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 203 毫秒
1.
Here, we describe simple B(C6F5)3-catalyzed mono- and dihydrosilylation reactions of terminal alkynes by using a silane-tuned chemoselectivity strategy, affording vinylsilanes and unsymmetrical geminal bis(silanes). This strategy is applicable to the dihydrosilylation of both aliphatic and aryl terminal alkynes with different silane combinations. Gram-scale synthesis and conducting the reaction without the exclusion of air and moisture demonstrate the practicality of this methodology. The synthetic utility of the resulting products was further highlighted by the structural diversification of geminal bis(silanes) through transforming the secondary silane into other silyl groups. Comprehensive theoretical calculations combined with kinetical isotope labeling studies have shown that a prominent kinetic differentiation between the hydrosilylation of alkynes and vinylsilane is responsible for the chemoselective construction of unsymmetrical 1,1-bis(silanes).

A B(C6F5)3/silane-based system enables the chemoselective dihydrosilylation of terminal alkynes. Using a combination of different types of hydrosilanes, a series of unsymmetrical or symmetrical 1,1-bis(silanes) could be constructed.  相似文献   

2.
Pd-catalyzed C(sp3)–H oxygenation has emerged as an attractive strategy for organic synthesis. The most commonly proposed mechanism involves C(sp3)–H activation followed by oxidative addition of an oxygen electrophile to give an alkylpalladium(iv) species and further C(sp3)–O reductive elimination. In the present study of γ-C(sp3)–H acyloxylation of amine derivatives, we show a different mechanism when tert-butyl hydroperoxide (TBHP) is used as an oxidant—namely, a bimetallic oxidative addition-oxo-insertion process. This catalytic model results in an alkoxypalladium(ii) intermediate from which acyloxylation and alkoxylation products are formed. Experimental and computational studies, including isolation of the putative post-oxo-insertion alkoxypalladium(ii) intermediates, support this mechanistic model. Density functional theory reveals that the classical alkylpalladium(iv) oxidative addition pathway is higher in energy than the bimetallic oxo-insertion pathway. Further kinetic studies revealed second-order dependence on [Pd] and first-order on [TBHP], which is consistent with DFT analysis. This procedure is compatible with a wide range of acids and alcohols for γ-C(sp3)–H oxygenation. Preliminary functional group transformations of the products underscore the great potential of this protocol for structural manipulation.

Alkoxypalladium(ii) species lead to γ-C(sp3)–H acyloxylation and alkoxylation products using tert-butyl hydroperoxide as the oxidant.  相似文献   

3.
Main group organometallic compounds can exhibit unusual optical properties arising from hybrid σ,π-conjugation. While linear silanes are extensively studied, the shortage of methods for the controlled synthesis of well-defined cyclic materials has precluded the study of cyclic conjugation. Herein we report that Ru-catalyzed addition of cyclosilanes to aryl acetylenes (hydrosilylation) proceeds with high chemoselectivity, regioselectivity, and diastereoselectivity, affording complex organosilanes that absorb visible light. We further show that the hydrosilylation products are useful building blocks towards novel conjugated polymers.

Hybrid σ,π-conjugated cyclosilanes were synthesized via highly selective hydrosilylation and have shown great potentials as building blocks to construct novel conjugated polymers with control of tacticity.  相似文献   

4.
Photochemically activated [Mo(CO)6] and [Mo(CO)44-nbd)] have been demonstrated to be very effective catalysts for hydrosilylation of norbornadiene (nbd) by tertiary (Et3SiH, Cl3SiH) and secondary (Et2SiH2 and Ph2SiH2) silanes to give 5-silyl-2-norbornene, which under the same reaction conditions transform in ring-opening metathesis polymerization (ROMP) to unsaturated polymers and to a double hydrosilylation product, 2,6-bis(silyl)norbornane. The yield of a particular reaction depends very strongly on the kind of silane involved. The reaction products were identified by means of chromatography (GC–MS) and 1H and 13C NMR spectroscopy. In photochemical reaction of [Mo(CO)44-nbd)] and Ph2SiH2 in cyclohexane-d12, η2-coordination of the SiH bond to the molybdenum atom is supported by 1H NMR spectroscopy due to the detection of two equal-intensity doublets with 2JHH = 5.4 Hz at δ 6.12 and −5.86 ppm.  相似文献   

5.
The photoactivated (350 nm) hydrosilylation of alkynes by silanes catalyzed by platinum(II) bis(acetylacetonato) has been studied. Platinum(II) bis(acetylacetonato) is an efficient catalyst. High yields of adducts (>98% for terminal alkynes) can be obtained in 2–3 h after a short induction period with a catalyst–reactant molar ratio of 10−3/1. The reaction rate depends on the choice of silane, irradiation time and the concentration of catalyst. The major product is the β-trans adduct. Minor products are the α isomer with a trace of β-cis isomer. Comparisons of hydrosilylation reactions of alkynes with hydrosilylation reactions of alkenes are reported.  相似文献   

6.
Hydrosilylation of terminal alkynes with a variety of silanes catalyzed by Cl2(PCy3)2RuCHPh (1) affords mainly the Z-isomer via trans addition in excellent yields. The presence of a hydroxyl group in close proximity to the triple bond was observed to exert a strong directing effect, resulting in the highly selective formation of the α-isomer. Intramolecular hydrosilylation of a homopropargylic silyl ether was demonstrated to give the cis addition product.  相似文献   

7.
Silylcoppers function as convenient and effective sources of silicon functional groups. Commonly used precursors for those species have been limited to certain symmetric disilanes and silylboranes. This fact renders the development of silylcopper precursors desirable so that more diverse silyl groups could be efficiently delivered. Here we extend the utility of sodium silylsilanolates as competent precursors of silylcoppers. A silanolate unit operates as an auxiliary to transfer a variety of silyl groups to the copper centre, which was demonstrated in the copper-catalysed hydrosilylation of internal alkynes, α,β-unsaturated ketones, and allenes. Our mechanistic studies through DFT calculation suggested that a copper silylsilanolate undergoes intramolecular oxidative addition of the Si–Si bond to the copper centre to generate a silylcopper, in contrast to the typical formal σ-bond metathesis mechanism for conventional disilanes or silylboranes and copper alkoxides. Accordingly, sodium silylsilanolate has been established as an expeditious precursor of a variety of silylcopper species.

Sodium silylsilanolates are demonstrated as useful silylating reagents for copper-catalysed hydrosilylation of unsaturated bonds via the formation of reactive silylcopper species that can deliver a series of silyl groups.  相似文献   

8.
Summary Trisubstituted silanes, HSiR3-n X n (R = Me or Et, X = Cl, OEt, or Ph; n = 0–3) oxidatively add to the complex [RhCl(cod)(1-hexene)] (cod = cycloocta-1,5-diene) to yield [RhCl(cod)(1-hexene)(H)(SiR3)] [(1)]. Subsequent steps of hydrosilylation follow, i.e. cis-insertion of the alkene (- rearrangement) and then reductive elimination of the product, according to the general Chalk and Harrod scheme. A quantitative correlation between the second order rate constant, k 1, of the oxidative addition (followed spectrophotometrically) at 20°C in benzene solution and the structure of the trisubstituted silane represented by Stereoelectronic parameters , and E' for the SiR3-n X n groups was established. The maximal hydrosilylation rate followed by g.l.c., is strongly retarded by highly electronegative substituents X on silicon and results from the elimination rate of the hydrosilylation product from (1) and the maximal concentration of (1) in solution.Dedicated to Professor K. Rühlmann on the occasion of his 65th birthday. Part XXVII in the series Catalysis of Hydrosilylation; for Part XXVI see Polish J. Appl. Chem., 38, 169 (1994).  相似文献   

9.
Regio- and stereoselective hydrosilylation of terminal alkynes on solid support using diisopropyl hydrosilanes yielding β-(E)-vinyl silanes with excellent selectivity is reported. The hydrosilylation is catalyzed by Pt(DVDS)/P(iBuNCH2CH2)3N (DVDS = 1,3-divinyl-1,1,3,3-tetramethyl-disiloxane), in which the bulky proazaphosphatrane ligand plays a key role for the selectivity. The immobilized products are characterized with gel phase13C NMR and 1H high resolution magic angle spinning NMR.  相似文献   

10.
Photoaffinity labeling (PAL) is a powerful tool for the identification of non-covalent small molecule–protein interactions that are critical to drug discovery and medicinal chemistry, but this approach is limited to only a small subset of robust photocrosslinkers. The identification of new photoreactive motifs capable of covalent target capture is therefore highly desirable. Herein, we report the design, synthesis, and evaluation of a new class of PAL warheads based on the UV-triggered 1,2-photo-Brook rearrangement of acyl silanes, which hitherto have not been explored for PAL workflows. Irradiation of a series of probes in cell lysate revealed an iPr-substituted acyl silane with superior photolabeling and minimal thermal background labeling compared to other substituted acyl silanes. Further, small molecule (+)-JQ1- and rapamycin-derived iPr acyl silanes were shown to selectively label recombinant BRD4-BD1 and FKBP12, respectively, with minimal background. Together, these data highlight the untapped potential of acyl silanes as a novel, tunable scaffold for photoaffinity labeling.

Irradiation initiated 1,2-photo Brook rearrangement of acyl silanes generated α-siloxycarbene intermediates that were used for photoaffinity labeling. Optimization of the acyl silane group produced a probe capable of capturing small molecule–protein interactions.  相似文献   

11.
Novel P,O-type ligands, N-disulfonyl bicyclic bridgehead phosphorus triamides, were synthesized and utilized in Pd-catalyzed hydrosilylation involving tertiary silanes, unactivated alkenes, and conjugated dienes. The ligand displayed a remarkable level of reactivity for alkene hydrosilylation with tertiary silanes and its use resulted in a significant improvement in the regioselectivity of diene hydrosilylation towards 1,2-hydrosilylation. X-ray crystallographic analysis confirmed the bidentate nature of the ligand, with coordination of phosphorus and oxygen. Control experiments provided evidence for the formation of Pd0 species and the reversibility of Pd−H insertion in the reaction mechanism. Density functional theory (DFT) computations supported the importance of the hemilabile P,O ligand in stabilizing both the rate-determining transition state of Pd−H insertion and the transition state of reductive elimination that determines the regioselectivity.  相似文献   

12.
The super acidity of the unsolvated Al(C6F5)3 enabled isolation of the elusive silane–alane complex [Si? H???Al], which was structurally characterized by spectroscopic and X‐ray diffraction methods. The Janus‐like nature of this adduct, coupled with strong silane activation, effects multifaceted frustrated‐Lewis‐pair‐type catalysis. When compared with the silane–borane system, the silane–alane system offers unique features or clear advantages in the four types of catalytic transformations examined in this study, including: ligand redistribution of tertiary silanes into secondary and quaternary silanes, polymerization of conjugated polar alkenes, hydrosilylation of unactivated alkenes, and hydrodefluorination of fluoroalkanes.  相似文献   

13.
The super acidity of the unsolvated Al(C6F5)3 enabled isolation of the elusive silane–alane complex [Si H⋅⋅⋅Al], which was structurally characterized by spectroscopic and X‐ray diffraction methods. The Janus‐like nature of this adduct, coupled with strong silane activation, effects multifaceted frustrated‐Lewis‐pair‐type catalysis. When compared with the silane–borane system, the silane–alane system offers unique features or clear advantages in the four types of catalytic transformations examined in this study, including: ligand redistribution of tertiary silanes into secondary and quaternary silanes, polymerization of conjugated polar alkenes, hydrosilylation of unactivated alkenes, and hydrodefluorination of fluoroalkanes.  相似文献   

14.
The transient directing group (TDG) strategy allowed long awaited access to the direct β-C(sp3)–H functionalization of unmasked aliphatic aldehydes via palladium catalysis. However, the current techniques are restricted to terminal methyl functionalization, limiting their structural scopes and applicability. Herein, we report the development of a direct Pd-catalyzed methylene β-C–H arylation of linear unmasked aldehydes by using 3-amino-3-methylbutanoic acid as a TDG and 2-pyridone as an external ligand. Density functional theory calculations provided insights into the reaction mechanism and shed light on the roles of the external and transient directing ligands in the catalytic transformation.

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

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

15.
In this Perspective we discuss the ability of transition metal complexes to activate and cleave the Si–H and B–H bonds of hydrosilanes and hydroboranes (tri- and tetra-coordinated) in an electrophilic manner, avoiding the need for the metal centre to undergo two-electron processes (oxidative addition/reductive elimination). A formal polarization of E–H bonds (E = Si, B) upon their coordination to the metal centre to form σ-EH complexes (with coordination modes η1 or η2) favors this type of bond activation that can lead to reactivities involving the formation of transient silylium and borenium/boronium cations similar to those proposed in silylation and borylation processes catalysed by boron and aluminium Lewis acids. We compare the reactivity of transition metal complexes and boron/aluminium Lewis acids through a series of catalytic reactions in which pieces of evidence suggest mechanisms involving electrophilic reaction pathways.

In this Perspective we compare the ability of transition metals and p-block Lewis acids to activate electrophilically hydrosilanes and hydroboranes. The mechanistic similarities and dissimilarities in different catalytic transformations are analyzed.  相似文献   

16.
Regioselective catalytic multi-functionalization reactions enable the rapid synthesis of complexed products from the same precursors. In this communication, we present a method for the regiodivergent borocarbonylation of benzylidenecyclopropanes with aryl iodides. Various γ-vinylboryl ketones and β-cyclopropylboryl ketones were produced in moderate to good yields with excellent regioselectivity from the same substrates. The choice of the catalyst is key for the regioselectivity control: γ-vinylboryl ketones were produced selectively with IPrCuCl and Pd(dppp)Cl2 as the catalytic system, while the corresponding β-cyclopropylboryl ketones were obtained in high regioselectivity with Cu(dppp)Cl, [Pd(η3-cinnamyl)Cl]2 and xantphos as the catalytic system. Moreover, γ-vinylboryl ketones and β-cyclopropylboryl ketones were successfully transformed into several other value-added products.

A novel procedure for regiodivergent borocarbonylation of benzylidenecyclopropanes has been developed. A variety of valuable γ-vinylboryl ketones and β-cyclopropylboryl ketones can be obtained selectively in excellent yields.  相似文献   

17.
Double hydrosilylation of alkynes represents a straightforward method to synthesize bis(silane)s, yet it is challenging if α-substituted vinylsilanes act as the intermediates. Here, a cobalt-catalyzed regiodivergent double hydrosilylation of arylacetylenes is reported for the first time involving this challenge, accessing both vicinal and geminal bis(silane)s with exclusive regioselectivity. Various novel bis(silane)s containing Si−H bonds can be easily obtained. The gram-scale reactions could be performed smoothly. Preliminarily mechanistic studies demonstrated that the reactions were initiated by cobalt-catalyzed α-hydrosilylation of alkynes, followed by cobalt-catalyzed β-hydrosilylation of the α-vinylsilanes to deliver vicinal bis(silane)s, or hydride-catalyzed α-hydrosilylation to give geminal ones. Notably, these bis(silane)s can be used for the synthesis of high-refractive-index polymers (nd up to 1.83), demonstrating great potential utility in optical materials.  相似文献   

18.
Oxidant-free Au-catalyzed reactions are emerging as a new synthetic tool for innovative organic transformations. Oxidant-free Au-catalyzed reactions are emerging as a new synthetic tool for innovative organic transformations. Still, a deeper mechanistic understanding is needed for a rational design of these processes. Here we describe the synthesis of two Au(i) complexes bearing bidentated hemilabile MIC^N ligands, [AuI(MIC^N)Cl], and their ability to stabilize square-planar Au(iii) species (MIC = mesoionic carbene). The presence of the hemilabile N-ligand contributed to stabilize the ensuing Au(iii) species acting as a five-membered ring chelate upon its coordination to the metal center. The Au(iii) complexes can be obtained either by using external oxidants or, alternatively, by means of feasible oxidative addition with strained biphenylene Csp2–Csp2 bonds as well as with aryl iodides. Based on the fundamental knowledge gained on the redox properties on these Au(i)/Au(iii) systems, we successfully develop a novel Au(i)-catalytic procedure for the synthesis of γ-substituted γ-butyrolactones through the arylation-lactonization reaction of the corresponding γ-alkenoic acid. The oxidative addition of the aryl iodide, which in turn is allowed by the hemilabile nature of the MIC^N ligand, is an essential step for this transformation.

A novel hemilabile MIC^N ligand-based Au(i)-catalytic procedure for the synthesis of γ-substituted γ-butyrolactones through the arylation-lactonization reaction of the corresponding γ-alkenoic acid is presented.  相似文献   

19.
The Fujiwara–Moritani reaction is a powerful tool for the olefination of arenes by Pd-catalysed C–H activation. However, the need for superstoichiometric amounts of toxic chemical oxidants makes the reaction unattractive from an environmental and atom-economical view. Herein, we report the first non-directed and regioselective olefination of simple arenes via an electrooxidative Fujiwara–Moritani reaction. The versatility of this operator-friendly approach was demonstrated by a broad substrate scope which includes arenes, heteroarenes and a variety of olefins. Electroanalytical studies suggest the involvement of a Pd(ii)/Pd(iv) catalytic cycle via a Pd(iii) intermediate.

The Fujiwara–Moritani reaction using electric current is a powerful tool for the olefination of arenes by Pd-catalysed C–H activation.

Transition metal-catalysed C–H functionalisation reactions have increasingly gained importance over the last few decades since they allow direct and rapid installation of functionality. Regardless of the undeniable synthetic value of such transformations, the need for superstoichiometric quantities of expensive and hazardous oxidants (e.g., silver and copper salts) remains a major drawback from a sustainable chemistry perspective.1,2 Additionally, chemical oxidants often lead to the formation of by-products, hindering purification and decreasing atom economy. Nevertheless, very few reports were also reported in the literature wherein mild oxidant such as molecular oxygen can also serve as the oxidising agent.2j To make chemical processes and transformations intrinsically sustainable, organic chemists re-discovered synthetic electrochemistry as an environmentally friendly approach.3–6 In the domain of synthetic electrochemistry, the Lei group achieved a significant milestone and installed C–C bonds through a different cross-coupling strategy.1k,2fh Electroorganic synthesis utilizes electric current to realize redox processes and thereby avoids the use of dangerous, expensive, and polluting chemical oxidising or reducing agents. Precise control of electrochemical reaction parameters often leads to commendable reactivity and chemoselectivity and hence to an improved atom economy. In addition, electrochemical processes fulfil the expectations of sustainability since electricity can be generated from renewable energy sources, such as wind, sunlight or biomass. Recent efforts in the field of electrochemical C–H activation resulted in significant progress towards efficient C–C and C–heteroatom bond formations.7–10 Hence, the utilization of electric current as an alternative oxidant in Pd-catalysed C–H functionalisations is emerging as an attractive alternative to stoichiometric reagents.11–13The Fujiwara–Moritani reaction is one of the earliest known examples of Pd-catalysed oxidative C–H functionalisations for C–C bond formation.14 This extraordinary C(sp2)-H alkenylation reaction avoids the use of prefunctionalised starting materials; however, it suffers from the drawbacks of regioselectivity, reactivity and use of excess arenes.15 Since its development, a number of modified strategies have been reported by different research groups to address the issue of reactivity and selectivity.16–21 In recent times, the ligand assisted oxidative C–H alkenylation of arenes without directing substituents has been established as one of the major strategies to overcome the reactivity issue and to elaborate the substrate scope.However, regioselectivity for most of the sterically and electronically unbiased arenes is still not up to the mark. The most recent studies on the non-directed oxidative C–H olefination of arenes were reported independently by Yu and van Gemmeren (Scheme 1). The Yu group employed electron-deficient 2-pyridone as an X-type ligand for the olefination of both electron-rich and electron-poor arenes including heteroarenes as the limiting reagent (Scheme 1a).18 The pyridone ligand improves the selectivity in a non-directed approach as compared to the directed C–H olefination reaction by enhancing the influence of steric effects. On the other hand, the van Gemmeren group utilizes two complementary ligands N-Ac–Gly–OH and a 6-methylpyridine derivative in a 1 : 1 ratio to accomplish the non-directed olefination reaction of arenes (Scheme 1b).20 Despite the indisputable advances made by these research groups in the area of non-directed oxidative C–H olefination of arenes, the use of superstoichiometric amounts of toxic and waste-generating oxidants (Ag salts) deciphers into a strong call for an environmentally responsive and atom-economic protocol. To address these shortcomings, we recently introduced Pd-photoredox catalysed olefination of non-directed arenes with excellent site selectivity under oxidant free conditions.21Open in a separate windowScheme 1Recent approaches to sustainable C–H alkenylation reactions.In 2007, Jutand reported the directing group assisted Pd-electracatalysed ortho-olefination of acetyl protected aniline in a divided cell by utilizing catalytic amounts of benzoquinone as a redox mediator (Scheme 1c).22a A Rh-catalysed ortho-C–H olefination of benzamide was developed through an electrooxidative pathway by the Ackermann group (Scheme 1d).22b Simple arenes that bear no directing groups are cheap, easily available and very desirable starting materials. However, the use of such arenes is significantly more challenging for selective functionalisation as transformations often result in the formation of complex product mixtures. With no report of an electrooxidative Pd-catalysed C(sp2)-H alkenylation of simple arenes present, we wish to present such a variant of the Fujiwara–Moritani reaction (Scheme 1e). The developed method proceeds through a non-directed pathway and is controlled by stereoelectronic factors. This protocol does not require additional chemical oxidizing agents and is executed using an operator-friendly undivided cell setup.To start our study, naphthalene was chosen as a challenging substrate because of its ability to form α- and β-products. We examined various reaction conditions for the desired Pd-catalysed electrooxidative C–H alkenylation in a simple undivided cell setup () with n-butyl acrylate as the coupling partner. After rigorous optimisation, we found that naphthalene reacts with n-butyl acrylate in dichloroethane (DCE) in the presence of Pd(OAc)2 (10 mol%), ligand L1 (20 mol%), and the electrolyte tetra-n-butylammonium hexafluorophosphate (TBAPF6, 0.5 equiv.) while employing a graphite felt anode and a platinum cathode maintaining constant current electrolytic conditions (j = 2.5 mA cm−2,
EntryAlteration from standard conditionsYield of 1b (%)Selectivity (β : α)
1None70>25 : 1
2Co(OAc)2·4H2O instead of Pd(OAc)291 : 1
3[Ru(p-cymene)Cl2]2 instead of Pd(OAc)2NR
4Pd(OAc)2·(5 mol%)51>25 : 1
5Pd(OAc)2·(20 mol%)71>25 : 1
6L2 instead of L1458 : 1
7L3 instead of L15920 : 1
8L4 instead of L1195 : 1
9L5 instead of L181 : 1
10Benzoquinone (10 mol%)68>25 : 1
11PivOH (1.0 equiv.)6120 : 1
12Ni foam instead of Pt64>25 : 1
13GF instead of Pt4915 : 1
14Steel instead of Pt3113 : 1
156 mA cm−2 instead of 2.5 mA cm−22711 : 1
1624 h reaction time4720 : 1
1712 h reaction time5621 : 1
18No electricityNR
19No Pd(OAc)2NR
Open in a separate windowaStandard reaction conditions: undivided cell, GF anode, Pt cathode, j = 2.5 mA cm−2, naphthalene (0.2 mmol), n-butyl acrylate (0.5 mmol), Pd(OAc)2 (10 mol%), L1 (20 mol%), TBAPF6 (0.5 equiv.), DCE (3 mL), 15 h, under air.bYield determined by 1H-NMR of crude reaction mixture. NR = no reaction; TBAPF6 = tetra-n-butylammonium hexafluorophosphate. GF = graphite felt. Surface area of electrodes dipped in solution = 0.7 cm × 0.7 cm, current = 1.225 mA and current density = 2.5 mA cm−2 (electrochemical surface area = 1.23 cm2).Notably, in the present transformation the ligand has a major influence on the reactivity and selectivity aspects (see the ESI, Table S4). After studying a series of 2-pyridone, pyridine and amino acid-based ligands L2–L5 it was found that L1 is the optimal ligand since it provided superior yield and selectivity (entries 6–9). Addition of catalytic amounts of p-benzoquinone as a redox mediator (entry 10) or pivalic acid as an additive (entry 11, Scheme 2). Following the olefination of naphthalene (68%, >25 : 1 β : α selectivity), 1,2,3,4-tetrahydronaphthalene was successfully reacted (52%, 11 : 1 β : α-selectivity). Next, we applied our standard reaction conditions to benzene and found them not to be equally effective as only 25% of the olefinated product 3 was obtained. As a result, further optimizations of electric current density and solvent were carried out to enhance the yield (see the ESI, Table S9). To our satisfaction, the yield of product 3 increased to 63% when the electrolysis was carried out with an electric current density of j = 1.5 mA cm−2 and in a solvent mixture of DCE/HFIP (5 : 1). These modified reaction conditions were applied to the electrosynthesis of all other olefinated products 4–26 (Scheme 2). The olefination of 1,3,5-trimethoxybenzene and mesitylene with n-butyl acrylate proceeded smoothly under the revised reaction conditions to afford products 4–5 in up to 65% yield. The regioselectivity issue was more prominent for arenes bearing two or more electronically similar C–H bonds (e.g., electron-rich arenes: ortho vs. para). Dimethoxy benzene gives β-selective olefinated product 6 (β : α; 7 : 1). While toluene was converted with para-selectivity (7 : 1) to 7, phenol afforded olefinated product 8 with ortho-selectivity (o : others; 9 : 1, Scheme 2). On the other hand, subjecting TBDMS (tert-butyldimethylsilyl) protected phenol to the established protocol furnished 9 with 8 : 1 para-selectivity (Scheme 2). The TBDPS (tert-butyldiphenylsilyl) protected phenol afforded exclusively the para-olefinated product 10 which might be due to the steric repulsion caused by the bulky protecting group. Conversion of 2,6-diiso-propylphenol provided olefinated product 11 as a single para-olefinated isomer with 67% yield. Anisole and ethoxybenzene both reacted smoothly to produce 12 (72%, 15 : 1) and 13 (70%, 10 : 1) with ortho-selectivity (Scheme 2). The compatibility of the present transformation was further showcased by the olefination of N,N-dimethyl aniline in 70% yield (14) and 8 : 1 ortho-selectivity. Similarly, methyl ferrocene carboxylate and biologically active caffeine reacted smoothly with n-butyl acrylate to produce olefinated products 15 and 16 in good yields (Scheme 2). Moderately electron-withdrawing arenes such as a phenyl acetic acid derivative (17, 51%, o : others = 7 : 1), a homoveratric acid derivative (18, 48%, o : others = 15 : 1) or 4-methoxy acetophenone (19, 59%, m : others = 7 : 1) gave the corresponding products in satisfactory yields. The coupling of unsubstituted thiophene and furan with n-butyl acrylate afforded the olefinated products 20 and 21 (64% and 68%) with synthetically useful C2-selectivity, respectively (C2 : others; 18 : 1 and C2 : others; 19 : 1, Scheme 2). In contrast, thiophenes bearing a substituent at the C2 position such as 2-phenylthiophene and 1-(4-(thien-2-yl)phenyl)ethan-1-one reacted with high C5-selectivity (>20 : 1) to afford the arylated α,β-unsaturated esters 22 and 23 (76% and 73% yield). Conversion of 2-(2-nitrophenyl)thiophene delivered the desired product 24 in 64% yield with exclusive C5-selectivity. A C3-substituted thiophene also reacted with the acrylate to afford 25 in 72% yield (C5 : others; 6 : 1 selectivity). Heteroarenes bearing electron-withdrawing substituents such as 2-acetyl thiophene (26) afforded the C5-olefinated product in moderate yield and selectivity (60%, C5 : others = 8 : 1). However, aromatic rings bearing strong electron-withdrawing groups (–NO2, –CHO, –CF3, –F etc.) are not compatible under our present reaction conditions (see details in the ESI, Section 4.3).Open in a separate windowScheme 2Evaluation of simple arenes and heteroarenes in the electrochemical olefination.a Reaction conditions: undivided cell, GF anode, Pt cathode, j = 2.5 mA cm−2 or j = 1.5 mA cm−2, corresponding arenes or heteroarenes (0.2 mmol), n-butyl acrylate (0.5 mmol), Pd(OAc)2 (10 mol%), L1 (20 mol%), TBAPF6 (0.5 equiv.), DCE (3 mL) or 5 : 1 ration of dichloroethane (DCE) and 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP), 15 h, under air. bYields of isolated products are reported.Next, we investigated the applicability of other olefins by reacting them with simple arenes (Scheme 3). In addition to other acrylates (methyl 27, ethyl 28 and tert-butyl 29), acrylic acid was successfully converted with naphthalene to its arylated product 30. Moderate yields (54–60%) and moderate to high β : α selectivities (up to >25 : 1) were obtained for all reactions. Coupling of methyl acrylate with benzene under adjusted electrochemical conditions (j = 1.5 mA cm2; DCE/HFIP mixtures) gave 62% of olefinated product 31. Other activated olefins such as methyl vinyl sulfone, and acrylonitrile were also amenable to the present olefination protocol. Subjecting these substrates in combination with different arenes to our protocol led to a variety of arylated products 32–35 in good yields and regioselectivities. α,β-Unsaturated ester derivatives of bioactive molecules such as δ-tocopherol and cholesterol were efficiently reacted with naphthalene to the olefinated products 36–37 in moderate yields. To further elaborate the scope of present protocol, un-activated olefins such as aliphatic olefins and styrene derivatives were tested. However, none of them afford olefinated products under our reaction conditions (see details in the ESI, Section 4.3). To monitor the scalability of the present transformation, two reactions were performed with the model reaction at scales of 0.504 g (46%, β : α = 7 : 1) and 1.08 g (41%, β : α = 7 : 1; see ESISection 4.2).Open in a separate windowScheme 3Evaluation of other α,β-unsaturated systems in the electrochemical olefination of arenes. aReaction conditions: undivided cell, GF anode, Pt cathode, j = 2.5 mA cm−2 or j = 1.5 mA cm−2, corresponding arenes or heteroarenes (0.2 mmol), activated olefins (0.5 mmol), Pd(OAc)2 (10 mol%), L1 (20 mol%), TBAPF6 (0.5 equiv.), DCE (3 mL) or 5 : 1 ratio of dichloroethane (DCE) and 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP), 15 h, under air. bYields of isolated products are reported.To gain insights into the catalytic mode of action, electrochemical and spectroelectrochemical experiments were performed. Cyclic voltammetry (CV) of Pd(OAc)2 in DCE revealed two oxidation waves at +1.42 V vs. NHE (Normal Hydrogen Electrode) and at +2.47 V vs. NHE (Fig. 1a) which might refer to the redox conversion of Pd(ii/iii) and Pd(iii/iv).23Fig. 1b shows the CVs of naphthalene (substrate), ligand L1, n-butyl acrylate, and Pd(OAc)2. In comparison to the Pd(ii/iii) redox pair, a significantly higher oxidation potential (+2.16 V vs. NHE) was observed for naphthalene, which suggests that substrate activation is potentially induced by a Pd species with an oxidation state greater than +II (Fig. 1b). The CVs of other substrates followed the same pattern (see the ESI, Fig. S2). According to an electrochemical study on approximate ranges of standard redox potentials for Pd intermediates in catalytic reactions, the oxidation of Pd(ii) to Pd(iv) is usually observed in the range of +1.00–2.00 V (vs. Fc/Fc+ = ferrocene) or 1.63–2.63 V (vs. NHE).23 The CV profile of Pd(OAc)2 in the negative scan revealed two reduction waves at −0.23 V and at −1.06 V vs. NHE (Fig. S6) which might refer to the redox conversion of Pd(ii/i) and Pd(i/0). Taking these results into account, involvement of a Pd(ii/iv) catalytic cycle during the present transformation appears to be likely as the negative scan rules out a Pd(ii/0) cycle.23d,eOpen in a separate windowFig. 1(a) Cyclic voltammograms of Pd(OAc)2 and L1-Pd(OAc)2 (1 mM, 100 mV s−1 scan rate, glassy carbon, potential vs. NHE, 0.1 M TBAPF6 in DCE); (b) cyclic voltammogram of reactants (1 mM, 100 mV s−1 scan rate, glassy carbon, potential vs. NHE, 0.1 M TBAPF6 in DCE); (c) in situ UV-Vis spectroelectrochemical spectra of the reaction mixture during bulk electrolysis at +2.61 V vs. NHE; (d) in situ UV-Vis spectroelectrochemical spectra of the Pd-ligand complex during bulk electrolysis at +2.61 V vs. NHE.In order to obtain further evidence for this hypothesis, we examined the reaction mixture at a constant potential of +2.61 V (vs. NHE) spectroelectrochemically (SEC) to check any changes in optical features during the reaction. This in situ UV-visible analysis of the reaction mixture revealed the gradual decrease of an absorption band at 379 nm and a new peak (∼350 nm) appeared over time (Fig. 1c). Similar behaviour was observed for the Pd-ligand complex as a blue shift of optical bands was found from 368 nm to 352 nm at the same potential of +2.61 V (vs. NHE, Fig. 1d). The differences in the observed UV-Vis peak positions are presumably due to a change in the geometry of the Pd-complex upon oxidation in the analysed reaction mixtures.To further consolidate this hypothesis, the same SEC experiment was repeated with only Pd(OAc)2 which showed an absorption peak at 404 nm (Fig. S3). Electrolysis of Pd(OAc)2 at +2.61 V (vs. NHE) also resulted in a blue shift with a new peak appearing at almost the same wavelength of 349 nm (Fig. S4). All these results led us to postulate that the new peak was associated with a change in the oxidation state of the Pd(ii) center. Moreover understand the nature of intermediates involved in the catalytic cycle, a series of electron paramagnetic resonance (EPR) experiments of the reaction mixture were conducted at different time intervals employing optimised reaction conditions. The EPR spectra (273 K) after 1 h showed a strong peak at g = 2.005 which was presumably due to the formation of an organic radical (Fig. 2a), however no naphthalene homo-coupled product was detected after different time intervals or under different conditions. At longer time intervals (4 h and 7 h), weak peaks at gx = 2.139, gy = 2.081 and gz = 2.055 arose due to the asymmetry of the electronic distribution. The appearance of rhombic signals suggested the formation of a Pd(iii) intermediate having a d7 center (Fig. 2a).24 An enlarged version of the spectra for Pd(iii) after 7 h is shown with simulated data in Fig. 2b. Time-dependent EPR spectra highlight that the build-up of Pd(iii) was concomitant with the decreased formation of an organic radical (PdIII–R to PdIIR.) as the corresponding peak diminished. This implied that the catalytically active Pd(iii) species got accumulated as the reaction approached towards completion. Furthermore, the EPR data in the absence of n-butyl acrylate (after 2 h) also revealed a very strong peak at g = 2.005; hence the formation of a radical species from the olefin was ruled out (Fig. S5).Open in a separate windowFig. 2(a) EPR spectrum of the reaction mixture under the standard reaction conditions at different time intervals (273 K); (b) enlarged EPR spectra of Pd(iii) after 7 h of experiment at 273 K (experimental vs. simulated).Additionally, radical quenching experiments with TEMPO did not show any effects under the standard reaction conditions. Furthermore, electrochemical arene oxidation to generate organic radicals has been well reported in the literature.5f All these control experiments suggest that a phenoxy radical from L1 (C′) might be formed from intermediate C (Scheme 4).Open in a separate windowScheme 4Proposed catalytic cycle for the electrooxidative olefination of arenes.All of the performed experiments give a strong indication that a Pd(ii)/Pd(iv) cycle is involved in this electrochemical variant of the Fujiwara–Moritani reaction. Also, a palladium complex PdII(L1)4 was synthesised and characterised by X-ray crystallography (Fig. 3). This PdII(L1)4 complex was found to be a competent intermediate for the Pd-catalysed electrooxidative olefination of arenes.Open in a separate windowFig. 3Single X-ray crystal structure of Pd-complex [Pd(L1)4].25Based on these results and literature precedence,23 a plausible Pd(ii/iv)-catalytic cycle is proposed for the electro-oxidative olefination of simple arenes (Scheme 4). The catalytic cycle starts with the anodic oxidation of the Pd(ii) catalyst A to form a Pd(iii) intermediate B. Arene C (sp2)–H bond activation delivers the organopalladium complex C which is converted to the Pd(iv) species D by anodic oxidation. Next, olefin coordination to form E followed by migratory insertion results in the formation of another organopalladium intermediate F. Finally, β-hydride elimination followed by reduction of Pd furnishes the olefinated product 1 and the Pd(ii) catalyst A is regenerated.  相似文献   

20.
Pd/Cu-Catalyzed amide-enabled selectivity-reversed borocarbonylation of unactivated alkenes     
Fu-Peng Wu  Xiao-Feng Wu 《Chemical science》2021,12(30):10341
The addition reaction between CuBpin and alkenes to give a terminal boron substituted intermediate is usually fast and facile. In this communication, a selectivity-reversed procedure has been designed and established. This selectivity-reversed borocarbonylation reaction is enabled by a cooperative action between palladium and copper catalysts and proceeds with complete regioselectivity. The key to the success of this transformation is the coordination of the amide group and slower CuBpin formation by using KHCO3 as the base. A wide range of β-boryl ketones were produced from terminal unactivated aliphatic alkenes and aryl iodides. Further synthetic transformations of the obtained β-boryl ketones have been developed as well.

A selectivity-reversed borocarbonylation reaction has been developed with complete regioselectivity.

The catalytic borocarbonylation of alkenes represents a novel synthetic tool for the simultaneous installation of boron and carbonyl groups across alkenes, enabling rapid construction of molecules with high complexity from abundant alkenes. In particular, the obtained organoboron compounds are versatile synthetic intermediates that can be readily converted into a wide range of functional groups with complete stereospecificity.1 Consequently, several catalytic systems have been developed to diversify the molecular frameworks through carbonylative borofunctionalization.2 In general, carbonylative borofunctionalization of alkenes proceeds via an alkyl-copper intermediate, which was produced by the addition of CuBpin to the terminal position of the alkene starting material,3 followed by CO insertion and other related steps. A new C–B bond is formed at the terminal position of the alkene and a carbonyl group has been installed at the β-position simultaneously (Scheme 1a). However, in contrast to the progress in the borocarbonylation, a selectivity-reversed procedure (the boryl group is installed at the internal position) to give β-boryl ketone products is still unprecedented.Open in a separate windowScheme 1Strategies for borocarbonylation of activated alkenes.Recently, several attractive strategies have emerged for the borofunctionalization of unactivated alkenes to give β-boryl products.4–7 In 2015, Fu, Xiao and their co-workers established a copper-catalyzed regiodivergent alkylboration of alkenes.4a In the same year, Miura and Hirano''s group reported a copper-catalyzed aminoboration of terminal alkenes.4b In these two attractive procedures, the regioselectivity was controlled by the ligand applied. More recently, an intermolecular 1,2-alkylborylation of alkenes was described by Ito''s research group.5 A radical-relay strategy was used to achieve the targeted regioselective addition. Furthermore, Engle and co-workers explored a palladium-catalyzed 1,2-carboboration and -silylation reaction of alkenes.6 Stereocontrol can be achieved in this new procedure with the assistance of a chiral auxiliary which is a coordinating group in this case.Inspired by these pioneering studies, we assumed that if the reaction could be initiated by the insertion of an acylpalladium complex into alkenes, followed by transmetalation with CuBpin before reductive elimination, β-boryl ketones can finally be produced (Scheme 1b). However, due to the inherent reactivity of the palladium species toward alkenes, olefin substrates were usually restricted to styrenes and a large excess of them is typically required (>6 equivalents).8,9 Therefore, the critical part of the reaction design is to promote the reaction of the acylpalladium intermediate with alkenes faster than the insertion of CuBpin into olefins. One of the ideas is taking advantage of the coordinating group to transform the reaction from intermolecular to intramolecular. Among the developed directing groups,10 8-aminoquinoline (AQ) is interesting and has been relatively well studied by various groups in a number of novel transformations.11–13 Although the AQ directing group contains a NH group which can participate in intramolecular C–N bond formation,14 we believe that the selectivity-reversed borocarbonylation of alkenes can potentially be achieved through cooperative Pd/Cu catalysis. Then, valuable β-boryl ketones can be produced from readily available substrates directly and effectively.To test the viability of our design on selectivity-reversed borocarbonylation of alkenes, N-(quinolin-8-yl)pent-4-enamide (1a), iodobenzene (2a), and bis(pinacolato)diboron (B2pin2) were chosen as model substrates for systematic studies. As shown in 15 In the testing of palladium precursors, allylpalladium chloride dimer proved to be the best palladium catalyst for this reaction, affording 3a in 41% yield () and tend to generate the by-product β-aminoketone. Xantphos was found to be superior to the other tested bidentate ligands ( Entry[Pd]LigandCuBaseYield of 3a (%)1Pd(TFA)2 L1 IMesCuClK2CO3292Pd(OAc)2 L1 IMesCuClK2CO3343[Pd(η3-C3H5)Cl]2 L1 IMesCuClK2CO3414[Pd(cinnamyl)Cl]2 L1 IMesCuClK2CO3365[Pd(η3-C3H5)Cl]2 L1 IPrCuClK2CO306[Pd(η3-C3H5)Cl]2 L1 CuClK2CO3337[Pd(η3-C3H5)Cl]2 L1 CuBrK2CO3418[Pd(η3-C3H5)Cl]2 L1 CuIK2CO3509[Pd(η3-C3H5)Cl]2 L2 CuIK2CO33810[Pd(η3-C3H5)Cl]2 L3 CuIK2CO34711[Pd(η3-C3H5)Cl]2 L4 CuIK2CO3012[Pd(η3-C3H5)Cl]2 L5 CuIK2CO3013[Pd(η3-C3H5)Cl]2 L6 CuIK2CO3<214[Pd(η3-C3H5)Cl]2 L7 CuIK2CO31015[Pd(η3-C3H5)Cl]2 L8 CuIK2CO31216[Pd(η3-C3H5)Cl]2 L1 CuIKHCO358 (51)b17[Pd(η3-C3H5)Cl]2 L1 CuIK2HPO42618[Pd(η3-C3H5)Cl]2 L1 CuINaHCO3019[Pd(η3-C3H5)Cl]2 L1 CuINaOtBu1120c[Pd(η3-C3H5)Cl]2 L1 CuIKHCO3<521[Pd(η3-C3H5)Cl]2 L7 CuIKHCO340 Open in a separate windowaAll reactions were carried out on a 0.1 mmol scale with alkene (0.1 mmol) and aryl iodide (2.0 equiv.). Yields were determined by 1H NMR analysis of the crude reaction mixture using 1,3,5-trimethoxybenzene as the internal standard.bIsolated yield.cXantphos (10 mol%).With the optimized reaction conditions in hand, we examined the scope of this selectivity-reversed borocarbonylation with various unactivated alkenes and aryl iodides toward the synthesis of β-boryl ketones ( Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond even when there are two C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bonds in the amide substrates (4i and 4j). In addition, sterically hindered 4-pentenoic amide was subjected to the optimized reaction conditions, and the corresponding product was formed in 42% yield (4k). Furthermore, mono-substitution at the β-position of 4-pentenoic amides could also be employed, affording the corresponding products in moderate yields (4l and 4m). Iodoarenes containing more complex substrates such as L-menthol, L-borneol, vitamin E, diacetonfructose and nerol were also competent substrates and gave moderate to good yields of the corresponding products. Finally, no desired product could be detected when 3-butenoic amide, 2-vinylbenzamide or internal alkene was tested under our standard conditions.Substrate scope for the synthesis of β-boryl ketonesa
Open in a separate windowaAll reactions were carried out on a 0.1 mmol scale. Alkenes (0.1 mmol), aryl iodides (2.0 equiv.), B2pin2 (1.5 equiv.), CuI (10 mol%), [Pd(η3-C3H5)Cl]2 (2.5 mol%), xantphos (5 mol%), KHCO3 (2.0 equiv.), CO (10 bar), and DMSO (0.2 M) were stirred at 70 °C for 18 h. The dr value given was determined by 1H NMR.To demonstrate the synthetic utilities of the obtained borocarbonylation products, a series of further synthetic transformations of the β-boryl ketones were performed (Scheme 2). From a practical point of view, the reaction can be easily performed on the gram-scale and gave the target product 3m in 67% yield. β-Hydroxyl ketone 6a (CCDC: 2079475; determined by X-ray crystallography and the ORTEP drawing with 50% thermal ellipsoids) was produced in 95% yield by oxidation of the parent β-boryl ketone 3a. Furthermore, the C–B bond can be easily converted into a C–N bond, affording β-aminoketone 6b in 60% yield. Upon the reduction reaction of 3m with NaBH4, the corresponding reduced oxaborole amide 6c could be isolated in 70% yield. Finally, a two-step transamination process was performed to remove the AQ group.16Open in a separate windowScheme 2Diversification of β-boryl ketones.To gain some insight into the mechanism of this selectivity-reversed borocarbonylation of alkenes, several control experiments were performed. The target product 3a was not formed, instead byproduct 6b was obtained in 40% yield, in the case without xantphos. Possible explanations for this result are: (i) the bidentate directing group AQ increases the stability of Pd(ii) species and promotes the carbonylation step; (ii) the role of xantphos is to coordinate to C(sp3)–Pd(ii) species after its formation and inhibit the formation of the C–N bond to give byproduct 6b (Scheme 3a). In addition, copper and B2pin2 were proven to be important, and KHCO3 was essential for the carbonylation step (Scheme 3b). Analysis of the copper system in the absence of palladium and iodobenzene revealed that alkenes failed to undergo CuBpin insertion under this condition and no hydroboration products could be detected after work-up (Scheme 3c). Additionally, alkenes without the directing group were also tested under our standard conditions, and no reaction occurred.Open in a separate windowScheme 3Control experiments.Although we did not observe compound 7a during our optimization and substrate scope processes, even after stopping the reaction after 8 hours, we tested the possibility that 7a might act as an intermediate. When 7a was subjected to this transformation, the product 3a was delivered in 24% yield and 6b was generated in 37% (Scheme 3d). No significant difference in the yield outcome was observed when xantphos was added. Additionally, in our deuterated substrate testing, the amount of the deuterated product obtained is lower than the theoretical value (Scheme 3e). Thus, a pathway of β-H elimination followed by hydroboration could be involved as well. However, we believe the direct reaction between palladium and copper intermediates is the main one for this procedure due to the proven importance of the AQ group and the known achievements of copper-catalyzed hydroboration of enones, even with enantioselective versions.17On the basis of the above results and related literature studies,7,11–14 a possible reaction pathway is proposed (Scheme 4). Initially, the AQ directing group coordinates with Pd0, which produces the active AQ-Pd0 catalyst I. This is followed by oxidative addition to aryl iodides to generate PdII species II, and then by base promoted iodine dissociation to form complex III. After the CO insertion step, the acyl-PdII species IV coordinates with the alkene and undergoes migratory insertion to generate C(sp3)–PdII intermediate V, which is stabilized by the xantphos ligand and AQ directing group. Subsequently, C(sp3)–PdII complex V reacts especially with CuBpin to give the desired product β-boryl ketone and regenerate the Pd(0) complex. Finally, ligand exchange of Pd0Ln regenerates AQ-Pd0I for the next catalytic cycle. Additionally, another minor pathway that involves the carbonylative Heck reaction to give an enone derivative, followed by its hydroboration to give the final product could be included as well.Open in a separate windowScheme 4Proposed catalytic cycle.In summary, a novel Pd/Cu catalyzed amide-directed selectivity-reversed borocarbonylation for the selective synthesis of β-boryl ketones from terminal alkenes has been developed. Various aryl iodides and aliphatic alkenes were transformed into the desired β-boryl ketones in moderate to excellent yields. In this catalyst system, the assistance from the AQ directing group is essential for successful reaction design.  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