mechanism for plasmon-assisted photocatalysis”, Dubi et al.1 argue that the results of multiple works on plasmon-excited-induced bond dissociation reactions can be explained by a purely photothermal enhancement of the reaction rates and that no non-thermal effects are required to explain the enhanced rates resulting from plasmonic excitation. Their argument rests on a reproduction of the reaction rate data by an Arrhenius expression with a light-intensity-dependent local temperature at the surface of the nanoparticles.Dubi et al.‘s straightforward analysis may have general appeal for explaining rate enhancements in bond dissociation reactions observed under plasmonic excitation of metal nanostructures without invoking hot electron contributions. But there is one caveat that deserves recognition when undertaking such an analysis. As shown below, under certain common scenarios, it is practically impossible to distinguish between a photochemical (non-thermal) effect of light excitation and a purely photothermal one using a phenomenological Arrhenius fitting of the data alone.As per the Arrhenius equation, the rate of a reaction depends on the set temperature Ts as:1where R0 is a constant for a given reaction and reaction conditions and Ea is the apparent activation energy barrier for the reaction. As an aside, one should note that unlike the Eyring equation, which is preferred for non-gas-phase reaction kinetics involving a vibrational reaction co-ordinate, the pre-exponential factor in the Arrhenius equation is assumed to have a negligible temperature dependence.A photochemical explanation of plasmon-enhanced catalysis is that the apparent activation energy Ea is lower under plasmonic excitation as compared to its value, Edarka, in the dark. Thus, as per eqn (1), at a fixed temperature Ts, R will be higher under light excitation. In fact, the measured apparent activation barrier has been found to be dependent on the light intensity I. For the sake of the following argument, let us assume that the decrease in Ea is linearly dependent on the light intensity:Ea = EdarkaBI2where B is a proportionality constant with units of eV cm2 W−1 when Ea is expressed in units of eV and I in units of W cm−2. Note that B is expected to be wavelength-dependent. Eqn (2) can be written alternatively as:Ea = Edarka(1 − bI)3where b is simply B/Edarka and has units of cm2 W−1. From eqn (1) and (3):4Using a Taylor''s expansion around I = 0 (dark condition),5For the light-intensity regime (I ≪ 1/b), the higher order terms can be neglected, so one gets from eqn (4) and (5):6Thus, if one simply uses an Arrhenius analysis of the reaction rate, the reaction appears to be carried out at a hypothetical temperature that is higher than the actual temperature Ts by an amount proportional to the light intensity I:Tdummy = Ts(1 + bI)7where this hypothetical temperature is referred to as Tdummy. Eqn (7) is equivalently expressed as:Tdummy = Ts + aI8where a = bTs is the photothermal conversion coefficient with units of K cm2 W−1. Eqn (8) is identical to the expression used by Dubi et al. in their argument in favor of a purely photothermal effect. In other words, it would appear as if plasmonic excitation led to an increase in the temperature, but led to no change in the apparent activation barrier. Effectively, in a phenomenological Arrhenius analysis, the photochemical (non-thermal) effect of plasmonic excitation on the reaction is simply masked as a temperature increase.Thus, as shown in Fig. 1, an Arrhenius analysis with a as an adjustable fit parameter may be futile for practically distinguishing the photochemical action of plasmonic excitation, (i.e., a rate enhancement caused by a decrease in the activation barrier) from a purely photothermal effect (i.e., a rate enhancement caused by an increase in the surface temperature). Under such a scenario, for distinguishing these effects, it is necessary to have precise knowledge and/or control over the temperature at the surface of the nanoparticles, as correctly argued by Dubi et al.,1 but also acknowledged by practitioners2–4 in the field. It is well appreciated that the localized inhomogeneous nature of photothermal heating results in a temperature gradient extending out from the surface of the nanoparticles to the bulk of the medium. These gradients are small in magnitude under conditions where the heat dissipation rate can keep up with the energy deposition rate. However, in systems where heat transfer rates are limiting, significant non-uniformities in temperature and thermal bottlenecks can arise. Such cases necessitate spatially precise temperature-probing localized to the nanoparticle surface.Open in a separate windowFig. 1The reaction rate under plasmonic excitation, R, relative to that in the dark, Rdark, is plotted as a function of light intensity for (i) the photochemical case (red dots), where the activation barrier is decreased by plasmonic excitation (eqn (1) and (2) with B = 0.1 eV cm2 W−1) while the temperature is kept fixed and (ii) the purely photothermal model (black line), where the temperature is increased by plasmonic excitation (eqn (1) and (8)) with a = 54 K cm2 W−1) but the activation barrier remains unchanged. In both cases, Edarka = 1.21 eV and Ts = 600 K. The two models yield trends that are practically indistinguishable.  相似文献   

9.
An efficient cooperative chiral Lewis acid/photoredox catalytic system for engaging highly reactive radicals in highly enantioselective conjugate addition to α,β-unsaturated carbonyls is highly desirable. Direct photoexcitation of unbound substrates typically induces undesired background pathways for racemic products and remains a formidable challenge to be addressed in the area of enantioselective photocatalysis. Herein, we report a cooperative catalytic system comprising a chiral Cu(i) complex and an Ir(iii) photocatalyst fueled by visible-light irradiation that allows for seamless integration of the catalytic formation of α-amino alkyl radicals and subsequent enantioselective addition to α,β-unsaturated amides. A 7-aza-6-MeO-indoline attachment on the amide substrates plays a pivotal role in suppressing the undesired pathways, resulting in excellent enantioselectivity and enabling expedited access to valuable γ-aminobutyramides. The indoline amide was readily diversified with full recovery of the azaindoline attachment, highlighting the synthetic utility of this cooperative catalytic system.

An efficient cooperative chiral Lewis acid and photoredox catalytic system towards the highly enantioselective radical conjugate addition of α-amino radicals to α,β-unsaturated amides is developed with the implementation of unique auxiliaries.  相似文献   

10.
A new type of amido-oxazoline ligands was conveniently synthesized from inexpensive and commercially available materials in high yields and enantiomeric excesses. The corresponding chiral copper complexes with this class of ligands [C2 symmetric S,S-bis(amido-oxazoline-Cu(II) complex] were synthesized accordingly. The ORTEP diagram of ligand 6a and complex 6a-copper were compared and characterization of the complex confirmed the involvement of both dentate parts of the ligands, the oxygen and nitrogen atoms, in complexation with copper. The utilization of this amido-oxazoline ligands in the copper-catalyzed enantioselective esterification of allylic CH bonds of cyclic olefins with tert-butyl-4-nitrobenzoperoxoate resulted in the highest activities, yields (up to 95%) and enantioselectivities (up to 96%) in the presence of HZSM-5 zeolite. These new findings highlight the protocol as one of the most attractive and useful methods for the oxidation of the asymmetric allylic CH bond of cycloalkenes compared to other methodologies reported in the literature.  相似文献   

11.
The synthesis of diverse N-fused heterocycles, including the pyrido[1,2-a]indole scaffold, using an efficient pyrone remodeling strategy is described. The pyrido[1,2-a]indole core was demonstrated to be a versatile scaffold that can be site-selectively functionalized. The utility of this novel annulation strategy was showcased in a concise formal synthesis of three fascaplysin congeners.

The synthesis of diverse N-fused heterocycles, including the pyrido[1,2-a]indole scaffold, using an efficient pyrone remodeling strategy is described.  相似文献   

12.
Pyrido[1,2-a]-1H-indoles are important scaffolds found in many biologically active compounds. Herein, we first developed an IPrAuCl/AgSbF6-catalyzed cycloisomerization of N-1,3-disubstituted allenyl indoles affording pyrido[1,2-a]-1H-indoles. Then the axial-to-central chirality transfer starting from enantio-enriched N-1,3-disubstituted allenylindoles affording optically active pyrido[1,2-a]-1H-indoles has been realized in excellent yields and enantioselectivities. A mechanism has been proposed based on mechanistic studies. Synthetic applications have also been demonstrated.

We reported an IPrAuCl/AgSbF6-catalyzed cycloisomerization of enantio-enriched N-1,3-disubstituted allenylindoles affording optically active pyrido[1,2-a]-1H-indoles in excellent yields and enantioselectivities.  相似文献   

13.
We describe the synergistic utilization of titanocene/photoredox dual catalysis driven by visible light for the radical opening/spirocyclization of easily accessible epoxyalkynes. This environmentally benign process uses the organic donor–acceptor fluorophore 2,4,5,6-tetra(9H-carbazol-9-yl)isophthalonitrile (4CzIPN) as a photocatalyst and Hantzsch ester (HE) as an electron donor instead of stoichiometric metallic reductants. The photocatalytic conditions showed exceptionally high reactivity for the synthesis of privileged and synthetically challenging spirocycles featuring a spiro all-carbon quaternary stereocenter. Cyclic voltammetry (CV) studies suggest that Cp2TiIIICl is the catalytically active species.

We describe the synergistic utilization of titanocene/photoredox dual catalysis driven by visible light for radical opening/spirocyclization of easily accessible epoxyalkynes.

Over the last few decades, radical-based transformations have been increasingly used in organic synthesis due to their salient features, such as ease of generation, mild reaction conditions, and broad functional group compatibility.1,2 As a mild single-electron-transfer (SET) reagent, titanocene monochloride (Cp2TiIIICl) is considered a formidable tool in contemporary radical chemistry due to its ability to promote various fundamental radical-based transformations.3–7 Cp2TiIIICl was first introduced by Nugent and RajanBabu as a very mild stoichiometric reagent for the reductive opening of epoxides.8–11 Later, the catalytic conditions developed by Gansäuer et al. (Scheme 1a)12 employing stoichiometric amounts of active metals in combination with 2,4,6-collidine·HCl further expanded its applications and led to the discovery of a number of novel transformations.13–16 The key to success was the formation of a stable complex A in reactions while decreasing the concentration of active Cp2TiIIICl.17,18 We were interested in the radical opening/cyclization reaction of epoxides which has attracted considerable attention from the synthetic community and has been used numerous times in the synthesis of natural products.19,20 Nevertheless, this reaction required stoichiometric metallic reductants and proceeded slowly particularly with sterically hindered substrates even with high catalyst loading.21 Therefore, the development of an eco-friendly and efficient catalytic system with an expanded substrate scope is highly desirable.Open in a separate windowScheme 1Cp2TiIIICl mediated radical opening/spirocyclization of epoxides; (a) generation of TiIIIvia a metal reduction approach; (b) dual titanocene/photoredox catalysis; (c) examples of drugs and natural products containing heterospirocycles.In recent years metallaphotoredox catalysis has been a new and rapidly growing research subject.22–29 Photoredox processes can directly modulate the oxidation state of metals by electron transfer (ET).30–33 Given that the generation of TiIII is a SET process, we envisioned that the reduction could be facilitated by a photoredox-controlled process while overcoming the aforementioned limitations. On the other hand, spirocycles bearing a chiral spiro all-carbon quaternary carbon are particularly attractive synthetic targets in pharmaceutical development (Scheme 1c).34–36 Such privileged rigid 3D structures offer the concomitant ability to project functionalities in all three-dimensional orientations and led to enhanced pharmacological activities of molecules. Thus significant attention has been paid to their synthesis.37,38 Against this backdrop, here we describe our efforts on the synthesis of various heterospirocycles with the aid of photoredox catalysis.We chose epoxyalkyne 2a as a model substrate for optimization of reaction conditions. After a systematic variation of different reaction parameters, we were pleased to identify the optimal reaction conditions in which a mixture of Cp2TiCl2 (5.0 mol%), [Ir(dtbbpy)(ppy)2]PF6 (1a, 1.0 mol%, EIII/II1/2 = −1.51 V vs. SCE in MeCN), HE (1.2 equiv.) and 2a (1.0 equiv.) in THF at room temperature under the irradiation of a 10 W 450 nm light emitting diode (LED) lamp for 12 hours afforded the desired product 3a in an excellent yield of 96% (13 : 1 d.r.) upon isolation (entry 1). Using a commercial 23 W compact fluorescent lamp (CFL) instead of the 10 W 450 nm LED did not compromise the overall yield of the reaction (entry 2). Notably, when the loading of Cp2TiCl2 was decreased to as low as only 2.0 mol%, the reaction still led to full conversion and produced 3a in 95% yield (entry 3). Further screening of other photosensitizers revealed that the cheap and readily obtained organic dye 4CzIPN 1b is a competent alternative, which led to full conversion with 94% isolated yield (entry 4). Importantly, the reaction did not proceed in the absence of Cp2TiCl2, HE, the photocatalyst, or visible light (entries 5–8). Various solvents, including DMF, MeOH, DMSO, and MeCN, were screened, and they all resulted in poor conversion. The use of other organic electron donors, such as triethylamine, triethanolamine, and ascorbic acid, afforded the product in poor yield.With satisfactory reaction conditions established, we then explored the scope of the cyclization reaction using 4CzIPN as the photosensitizer. Positively, the cyclization reaction worked well and afforded the desired variably heterospirocyclic products in good to excellent yield (Tables 2 and and3).3). The reaction allows the rapid construction of various 5/5, 5/6, 5/7 and 5/8 spiro-ring fused systems (3a–3k) bearing tetrahydrofuran or pyrrolidine motifs via the 5-exo cyclization pathway. Interesting, the diastereoselectivity of the cyclization reaction is highly correlated with the ring size in the substrates. Heterospirocycles containing a 5/5 spiro-ring fused system (3a–3f) were obtained with surprisingly high diastereoselectivity. In some cases (3b, 3c, and 3e) only a single isomer was obtained. The product 3d with a sterically hindered t-butyloxy carbonyl (Boc) protecting group on the N atom was obtained with reduced diastereoselectivity (5 : 1 d.r.). The diastereoselectivities dropped in 5/6, 5/7 and 5/8 spiro-ring fused systems. Given that enantioenriched epoxides could be easily obtained (e.g. via sharpless asymmetric epoxidation), this strategy provides access to optically active spirocycles featuring an all-carbon quaternary stereocenter with the transfer of stereochemical information from epoxides (3c, 3e and 3f). Bis-heterospirocyclic scaffolds were frequently employed in pharmaceutical chemistry. For example, bis-heterospirocyclic 3d is the core structure of DLK inhibitors39 and XEN402 (ref. 40) (scheme 1c), which are used for treating neurodegeneration and congenital erythromelalgia respectively. Furthermore, 6-exo cyclization was also investigated under the standard conditions and smoothly produced a serious of drug-like 6-(trifluoromethyl)-3-pyridinesulfonyl piperidine derivatives including 6/5, 6/6 and 6/7 spiro-ring fused systems (5a–5k) in generally excellent yields. Moreover, cyclization reactions with epoxy-alkynes afforded products containing exocyclic-alkenes and free alcohols which were suitable for further functionalization. This approach provides access to a broad range of novel spirocyclic piperidine and pyrrolidine spirocycles which could be of interest to synthetic and medicinal chemists.Scope of 5-exo and 6-exo cyclizationa,b,c,d
Open in a separate windowaReaction conditions: 2 and 4 (100 mg, 0.1 M in THF).bIsolated yield.c 3c, 3e and 3f were synthesized from enantiomer pure epoxides.dYields within parentheses are based on catalytic conditions using metal as a reductant: CpTi2Cl2 (5 mol%), Zn (2.0 eq.), coll·HCl (2.5 eq.), THF, 20 hours.Additive effect on Ti-catalyzed cyclizationa,b
Open in a separate windowaIn all cases, 2l was used as the substrate and the yield of 3l was determined with 1H NMR.bValues within parentheses are recovery yields of the additives determined with 1H NMR.To examine the scalability of the reaction, gram-scale synthesis of 3a and 3c was performed under the standard conditions with 23 W CFL irradiation. Pleasingly, 92% (3a) and 89% (3c) isolated yields were obtained respectively without any deterioration. Furthermore, an additive-based investigation41 was performed and the results are summarized (Table 3). From this screening, we found that 11 out of 12 additives have no adverse impact on the yield of the reaction. The additives were recovered after the reaction, including benzoxazole 6a, quinazolinone 6b, collidine 6c, tetrahydroquinoline 6d, benzothiazole 6e, indole 6f, and benzofuran 6h. However, quinoxaline 6l strongly inhibited Ti catalysis and 3l was produced in only 23% yield. Notably, the reaction is compatible with various functional groups including phenols 6g, free alcohols and alkene 6j, iodobenzene 6i, ester 3m, ether 3n, dioxolane 5i, lactone 5k, and alkyne 3e. The generality of the dual Ti/4CzIPN catalysis system was further demonstrated by a highly related hydrogen transfer reaction of epoxides which could exclusively provide anti-Markovnikov alcohols (see the ESI S8). The low-cost of 4CzIPN, broad compatibility with sensitive functional groups, and simple operation conditions are appealing for laboratory and industrial applications.Importantly, the reactions were re-subjected to metal reduction catalytic conditions for comparison with the photocatalytic conditions and the yields are shown within parentheses (Table 2d). It clearly showed that the yields were generally lower. Particularly, we found that the yields dropped dramatically as the steric encumbrance of the substrates increased. For example, substrates containing dioxolane (2g, 4i) or a 7/8-membered-ring (2j, 2k, 4j) afforded the products (3g, 5i, 3j, 3k, 5j) in less than 10% yield with the recovery of the starting materials.  相似文献   

14.
A catalytic asymmetric conjugate addition/Schmidt-type rearrangement of vinyl azides and (E)-alkenyloxindoles was realized. It afforded a variety of optically active 3,2′-pyrrolinyl spirooxindoles with high yields (up to 98%), and excellent diastereo- and enantioselectivities (up to 98% ee, >19 : 1 dr), even at the gram-scale in the presence of a chiral N,N′-dioxide–nickel(ii) complex. In addition, a possible catalytic cycle and transition state model were proposed to rationalize the stereoselectivity.

Lewis acid catalyzed asymmetric synthesis of 3,2′-pyrrolinyl spirooxindole skeletons via conjugate addition/Schmidt-type rearrangement of vinyl azides and (E)-alkenyloxindoles.  相似文献   

15.
A novel and efficient desymmetrizing asymmetric ortho-selective mono-bromination of bisphenol phosphine oxides under chiral squaramide catalysis was reported. Using this asymmetric ortho-bromination strategy, a wide range of chiral bisphenol phosphine oxides and bisphenol phosphinates were obtained with good to excellent yields (up to 92%) and enantioselectivities (up to 98.5 : 1.5 e.r.). The reaction could be scaled up, and the synthetic utility of the desired P-stereogenic compounds was proved by transformations and application in an asymmetric reaction.

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

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

16.
Borrowing hydrogen from alcohols, storing it on a catalyst and subsequent transfer of the hydrogen from the catalyst to an in situ generated imine is the hallmark of a transition metal mediated catalytic N-alkylation of amines. However, such a borrowing hydrogen mechanism with a transition metal free catalytic system which stores hydrogen molecules in the catalyst backbone is yet to be established. Herein, we demonstrate that a phenalenyl ligand can imitate the role of transition metals in storing and transferring hydrogen molecules leading to borrowing hydrogen mediated alkylation of anilines by alcohols including a wide range of substrate scope. A close inspection of the mechanistic pathway by characterizing several intermediates through various spectroscopic techniques, deuterium labelling experiments, and DFT study concluded that the phenalenyl radical based backbone sequentially adds H+, H˙ and an electron through a dearomatization process which are subsequently used as reducing equivalents to the C–N double bond in a catalytic fashion.

An efficient method is developed for harvesting hydrogen, its storage and catalytic transfer by an odd alternant hydrocarbon. The strategy is reminiscent of transition metals in borrowing hydrogen mediated processes.  相似文献   

17.
《Tetrahedron letters》2019,60(39):151063
The 1,3-dipolar cycloaddition of keto- and aldonitrones with N-arylitaconimides proceeds regioselectively giving only 5-spiroisoxazolidines. In the case of aldonitrones the reaction proceeds with high diastereoselectivity. A range of the obtained adducts were subjected to reductive cleavage of the NO bond using zinc powder in acetic acid to give the corresponding spirolactones and 1,3-amino alcohols.  相似文献   

18.
Although noble metal nanocatalysts show superior performance to conventional catalysts, they can be problematic when balancing catalytic efficiency and reusability. In order to address this dilemma, we developed a smart paper transformer (s-PAT) to support nanocatalysts, based on easy phase conversion between paper and pulp, for the first time. The pulp phase was used to maintain the high catalytic efficiency of the nanocatalysts and the transformation to paper enabled their high reusability. Herein, as an example of smart paper transformers, a novel chromatography paper-supported Au nanosponge (AuNS/pulp) catalyst was developed through a simple water-based preparation process for the successful reduction of p-nitrophenol to demonstrate the high catalytic efficiency and reusability of the noble metal nanocatalyst/pulp system. The composition, structure, and morphology of the AuNS/pulp catalyst were characterized by XRD, TGA, FE-SEM, ICP, TEM, FT-IR, and XPS. The AuNS/pulp catalyst was transformed into the pulp phase during the catalytic reaction and into the paper phase to recover the catalysts after use. Owing to this smart switching of physical morphology, the AuNS/pulp catalyst was dispersed more evenly in the solution. Therefore, it exhibited excellent catalytic performance for p-nitrophenol reduction. Under optimal conditions, the conversion rate of p-nitrophenol reached nearly 100% within 6 min and the k value of AuNS/pulp (0.0106 s−1) was more than twice that of a traditional chromatography paper-based catalyst (0.0048 s−1). Additionally, it exhibited outstanding reusability and could maintain its high catalytic efficiency even after fifteen recycling runs. Accordingly, the unique phase switching of this smart paper transformer enables Au nanosponge to transform into a highly efficient and cost-effective multifunctional catalyst. The paper transformer can support various nanocatalysts for a wide range of applications, thus providing a new insight into maintaining both high catalytic efficiency and reusability of nanocatalysts in the fields of environmental catalysis and nanomaterials.

A smart paper transformer supported nanocatalyst platform is developed based on the facile phase conversion between paper and pulp for both high-efficiency and high-reusability catalysis, with wide applications demonstrated by using Au nanosponge.  相似文献   

19.
How would acidic bond dissociation be affected by adding a small quantity of a weakly polar ionic liquid IL (the “apparent” or “measured” dielectric constant ε of the IL is around 10–15) into a strongly polar molecular solvent (e.g., ε of DMSO: 46.5), or vice versa? The answer is blurred, because no previous investigation was reported in this regard. Toward this, we, taking various IL/DMSO mixtures as representatives, have thoroughly investigated the effects of the respective solvent in ionic–molecular binary systems on self-dissociation of C–H acid phenylmalononitrile PhCH(CN)2via pKa determination. As disclosed, in this category of binary media, (1) no linear correspondence exists between pKa and molar fractions of the respective solvent components; (2) only ∼1–2 mol% of weakly polar ILs in strongly polar DMSO make C–H bonds even more dissociative than in neat DMSO; (3) a small fraction of DMSO in ILs (<10 mol%) can dramatically ease acidic C–H-dissociation; and (4) while the DMSO fraction further increases, its acidifying effect becomes much attenuated. These findings, though maybe counterintuitive, have been rationalized on the basis of the precise pKa measurement of this work in relation to the respective roles of each solvent component in solvation.

The dependence of PhCH(CN)2 pKa on the molar fraction of ionic liquids in ionic–molecular binary mixtures showed a nonlinear three-fragment plot, which was rationalized for the first time by the respective roles of each solvent component for solvation.  相似文献   

20.
Convergent paired electrosynthesis is an energy-efficient approach in organic synthesis; however, it is limited by the difficulty to match the innate redox properties of reaction partners. Here we use nickel catalysis to cross-couple the two intermediates generated at the two opposite electrodes of an electrochemical cell, achieving direct arylation of benzylic C–H bonds. This method yields a diverse set of diarylmethanes, which are important structural motifs in medicinal and materials chemistry. Preliminary mechanistic study suggests oxidation of a benzylic C–H bond, Ni-catalyzed C–C coupling, and reduction of a Ni intermediate as key elements of the catalytic cycle.

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

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

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