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 91 毫秒
1.
Aurone-derived azadienes are well-known four-atom synthons for direct [4 + n] cycloadditions owing to their s-cis conformation as well as the thermodynamically favored aromatization nature of these processes. However, distinct from this common reactivity, herein we report an unusual formal migrative annulation with siloxy alkynes initiated by [2 + 2] cycloaddition. Unexpectedly, this process generates benzofuran-fused nitrogen heterocyclic products with formal substituent migration. This observation is rationalized by less common [2 + 2] cycloaddition followed by 4π and 6π electrocyclic events. DFT calculations provided support to the proposed mechanism.

A HNTf2-catalyzed formal migrative cycloaddition of aurone-derived azadienes with siloxy alkynes has been developed to provide access to benzofuran-fused dihydropyridines.

Benzofuran is an important scaffold in biologically important natural molecules and therapeutic agents.1 Among them, benzofuran-fused nitrogen heterocycles are particularly noteworthy owing to their broad spectrum of bioactivities for the treatment of various diseases (Fig. 1).2 Consequently, the development of efficient methods for their assembly has been a topic receiving enthusiastic attention from synthetic chemists.3 Notably, aurone-derived azadienes (e.g., 1) have been extensively employed as precursors toward these skeletons owing to their easy availability and versatile reactivity (Scheme 1a).3 The polarized conjugation system, combined with the preexisting s-cis conformation, has enabled them to serve as ideal annulation partners for the synthesis of nitrogen heterocycles of variable ring sizes. Moreover, the aromatization nature of these processes by forming a benzofuran ring provides additional driving force for them to behave as a perfect four-atom synthon for [4 + n] cycloaddition.3 In contrast, the use of such species as a two-atom partner for [2 + n] cycloaddition has been less developed.3c,k,4 Herein, we report a new migrative annulation leading to benzofuran-fused dihydropyridines of unexpected topology (Scheme 1b, with formal R2 migration), which is initiated by the less common [2 + 2] cycloaddition.Open in a separate windowFig. 1Benzofuran-fused N-heterocyclic natural and bioactive molecules.Open in a separate windowScheme 1Synthesis of benzofuran-fused nitrogen heterocycles.Siloxy alkynes are another important family of building blocks in organic synthesis.5–8 The presence of a highly polarized C–C triple bond enables such molecules to serve as versatile two-carbon cycloaddition partners in various annulation reactions.5–7 In the above context and in continuation of our interest in the study of such electron-rich alkynes,7 we envisioned that the reaction between aurone-derived azadienes 1 and siloxy alkynes 2 should lead to facile electron-inversed [4 + 2] cycloaddition to form benzofuran-fused dihydropyridine products (Scheme 1b). Interestingly, the expected product 3′ from direct [4 + 2] cycloaddition was not observed. Instead, a dihydropyridine product 3 with formal R2 migration was observed. Careful analysis of the mechanism suggested that a [2 + 2] cycloaddition followed by 4π and 6π electrocyclic steps might be responsible for this unexpected product topology (vide infra).We began our investigation with the model substrates 1a and 2a, which were easily prepared in one step from aurone and 1-hexyne, respectively.8 Various Lewis acids were initially examined as potential catalysts for this cycloaddition (Table 1). Unfortunately, common Lewis acids (e.g., TiCl4, BF3·OEt2, Sc(OTf)3, In(OTf)3, and AgOTf) were all ineffective (entries 1–5). Substrate decomposition into an unidentifiable mixture was typically observed. However, further screening indicated that AgNTf2 served as an effective catalyst, leading to benzofuran-fused dihydropyridine 3a in 44% yield (entry 6). Careful analysis by X-ray crystallography confirmed that it was not formed by simple [4 + 2] cycloaddition, as the positions of the phenyl and the siloxy groups were switched (vs. the expected topology). The distinct catalytic performance of AgNTf2 (vs. AgOTf) suggested that the triflimide counter anion Tf2N might be important. However, further screening of various metal triflimide salts did not improve the reaction efficiency (entry 7). Instead, we were delighted to find that the corresponding Brønsted acid HNTf2 served as a better catalyst (57% yield, entry 8). However, triflic acid (TfOH) led to no desired product in spite of complete conversion (entry 9). After considerable efforts in the optimization of other reaction parameters, an improved yield of 75% was obtained with 2.5 mol% of HNTf2 and 2.5 equivalents of 2a at 60 °C (entry 10). Solvent screening indicated that the reaction proceeded faster in DCE with comparable yield (entry 11). However, other solvents were all inferior (entries 12–15). Finally, with a reversed order of addition of the two reactants, the yield was slightly improved (entry 16). We believe that this might be related to the relative decomposition rates of the substrates.Reaction conditionsa
EntryCatalystSolventTime (h)Yield (%)
1TiCl4DCM90
2BF3·OEt2DCM90
3Sc(OTf)3DCM90
4In(OTf)3DCM90
5AgOTfDCM90
6AgNTf2DCM944
7Sc(NTf2)3DCM90
8HNTf2DCM957
9HOTfDCM90
10bHNTf2DCM4275
11bHNTf2DCE1872
12bHNTf2CHCl31820
13bHNTf2THF180
14bHNTf2MeCN180
15bHNTf2EtOAc180
16b,cHNTf2DCE1881 (76)d
Open in a separate windowa 2a (0.06 mmol) was added to the solution of 1a (0.05 mol) and the catalyst (10 mol%). Yield was determined by analysis of the 1H NMR spectrum of the crude mixture using CH2Br2 as an internal standard.bRun with 2.5 mol% catalyst and 2.5 equiv. of 2a at 60 °C.c 1a was added into the solution of 2a and the catalyst.dYield in parentheses was isolated yield.With the optimized conditions, we examined the reaction scope. A range of aurone-derived azadienes with different electron-donating and electron-withdrawing substituents at various positions smoothly participated in this formal migrative cycloaddition process with siloxy alkyne 2a (Scheme 2). The corresponding benzofuran-fused dihydropyridine products 3 were formed with excellent selectivity and moderate to good efficiency. A thiophene unit was also successfully incorporated into the product (3h). However, substitution with a pyridinyl group shut down the reactivity, even with 1.1 equivalents of HNTf2. Other siloxy alkynes bearing different alkyl substituents on the triple bond were also good reaction partners, except that these reactions were more efficient when the catalyst loading was increased to 10 mol% (Table 2). Unfortunately, direct aryl substitution on the alkyne triple bond resulted in essentially no reaction (entry 7). Notably, in spite of the strong acidic conditions, various functional groups, such as TIPS-protected alcohol (3p) and acetal (3c), were tolerated. Moreover, increasing steric hindrance in close proximity to the reaction centers (e.g., tBu group in 3i and 3r) did not obviously affect the reaction efficiency.Scope of siloxyl alkynesa
EntryR 3 Yield (%)
1 3m 66
2 3n 74
3 3o 53b
4 3p 64
5 3q 58
6 3r 62
7 3s <5
Open in a separate windowaConditions: 1d (0.3 mmol), 2 (0.75 mmol), HNTf2 (10 mol%), DCE (3 mL), 60 °C. Isolated yield.bRun with 2.5 mol% of HNTf2.Open in a separate windowScheme 2Scope of aurone-derived azadienes. Conditions: 1 (0.3 mmol), 2a (0.75 mmol), HNTf2 (2.5 mol%), DCE (3.0 mL), 60 °C. Isolated yield.Owing to the electron-rich silyl enol ether motif, the benzofuran-fused dihydropyridine products can be transformed into other related heterocycles upon treatment with electrophiles. For example, deprotection of the silyl group in 3d with TBAF in the presence of water produced ketone 4a (eqn (1)). In the presence of NBS or NCS, the corresponding bromoketone 4b and chloroketone 4c were obtained, respectively (eqn (2)). These reactions were both efficient and highly diastereoselective. The structures of 4b and 4c were also confirmed by X-ray crystallography. Moreover, deprotection of the N-tosyl group with Li/naphthalene followed by air oxidation led to the highly-substituted benzofuran-fused pyridine 5, the core structure of a family of bioactive molecules (eqn (3)).2A possible mechanism is proposed to rationalize the unusual formal migrative process (Scheme 3). The reaction begins with LUMO-lowering protonation of the aurone-derived azadiene 1 by HNTf2.9 Then, the electron-rich alkyne attacks the resulting activated iminium intermediate I, leading to ketenium ion II after intermolecular C–C bond formation. Subsequent intramolecular cyclization from the electron-rich enamine motif to the electrophilic ketenium unit forms oxetene III. The formation of this oxetene can also be considered as a [2 + 2] cycloaddition of the two reactants.6ad,11 Subsequent 4π-electrocyclic opening of oxetene III affords azatriene IV. Further 6π-electrocyclic closing leads to the observed product 3. This observed product topology is fully consistent with this pathway. It is worth noting that the excellent performance with HNTf2 might be attributed to the low nucleophilicity and good compatibility of its counter anion with the highly electrophilic cationic intermediates (e.g., ketenium II) in this process. We have also carried out DFT studies. The results indicated that the proposed pathway is energetically viable and consistent with the experimental data (Scheme 3 and Fig. S1). Moreover, some other possible pathways that engage the nitrogen atom in intermediate II to directly attack the ketenium in a [4 + 2] mode were explored. However, no reasonable transition state could be located (Fig. S2). Thus, the origin of preference toward [2 + 2] cycloaddition remains unclear.Open in a separate windowScheme 3Proposed mechanism and free energies (in kcal mol−1) computed at the M06-2X(D3)/6-311G(d,p)-SMD//M06-2X/6-31G(d) level of theory.We also prepared TIPSNTf2 and examined its catalytic activity in this reaction since it is known that such a Lewis acid might be generated in situ.10 However, no reaction was observed when TIPSNTf2 was used in place of HNTf2, suggesting that it is unlikely the actual catalyst. Finally, in order to probe the nature of the substituent migration (intermolecular vs. intramolecular), we carried out a cross-over experiment (Scheme 4). Under the standard conditions, the reaction using a 1 : 1 mixture of 1d and 1k led to exclusive formation of 3d and 3k, without detection of any cross-over products. This result is consistent with the proposed intramolecular migration pathway.Open in a separate windowScheme 4Cross-over experiment.In conclusion, we have discovered an unusual formal migrative cycloaddition of aurone-derived azadienes with siloxy alkynes. In the presence of a catalytic amount of HNTf2, this reaction provided expedient access to a range of useful benzofuran-fused dihydropyridine products with unexpected topology, distinct from normal [4 + 2] cycloaddition. Although aurone-derived azadienes are ideal four-atom synthons for direct [4 + n] cycloaddition, the present process is initiated by less common [2 + 2] cycloaddition, which is critical for the observed product formation. Subsequent electrocyclic opening and cyclization steps provide a reasonable rationale. The heterocyclic products generated from this process are precursors toward other useful structures, such as benzofuran-fused pyridines.  相似文献   

2.
Palladium-catalyzed regioselective di- or mono-arylation of o-carboranes was achieved using weakly coordinating amides at room temperature. Therefore, a series of B(3,4)-diarylated and B(3)-monoarylated o-carboranes anchored with valuable functional groups were accessed for the first time. This strategy provided an efficient approach for the selective activation of B(3,4)–H bonds for regioselective functionalizations of o-carboranes.

B–H: site-selective B(3,4)–H arylations were accomplished at room temperature by versatile palladium catalysis enabled by weakly coordinating amides.

o-Carboranes, icosahedral carboranes – three-dimensional arene analogues – represent an important class of carbon–boron molecular clusters.1 The regioselective functionalization of o-carboranes has attracted growing interest due to its potential applications in supramolecular design,2 medicine,3 optoelectronics,4 nanomaterials,5 boron neutron capture therapy agents6 and organometallic/coordination chemistry.7 In recent years, transition metal-catalyzed cage B–H activation for the regioselective boron functionalization of o-carboranes has emerged as a powerful tool for molecular syntheses. However, the 10 B–H bonds of o-carboranes are not equal, and the unique structural motif renders their selective functionalization difficult, since the charge differences are very small and the electrophilic reactivity in unfunctionalized o-carboranes reduces in the following order: B(9,12) > B(8,10) > B(4,5,7,11) > B(3,6).8 Therefore, efficient and selective boron substitution of o-carboranes continues to be a major challenge.Recently, transition metal-catalyzed carboxylic acid or formyl-directed B(4,5)–H functionalization of o-carboranes has drawn increasing interest, since it provides an efficient approach for direct regioselective boron–carbon and boron–heteroatom bond formations (Scheme 1a),9 with major contributions by the groups of Xie,10 and Yan,11 among others.12 Likewise, pyridyl-directed B(3,6)–H acyloxylations (Scheme 1b),13 and amide-assisted B(4,7,8)–H arylations14 (Scheme 1c) have been enabled by rhodium or palladium catalysis, respectively.15,16 Despite indisputable progress, efficient approaches for complementary site-selective functionalizations of o-carboranes are hence in high demand.17 Hence, metal-catalyzed position-selective B(3,4)–H functionalizations of o-carboranes have thus far not been reported.Open in a separate windowScheme 1Chelation-assisted transition metal-catalyzed cage B–H activation of o-carboranes.Arylated compounds represent key structural motifs in inter alia functional materials, biologically active compounds, and natural products.18 In recent years, transition metal-catalyzed chelation-assisted arylations have received significant attention as environmentally benign and economically superior alternatives to traditional cross-coupling reactions.19 Within our program on sustainable C–H activation,20 we have now devised a protocol for unprecedented cage B–H arylations of o-carboranes with weak amide assistance, on which we report herein. Notable features of our findings include (a) transition metal-catalyzed room temperature B–H functionalization, (b) high levels of positional control, delivering B(3,4)-diarylated and B(3)-monoarylated o-carboranes, and (c) mechanistic insights from DFT computation providing strong support for selective B–H arylation (Scheme 1d).We initiated our studies by probing various reaction conditions for the envisioned palladium-catalyzed B–H arylation of o-carborane amide 1a with 1-iodo-4-methylbenzene (2a) at room temperature (Tables 1 and S1). We were delighted to observe that the unexpected B(3,4)-di-arylated product 3aa was obtained in 59% yield in the presence of 10 mol% Pd(OAc)2 and 2 equiv. of AgTFA, when HFIP was employed as the solvent, which proved to be the optimal choice (entries 1–5).21 Control experiments confirmed the essential role of the palladium catalyst and silver additive (entries 6–7). Further optimization revealed that AgOAc, Ag2O, K2HPO4, and Na2CO3 failed to show any beneficial effect (entries 8–11). Increasing the reaction temperature fell short in improving the performance (entries 12 and 13). The replacement of the amide group in substrate 1a with a carboxylic acid, aldehyde, ketone, or ester group failed to afford the desired arylation product (see the ESI). We were pleased to find that the use of 1.0 equiv. of trifluoroacetic acid (TFA) as an additive improved the yield to 71% (entry 14). To our delight, replacing the silver additive with Ag2CO3 resulted in the formation of B(3)–H mono-arylation product 4aa as the major product (entries 15–16).Optimization of reaction conditionsa
EntryAdditiveSolventYield of 3aa/%Yield of 4aa/%
1AgTFAPhMe00
2AgTFADCE00
3AgTFA1,4-Dioxane00
4AgTFATFE213
5AgTFAHFIP594
6AgTFAHFIP00b
7HFIP00
8AgOAcHFIP5<3
9Ag2OHFIP<3<3
10K2HPO4HFIP00
11Na2CO3HFIP00
12AgTFAHFIP534c
13AgTFAHFIP423d
14 AgTFA HFIP 71 <3 e
15Ag2CO3HFIP934f
16 Ag 2 CO 3 HFIP 5 55 f , g
Open in a separate windowaReaction conditions: 1a (0.20 mmol), 2 (0.48 mmol), Pd(OAc)2 (10 mol%), additive (0.48 mmol), solvent (0.50 mL), 25 °C, 16 h, and isolated yield.bWithout Pd(OAc)2.cAt 40 °C.dAt 60 °C.eTFA (0.2 mmol) was added.f 1a (0.20 mmol), 2a (0.24 mmol), Pd(OAc)2 (5.0 mol%), and Ag2CO3 (0.24 mmol).g 2a was added in three portions every 4 h. DCE = dichloroethane, TFE = 2,2,2-trifluoroethanol, HFIP = hexafluoroisopropanol, and TFA = trifluoroacetic acid.With the optimized reaction conditions in hand, we probed the scope of the B–H di-arylation of o-carboranes 1a with different aryl iodides 2 (Scheme 2). The versatility of the room temperature B(3,4)–H di-arylation was reflected by tolerating valuable functional groups, including bromo, chloro, and enolizable ketone substituents. The connectivity of the products 3aa and 3ab was unambiguously verified by X-ray single crystal diffraction analysis.22Open in a separate windowScheme 2Cage B(3,4)–H di-arylation of o-carboranes.Next, we explored the effect exerted by the N-substituent at the amide moiety (Scheme 3). Tertiary amides 1b–1f proved to be suitable substrates with optimal results being accomplished with substrate 1a. The effect of varying the cage carbon substituents R1 on the reaction''s outcome was also probed, and both aryl and alkyl substituents gave the B–H arylation products and the molecular structures of the products 3dd, 3ea and 3fa were fully established by single-crystal X-ray diffraction.Open in a separate windowScheme 3Effect of substituents on B–H diarylation. aAt 50 °C.The robustness of the palladium-catalyzed B–H functionalization was subsequently investigated for the challenging catalytic B–H monoarylation of o-carboranes (Scheme 4). The B(3)–H monoarylation, as confirmed by single-crystal X-ray diffraction analysis of products 4aa and 4ai, proceeded smoothly with valuable functional groups, featuring aldehyde and nitro substituents, which should prove invaluable for further late-stage manipulation.Open in a separate windowScheme 4Cage B(3)–H mono-arylation of o-carboranes.To elucidate the palladium catalysts'' working mode, a series of experiments was performed. The reactions in the presence of TEMPO or 1,4-cyclohexadiene produced the desired product 3aa, which indicates that the present B–H arylation is less likely to operate via radical intermediates (Scheme 5a). The palladium catalysis carried out in the dark performed efficiently (Scheme 5b). Compound 4aa could be converted to di-arylation product 3aa with high efficiency, indicating that 4aa is an intermediate for the formation of the diarylated cage 3aa (Scheme 5c).Open in a separate windowScheme 5Control experiments.To further understand the catalyst mode of action, we studied the site-selectivity of the o-carborane B–H activation for the first B–H activation at the B3 versus B4 position and for the second B–H activation at the B4 versus B6 position using density functional theory (DFT) at the PBE0-D3(BJ)/def2-TZVP+SMD(HFIP)//TPSS-D3(BJ)/def2-SVP level of theory (Fig. 1). Our computational studies show that the B3 position is 5.8 kcal mol−1 more favorable than the B4 position for the first B–H activation, while the B4 position is 3.4 kcal mol−1 more favorable than the B6 position for the second B–H activation. It is noteworthy that here the interaction between AgTFA and a cationic palladium(ii) complex was the key to success, being in good agreement with our experimental results (for more details, see the ESI).Open in a separate windowFig. 1Computed relative Gibbs free energies in kcal mol−1 and the optimized geometries of the transition states involved in the B–H activation at the PBE0-D3(BJ)/def2-TZVP+SMD(HFIP)//TPSS-D3(BJ)/def2-SVP level of theory. (a) First B–H activation transition states at the B3 and B4 positions. (b) Second B–H activation transition states at the B4 and B6 positions. Irrelevant hydrogen atoms in the transition states are omitted for clarity and the bond lengths are given in Å.A plausible reaction mechanism is proposed which commences with an organometallic B(3)–H activation of 1a with weak assistance of the amide group and assistance by AgTFA to form the cationic intermediate I (Scheme 6). Oxidative addition with the aryl iodide 2 affords the proposed cationic palladium(iv) intermediate II, followed by reductive elimination to give the B(3)-mono-arylation product 4aa. Subsequent B(4)-arylation occurs assisted by the weakly coordinating amide to generate the B(3,4)-di-arylation product 3aa. Due to the innate higher reactivity of the B(4)–H bond in intermediate 4aa – which is inherently higher than that of the B(6)–H bond – the B(3,6)-di-arylation product is not formed.Open in a separate windowScheme 6Proposed reaction mechanism.In summary, room temperature palladium-catalyzed direct arylations at cage B(3,4) positions in o-carboranes have been achieved with the aid of weakly coordinating, synthetically useful amides. Thus, palladium-catalyzed B–H activations enable the assembly of a wealth of arylated o-carboranes. This method features high site-selectivity, high tolerance for functional groups, and mild reaction conditions, thereby offering a platform for the design and synthesis of boron-substituted o-carboranes. Our findings offer a facile strategy for selective activations of B(3,4)–H bonds, which will be instrumental for future design of optoelectronics, nanomaterials, and boron neutron capture therapy agents.  相似文献   

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

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

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

4.
5.
Sulfuric chloride is used as the source of the –SO2– group in a palladium-catalyzed three-component synthesis of sulfonamides. Suzuki–Miyaura coupling between the in situ generated sulfamoyl chlorides and boronic acids gives rise to diverse sulfonamides in moderate to high yields with excellent reaction selectivity. Although this transformation is not workable for primary amines or anilines, the results show high functional group tolerance. With the solving of the desulfonylation problem and utilization of cheap and easily accessible sulfuric chloride as the source of sulfur dioxide, redox-neutral three-component synthesis of sulfonamides is first achieved.

Sulfuric chloride is used as the source of the –SO2– group in a palladium-catalyzed three-component synthesis of sulfonamides.

Since its development in the 1970s,1 Suzuki–Miyaura coupling has become a widely used synthetic step in diverse areas. With two of the most widely sourced materials, organoborons and alkyl/aryl halides, a number of C–C coupling reactions are established and the Suzuki–Miyaura reaction has successfully acted as the key step in the synthesis of medicines and agrochemicals.2In addition to the well-known aryl halides and esters, various other substrates such as acid chlorides,3 anhydrides,4 diazonium salts5 and sulfonyl chlorides6 were also reported for the coupling in the past decades. As far as acid chlorides are concerned, carbamoyl chlorides were successfully transformed to the corresponding benzamides in the early years of the 21st century.7 However, the use of sulfamoyl chlorides as coupling partners is challenging due to the strong electron-withdrawing properties of the sulfonyl group, which cause the tendency of desulfonylation to form tertiary amines.Synthesis of sulfonyl-containing compounds, especially sulfones and sulfonamides, via the insertion of sulfur dioxide has been extensively studied during the last decade.8 A series of sulfur-containing surrogates have been developed as the source of the –SO2– group. Willis and co-workers first reported the use of DABCO·(SO2)2, a bench-stable solid adduct of DABCO and gaseous SO2 discovered by Santos and Mello,9a as the source of sulfur dioxide in the synthesis of sulfonylhydrazines.9b Soon after, alkali metal metabisulfites were found to provide sulfur dioxide for the formation of sulfonyl compounds.10 In the recent developments in this field, DABCO·(SO2)2 and metabisulfites have become the most popular SO2 surrogates for the insertion of sulfur dioxide.8 However, the practical applications of sulfur dioxide insertion reactions are limited by atom-efficiency problems and the unique properties of reactants. For instance, the three-component synthesis of aryl sulfonylhydrazines using aryl halides, SO2 surrogates and hydrazines by a SO2-doped Buchward–Hartwig reaction was realized in the earliest developments in this field.10 However, similar transformations from aryl halides and amines to the corresponding sulfonamides still remain unresolved (Scheme 1a).11,12Open in a separate windowScheme 1Synthetic approaches to sulfonamides.In order to provide a simple and efficient method for the three-component synthesis of aryl sulfonamides without the pre-synthesis of sulfonyl chlorides, many scientists have made various attempts. Interestingly, the use of arylboronic acids instead of aryl halides provided an alternative route. An oxidative reaction between boronic acids, DABCO·(SO2)2 and amines for the preparation of aryl sulfonamides at high temperature was realized,12 while reductive couplings of boronic acids, SO2 surrogates and nitroarenes were also reported (Scheme 1b).13 However, due to the reversed electronic properties of boronic acids from halides, additional additives and restrictions had to be considered. Extra oxidants and harsh conditions were usually used, and some of the transformations required “oxidative” substrates, such as nitroarenes and chloroamines.14Early in 2020, a reductive hydrosulfonamination of alkenes by sulfamoyl chlorides was reported,15 which gave us the inspiration to use in situ generated sulfamoyl chlorides as the electrophile for the synthesis of aryl sulfonamides by Suzuki–Miyaura coupling. In this way, sulfamoyl chlorides could be formed by nucleophilic substitution of an amine to sulfuric chloride, and the S(vi) central atom introduced into the reaction could reverse the electronic properties of the amine, which would eliminate the addition of oxidants (Scheme 1c). With the utilization of boronic acids as the coupling partner, a palladium-catalyzed Suzuki–Miyaura coupling could provide the sulfonamide products. Compared with traditional attempts, reversing the electronic properties of an amine from nucleophilic to electrophilic could reverse the whole reaction process, and two-step synthesis starting from the amine side could bypass the existing difficulty of S–N bond forming reductive elimination.12 Instead, a C–S bond formation could be the key for success (Scheme 2). In this proposed route, the presence of a base would be essential to remove the acid generated in situ during the reaction process. Additionally, we expected that the addition of a ligand would improve the oxidative addition of Pd(0) to sulfamoyl chloride, thus leading to the desired sulfonamide product.Open in a separate windowScheme 2Comparison between the traditional route and designed work.As designed based on our assumption, we used a commercialized sulfamoyl chloride intermediate A, which would be generated from morpholine 1a and SO2Cl2, to start our early investigations. The results showed that the direct Suzuki–Miyaura coupling of sulfamoyl chloride intermediate A and 2-naphthaleneboronic acid 2a mostly led to the generation of byproduct 3a′ with traditional phosphine ligands added to the reaction, and the desired product 3a was obtained in poor yields (Table 1, entries 1 and 2). It is known that an electron-rich ligand would enhance the oxidative addition of Pd(0) to the electrophile, and the bulky factor would facilitate the reductive elimination process. As expected, the yield of product 3a was increased significantly when electron-rich and bulky tris-(2,6-dimethoxyphenyl)phosphine was used as the ligand (Table 1, entry 3). Moreover, the reaction could proceed more efficiently by using a mixture of THF and MeCN as the co-solvent (Table 1, entry 4).Early investigations using morpholine-4-sulfonyl chloride A as the starting material
EntrySolventLigandYielda (%)
11,4-DioxanePtBu3·HBF414
2THFPtBu3·HBF423
3THFPAr3·Ar = 2,6-di-OMe–C6H357
4THF/MeCNPAr3·Ar = 2,6-di-OMe–C6H372
Open in a separate windowa 1H NMR yield obtained using 1,3,5-trimethoxybenzene as the internal standard.With that brief conclusion in hand, we then shifted our focus to the in situ generation of sulfamoyl chloride intermediate A in the reaction process, and a number of attempts were made with morpholine 1a and SO2Cl2 (for details, see the ESI). After careful measurement of product 3a and desulfonylated byproduct 3a′ generated during the transformation, the selective formation of compound 3a was realized and “standard conditions” were identified. By using PdCl2(PhCN)2 as the catalyst and Na2HPO4 as the base, the desired product 3a was isolated in 71% yield, giving the least amount of desulfonylated product 3a′ (Table 2, entry 1). The control experiment showed that 3a or 3a′ was not detected in the absence of the palladium catalyst (Table 2, entry 2). It was also observed that compound 3a′ could not be generated when SO2Cl2 was omitted (Table 2, entry 3), indicating that the byproduct wasn''t produced by the direct coupling of boronic acid and amine. Other changes to the catalyst, ligand, base or solvent all resulted in lower yields of compound 3a or higher yields of desulfonylated product 3a′ (Table 2, entries 4–7).Effects of variation of reaction parametersa
EntryVariation from “standard conditions”Yield of 3a′b (%)Yield of 3ab (%)
1None580 (69)
2No PdCl2(PhCN)2n.d.n.d.
3No SO2Cl2n.d.n.d.
4Pd(OAc)2 instead of PdCl2(PhCN)21380
5PPh3 instead of PAr31568
6K2CO3 instead of Na2HPO44323
7MeCN instead of THF/MeCN1663
Open in a separate windowaStandard conditions: morpholine 1a (0.2 mmol, 1.0 equiv.), SO2Cl2 (0.5 mmol, 2.5 equiv.), Et3N (0.53 mmol, 2.65 equiv.), 2-naphthaleneboronic acid 2a (0.4 mmol, 2.0 equiv.), Na2HPO4 (0.6 mmol, 3.0 equiv.), PdCl2(PhCN)2 (10 mol%), tris-(2,6-dimethoxyphenyl)phosphine (20 mol%), THF (1.0 mL)/MeCN (1.5 mL), 70 °C, 16 h. See the ESI for the detailed procedure.b 1H NMR yield obtained using 1,3,5-trimethoxybenzene as the internal standard. The isolated yield of entry 1 is shown in parentheses.With the “standard conditions” in hand, various secondary amines 1 and arylboronic acids 2 were subjected to the reaction for the exploration of substrate adaptability (Scheme 3). To our delight, most of the reactions proceeded smoothly, giving rise to the desired product 3 in moderate to high yields. Considering the scope of boronic acids, a number of para-, meta- and ortho-(3t) substituted boronic acids showed good reactivities. However, lower yields were observed for some substrates with electron-withdrawing substituents, providing more desulfonylated byproducts due to the electron-deficiency of the palladium intermediate. Aryl boronic acids with acid-sensitive Boc-substituted amine, oxidation-sensitive phenol, sulfide and vinyl substitution were all tolerated. It is noteworthy that bromo- and acetoxy-substrates could also be efficiently converted to the corresponding products 3f and 3r, showing quite high selectivity during the reaction process. A series of heteroaromatic products were afforded successfully as well, and compounds with indole, indazole, dibenzothiophene and pyridine were all compatible (3aa–3af).Open in a separate windowScheme 3Synthesis of sulfonamides via a palladium-catalyzed Suzuki–Miyaura coupling. Isolated yields.Subsequently, with respect to amines, 4-phenylboronic acid and 4-(methylthio)phenylboronic acid were selected as coupling partners based on their electronic properties and cost. Saturated cyclic products 3ah–3an were obtained in moderate yields, among which an α-amino acid derivative showed high reactivity, giving rise to product 3aj in 71% yield. Methylallylamine was transformed to the corresponding product 3ao smoothly, and thiomorpholine 1,1-dioxide was also tolerated under the conditions (3ap). Various sensitive groups including acetyl, Boc, Cbz and cyclopropylcarbonyl (3aq–3at) on amines remained intact during the transformation. However, the amine scope was limited, since the transformation failed to provide the corresponding products when primary amines or anilines were used as the substrates. We assumed that during the reaction process for the oxidative addition of the sulfamoyl chloride intermediate to the palladium catalyst, Pd–SO2–NHR would be formed when a primary amine was used. Thus, β-hydride elimination would occur instead of the desired process.Furthermore, the practicality of this method was also verified by gram-scale synthesis and late-stage functionalization (Scheme 4). The reaction worked smoothly on the 4.0 mmol scale, and reducing the loading amount of the palladium catalyst to 1 mol% showed no obvious impact on the transformation. With a boronic acid synthesized from estrone and desloratadine, an antihistamine drug used as the substrate, the target products 4a and 4b were achieved in moderate to good yields, showing potential possibilities for synthetic applications.Open in a separate windowScheme 4Gram-scale synthesis and late-stage functionalization.In conclusion, a redox-neutral three-component synthesis of sulfonamides is established through a palladium-catalyzed Suzuki–Miyaura coupling of sulfuric chloride, secondary amines and arylboronic acids. Sulfuric chloride is used as the source of sulfur dioxide, and the S(vi) linchpin makes the transformation possible without the assistance of oxidants. Although this transformation is not workable for primary amines or anilines, the results show high functional group tolerance and good selectivity. A clear reaction process is described, in which the in situ generated sulfamoyl chloride undergoes a palladium-catalyzed Suzuki–Miyaura reaction with boronic acids, giving rise to the corresponding sulfonamide products. Additionally, the desulfonylation problem is surmounted during the reaction process. With a boronic acid synthesized from estrone and an antihistamine drug, desloratadine, used as the substrate, the target products are achieved in moderate to good yields, showing potential possibilities for synthetic applications in organic chemistry and medicinal chemistry.  相似文献   

6.
A new catalytic asymmetric formal cross dehydrogenative coupling process for the construction of all-aryl quaternary stereocenters is disclosed, which provides access to rarely explored chiral tetraarylmethanes with excellent enantioselectivity. The suitable oxidation conditions and the hydrogen-bond-based organocatalysis have enabled efficient intermolecular C–C bond formation in an overwhelmingly crowded environment under mild conditions. para-Quinone methides bearing an ortho-directing group serve as the key intermediate. The precise loading of DDQ is critical to the high enantioselectivity. The chiral products have also been demonstrated as promising antiviral agents.

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

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

7.
Ruthenium-catalyzed formal sp3 C–H activation of allylsilanes/esters with olefins: efficient access to functionalized 1,3-dienes     
Dattatraya H. Dethe  Nagabhushana C. Beeralingappa  Saikat Das  Appasaheb K. Nirpal 《Chemical science》2021,12(12):4367
Ru-catalysed oxidative coupling of allylsilanes and allyl esters with activated olefins has been developed via isomerization followed by C(allyl)–H activation providing efficient access to stereodefined 1,3-dienes in excellent yields. Mild reaction conditions, less expensive catalysts, and excellent regio- and diastereoselectivity ensure universality of the reaction. In addition, the unique power of this reaction was illustrated by performing the Diels–Alder reaction, and enantioselective synthesis of highly functionalized cyclohexenone and piperidine and finally synthetic utility was further demonstrated by the efficient synthesis of norpyrenophorin, an antifungal agent.

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

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

8.
Redox-active benzimidazolium sulfonamides as cationic thiolating reagents for reductive cross-coupling of organic halides     
Weigang Zhang  Mengjun Huang  Zhenlei Zou  Zhengguang Wu  Shengyang Ni  Lingyu Kong  Youxuan Zheng  Yi Wang  Yi Pan 《Chemical science》2021,12(7):2509
Redox-active benzimidazolium sulfonamides as thiolating reagents have been developed for reductive C–S bond coupling. The IMDN-SO2R reagent provides a bench-stable cationic precursor to generate a portfolio of highly active N–S intermediates, which can be successfully applied in cross-electrophilic coupling with various organic halides. The employment of an electrophilic sulfur source solved the problem of catalyst deactivation and avoided odorous thiols, featuring practical conditions, broad substrate scope, and excellent tolerance.

Redox-active benzimidazolium sulfonamides as thiolating reagents have been developed for reductive C–S bond coupling.

The high frequency of sulfur-containing moieties in natural products,1 bioactive molecules,2 pharmaceuticals,3 organic materials,4 fragrances5 and asymmetric catalysis as chiral catalysts/ligands6 has triggered the best endeavours for the selective construction of C–S bonds. The conventional cross-coupling of thiols with aryl halides generally relies on the conversion of mercaptans to thiolates by means of transition-metal catalysis7 (such as Pd,8 Cu,9 and Ni10) and other metals,11 although these efforts were plagued by several drawbacks. The strong coordination of thiolates to metals often leads to catalyst deactivation and displays low efficiencies. Therefore, high catalyst loading, specific ligands, excessive heating and strong bases are often required to facilitate this transformation (Scheme 1a, left). Recent development using photochemical12 and electrochemical13 induced thiol radicals as a sulfur source could avoid the problem of catalyst poisoning, although restricted substrate scope was displayed (Scheme 1a, middle). Despite the progress made for C–S bond construction,14 the longstanding issues that exist in the above-mentioned strategies should not be overlooked.Open in a separate windowScheme 1Origin of the reaction design. (a) C–S formation from organic halides. (b) The preparation of the electrophilic thiolating reagent (R–S+). (c) This work: cationic active reagent for cross electrophilic coupling.Compared to classical cross-coupling processes, the nickel-catalyzed reductive cross-coupling of two electrophilic partners has emerged as a powerful tool for the replacement of air- and moisture-sensitive organometallic reagents.15 The cross electrophilic coupling for the construction of C–S bonds is more challenging due to the lack of sulfur sources and homo-coupling of organic halides (Scheme 1a, right). Several examples of reductive thiolation16 have been described using thiol derivatives as S+ sources (Scheme 1b). We speculated that readily accessible sulfonyl chloride as an electrophilic sulfur source could avoid the use of highly toxic thiols and significantly the substrate scope.17 However, the direct reduction of sulfonyl chloride inevitably resulted in dimerization to disulfide.17b Putatively, activated by an electron-deficient heterocycle such as imidazole, sulfonyl chloride could be assembled into a bench-stable cationic reagent (Scheme 1b). Benzimidazolium sulfonamides would be better electron acceptors18 and easily reduced by PPh3 to generate a highly reactive N–S+ species in situ. The positive charge of this intermediate is delocalized on both nitrogens of imidazole. Followed by the cleavage of the weak N–S bond (BDE ≈ 70 kcal mol−1),19 the electrophilic sulfur species can be captured by metal catalysts for cross-coupling. Herein, we have described a redox-active benzimidazolium sulfonamide reagent (IMDN-SO2R) for Ni-catalyzed reductive coupling of organic halides for a portfolio of C(sp)–S, C(sp2)–S and C(sp3)–S bond formations (Scheme 1c). This strategy avoids the formation of disulfide by-products and the use of organometallic reagents, which facilitates purification and enhances the functionality tolerance.To determine the suitable conditions for the reductive coupling of the cationic reagent with organic halides, we first studied the reductive thiolation of p-iodo-methoxybenzene (2a) and 1a (2 equiv.) with a survey of Ni catalysts in the presence of dtbbpy (20 mol%), PPh3 (2.5 equiv.), Zn powder (3.0 equiv.) and MgCl2 (2.0 equiv.) in DMA at 60 °C (Table 1). Ni(OTf)2 as a catalyst was able to promote the reductive process to afford the desired aryl thioether product 3a in 92% isolated yield (entry 1). When using Ni(cod)2, Ni(acac)2, Ni(OAc)2·4H2O and Ni(PCy3)2Cl2, lower yields were obtained (entries 2–5). Decreasing the 1a loading to 1.5 equiv., the yield of 3a reduced to 75% (entry 6). Switching 1a to a 2-phenyl substituted imidazolium sulfonamide reagent 1b, similar yields were achieved (entry 7). When decreasing both Ni(OTf)2 and dtbbpy loading to 10 mol%, the yield reduced to 75%. In the absence of Ni(OTf)2 or Zn powder, no desired product was observed (entries 9–10). Switching Zn powder to an organic reductant tetrakis-(dimethylamino)ethylene (TDAE), still a low conversion of the reaction was observed (entry 11). The results demonstrated that no organozinc reagents were generated. In the absence of the ligand, MgCl2 or PPh3, the reaction resulted in low yields, indicating that these ingredients were essential for this catalytic system (entries 12–14). Thus, the optimized conditions were selected for further investigation of the reaction scope.Optimization of the reaction conditions. Reaction condition: 2a (0.20 mmol), 1a (0.40 mmol, 2.0 equiv.), MgCl2 (2.0 equiv.), PPh3 (2.5 equiv.), Zn (3.0 equiv.), Ni(OTf)2 (20 mol%) and dtbbpy (20 mol%) in DMA (2 mL) at 60 °C under Ar. aCrude yields determined by 1H NMR spectroscopy using dibromomethane as an internal standard. bIsolated yield. dtbbpy = 4,4′-di-tert-butyl-2,2′-bipyridine. TDAE = tetrakis(dimethylamino)ethylene
EntryVariationYielda/%EntryVariationYielda/%
1None96 (92)b810 mol% [Ni]83
2Ni(acac)2429No Ni(OTf)2nd
3Ni(cod)26310No Znnd
4Ni(OAc)2·4H2O2111TDAE for Zn21
5Ni(PCy3)2Cl25712No dtbbpy46
61.5 equiv. 1a7513No MgCl239
72.0 equiv. 1b8814No PPh320
Open in a separate windowThe generality of the reductive thiolation was evaluated under the optimized conditions (Scheme 2). A wide range of aryl iodides containing either electron-withdrawing or electron-donating functionality were tolerated, delivering products with methoxy (3a, 3c), trifluoromethoxy (3b), methyl (3d), acetyl (3e), ester (3f), boronate (3g), and phenyl (3h) groups in good to excellent yields (63–94%). 5-Iodobenzo[d][1,3]dioxole and 2-iodofluorene also furnished the corresponding adducts (3i, 3j) in 91% and 82% yields, respectively. Notably, aryl iodides bearing sensitive groups such as amine (3k) and hydroxyl (3l) were engaged in the cross-coupling to forge the C–S bond in good yields. Aryl bromides (3a, 3m) were also found compatible. The scope of the heteroaryl halide coupling partner was also explored. Various five- and six-membered heterocycles, including thiophene (3n), pyrazole (3o), quinoline (3p), carbazole (3q), indole (3r), isothiazole (3s), benzofuran (3t), benzothiophene (3u), benzothiazole (3v), pyrimidine (3w), pyrazine (3x), and pyridine (3y, 3z) derived heteroaryl bromides and iodides were treated with 1a to produce the corresponding sulfides in good yields. In the cases of relatively unreactive organic chlorides, the corresponding coupling products (3aa–3cc) could be obtained in low yields under the standard conditions. Subsequently, we assessed the scope of aryl sulfonamides. Reagents 1c–1k containing electron-withdrawing and electron-donating substitutions on the benzene ring afforded the desired products in good to excellent yields (3dd–3ll). Sterically hindered imidazolium sulfonamides 2,4,6-trimethylated 1j and 2,4,6-triisopropylated 1k showed good compatibility in this reaction to give the corresponding products 3kk and 3ll in high yields.Open in a separate windowScheme 2Substrate scope of (hetero)aryl and benzimidazolium sulfonamides. aReaction was performed at 80 °C.The more challenging aliphatic halides have also been examined with the redox-active reagent 1a (Scheme 3). Primary and secondary alkyl halides yielded alkyl sulfides (5a–5k and 5l) in good to excellent yields. No dimerization side-product was observed. Functionalities including esters (5b, 5i, and 5j), sulfide (5c), alkene (5f), acetal (5h) and ether (5k) are tolerated. Some sensitive functional groups including silyl ether (5e) and organoboronate (5g) were well tolerated, provided in 75% and 57% yields, respectively. Alkenyl halides were also tolerated to afford 7a and 7b in good yields. In addition, alkynyl bromides were employed for reductive thiolation with benzimidazolium sulfonamides 1a and 1c–1i to afford the corresponding C(sp)–S bond coupling products (9a–9h) in moderate to good yields.Open in a separate windowScheme 3Substrate scope of alkyl, alkenyl and alkynyl bromide. aNi(PCy3)2Cl2 instead of Ni(OTf)2 and without dtbbpy. b4-(Trifluoromethyl)pyridine (0.2 mmol) and THF (2 mL) were used instead of dtbbpy and DMA.To demonstrate the synthetic potential of this cross-electrophile coupling, a gram scale reaction was performed with 1a and p-iodo-methoxybenzene 2a under the standard conditions (Scheme 4a). The reductive thiolation was also applied in the late-stage modification of biologically active molecules and synthesis of pharmaceutically active molecules. d-Glucose (10a), (+)-α-tocopherol (10b), rosin amine (10c), and 17a-methyl-drostanolone-derived (10d) alkyl halides were compatible to afford thiolated products in good yields (Scheme 4b). Treating benzimidazolium sulfonamide 1n with piperazine-derived iodobenzene 11 generated the thiolated intermediate 12 and deprotection with TFA afforded anti-depressive vortioxetine 13 (ref. 20) in 50% overall yields (Scheme 4c).Open in a separate windowScheme 4Further transformations. (a) Gram-scale experiment. (b) Late-stage modification of natural products. (c) The synthesis of vortioxetine using our protocol.To investigate the mechanism for this reductive thiolation, a series of control experiments have been carried out. First, the treatment of benzimidazolium sulfonamide 1a with PPh3 furnished the key N–S+ intermediate Int-I and Ph3P = O, which were confirmed by 1H NMR, HRMS data and 31P NMR (Scheme 5a, see the ESI). A mixture of Int-I and 2a was able to afford 3a in 80% yield under the standard conditions, indicating that the reaction did go through this route (Scheme 5b). Furthermore, the key nickel-complex 14 was prepared and reacted with Int-I to furnish thioether 3d in 15% yield (Scheme 5c). Finally, sulfone 15 was used in the absence of benzimidazolium sulfonamide 1a and 2a under standard conditions. No reduced product 3a was detected, which indicates that the PPh3 reduction occurs before the cross-coupling (Scheme 5d). On the basis of the experimental results and previous reports,16c,g a plausible mechanism for the reductive thiolation of organic halides is proposed (Scheme 5e). Initially, the reduction of the Ni(ii) salt by zinc affords the active Ni(0) catalyst A, which undergoes oxidative addition into the C–X bond of organic halides to give the intermediate R–Ni(ii)X B. The following reduction of B forms R–Ni(i) intermediate C.15b,16g Then C and Int-I undergo stepwise single-electron transfer with the possibility of radical trapping within a solvent cage to afford intermediate D before the generation of R–Ni(iii)-SAr E and benzoimidazole residue 16. Finally, the reductive elimination of E furnishes the desired thioether product and Ni(i) species F which is reduced by zinc to facilitate the next catalytic cycle.Open in a separate windowScheme 5Mechanistic studies and proposed mechanism.  相似文献   

9.
Access to P-stereogenic compounds via desymmetrizing enantioselective bromination     
Qiu-Hong Huang  Qian-Yi Zhou  Chen Yang  Li Chen  Jin-Pei Cheng  Xin Li 《Chemical science》2021,12(12):4582
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   

10.
Catalytic asymmetric hydrogenation of (Z)-α-dehydroamido boronate esters: direct route to alkyl-substituted α-amidoboronic esters     
Yazhou Lou  Jun Wang  Gelin Gong  Fanfu Guan  Jiaxiang Lu  Jialin Wen  Xumu Zhang 《Chemical science》2020,11(3):851
The direct catalytic asymmetric hydrogenation of (Z)-α-dehydroamino boronate esters was realized. Using this approach, a class of therapeutically relevant alkyl-substituted α-amidoboronic esters was easily synthesized in high yields with generally excellent enantioselectivities (up to 99% yield and 99% ee). The utility of the products has been demonstrated by transformation to their corresponding boronic acid derivatives by a Pd-catalyzed borylation reaction and an efficient synthesis of a potential intermediate of bortezomib. The clean, atom-economic and environment friendly nature of this catalytic asymmetric hydrogenation process would make this approach a new alternative for the production of alkyl-substituted α-amidoboronic esters of great potential in the area of organic synthesis and medicinal chemistry.

The direct catalytic asymmetric hydrogenation of (Z)-α-dehydroamino boronate esters was realized.

Since FDA approval of bortezomib1 for the treatment of multiple myeloma, chiral α-aminoboronic acids have been recognized as key pharmacophores for the design of proteasome inhibitors.2 The incorporation of chiral α-aminoboronic acid motifs at the C-terminal position of a peptide3 to develop potential clinical drug candidates has drawn increasing interest4 (Fig. 1). Meanwhile, chiral α-amidoboronic acids and their derivatives are useful synthetic building blocks for the stereospecific construction of chiral amine compounds.5 The biological and synthetic value of α-amidoboronates has led to considerable efforts for the development of efficient synthetic methods. However, up to now, limited transition-metal-catalyzed asymmetric approaches have been reported. The widely used strategies to synthesize these compounds are stepwise Matteson homologation/N-nucleophilic replacement,6 borylation of imines,7 and alkene functionalization.8 Recently, two other elegant approaches, Ni-catalyzed decarboxylative borylation of α-amino acid derivatives9 and enantiospecific borylation of lithiated α-N-Boc species,10 were reported by the Baran and Negishi groups, respectively. To the best of our knowledge, the majority of the methods relied on either stoichiometric amounts of chiral auxiliaries6,7a,b or substrate-control strategies9 and most of these methods enable the construction of aryl-substituted α-aminoboronates. Enantioselective methods to access unfunctionalized alkyl-substituted α-aminoboronic esters are still rarely developed and so far only two examples have been realized by the Miura8a and Scheidt7f groups, respectively. Considering that most therapeutically relevant α-amidoboronic acid fragments contain an alkyl subunit and the fact that the options for the synthesis of alkyl-substituted α-amidoboronic esters in an enantioselective manner are still rare, the development of other distinct approaches would be highly desirable. Herein, we report a new alternative to access these compounds by catalytic asymmetric hydrogenation of (Z)-α-dehydroamidoboronate esters. With this approach, the desired chiral alkyl-substituted α-amidoboronic esters could be obtained in high yields and generally excellent enantioselectivities (up to 99% yield and 99% ee) with simple purification.Open in a separate windowFig. 1Selected inhibitors containing chiral alkyl-substituted α-amidoboronic acids.Catalytic asymmetric hydrogenation of olefins is an atom-economic, environmentally friendly and clean process for the synthesis of valuable pharmaceuticals, agricultural compounds and feedstock chemicals.11 Recently, hydrogenation of vinylboronic compounds has emerged for the preparation of chiral boronic compounds in a regiodefined manner.12,13 However, surprisingly α-dehydroamido boronate esters and their derivatives, as elegant precursors to access alkyl-substituted α-amidoboronic compounds, have never been used as substrates in asymmetric hydrogenation and remain a challenging project. To our knowledge, only one efficient hydrogenation approach to (1-halo-1-alkenyl) boronic esters was reported for indirect synthesis of alkyl-substituted α-aminoboronic esters but it was accompanied by inevitable de-halogenated by-products14 (Scheme 1). Given the catalytic efficiency and atom economy of the hydrogenation method, the development of a new direct hydrogenation approach to construct these important chiral alkyl-substituted α-amidoboronic esters would be very appealing.Open in a separate windowScheme 1Approaches towards the synthesis of chiral alkyl-substituted α-aminoboronic esters.The inspiration for our approach to the hydrogenation of α-dehydroamido boronates came from the molecular structures of relevant biologically active inhibitors containing alkyl-substituted α-amidoboronic acid fragments. Due to the limited stability of free α-aminoboronic acids, an electron-withdrawing carboxylic N-substituent is often required.15 Thus, we envisaged that N-carboxyl protected α-dehydroamido boronate esters could serve as a potential precursor for the synthesis of alkyl-substituted α-amidoboronates through Rh-catalyzed asymmetric hydrogenation of the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond16 (Fig. 1), a strategically distinct approach to the construction of unfunctionalized alkyl-substituted α-amidoboronic esters. However, challenges still remain, including: (1) how to synthesize α-dehydroamido boronates; (2) the facile transmetalation process of the starting materials leading to deboronated by-products in the hydrogenation process;17 (3) the unknown stability of α-amidoboronic compounds in the presence of a transition-metal catalyst and hydrogen molecules. As part of our continuous efforts to develop efficient hydrogenation approaches to construct valuable motifs,18 here we present the results of the investigation to address the aforementioned challenges.The desired aryl-substituted (Z)-α-dehydroamido boronates could be obtained by Cu-catalyzed regioselective hydroborylation of ynamide according to a previous report.19 However, different α/β-regioselectivity was observed for the preparation of alkyl-substituted (Z)-α-dehydroamido boronate esters and a new synthetic route was developed (Scheme 2, see the ESI for details). Of note, (Z)-α-dehydroamido boronate esters should be purified with deactivated silica gel,7c or else protodeborylation would occur readily with flash chromatography.Open in a separate windowScheme 2Synthetic route to (Z)-α-dehydroamino boronates.In order to check the feasibility of our hypothesis, three substrates were prepared with Rh(NBD)2BF4 and examined and our group prepared (Rc,Sp)-DuanPhos under 50 atm hydrogen pressure (Table 1). Gratifyingly, substrate 1b reacted smoothly to provide the desired product 2b in high yield and enantioselectivity (>99% conv., 98% ee, entry 2) whilst the reaction with substrate 1a yielded a mixture of deborylation products and 1c did not work at all (entries 1 and 3). Of note, we did not observe deborylation products with 1b under the current reaction conditions and we did not select (Z)-α-dehydroamido boronic acid 1a as the model substrate because of its poor solubility in most solvents. Then, a variety of chiral diphosphine ligands were investigated along with Rh(NBD)2BF4 and the results are shown in Table 1. In most cases, the reaction proceeded smoothly to furnish the desired products and the best results were obtained when (Rc,Sp)-DuanPhos was used as the ligand (entries 2 and 4–12). Poor results were obtained with axially bidentate phosphine ligands (entries 5, 6 and 9). (R,R)-QuinoxP* and (R,R)-Ph-BPE also gave good conversion with a slightly decreased ee whilst (R,R)-iPr-DuPhos exhibited poor results (entries 4, 7 and 10). Subsequent solvent screening revealed that the desired products could be obtained in most of the solvents and 1,2-DCE was the best solvent. (Entry 13, see the ESI).Condition optimization for catalytic asymmetric hydrogenation of 1a
EntrySubLigandConv.b (%)eec (%)
1d 1a (Rc,Sp)-DuanPhos89n.d.
2e 1b (Rc,Sp)-DuanPhos>9998
3 1c (Rc,Sp)-DuanPhosn.r.n.d.
4 1b (R,R)-QuinoxP*>9997
5 1b (S)-SegPhos>9917
6 1b (S)-BINAP>9910
7 1b (R,R)-iPr-DuPhos>993
8 1b (R,S)-Cy-JosiPhos>9914
9 1b (R)-BIPHEP>99−30
10 1b (R,R)-Ph-BPE>99−86
11 1b (S,S)-f-Binaphane>9961
12 1b (2S,4S)-BDPP>9959
13e,f,g 1b (Rc,Sp)-DuanPhos>99(99)99
Open in a separate windowaUnless otherwise mentioned, the reactions were performed with 1 (0.1 mmol), Rh(NBD)2BF4 (10 mol%), and a ligand (11 mol%) in 1.0 mL THF at 50 °C for 15 h.bDetermined by crude 1H NMR.cDetermined with chiral HPLC.dThe reaction was performed in iPrOH.eRh(NBD)2BF4 (1.0 mol%) and ligand (1.05 mol%) were used.fIsolated yield in parentheses.g1,2-DCE was used as the solvent. Pin = 2,3-dimethyl-2,3-butanediol; dan = 1,8-diaminonaphthalene.With the optimized reaction conditions in hand, a series of (Z)-α-dehydroamido boronate esters were tested and the results are summarized in Table 2. All the substrates reacted smoothly to give the corresponding alkyl-substituted α-amidoboronates in high yields with good to excellent enantioselectivities (2b, 2d–2r, and 2u, 99% yield, 57–99% ee). Alkyl-substituted (Z)-α-dehydroamido boronate esters were well tolerated in the current reaction, providing the corresponding α-amidoboronates in high yields and excellent enantioselectivities (2d–2i, 99% yield, 96–99% ee). Aryl-substituted (Z)-α-dehydroamido boronate esters with electron-donating (2j–l, 2n and 2p–r) and withdrawing (2m and 2o) substituents could also give the desired products in excellent yield with excellent enantioselectivities (90–99% ee). The ortho-methyl-substituted substrate 1r reacted smoothly to give the desired product with excellent enantioselectivity, but the 2,6-dimethyl-substituted substrate 1z could not react at all. Functional groups such as ether, halo and benzyl were well tolerated in the current reaction (2k, 2l, 2m and 2o–q). Replacement of the N-substituents with acyclic carbamate was also tolerated but with a decreased ee (2u and 2z). Substrates containing a chiral oxazolidin-2-one unit bearing bulky Ph-substituents around the nitrogen and oxygen were also competent, yielding the desired products with good to excellent diastereoselectivities (2s, 2t, and 2v–y). Of note, the substrate 1s bearing an N-Ms substituent and the cyclic substrate 1t did not work in the current reaction. The absolute configuration of generated α-amidoboronates was assigned as (S) by X-ray crystallographic analysis of 2i (Scheme 3).20Substrate scope.a,b,c
Open in a separate windowaUnless otherwise mentioned, the reactions were performed with 1 (0.1 mmol), Rh(NBD)2BF4 (1.0 mol%), and a ligand (1.05 mol%) in 1.0 mL 1,2-DCE at 50 °C under 50 atm H2 for 15 h.bIsolated yield.cDetermined with chiral HPLC.dDetermined by crude 1H NMR.Open in a separate windowScheme 3Scale-up synthesis and synthetic utility.To demonstrate the utility of the products, a scale-up reaction (0.62 g) was successfully performed with 0.1 mol% catalytic loading, giving 2b in 99% yield and 98% ee, and 2b could be easily transformed to a more stable α-amidoborate 3b with KHF2,6d,21 followed by hydrolysis with TMSCl to yield α-amido boronic acid 4b in 46% yield,22 which could also be obtained from 2b by treating it with BCl3 in 84% yield, without loss of the optical purity.8b2m could easily be transformed to 4m in 68% yield by a Pd-catalyzed borylation reaction. Meanwhile, after hydrogenation of 1x to 2x′ and transformation of 2x′ to its trifluoroborate derivative 3x′, removal of the benzyl group of 3x′ with Pd/C under hydrogenation conditions23 yielded the primary α-aminoborate 4x in 62% yield in three steps, which could serve as a potential precursor15 to synthesize bortezomib.  相似文献   

11.
Asymmetric total synthesis of (+)-xestoquinone and (+)-adociaquinones A and B     
Xiao-Long Lu  Yuanyou Qiu  Baochao Yang  Haibing He  Shuanhu Gao 《Chemical science》2021,12(13):4747
The asymmetric total synthesis of (+)-xestoquinone and (+)-adociaquinones A and B was achieved in 6–7 steps using an easily accessible meso-cyclohexadienone derivative. The [6,6]-bicyclic decalin B–C ring and the all-carbon quaternary stereocenter at C-6 were prepared via a desymmetric intramolecular Michael reaction with up to 97% ee. The naphthalene diol D–E ring was constructed through a sequence of Ti(Oi-Pr)4-promoted photoenolization/Diels–Alder, dehydration, and aromatization reactions. This asymmetric strategy provides a scalable route to prepare target molecules and their derivatives for further biological studies.

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

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

12.
Benzannulation of isobenzopyryliums with electron-rich alkynes: a modular access to β-functionalized naphthalenes     
An Wu  Hui Qian  Wanxiang Zhao  Jianwei Sun 《Chemical science》2020,11(30):7957
Described here is a modular strategy for the rapid synthesis of β-functionalized electron-rich naphthalenes, a family of valuable molecules lacking general access previously. Our approach employs an intermolecular benzannulation of in situ generated isobenzopyrylium ions with various electron-rich alkynes, which were not well utilized for this type of reaction before. These reactions not only feature a broad scope, complete regioselectivity, and mild conditions, but also exhibit unusual product divergence depending on the substrate substitution pattern. This divergence allows further expansion of the product diversity. Control experiments provided preliminary insights into the reaction mechanism.

A substituent-controlled divergent benzannulation provides rapid access to various β-functionalized naphthalenes from electron-rich alkynes.

Functionalized naphthalenes are important structural motifs widely present in bioactive natural products, pharmaceuticals and functional materials.1,2 They also serve as valuable intermediates in organic synthesis.3 Among them, naphthalenes bearing an electron-donating group (EDG) at the β-position (e.g., β-naphthols, β-naphthylamines, and β-naphthyl thioethers) are particularly versatile (Fig. 1).2,3 For example, BINOL, which is derived from β-naphthol, is an extensively utilized synthetic precursor toward a wide range of privileged chiral ligands and catalysts.3 While various approaches have been developed for the synthesis of naphthalenes, efficient and selective de novo strategies toward these β-functionalized ones still remain in high demand. Moreover, to the best of our knowledge, a general and modular approach for rapid access to all these electron-rich β-functionalized naphthalenes remains unknown.4Open in a separate windowFig. 1Useful β-naphthol, β-naphthylamine, and β-naphthyl thioether units.Isobenzopyrylium ions are readily accessible and versatile intermediates in organic synthesis (Scheme 1).5 They are known to participate in cycloaddition reactions with diverse carbon–carbon double bonds or triple bonds for the synthesis of naphthalenes.5,6 While extensive progress has been achieved in this topic, challenges still remain to be addressed. For example, the majority of these benzannulations with alkynes have to be executed at high temperature, except those intramolecular cases or with stoichiometric activators. Moreover, electron-rich alkynes have not been well explored as reaction partners for the benzannulations with isobenzopyrylium ions. Nevertheless, such reactions would provide expedient access to the valuable β-naphthol and β-napthylamine derivatives with diverse substitution patterns (Scheme 1). In this context, here we report our effort in achieving a general and modular strategy toward these electron-rich naphthalenes with high efficiency and regioselectivity under mild conditions. We also observed interesting selectivity divergence controlled by the substituent of the isobenzopyrylium substrates, giving rise to different types of naphthalene products (e.g., 2-hydroxy-1-naphthaldehydes, Scheme 1).Open in a separate windowScheme 1Benzannulation of isobenzopyryliums with electron-rich alkynes.We began our study with isochromene 1a as the isobenzopyrylium precursor. Siloxy alkyne 2a was first employed as the model electron-rich alkyne in view of the extraordinary versatility of this type of alkyne in various cyclization reactions, including benzannulation.7–9 Various Brønsted and Lewis acids were evaluated as catalysts for this reaction. Unfortunately, most of them did not show obvious catalytic activity toward the formation of a naphthalene product (Table 1). These reactions either had little conversion or resulted in a mixture of undesired products. Nevertheless, further screening led to the identification of AgOTf and BF3·OEt2 as capable catalysts (entries 5–6), with the latter being superior, leading to the formation of naphthalene 3a as the major product (75% yield, entry 6). Increasing the catalyst loading to 20 mol% further improved the product yield to 85% (with an isolated yield of 81%, entry 7). Further increasing the catalyst loading proved to be not beneficial (entry 8). Notably, the use of substrate 1a′ bearing an ethoxy leaving group also provided an equally good result. Other solvents, such as toluene and MeCN, were also evaluated, but all gave lower product yields. Finally, it is worth noting that the mild conditions used here are in sharp contrast to the typical high temperature required for the previously known catalytic intermolecular benzannulations involving isobenzopyryliums and alkynes.5dEvaluation of reaction conditions
EntryCatalystYielda (%)
1MeSO3H<5b
2HNTf2<5c
3Sc(OTf)3<5c
4TiCl4<5b
5AgOTf30c
6BF3·OEt275
7 BF 3 ·OEt 2 (20 mol%) 85 (81) d
8BF3·OEt2 (1.0 equiv.)83
9BF3·OEt2 (20 mol%), with 1a′(75)d
Open in a separate windowaReaction scale: 1a (0.05 mmol), 2a (0.06 mmol), catalyst (10 mol%), and DCM (0.5 mL). Yield is based on the analysis of the 1H NMR spectrum of the crude product using CH2Br2 as the internal standard.bUnreacted substrates account for the major remainder of the mass balance.cAn unidentifiable mixture accounts for the major remainder of the mass balance.dYield in parentheses is isolated yield.With the optimized conditions, the generality of this process was examined (Scheme 2). A range of isochromenes 1 and siloxy alkynes 2 participated in this benzannulation to form the desired silylated β-naphthols 3 with moderate to good efficiency. Notably, these products were all generated as a single isomer. This is particularly noteworthy when compared with the cases using phthalazines, which gave a mixture of regioisomers if the phthalazine was unsymmetrically substituted.8g The excellent regioselectivity observed in our reaction is likely attributed to the significant polarization of both cycloaddition partners. While alkyl-substituted siloxy alkynes reacted efficiently, unfortunately aryl-substituted ones led to low yield.Open in a separate windowScheme 2Reaction scope. Reaction scale: 1 (0.5 mmol), 2 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 24 h.During the above scope study, we found that the reaction of 3-unsubstituted isochromene 1g and siloxy alkyne 2a did not form the desired product 3a. Instead, 2-hydroxy-1-naphthaldehyde 4a was formed as the major product (64% yield, eqn (1)). Careful analysis of this product structure indicated that the original leaving part in 1a (in the case of 3a) was not completely cleaved from the molecule. Instead, only the C–O bond cleavage took place, which led to a dangling aldehyde group. Another important observation is that the TIPS group was lost in the naphthalene product. It was proposed that the methoxide leaving group in 1a might help remove this silyl group and assist the above C–O bond cleavage. In contrast, in the formation of 3a, this methoxide group was a part of the whole leaving unit during rearomatization and thus unavailable for desilylation (vide infra).In fact, such 2-hydroxy-1-naphthaldehyde products, resembling the highly versatile salicylaldehyde family, are indeed a useful substructure of bioactive molecules, such as gossypol (Fig. 1).2b They can also serve as key intermediates toward functional materials, such as molecular sensors (e.g., AHN, Fig. 1).2e In view of these important applications, considerable efforts were next devoted to further improving the reaction efficiency (see the ESI for details). Finally, we found that the reaction yield of 4a could be improved to 85% with 2.0 equivalents of BF3·OEt2 and one equivalent of 2,4,6-collidine as the additive. We believe that the role of collidine is to help reversibly stabilize the isobenzopyrylium intermediate and prevent its decomposition during the reaction progress.9b1The above protocol for the synthesis of 2-hydroxy-1-naphthaldehydes is general for a range of 3-substituted isobenzopyryliums (Scheme 3). The reaction efficiency was not affected by electron-donating or electron-withdrawing groups. Various siloxy alkynes were also suitable reaction partners, including aryl- and alkyl-substituted ones. Again, these products were all formed as a single regioisomer. While the majority of these reactions were highly chemoselective toward the formation of aldehydes 4, it is worth noting that tBu- and TBS-substituted alkynes resulted in a mixture of 3 and 4, with the former being major. This is likely due to the substantial steric hindrance in close proximity to the silyl group, whose cleavage was obstructed and thus the driving force for the subsequent C–O bond cleavage was weakened, thereby altering the pathway to preferentially form 3 (vide infra). Finally, the structures of 4a and 4b were unambiguously confirmed by X-ray crystallography.Open in a separate windowScheme 3Scope for the synthesis of 2-hydroxy-1-naphthaldehydes 4. Reaction scale: 1 (0.5 mmol), 2 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 16 h. b Run for 19 h.The divergent reaction patterns observed with siloxy alkynes prompted us to explore other electron-rich alkynes. Thioalkyne 5a was next employed for the reaction with 1a. Indeed, direct extension with BF3·OEt2 as the catalyst successfully resulted in regioselective formation of β-naphthyl thioethers 6a in about 70% yield (Table 2, entries 1 and 2). Further screening of other catalysts was performed to further improve the reaction efficiency (see the ESI for details). While Brønsted acids led to a significant drop in the yield, we were pleased to find that the Lewis acid AgNTf2 exhibited excellent catalytic activity, furnishing 6a in 89% yield (entry 6).Condition optimizationa
EntryCatalystYieldb (%)
1BF3·OEt270
2BF3·OEt2 (20 mol%)72 (68)b
3HNTf220
4MeSO3H7
5AgOTf80
6 AgNTf 2 89 (83) b
Open in a separate windowaReaction scale: 1a (0.05 mmol), 5a (0.06 mmol), catalyst (10 mol%), and DCM (0.5 mL). The yields are based on NMR analysis of the crude product using CH2Br2 as the internal standard.bIsolated yield.With the above conditions, we carried out a brief examination of the reaction scope (Scheme 4). Interestingly, a similar selectivity divergence was also observed with thioalkynes. Indeed, under identical conditions, the 3-butyl-substituted and 3-unsubstituted isobenzopyryliums all reacted efficiently, with the former leading to only 2-naphthyl thioethers 6, but the latter preferentially to 2-sulfenyl-1-naphthaldehydes 7. It is worth noting that only one regioisomer was observed in all these products.Open in a separate windowScheme 4Scope for the synthesis of 2-naphthyl thioethers. Reaction scale: 1 (0.3 mmol), 5 (0.36 mmol), 12 h, and isolated yield. a Run for 37 h. b Run for 19 h. The unreacted alkyne accounted for the major mass balance.Finally, we were curious about the reactivity of ynamides in such benzannulations.10 To our delight, in the presence of 20 mol% of BF3·OEt2, the reaction between isochromene 1a and different ynamides 8 successfully afforded the desired 2-naphthylamine products 9 with excellent regioselectivity and good efficiency (Scheme 5). Different electron-withdrawing groups on the ynamides were all compatible with this protocol. However, to our surprise, the use of 3-unsubstituted isochromene 1g did not lead to the expected naphthaldehyde 10 at all. Instead, the same product 9a was obtained as the only product, even if two equivalents of BF3·OEt2 were used together with collidine as the additive (conditions in Scheme 3). This result is in dramatic contrast to the cases of siloxy alkynes and thioalkynes.Open in a separate windowScheme 5Scope of ynamides. Reaction scale: 1 (0.5 mmol), 8 (0.6 mmol), DCM (5 mL), 12 h, and isolated yield. a Run for 30 h. b Yield in parentheses is obtained with BF3·OEt2 (2.0 equiv.), 2,4,6-collidine (1.0 equiv.), 0 °C, and DCM.Next, further studies were performed to help understand the unusual selectivity divergence observed with siloxy alkynes and thioalkynes. It is obvious that the substituent at the 3-position of the isochromene substrates has a crucial impact on the product distribution. To further probe the possible steric and electronic effects, we incorporated various other substituents, such as different aryl groups and the bulky tBu group (Table 3). We found that these reactions afforded products 3a and 4′ with variable ratios. With the tBu group, almost the same product distribution as the nBu group was observed, exclusively forming 3a (Table 3, entry 2). However, aryl substitution reversed this ratio, giving product 4′ as the major product (entries 3–5). More importantly, the preference toward 4′ is more pronounced with electron-deficient aryl groups. p-Fluorophenyl substitution led to almost only 4′. These results suggested that the product distribution appeared to be more related to the electronic effect than the steric effect.Influence of the 3-substituents in isochromenesa
Open in a separate windowaReaction scale: 1 (0.05 mmol), 2a (0.06 mmol), and 12 h. Yield is based on the analysis of the NMR yield of the crude mixture using CH2Br2 as an internal standard.bRun with 10 mol% of BF3·OEt2.Possible mechanisms are depicted in Scheme 6 using a siloxy alkyne as an example. The reaction begins with Lewis acid activation on the acetal motif to form the isobenzopyrylium intermediate I. The subsequent cycloaddition with the alkyne triple bond forms the key bicyclic oxonium intermediate II, which has a resonance form II′. This step might also be stepwise by forming one C–C bond first via ketenium III. In either case, the regioselectivity is precisely controlled by matching the polarity of the two partners, which explains the exclusive formation of only one regioisomer in all the cases. Depending on the R substituent, oxonium II can proceed via three possible pathways. In path a (R = alkyl), the methoxide attacks the oxonium carbon to form intermediate IV, which then undergoes retro-[4 + 2] cycloaddition with concomitant rearomatization to form naphthalene 3, together with the release of ester RCO2Me. However, if the oxonium carbon is unsubstituted (R = H), this intermediate is relatively unstable due to less stabilization of the positive charge (also viewed as a secondary cation in II′, versus tertiary carbocation if R ≠ H). Thus, the silyl enol ether motif tends to push the electron toward this unstable oxonium (a “push–pull” scenario) to cleave the bridging C–O bond. This step, likely assisted by the methoxide attack on the silyl unit, leads to 1,3-dicarbonyl intermediate V. The subsequent tautomerization/aromatization leads to the observed 2-hydroxy-1-naphthaldehyde 4.6b As an exception, if the siloxy alkyne is bulky (e.g., R′ = tBu or TBS), the reaction preferentially gives product 3 (Scheme 3, 4l–m). We believe that the increased steric repulsion by bulky R′ significantly disfavors the approach of methoxide to the neighboring silyl group in IIb. Instead, the methoxide prefers to add to the oxonium carbon, which allows path a to operate and forms 3 as the major product. Another exceptional case is the use of ynamides, which exclusively give products 3 even if R = H. Although actual rationalization would require more sophisticated mechanistic study, we currently reason that the enamide unit in the IIb analogue might not be electron-rich enough to exert sufficient driving force to cleave the C–O bond. This can also be viewed from the standpoint of the limited stabilization of the resulting iminium ion next to an electron-withdrawing group if this C–O bond is cleaved. Consequently, the methoxide prefers to attack the oxonium carbon to favor path a. Finally, with 3-aryl-substituted isochromenes (R = Ar), we believe that the oxonium intermediate IIc is well stabilized by aryl resonance. Thus, the effective delocalization of the positive charge makes this oxonium carbon less electrophilic for nucleophilic attack by methoxide. Instead, the methoxide serves as a base to deprotonate the bridgehead hydrogen, which triggers rearomatization and C–O cleavage to form product 4′. The more electron-deficient aryl group makes this bridgehead hydrogen more acidic, thus further favoring this pathway, which explains the trend in entries 3–5, Table 3.Open in a separate windowScheme 6Proposed mechanism.To further substantiate the above rationale, we carried out some control experiments. First of all, to confirm the fate of the leaving part in the isochromene substrates when β-naphthols 3 were formed, we used a large homobenzyl group in isochromene 1p (eqn (2)). After its reaction with 2a, we were able to isolate ester 11, which is consistent with path a of the proposed mechanism. Next, we also suspected that products 3 and 4 might interconvert to each other depending on the reaction conditions. To probe this possibility, a mixture of 3a and methyl formate was treated with BF3·OEt2 and 2,4,6-collidine, the standard conditions toward products 4 (eqn (3)). However, product 4a was not observed at all, indicating that 3 is not an intermediate toward 4. Similarly, subjecting 4a and MeOH (or MeONa) to BF3·OEt2 did not lead to the deformylation product β-naphthol 3a′, suggesting that 4 is unlikely an intermediate in the formation of 3 (eqn (4)). These observations are consistent with the proposed mechanism, in which the product distribution is kinetically controlled by the barrier in each case and these paths are likely irreversible, but not thermodynamically controlled by product stability.234In summary, we have developed a modular strategy for the rapid synthesis of valuable β-functionalized electron-rich naphthalenes, specifically, β-naphthol, β-naphthylamine, and β-naphthyl thioether derivatives. It also represents the first systematic study of the benzannulations of versatile isobenzopyryliums with general electron-rich alkynes. With suitable choice of the isochromene substrates and the Lewis acid catalysts, different types of electron-rich alkynes, such as siloxy alkynes, ynamides, and thioalkynes, participated in the intermolecular cycloaddition reactions under mild conditions with high efficiency and complete regioselectivity. Moreover, depending on the substitution pattern of the isochromene substrates, unusual divergence toward different naphthalene products was observed, thus allowing further diversification of the naphthalene products. Control experiments provided preliminary insights into the intriguing mechanism. Further detailed investigations toward a better understanding are ongoing.  相似文献   

13.
Reductive radical-initiated 1,2-C migration assisted by an azidyl group     
Xueying Zhang  Zhansong Zhang  Jin-Na Song  Zikun Wang 《Chemical science》2020,11(30):7921
We report here a novel reductive radical-polar crossover reaction that is a reductive radical-initiated 1,2-C migration of 2-azido allyl alcohols enabled by an azidyl group. The reaction tolerates diverse migrating groups, such as alkyl, alkenyl, and aryl groups, allowing access to n+1 ring expansion of small to large rings. The possibility of directly using propargyl alcohols in one-pot is also described. Mechanistic studies indicated that an azidyl group is a good leaving group and provides a driving force for the 1,2-C migration.

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

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

14.
Formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloaddition reactions of donor–acceptor cyclobutenes,cyclopropenes and siloxyalkynes induced by Brønsted acid catalysis     
Haifeng Zheng  Rui Wang  Kan Wang  Daniel Wherritt  Hadi Arman  Michael P. Doyle 《Chemical science》2021,12(13):4819
Brønsted acid catalyzed formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloadditions of donor–acceptor cyclobutenes, cyclopropenes, and siloxyalkynes with benzopyrylium ions are reported. [4 + 2]-cyclization/deMayo-type ring-extension cascade processes produce highly functionalized benzocyclooctatrienes, benzocycloheptatrienes, and 2-naphthols in good to excellent yields and selectivities. Moreover, the optical purity of reactant donor–acceptor cyclobutenes is fully retained during the cascade. The 1,3-dicarbonyl product framework of the reaction products provides opportunities for salen-type ligand syntheses and the construction of fused pyrazoles and isoxazoles that reveal a novel rotamer-diastereoisomerism.

Brønsted acid catalysis realizes formal [4 + 4]-, [4 + 3]-, and [4 + 2]-cycloadditions of donor–acceptor cyclobutenes, cyclopropanes, and siloxyalkynes with benzopyrylium ions.

Medium-sized rings are the core skeletons of many natural products and bioactive molecules,1 and a growing number of strategies have been developed for their synthesis.2 Because of their enthalpic and entropic advantages, ring expansion is a highly efficient methodology for these constructions.3–6 For example, Sun and co-workers have developed acid promoted ring extensions of oxetenium and azetidinium species formed from siloxyalkynes with cyclic acetals and hemiaminals (Scheme 1a).4 Takasu and co-workers have reported an elegant ring expansion with a palladium(ii) catalyzed 4π-electrocyclic ring-opening/Heck arylation cascade with fused cyclobutenes (Scheme 1b).5 Each transformation is initiated by the formation of fused bicyclic units followed by ring expansion or rearrangement to give medium-sized rings.Open in a separate windowScheme 1Cycloaddition/ring expansion background and this work.Strategies for the formation of fused bicyclic compounds rely on cycloaddition of dienes or dipoles with unsaturated cyclic compounds7 and, if the reactant cyclic compound is strained and chiral, the resulting bicyclic compound is activated toward ring opening that results in retention of chirality. We have recently reported access to donor–acceptor cycloalkenes by [3 + n] cycloaddition that have the prerequisites of unsaturated cyclic compounds suitable for cycloaddition.8 Donor–acceptor (D–A) cyclopropenes9 and cyclobutenes10 have sufficient strain in the resulting bicyclic compounds to undergo ring opening. We envision that the selection of a diene or dipolar reactant and suitable reaction conditions could realize cycloaddition and subsequent ring expansion. Benzopyrylium species,11,12 which are generated by metal or acid catalysis, have attracted our attention. We anticipated that their high reactivity would overcome the conventional unfavorable kinetic and/or thermodynamic factors that typically impede medium-sized ring formation. From a mechanistic perspective, benzopyrylium species are often formed by transition metal catalyzed reactions with 2-alkynylbenzaldehydes.11 Recently, the reaction between 1H-isochromene acetal and Brønsted acid catalyst forms the 2-benzopyrylium salts that could react with functional alkenes to give same cycloaddition products.12 Consequently, we believed that the [4 + 2]-cyclization between benzopyrylium species and donor–acceptor cyclobutenes, cyclopropenes, or siloxyalkynes would give bridged oxetenium intermediates, which contain high strain energy that should provide the driving force for ring expansion. The “push and pull” electronic effect of donor–acceptor functional groups facilitates deMayo-type ring-opening of the cyclobutane or cyclopropane skeletons (Scheme 1c).13Here we report bis(trifluoromethanesulfonyl)imide (HNTf2) catalyzed formal [4 + 4]-, [4 + 3]- and [4 + 2]-cycloaddition reactions of D–A cyclobutenes, cyclopropenes, and siloxyalkynes with benzopyrylium salts. Polysubstituted benzo-cyclooctatrienes, benzocycloheptatrienes and 2-naphthols, are produced in good to excellent yields and selectivities. Complete retention of configuration occurs using chiral cyclobutenes, and opportunities for further functionalization are built into these constructions.Initially, we conducted transition metal catalyzed reactions with 2-alkynylbenzaldehyde intending to produce the corresponding benzopyrylium ion and explore the possibility of cycloaddition/ring opening with donor–acceptor cyclobutene 2a. Use of Ph3PAuCl/AgSbF6, Pd(OAc)2 and Cu(OTf)2, which were efficient catalysts in previous transformations,11 gave only a trace amount of cycloaddition product (Fig. 1). Spectral analysis showed that mostly starting material remained (Fig. 1-i and ii). Increasing the reaction temperature led only to decomposition of 2-alkynylbenzaldehyde (Fig. 1-iv) or donor–acceptor cyclobutene 2a (Fig. 1-iv).Open in a separate windowFig. 1Reaction of 2-alkynylbenzaldehyde with donor–acceptor cyclobutene 2a.From these disappointments we turned our attention to 1H-isochromene acetal 1a as the benzopyrylium ion precursor. Various Lewis acid catalysts were employed with limited success, but we were pleased to observe the formation of the desired formal [4 + 4]-cycloaddition benzocyclooctatriene product 3aa, albeit in low yields (Table 1, entries 1–6). Selection of the Brønsted super acid HNTf2 (ref. 14) proved to be the most promising, producing 3aa in 40% yield (Table 1, entry 7). Increasing the amount of D–A cyclobutene 2a by 30% gave a much higher yield of 3aa (Table 1, entry 8 vs. 7). Optimization of the stoichiometric reaction between 1a and 2a by increasing the reaction temperature from rt to 35 °C led to the formation of the desired product in 76% isolated yield (Table 1, entry 9 vs. 8). However, a further increase in the reaction temperature (Table 1, entry 10) or reducing the catalyst loading to 5 mol% did not improve the yield of 3aa (Table 1, entry 11).Optimization of reaction conditionsa
EntryCat (10 mol%)Temp. (°C)Yieldb
1Sc(OTf)3rt23
2Yb(OTf)3rtTrace
3In(OTf)3rt16
4TiCl4rtTrace
5BF3·OEt2rt20
6TMSOTfrt17
7HNTf2rt40
8cHNTf2rt67
9cHNTf23578(76)d
10c,eHNTf26062
11c,fHNTf23550
Open in a separate windowaReactions were performed by adding the catalyst (10 mol%) to 1a (0.1 mmol) and 2a (0.1 mmol) in CH2Cl2 (2 mL) at the corresponding temperature for 24 h.bYields were determined by 1H NMR spectroscopic analysis with CH2Br2 as the internal standard.c1.3 equiv. 2a was used.dIsolated yield.eReaction performed at 60 °C for 12 h.f5 mol% catalyst loading.With optimized conditions using 1a in hand, we examined the scope of the formal [4 + 4]-cycloaddition reactions of D–A cyclobutenes 2 with a diverse set of acetal compounds 1. As shown in Scheme 2, a wide range of acetal substrates (1a–1i) with different substituents at different positions all reacted smoothly with D–A cyclobutene 2a to form the corresponding benzocyclooctatriene products 3 in good to excellent yields. Structural variations in the acetals produced only modest changes in product yields which ranged from 55 to 87%. Similarly, both electron-withdrawing and electron-donating substituents at the 4-position of the cyclobutene phenyl ring produced the corresponding products (3db–3dd) in good yields, and 2-naphthyl (2e) and 2-thienyl (2f) substituted cyclobutenes were suitable substrates (85% and 51% product yields, respectively). trans-1,2,3,4-Tetrasubstituted (R2 = CH3) 2-sil-oxycyclobutenecarboxylate 2g also underwent [4 + 4]-cycloaddition with 1d in good yield and fully retained its diastereoselectivity. The structure of 3fa was confirmed by X-ray diffraction (Scheme 2).15Open in a separate windowScheme 2Scope of the [4 + 4]-cycloaddition reaction of D–A cyclobutenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10 mol%) to 1 (0.1 mmol) and 2 (0.13 mmol) in CH2Cl2 (2 mL) at 35 °C for 24 h. Isolated yields are reported.Chiral donor–acceptor cyclobutenes with high enantiomeric excess and diastereoselectivity are conveniently obtained by catalytic [3 + 1]-cycloaddition of enoldiazoacetates with acyl ylides of sulfur.10 To determine if optical purity is retained, chiral D–A cyclobutene 2a (80% ee) was reacted with 1d under the optimized conditions, and the corresponding benzocyclo-octatriene product 3da was obtained in good yield with complete retention of configuration (Scheme 3, eqn (1)). With trans-disubstituted 2g and 2h that have higher optical purity, however, 3dg and 3dh were obtained in moderate yields with full retention of diastereo- and enantioselectivities, but addition products 8dg and 8dh were formed competitively (Scheme 3, eqn (2)). These compounds resulted from initial addition then desilylation, indicating that the [4 + 2]-cyclization is a stepwise reaction. Attempts to suppress the competing pathway by changing solvents or using the isopropyl acetal substrate to form a bulky 2-propanol nucleophile (1j) failed (for details, see ESI). However, p-methoxy (R1) substituted 1g that would stabilize the incipient benzopyrylium ion gave higher selectivity (4 : 1 vs. 2 : 1), and the desired [4 + 4] cycloaddition product 3gg was isolated in 50% yield with 97% ee and >19 : 1 dr (Scheme 3, eqn (3)).Open in a separate windowScheme 3Stereochemical features of the [4 + 4]-cycloaddition of D–A cyclobutenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10 mol%) to 1 (0.1 mmol) and 2 (0.13 mmol) in CH2Cl2 (2 mL) at 35 °C for 24 h. Isolated yields are reported.To further expand the generality of this strategy, we investigated its use with donor–acceptor cyclopropenes 4. The desired formal [4 + 3]-cycloaddition products 5 were obtained in good to excellent yields (Scheme 4). Optimized conditions used 20 mol% HNTf2 catalyst with 4 Å molecular sieves at room temperature for 2 h. The scope of this [4 + 3]-cycloaddition reaction with cyclopropenes 4 showed that acetals bearing both electron-donating and electron-withdrawing substituents on the aromatic ring were tolerated. However, as with [4 + 4]-cycloaddition reactions, the [4 + 3] reactions of 1 with R1 = alkyl or H did not produce any of the desired products. Furthermore, 3-substituted cyclopropenes 4b–4d participated in this reaction, and their products (5ab–5ad) were obtained in 63%–81% yields.Open in a separate windowScheme 4Scope of the [4 + 3]-cycloaddition reaction of D–A cyclopropenes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (20 mol%) to 1 (0.2 mmol), 4 (0.24 mmol) and 4 Å (50 mg) in CH2Cl2 (2 mL) at rt for 2 h. Isolated yields are reported.Siloxyalkynes 6, as electron-rich alkynes, have been widely used in diverse cyclization reaction.4,16 We expected that 6 could also participate in [4 + 2]-cyclization/ring-expansion cascade processes, giving substituted 2-naphthol products. Interestingly, substituent controlled diverse products were obtained in good to excellent reactivity and selectivity (Scheme 5). Aryl (R1) substituted acetals (1a–1d, 1f, 1h, 1i) reacted with siloxyalkynes 6a–6d, giving 2-naphthols (7aa–7ad and 7ba–7ia) in 35%–96% yields. With electron-donating group (EDG) substituents (7ba and 7ca) on the aromatic ring, higher reactivity was observed relative to those with electron-withdrawing groups (7da–7ia). In addition, the acetal with R1 = H (1m) reacted with siloxyalkynes 6 to form the 2-naphthol-1-carboxaldehyde derivative in good yield, and the structure of 7ma was confirmed by X-ray diffraction (Scheme 5b).15 Intriguingly, the n-butyl (R1) substituted acetal 1n reacted with siloxyalkynes via a [4 + 2]-cyclization with loss of methyl pentanoate (BuCO2CH3), affording siloxy naphthalenes 7na–7nd that are important precursors to the widely used axially chiral 2,2′-binols.17 The substrate scope of siloxyalkynes 6 for their formal [4 + 2]-cycloaddition reaction with n-butyl substituted acetal 1n was also explored (Scheme 5c). In all cases methyl pentanoate was eliminated to form 2,3-disubstituted naphthalene products (7na–7nd). Alkyl substituted siloxyalkynes (6a–6c) showed higher reactivity compared with phenyl substituted siloxyalkynes 6d. It should be mentioned that, recently, a similar transformation using BF3·OEt2 as catalyst or in excess (2 equiv.) with 2,4,6-collidine (1 equiv.) was reported,18 and HNTf2 was stated to be much less effective. To clarify this discrepancy, we carefully repeated these transformations (7ma and 7na) and found that all starting materials are completely consumed in less than 10 min to deliver [4 + 2]-cycloaddition products in good yields. Prolonging the reaction time to 12 hours, which was the reaction time used by the authors, results in their decomposition to a complex mixture of materials.Open in a separate windowScheme 5Scope of the [4 + 2]-cycloaddition reaction of siloxyalkynes and 1H-isochromene acetals.a aReactions were performed by adding HNTf2 (10–30 mol%) to 1 (0.2 mmol) and 6 (0.3–0.4 mmol) in CH2Cl2 (2 mL) at rt for 10 min. Isolated yields are reported. b40 mol% HNTf2 catalyst was used.To illustrate the utility of this process, a large scale catalytic [4 + 4] cycloaddition was performed, and adduct 3da was obtained in 87% yield. Further transformations were conducted for the synthesis of pyrazole and isoxazole structures based on its 1,3-dicarbonyl skeleton (Fig. 2a). Compound 3da reacted with hydrazine and hydroxylamine in refluxed ethanol, affording pyrazole 14da and isoxazole 15da in 89% yield or 66% yield, respectively. The structure of 15da was confirmed by X-ray diffraction.15 Interestingly, two NMR distinguishable interconvertible diastereoisomers were detected for each of these eight-membered cyclic products (2.5 : 1 and 5 : 1 dr for 14da and 15da, respectively, in CDCl3). These diastereoisomers are rotamers (for details, see ESI) that exist at equilibrium with each other in solution but form one crystalline product (X-ray structure of 15da). In addition, the cycloaddition product 7ma reacted with chiral 1,2-cyclohexanediamine and 1,2-diphenylethylenediamine to give salen-type ligands L1 and L2 in 53% yield and 74% yield, respectively, which provides new opportunities for ligand screening (Fig. 2b).19Open in a separate windowFig. 2Large scale reaction, further transformations, and ligands synthesis.In the mechanistic possibility considered for these HNTf2 catalyzed cycloaddition reactions (Fig. 3), protonation of acetal 1 with HNTf2 gives the corresponding highly reactive benzopyrylium intermediate int-I, which reacts with donor–acceptor cyclobutenes 2, cyclopropenes 4, or siloxyalkynes 6 affording addition intermediates int-II that undergo ring closure to int-III. Ring expansion then occurs to deliver 3, 5 and 7 in good to excellent yields with fully retained stereoselectivities. Furthermore, the formed TIPSNTf2 ([4 + 4]- and [4 + 3]-cycloaddition) or HNTf2 {[4 + 2] cycloaddition} are active acid catalysts for the conversion of 1 to benzopyrylium intermediate int-I that continues the catalytic cycle. With sterically larger D–A cyclobutenes or when a less ring-strained benzocyclopentene 12 is employed (for details, see ESI), the competing direct desilylation of int-II occurs, delivering addition byproducts 8 or 13. Compounds 7na–7nd arise from the analog to int-III from which ketene formation or methanol displacement effects 1,4-elimination.Open in a separate windowFig. 3Proposed mechanism.  相似文献   

15.
Electrophilic fluoroalkylthiolation induced diastereoselective and stereospecific 1,2-metalate migration of alkenylboronate complexes     
Feng Shen  Long Lu  Qilong Shen 《Chemical science》2020,11(30):8020
A transition metal free process for conjunctive functionalization of alkenylboron ate-complexes with electrophilic fluoroalkylthiolating reagents is described, affording β-trifluoroalkylthiolated and difluoroalkylthiolated boronic esters in good yield and excellent diastereoselectivity. The potential applicability of the method was demonstrated by the preparation of a difluoromethylthiolated mimic 12 of a potential drug molecule PF-4191834 for the treatment of asthma.

A transition metal free process for conjunctive functionalization of alkenylboron ate-complexes with electrophilic fluoroalkylthiolating reagents affords β-tri- and difluoroalkylthiolated boronic esters in good yield and diastereoselectivity.

An electrophile-induced 1,2-metalate migration of an alkenylboron “ate” complex and subsequent base-promoted β-elimination to form a functionalized cis-alkene, now the so-called Zweifel reaction, was first reported by Zweifel and co-workers in 1967 (Fig. 1A).1–3 The reaction was proposed to proceed via an initial attack of the π electron of the alkene moiety to iodine to generate a zwitterionic iodonium ion, which then undergoes a stereospecific 1,2-metalate to afford a β-iodoboronic ester, followed by anti-elimination upon treatment with a base to afford a cis-olefin. Thus, if the iodine is replaced by an alternative electrophilic reagent and the use of a base is omitted, an interrupted-Zweifel reaction for the preparation of a stereospecific β-functionalized boronic ester could be realized. Toward this end, Aggarwal reported the first example of such a reaction by employing PhSeCl as the electrophilic reagent.4 It was proposed that PhSeCl first reacts with an alkenylboronate complex to form a zwitterionic seleniranium ion. Subsequent diastereospecific 1,2-metalate migration affords the stereospecific β-seleno-alkylboronate (Fig. 1B). Likewise, shortly after, Denmark and co-workers reported an analogous Lewis-base catalysed enantioselective and diastereoselective carbosulfenylation of an alkenylboronate complex using N-arylthiosaccharin as the electrophile (Fig. 1C).5Open in a separate windowFig. 1The interrupted Zweifel reaction.In light of these discoveries and our recent success in the development of a toolbox of electrophilic fluoroalkylthiolating reagents including three trifluoromethylthiolating reagents α-cumyltrifluoromethane sulfenate,6N-trifluoromethylthio-saccharin7 and N-trifluoromethylthiodibenzenesulfonimide,8 and two difluoromethylthiolating reagents N-difluoromethylthiophthalimide9 and S-(difluoromethyl)benzenesulfonothioate,10 we wondered whether these electrophilic fluoroalkylthiolating reagents could also trigger the proposed stereospecific 1,2-metalatation of the alkenylboronate complex to afford β-fluoroalkylthiolated borane derivatives (Fig. 1D). The trifluoromethylthio (–SCF3) and the difluoromethylthio (–SCF2H) groups have gained great attention recently, partially because of their high and tuneable lipophilicity11 that might improve the drug candidate''s cell membrane permeability and consequently, its overall pharmacokinetics.12 Thus, the development of new efficient reactions for the incorporation of the trifluoromethylthio13 or difluoromethylthio groups14 would be of vital importance in facilitating medicinal chemists'' endeavours in new drug discovery. Herein, we report that by employing electrophilic difluoromethylthiolating reagent PhSO2SCF2H 2a as the electrophile, the proposed difluoromethylthiolating induced stereospecific 1,2-metalate migration of alkenyl boronate complexes occurred smoothly to afford β-difluoromethylthiolated boronic esters in good yields and excellent diastereoselectivity. Likewise, when electrophilic trifluoromethylthiolating reagent N-trifluoromethylthiosaccharin 7 was used, an analogous reaction for the diastereoselective formation of β-trifluoromethylthiolated boronic esters was successfully achieved.We began our study by examining the reaction of the electrophilic difluoromethylthiolating reagent 2a with the alkenylboronate complex which was generated in situ by mixing 1a and PhLi in diethyl ether. It was found that the reaction in CH3CN occurred in full conversion after 12 hours at room temperature, affording the corresponding product 3a in 53% yield (Table 1, entry 1). When the amount of PhLi was increased to 1.3 equivalents, the yield was increased to 76%, while the yield decreased to 66% when 2.0 equivalents of PhLi were used, likely due to the decomposition of the product under strong basic conditions (Table 1, entries 1–5). We then further investigated the effect of the reaction temperature and the solvent. It was found that the temperature did not affect the reaction significantly since the yields of the desired products were decreased slightly to 72% and 70%, respectively, when the reactions were conducted at 0 °C or −15 °C (Table 1, entries 6 and 7). Likewise, the reaction was not sensitive to the polarity of the solvent since reactions conducted in less polar solvents such as THF or CH2Cl2 or nonpolar solvents like toluene occurred in slightly lower 60–73% yields (Table 1, entries 9–11). We also found that reaction using N-difluoromethylthiophthalimide as the electrophilic difluoromethylthiolating reagent gave the same product in a slightly lower yield (Table 1, entry 8).Optimization of conditions for the reaction of the alkenyl boronate complex with PhSO2SCF2Ha
EntryEquiv. of PhLiSolventTemp (°C)Yielda (%)
11.0CH3CNrt53
21.1CH3CNrt60
31.2CH3CNrt72
41.3CH3CNrt76(72)b
52.0CH3CNrt66
61.3CH3CN072
71.3CH3CN−1570
81.3CH3CNrt56c
91.3THFrt73
101.3CH2Cl2rt64
111.3Toluenert60
Open in a separate windowaReaction conditions: vinyl boronate 1a (0.10 mmol) and reagent 2a (0.15 mmol), in CH3CN (1.0 mL) at room temperature for 12 h; Yields were determined by 19F NMR spectroscopy using PhCF3 as an internal standard.bIsolated yield.c N-Difluoromethylthiophthalimide was used.With optimum reaction conditions established, a range of different alkenylboronate complexes were tested under standard conditions (Scheme 1). Alkenylboronate complexes obtained by treating 3,6-dihydro-2H-pyran-4-boronic acid pinacol ester with diverse aryl lithiums reacted efficiently with reagent 2a to give the corresponding β-difluoroalkylthionated boronic esters 3b–e and 3g–m in good yield and excellent diastereoselectivity. A range of aryllithiums with both the electron-donating methoxy group (3c) and electron-withdrawing groups such as a fluoride (3d) or a trifluoromethyl group (3g) or a bulky tert-butyl group at meta-position (3i) worked well. The reaction can also proceed smoothly for naphthyllithium (3h) and n-butyllithium (3j). Moreover, organolithiums generated from heteroaromatics, such as indole (3k), benzothiophene (3l), benzofuran (3m), could also be used. Notably, it is well-known that bromine is not compatible with butyl lithium. Yet, 3f with a para-bromophenyl moiety was obtained from the reaction of the alkenylboronate complex in situ generated by treating (3,6-dihydro-2H-pyran-4-yl)lithium with 4-bromophenylboronic acid pinacol ester. However, the alkenylboronate complex generated by treating (E)-4,4,5,5-tetramethyl-2-(5-phenylpent-1-en-1-yl)-1,3,2-dioxaborolane with tert-butyllithium, failed to react with reagent 2a to give the corresponding β-difluoroalkylthionated boronic esters (3r). Next, the scope with respect to the alkenyl boronic ester component was explored. 3,6-Dihydro-2H-thiopyran-4-ylboronic acid pinacol ester (3n), or N-Ts-3,6-dihydro-2H-pyran-4-boronic acid pinacol ester (3o) and 1-phenylvinylboronic acid pinacol ester (3q) could react well to afford the corresponding products. To demonstrate the scalability of the reaction, 3p was prepared on a gram scale in 75% yield. Furthermore, bridged cyclic boronate 3s could also be obtained in moderate yield, and the anti diastereoselectivity of the reaction was confirmed by X-ray diffraction of its single crystals.Open in a separate windowScheme 1Scope of 1,2-metalate migration of alkenyl boronates with reagent 2a.a a Reaction conditions: alkenyl or aryl boronic ester (0.30 mmol, 1.0 equiv.), R3Li (0.33 mmol, 1.1 equiv.) in Et2O (1.5 mL) at −78 °C to room temperature for 30 min; then the solvent was swapped with CH3CN (3.0 mL); 2a (0.45 mmol, 1.5 equiv.) was added. Isolated yield. b R3Li (0.39 mmol) in Et2O (1.5 mL) at 0 °C to room temperature for 30 min. c The mixture was treated with NaBO3 (0.9 mmol, 3.0 equiv.) in THF/H2O (v/v = 1 : 1, 6 mL) at room temperature for 6 h.Furthermore, it was found that the resultant boronic esters could be easily oxidized to alcohols, with the difluoromethylthio group remaining intact, by treatment with 3.0 equivalents of NaBO3 at room temperature for 6 h. For example, difluoromethylthiolated β-alcohols 4a–4d were obtained in moderate to good yields under these conditions (Scheme 1).In general, it is a common practice to use E or Z-alkenes in the reaction to probe whether the reaction is stereo-specific. Thus, we examined the reaction of E-(3′-phenylpropyl)vinyl boronic acid pinacol ester and Z-(3′-phenylpropyl)vinyl boronic acid pinacol ester under standard conditions. It was found that the reaction is stereospecific since the reactions of E- and Z-alkenyl boronic esters specifically produced corresponding anti- and cis-difluoromethylthiolated alcohols (4e and 4f) with excellent diasteroselectivity (>20 : 1), respectively (Scheme 2).Open in a separate windowScheme 2Reactions of E- and Z-alkenyl boronate complexes with reagent 2a.To further expand the scope of the reaction, we studied the difluoromethylthiolative triggered stereospecific 1,2-metalate migration of in situ generated vinyl boronate complexes from enantio-enriched secondary alkyl boronic esters with vinyl lithium. The resulting crude alkyl boronic esters were then sequentially oxidized by NaBO3 and Jone''s oxidation to give α-chiral ketone derivatives. It was found that chirality of the secondary alkyl boronic esters was stereospecifically transferred to the final products 6a–c with 100% es (Scheme 3).Open in a separate windowScheme 3Synthesis of α-chiral ketones by stereospecific 1,2-migration.a a Reaction conditions: alkyl boronic ester (0.30 mmol, 1.0 equiv.), R3Li (0.36 mmol, 1.2 equiv.) in Et2O (1.5 mL) at −78 °C to room temperature for 30 min; then the solvent was swapped with CH3CN (3.0 mL); 2a (0.45 mmol, 1.5 equiv.) was added; and then NaBO3 (0.9 mmol, 3.0 equiv.) in THF/H2O (v/v = 1 : 1, 6 mL) was used; and then Jone''s reagent (0.45 mmol, 1.5 equiv.) was used. Isolated yield.Encouraged by the excellent diastereoselective difluoromethylthiolation of alkenyl boronic acid pinacol esters, we then extended this highly selective reaction to analogous trifluoromethylthiolation triggered 1,2-metalate migration of alkenylboronate (Scheme 4). It was found that when N-trifluoromethylthiosaccharin 7 was used as the electrophilic trifluoromethylthiolating reagent, the reaction of alkenylboronate derived from PhLi occurred smoothly in CH3CN after 12 h at 0 °C to give β-trifluoroalkylthionated boronic ester 8a in 76% yield (8a). Likewise, a variety of other aryllithiums could be successfully employed in this reaction to afford the corresponding β-trifluoroalkylthionated boronic esters (8b–h) in high yields. This reaction appears to be compatible with labile functional groups such as chlorine (8b), trifluoromethyl (8c), ketal (8d), and acetal (8e). In addition, organolithiums generated from heteroaromatics, such as benzofuran (8g) and benzothiophene (8h) could also be employed. Lastly, it was found that a single diastereoisomer with an anti configuration (8i) was isolated in 75% yield when the corresponding E-alkenyl boronic ester was used. Yet, the scope of alkenyoboronate complexes for the reaction with N-trifluoromethylthiosaccharin 7 is not as broad as that with PhSO2SCF2H since alkenylboronate complexes generated by treating 3,6-dihydro-2H-pyran-4-boronic acid pinacol ester with n-butyllithium or by treating 2,2-dimethylethenylboronic acid pinacol ester with lithium benzothiophene failed to produce the desired β-trifluoroalkylthionated boronic esters 8j and 8k under the standard conditions.Open in a separate windowScheme 4Scope of 1,2-metalate migration of alkenyl boronates with electrophilic trifluoromethylthiolating reagent 7.a a Reaction conditions: alkenyl boronic ester (0.30 mmol, 1.0 equiv.), R3Li (0.33 mmol, 1.1 equiv.) in Et2O (1.5 mL) at −78 °C to room temperature for 30 min; then the solvent was swapped with CH3CN (3.0 mL); reagent 5 (0.45 mmol) was added. b R3Li (0.39 mmol, 1.3 equiv.) in Et2O (1.5 mL) at 0 °C to room temperature for 30 min. Isolated yield.To further demonstrate the great potential of this reaction, we applied this protocol as a key step in the synthesis of a difluoromethylthiolated mimic of PF-4191834, which is a potent competitive inhibitor of the 5-lipoxygenase (5 LOX) enzyme for the treatment of mild to moderate asthma15 (Fig. 2). Firstly, arylsulfide 11 was synthesized efficiently by deborylthiolation of organoboron 9 with thiosulfonate 10 in the presence of 5 mol% CuSO4 as the catalyst. Lithium halide exchange of compound 11 with t-butyllithium at −78 °C for 30 min generated the corresponding aryl lithium species in situ, which was treated with 3,6-dihydro-2H-pyran-4-boronic acid pinacol ester to afford the alkenyl boronate complex. Switching the solvent from diether ether to CH3CN, followed by the addition of 1.5 equivalents of PhSO2SCF2H 2a, and further reaction at room temperature for 12 h produced the difluoromethylthiolated mimic of PF-4191834 12 in 70% yield. This example showed the potential of the current protocol in the preparation of biological active compounds.Open in a separate windowFig. 2Construction of PF-4191834 mimic by conjunctive cross-coupling.In summary, a method of conjunctive three-component coupling between alkenyl boronic esters, organolithiums and electrophilic fluoroalkylthiolating reagents was successfully developed, affording β-trifluoroalkylthionated and difluoroalkylthionated boronic esters in good yield and excellent diastereoselectivity. The reaction is stereospecific since the reaction of the E-alkenyl boronic ester specifically gave an anti-difluoromethylthiolated β-alcohol and the reaction of the Z-alkenyl boronic ester specifically gave cis-difluoromethylthiolated β-alcohol 4f with excellent diasteroselectivity (>20 : 1). The potential applicability of the method was demonstrated by the preparation of a difluoromethylthiolated derivative of a potential drug molecule for the treatment of asthma PF-4191834 12. The reactions of the alkenyl boronate complexes with other electrophilic fluoroalkylating reagents are currently actively underway in our laboratory.  相似文献   

16.
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   

17.
Aluminum-catalyzed tunable halodefluorination of trifluoromethyl- and difluoroalkyl-substituted olefins     
Zhong Liu  Xian-Shuang Tu  Le-Tao Guo  Xiao-Chen Wang 《Chemical science》2020,11(42):11548
Herein, we report unprecedented aluminum-catalyzed halodefluorination reactions of trifluoromethyl- and difluoroalkyl-substituted olefins with bromo- or chlorotrimethylsilane. The interesting feature of these reactions is that one, two, or three fluorine atoms can be selectively replaced with bromine or chlorine atoms by modification of the reaction conditions. The generated products can undergo a variety of subsequent transformations, thus constituting a valuable stock of building blocks for installing fluorine-containing olefin motifs in other molecules.

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

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

18.
Three-component 1,2-carboamination of vinyl boronic esters via amidyl radical induced 1,2-migration     
Cai You  Armido Studer 《Chemical science》2021,12(47):15765
Three-component 1,2-carboamination of vinyl boronic esters with alkyl/aryl lithium reagents and N-chloro-carbamates/carboxamides is presented. Vinylboron ate complexes generated in situ from the boronic ester and an organo lithium reagent are shown to react with readily available N-chloro-carbamates/carboxamides to give valuable 1,2-aminoboronic esters. These cascades proceed in the absence of any catalyst upon simple visible light irradiation. Amidyl radicals add to the vinylboron ate complexes followed by oxidation and 1,2-alkyl/aryl migration from boron to carbon to give the corresponding carboamination products. These practical cascades show high functional group tolerance and accordingly exhibit broad substrate scope. Gram-scale reaction and diverse follow-up transformations convincingly demonstrate the synthetic potential of this method.

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

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

19.
Intermolecular oxidative amination of unactivated alkenes by dual photoredox and copper catalysis     
Xiangli Yi  Xile Hu 《Chemical science》2021,12(5):1901
  相似文献   

20.
Nickel catalysis enables convergent paired electrolysis for direct arylation of benzylic C–H bonds     
Lei Zhang  Xile Hu 《Chemical science》2020,11(39):10786
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