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 15 毫秒
1.
This study described palladium-catalyzed chemoselective direct α-arylation of carbonyl compounds with chloroaryl triflates in the Ar–Cl bond. The Pd/SelectPhos system showed excellent chemoselectivity toward the Ar–Cl bond in the presence of the Ar–OTf bond with a broad substrate scope and excellent product yields. The electronic and steric hindrance offered by the –PR2 group of the ligand with the C2-alkyl group was found to be the key factor affecting the reactivity and chemoselectivity of the α-arylation reaction. The chemodivergent approach was also successfully employed in the synthesis of flurbiprofen and its derivatives (e.g., –OMe and –F).

Palladium-catalyzed chemoselective direct α-arylation of carbonyl compounds with chloroaryl triflates in the Ar–Cl bond is reported. The effects of –PR2 and C2-alkyl groups of the ligands are investigated using experimental and computational methods.  相似文献   

2.
Nuclear Receptors (NRs) are highly relevant drug targets, for which small molecule modulation goes beyond a simple ligand/receptor interaction. NR–ligands modulate Protein–Protein Interactions (PPIs) with coregulator proteins. Here we bring forward a cooperativity mechanism for small molecule modulation of NR PPIs, using the Peroxisome Proliferator Activated Receptor γ (PPARγ), which describes NR–ligands as allosteric molecular glues. The cooperativity framework uses a thermodynamic model based on three-body binding events, to dissect and quantify reciprocal effects of NR–coregulator binding (KID) and NR–ligand binding (KIID), jointly recapitulated in the cooperativity factor (α) for each specific ternary ligand·NR·coregulator complex formation. These fundamental thermodynamic parameters allow for a conceptually new way of thinking about structure–activity-relationships for NR–ligands and can steer NR modulator discovery and optimization via a completely novel approach.

A cooperativity framework describes the formation of nuclear receptor ternary complexes and deconvolutes ligand and cofactor binding into intrinsic affinities and a cooperativity factor, providing a conceptually new understanding of NR modulation.  相似文献   

3.
Polyesters are important plastics, elastomers and fibres; efficient and selective polymerizations making predictable, high molar mass polymers are required. Here, a new type of catalyst for the ring-opening polymerization (ROCOP) of epoxides and anhydrides combines unusually high chain end-group selectivity, fast rates, and good molar mass control. The organometallic heterodinuclear Al(iii)/K(i) complex, applied with a diol, is tolerant to a range of epoxides/phthalic anhydride and produces only α,ω-hydroxyl telechelic polyesters with molar masses from 6–91 kg mol−1, in all cases with monomodal distributions. As proof of its potential, high molar mass poly(vinyl cyclohexene oxide-alt-phthalic anhydride) (91 kg mol−1) shows 5× greater flexural strain at break (εb = 3.7%) and 9× higher maximum flexural stress (σf = 72.3 MPa) than the previously accessed medium molar mass samples (24 kg mol−1). It is also enchains phthalic anhydride, vinyl cyclohexene oxide and ε-decalactone, via switchable catalysis, to make high molar mass triblock polyesters (81 kg mol−1, Đ = 1.04). This selective catalyst should be used in the future to qualify the properties of these ROCOP polyesters and to tune (multi)block polymer structures.

A heterodinuclear Al(iii)/K(i) organometallic ring-opening copolymerization catalyst shows exceptional rates, end-group selectivity and good loading tolerance to deliver monodisperse polyesters with molar masses up to 91 kg mol−1.  相似文献   

4.
We report here a sequential enantioselective reduction/C–H functionalization to install contiguous stereogenic carbon centers of benzocyclobutenols and cyclobutanols. This strategy features a practical enantioselective reduction of a ketone and a diastereospecific iridium-catalyzed C–H silylation. Further transformations have been explored, including controllable regioselective ring-opening reactions. In addition, this strategy has been utilized for the synthesis of three natural products, phyllostoxin (proposed structure), grandisol and fragranol.

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

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

5.
Identification of the Primary Structure of Selenium-Containing Polysaccharides Selectively Inhibiting T-Cell Proliferation     
Marzenna Klimaszewska  Sabina Grska  Grzegorz &#x;apienis  Beata Kaleta  Sandra Grska  Marta Kaszowska  Maciej Dawidowski  Andrzej Gamian  Radoslaw Zago d on  Andrzej Grski  Jadwiga Tur&#x;o 《Molecules (Basel, Switzerland)》2021,26(17)
We previously described the biosynthesis, isolation, and immunosuppressive activity of the selenium-containing polysaccharide fraction isolated from the mycelial culture of Lentinula edodes. Structural studies have shown that the fraction was a protein-containing mixture of high molar mass polysaccharides α- and β-glucans. However, which of the components of the complex fraction is responsible for the immunosuppressive activity non-typical for polysaccharides of fungal origin has not been explained. In the current study, we defined four-polysaccharide components of the Se-containing polysaccharide fraction determined their primary structure and examined the effect on T- and B-cell proliferation. The isolated Se-polysaccharides, α-1,4-glucan (Mw 2.25 × 106 g/mol), unbranched β-1,6-d-glucan, unbranched β-1,3-d-glucan and β-1,3-branched β-1,6-d-glucan (Mw 1.10 × 105 g/mol), are not typical as components of the cell wall of L. edodes. All are biologically active, but the inhibitory effect of the isolated polysaccharides on lymphocyte proliferation was weaker, though more selective than that of the crude fraction.  相似文献   

6.
Targeting antioxidant pathways with ferrocenylated N-heterocyclic carbene supported gold(i) complexes in A549 lung cancer cells     
J. F. Arambula  R. McCall  K. J. Sidoran  D. Magda  N. A. Mitchell  C. W. Bielawski  V. M. Lynch  J. L. Sessler  K. Arumugam 《Chemical science》2016,7(2):1245-1256
Ferrocene containing N-heterocyclic carbene (NHC) ligated gold(i) complexes of the type [Au(NHC)2]+ were prepared and found to be capable of regulating the formation of reactive oxygen species (ROS) via multiple mechanisms. Single crystal X-ray analysis of bis(1-(ferrocenylmethyl)-3-mesitylimidazol-2-ylidene)-gold(i) chloride (5) and bis(1,3-di(ferrocenylmethyl)imidazol-2-ylidene)-gold(i) chloride (6) revealed a quasi-linear geometry around the gold(i) centers (i.e., the C–Au–C bond angle were measured to be ∼177° and all the Au–Ccarbene bonds distances were in the range of 2.00 (7)–2.03 (1) Å). A series of cell studies indicated that cell proliferation inhibition and ROS generation were directly proportional to the amount of ferrocene contained within the [Au(NHC)2]+ complexes (IC50 of 6 < 5 < bis(1-benzyl-3-mesitylimidazol-2-ylidene)-gold(i) chloride (4)). Complexes 4–6 were also confirmed to inhibit thioredoxin reductase as inferred from lipoate reduction assays and increased chelatable intracellular zinc concentrations. RNA microarray gene expression assays revealed that 6 induces endoplasmic reticulum stress response pathways as a result of ROS increase.  相似文献   

7.
The Antifungal Action Mode of N-Phenacyldibromobenzimidazoles     
Monika Staniszewska  &#x;ukasz Kuryk  Aleksander Gryciuk  Joanna Kawalec  Marta Rogalska  Joanna Baran  Anna Kowalkowska 《Molecules (Basel, Switzerland)》2021,26(18)
Our study aimed to characterise the action mode of N-phenacyldibromobenzimidazoles against C. albicans and C. neoformans. Firstly, we selected the non-cytotoxic most active benzimidazoles based on the structure–activity relationships showing that the group of 5,6-dibromobenzimidazole derivatives are less active against C. albicans vs. 4,6-dibromobenzimidazole analogues (5e–f and 5h). The substitution of chlorine atoms to the benzene ring of the N-phenacyl substituent extended the anti-C. albicans action (5e with 2,4-Cl2 or 5f with 3,4-Cl2). The excellent results for N-phenacyldibromobenzimidazole 5h against the C. albicans reference and clinical isolate showed IC50 = 8 µg/mL and %I = 100 ± 3, respectively. Compound 5h was fungicidal against the C. neoformans isolate. Compound 5h at 160–4 µg/mL caused irreversible damage of the fungal cell membrane and accidental cell death (ACD). We reported on chitinolytic activity of 5h, in accordance with the patterns observed for the following substrates: 4-nitrophenyl-N-acetyl-β-d-glucosaminide and 4-nitrophenyl-β-d-N,N′,N″-triacetylchitothiose. Derivative 5h at 16 µg/mL: (1) it affected cell wall by inducing β-d-glucanase, (2) it caused morphological distortions and (3) osmotic instability in the C. albicans biofilm-treated. Compound 5h exerted Candida-dependent inhibition of virulence factors.  相似文献   

8.
Support stabilized PtCu single-atom alloys for propane dehydrogenation     
Xiaohe Liu  Xianhui Wang  Shiyu Zhen  Guodong Sun  Chunlei Pei  Zhi-Jian Zhao  Jinlong Gong 《Chemical science》2022,13(33):9537
PtCu single-atom alloys (SAAs) open an extensive prospect for heterogeneous catalysis. However, as the host of SAAs, Cu suffers from severe sintering at elevated temperature, resulting in poor stability of catalysts. This paper describes the suppression of the agglomeration of Cu nanoparticles under high temperature conditions using copper phyllosilicate (CuSiO3) as the support of PtCu SAAs. Based on quasi in situ XPS, in situ CO-DRIFTS, in situ Raman spectroscopy and in situ XRD, we demonstrated that the interfacial Cu+–O–Si formed upon reduction at 680 °C serves as the adhesive between Cu nanoparticles and the silicon dioxide matrix, strengthening the metal–support interaction. Consequently, the resistance to sintering of PtCu SAAs was improved, leading to high catalytic stability during propane dehydrogenation without sacrificing conversion and selectivity. The optimized PtCu SAA catalyst achieved more than 42% propane conversion and 93% propylene selectivity at 580 °C for at least 30 hours. It paves a way for the design and development of highly active supported single-atom alloy catalysts with excellent thermal stability.

This paper describes PtCu single-atom alloys supported on copper phyllosilicate via Cu+–O–Si. The catalyst exhibits sintering resistance in propane dehydrogenation reaction without sacrificing activity and selectivity.  相似文献   

9.
Exploring benzylic gem-C(sp3)–boron–silicon and boron–tin centers as a synthetic platform     
Wei W. Chen  Nahiane Pipaon Fernndez  Marta Díaz Baranda  Anton Cunillera  Laura G. Rodríguez  Alexandr Shafir  Ana B. Cuenca 《Chemical science》2021,12(31):10514
A stepwise build-up of multi-substituted Csp3 carbon centers is an attractive, conceptually simple, but often synthetically challenging type of disconnection. To this end, this report describes how gem-α,α-dimetalloid-substituted benzylic reagents bearing boron/silicon or boron/tin substituent sets are an excellent stepping stone towards diverse substitution patterns. These gem-dimetalloids were readily accessed, either by known carbenoid insertion into C–B bonds or by the newly developed scalable deprotonation/metallation approach. Highly chemoselective transformations of either the C–Si (or C–Sn) or the C–B bonds in the newly formed gem-Csp3 centers have been achieved through a set of approaches, with a particular focus on exploiting the synthetically versatile polarity reversal in organometalloids by λ3-aryliodanes. Of particular note is the metal-free arylation of the C–Si (or C–Sn) bonds in such gem-dimetalloids via the iodane-guided C–H coupling approach. DFT calculations show that this transfer of the (α-Bpin)benzyl group proceeds via unusual [5,5]-sigmatropic rearrangement and is driven by the high-energy iodine(iii) center. As a complementary tool, the gem-dimetalloid C–B bond is shown to undergo a potent and chemoselective Suzuki–Miyaura arylation with diverse Ar–Cl, thanks to the development of the reactive gem-α,α-silyl/BF3K building blocks.

This work explores divergent reactivity of the benzylic gem-boron–silicon and boron–tin double nucleophiles, including the arylation of the C–B bond with Ar–Cl, along with a complementary oxidative λ3-iodane-guided arylation of the C–Si/Sn moiety.  相似文献   

10.
Mutual functionalization of dinitrogen and methane mediated by heteronuclear metal cluster anions CoTaC2−     
Li-Hui Mou  Yao Li  Gong-Ping Wei  Zi-Yu Li  Qing-Yu Liu  Hui Chen  Sheng-Gui He 《Chemical science》2022,13(32):9366
The direct coupling of dinitrogen (N2) and methane (CH4) to construct the N–C bond is a fascinating but challenging approach for the energy-saving synthesis of N-containing organic compounds. Herein we identified a likely reaction pathway for N–C coupling from N2 and CH4 mediated by heteronuclear metal cluster anions CoTaC2, which starts with the dissociative adsorption of N2 on CoTaC2 to generate a Taδ+–Ntδ (terminal-nitrogen) Lewis acid–base pair (LABP), followed by the further activation of CH4 by CoTaC2N2 to construct the N–C bond. The N Created by potrace 1.16, written by Peter Selinger 2001-2019 N cleavage by CoTaC2 affording two N atoms with strong charge buffering ability plays a key part, which facilitates the H3C–H cleavage via the LABP mechanism and the N–C formation via a CH3 migration mechanism. A novel Nt triggering strategy to couple N2 and CH4 molecules using metal clusters was accordingly proposed, which provides a new idea for the direct synthesis of N-containing compounds.

A possible N–C bond formation directly from N2 and CH4 mediated by heteronuclear metal cluster anions CoTaC2 at room temperature was identified.  相似文献   

11.
Silica nanoparticle-loaded thermoresponsive block copolymer vesicles: a new post-polymerization encapsulation strategy and thermally triggered release     
Adam Czajka  Sarah J. Byard  Steven P. Armes 《Chemical science》2022,13(33):9569
A thermoresponsive amphiphilic diblock copolymer that can form spheres, worms or vesicles in aqueous media at neutral pH by simply raising the dispersion temperature from 1 °C (spheres) to 25 °C (worms) to 50 °C (vesicles) is prepared via polymerization-induced self-assembly (PISA). Heating such an aqueous copolymer dispersion from 1 °C up to 50 °C in the presence of 19 nm glycerol-functionalized silica nanoparticles enables this remarkable ‘shape-shifting’ behavior to be exploited as a new post-polymerization encapsulation strategy. The silica-loaded vesicles formed at 50 °C are then crosslinked using a disulfide-based dihydrazide reagent. Such covalent stabilization enables the dispersion to be cooled to room temperature without loss of the vesicle morphology, thus aiding characterization and enabling the loading efficiency to be determined as a function of both copolymer and silica concentration. Small-angle X-ray scattering (SAXS) analysis indicated a mean vesicle membrane thickness of approximately 20 ± 2 nm for the linear vesicles and TEM studies confirmed encapsulation of the silica nanoparticles within these nano-objects. After removal of the non-encapsulated silica nanoparticles via multiple centrifugation–redispersion cycles, thermogravimetric analysis indicated that vesicle loading efficiencies of up to 86% can be achieved under optimized conditions. Thermally-triggered release of the silica nanoparticles is achieved by cleaving the disulfide bonds at 50 °C using tris(2-carboxyethyl)phosphine (TCEP), followed by cooling to 20 °C to induce vesicle dissociation. SAXS is also used to confirm the release of silica nanoparticles by monitoring the disappearance of the structure factor peak arising from silica–silica interactions.

A loading efficiency of up to 86% is achieved for silica nanoparticles encapsulated within crosslinkable redox-sensitive thermoresponsive diblock copolymer vesicles in water at 50 °C; triggered release is also demonstrated for this system.  相似文献   

12.
Ti-catalyzed ring-opening oxidative amination of methylenecyclopropanes with diazenes     
Evan P. Beaumier  Amy A. Ott  Xuelan Wen  Zachary W. Davis-Gilbert  T. Alexander Wheeler  Joseph J. Topczewski  Jason D. Goodpaster  Ian A. Tonks 《Chemical science》2020,11(27):7204
The ring-opening oxidative amination of methylenecyclopropanes (MCPs) with diazenes catalyzed by py3TiCl2(NR) complexes is reported. This reaction selectively generates branched α-methylene imines as opposed to linear α,β-unsaturated imines, which are difficult to access via other methods. Products can be isolated as the imine or hydrolyzed to the corresponding ketone in good yields. Mechanistic investigation via density functional theory suggests that the regioselectivity of these products results from a Curtin–Hammett kinetic scenario, where reversible β-carbon elimination of a spirocyclic [2 + 2] azatitanacyclobutene intermediate is followed by selectivity-determining β-hydrogen elimination of the resulting metallacycle. Further functionalizations of these branched α-methylene imine products are explored, demonstrating their utility as building blocks.

The ring-opening oxidative amination of methylenecyclopropanes (MCPs) with diazenes catalyzed by py3TiCl2(NR) complexes is reported.  相似文献   

13.
Ring-opening and ring-expansion reactions of carborane-fused borirane     
Hanqiang Wang  Jie Zhang  Zuowei Xie 《Chemical science》2021,12(39):13187
Though the reaction chemistry of three-membered ring molecules such as cyclopropanes and their heteroatom-containing analogues has been extensively studied, the chemical properties of their boron analogues, boriranes, are little known thus far. This work describes the diverse reactivity patterns of carborane-fused borirane 2. This borirane engages in ring-opening reactions with different types of Lewis acids, such as BBr3, GeCl2, GaCl3, BH3(SMe2) and HBpin, affording a series of ring-opening products, in which M–X or B–H bonds add across the B–C(cage) bond of the three-membered ring in 2. On the other hand, borirane 2 can undergo ring-expansion reactions with unsaturated molecules such as PhCHO, CO2 and PhCN to give ring-expansion products, five-membered boracycles, via a concerted reaction mechanism as supported by DFT calculations. The results of this work not only enrich the reaction chemistry of boriranes, but also offer new routes to boron-containing compounds and heterocycles.

Carborane-fused borirane can not only engage in ring-opening reactions with different types of Lewis acids, but also undergo ring-expansion reactions with unsaturated molecules such as PhCHO, CO2 and PhCN to give five-membered boracycles.  相似文献   

14.
A faux hawk fullerene with PCBM-like properties     
Long K. San  Eric V. Bukovsky  Bryon W. Larson  James B. Whitaker  S. H. M. Deng  Nikos Kopidakis  Garry Rumbles  Alexey A. Popov  Yu-Sheng Chen  Xue-Bin Wang  Olga V. Boltalina  Steven H. Strauss 《Chemical science》2015,6(3):1801-1815
Reaction of C60, C6F5CF2I, and SnH(n-Bu)3 produced, among other unidentified fullerene derivatives, the two new compounds 1,9-C60(CF2C6F5)H (1) and 1,9-C60(cyclo-CF2(2-C6F4)) (2). The highest isolated yield of 1 was 35% based on C60. Depending on the reaction conditions, the relative amounts of 1 and 2 generated in situ were as high as 85% and 71%, respectively, based on HPLC peak integration and summing over all fullerene species present other than unreacted C60. Compound 1 is thermally stable in 1,2-dichlorobenzene (oDCB) at 160 °C but was rapidly converted to 2 upon addition of Sn2(n-Bu)6 at this temperature. In contrast, complete conversion of 1 to 2 occurred within minutes, or hours, at 25 °C in 90/10 (v/v) PhCN/C6D6 by addition of stoichiometric, or sub-stoichiometric, amounts of proton sponge (PS) or cobaltocene (CoCp2). DFT calculations indicate that when 1 is deprotonated, the anion C60(CF2C6F5) can undergo facile intramolecular SNAr annulation to form 2 with concomitant loss of F. To our knowledge this is the first observation of a fullerene-cage carbanion acting as an SNAr nucleophile towards an aromatic C–F bond. The gas-phase electron affinity (EA) of 2 was determined to be 2.805(10) eV by low-temperature PES, higher by 0.12(1) eV than the EA of C60 and higher by 0.18(1) eV than the EA of phenyl-C61-butyric acid methyl ester (PCBM). In contrast, the relative E 1/2(0/–) values of 2 and C60, –0.01(1) and 0.00(1) V, respectively, are virtually the same (on this scale, and under the same conditions, the E 1/2(0/–) of PCBM is –0.09 V). Time-resolved microwave conductivity charge-carrier yield × mobility values for organic photovoltaic active-layer-type blends of 2 and poly-3-hexylthiophene (P3HT) were comparable to those for equimolar blends of PCBM and P3HT. The structure of solvent-free crystals of 2 was determined by single-crystal X-ray diffraction. The number of nearest-neighbor fullerene–fullerene interactions with centroid···centroid (⊙···⊙) distances of ≤10.34 Å is significantly greater, and the average ⊙···⊙ distance is shorter, for 2 (10 nearest neighbors; ave. ⊙···⊙ distance = 10.09 Å) than for solvent-free crystals of PCBM (7 nearest neighbors; ave. ⊙···⊙ distance = 10.17 Å). Finally, the thermal stability of 2 was found to be far greater than that of PCBM.  相似文献   

15.
Indole- and Pyrazole-Glycyrrhetinic Acid Derivatives as PTP1B Inhibitors: Synthesis,In Vitro and In Silico Studies     
Ledy De-la-Cruz-Martínez  Constanza Duran-Becerra  Martin Gonzlez-Andrade  Jos C. Pez-Franco  Juan Manuel Germn-Acacio  Julio Espinosa-Chvez  J. Martin Torres-Valencia  Jaime Prez-Villanueva  Juan Francisco Palacios-Espinosa  Olivia Soria-Arteche  Francisco Corts-Benítez 《Molecules (Basel, Switzerland)》2021,26(14)
Regulating insulin and leptin levels using a protein tyrosine phosphatase 1B (PTP1B) inhibitor is an attractive strategy to treat diabetes and obesity. Glycyrrhetinic acid (GA), a triterpenoid, may weakly inhibit this enzyme. Nonetheless, semisynthetic derivatives of GA have not been developed as PTP1B inhibitors to date. Herein we describe the synthesis and evaluation of two series of indole- and N-phenylpyrazole-GA derivatives (4a–f and 5a–f). We measured their inhibitory activity and enzyme kinetics against PTP1B using p-nitrophenylphosphate (pNPP) assay. GA derivatives bearing substituted indoles or N-phenylpyrazoles fused to their A-ring showed a 50% inhibitory concentration for PTP1B in a range from 2.5 to 10.1 µM. The trifluoromethyl derivative of indole-GA (4f) exhibited non-competitive inhibition of PTP1B as well as higher potency (IC50 = 2.5 µM) than that of positive controls ursolic acid (IC50 = 5.6 µM), claramine (IC50 = 13.7 µM) and suramin (IC50 = 4.1 µM). Finally, docking and molecular dynamics simulations provided the theoretical basis for the favorable activity of the designed compounds.  相似文献   

16.
Shape-shifting thermoreversible diblock copolymer nano-objects via RAFT aqueous dispersion polymerization of 4-hydroxybutyl acrylate     
Oliver J. Deane  James Jennings  Steven P. Armes 《Chemical science》2021,12(41):13719
2-Hydroxypropyl methacrylate (HPMA) is a useful model monomer for understanding aqueous dispersion polymerization. 4-Hydroxybutyl acrylate (HBA) is an isomer of HPMA: it has appreciably higher aqueous solubility so its homopolymer is more weakly hydrophobic. Moreover, PHBA possesses a significantly lower glass transition temperature than PHPMA, which ensures greater chain mobility. The reversible addition–fragmentation chain transfer (RAFT) aqueous dispersion polymerization of HBA using a poly(ethylene glycol) (PEG113) precursor at 30 °C produces PEG113–PHBA200–700 diblock copolymer nano-objects. Using glutaraldehyde to crosslink the PHBA chains allows TEM studies, which reveal the formation of spheres, worms or vesicles under appropriate conditions. Interestingly, the partially hydrated highly mobile PHBA block enabled linear PEG113–PHBAx spheres, worms or vesicles to be reconstituted from freeze-dried powders on addition of water at 20 °C. Moreover, variable temperature 1H NMR studies indicated that the apparent degree of hydration of the PHBA block increases from 5% to 80% on heating from 0 °C to 60 °C indicating uniform plasticization. In contrast, the PHPMAx chains within PEG113–PHPMAx nano-objects become dehydrated on raising the temperature: this qualitative difference is highly counter-intuitive given that PHBA and PHPMA are isomers. The greater (partial) hydration of the PHBA block at higher temperature drives the morphological evolution of PEG113–PHBA260 spheres to form worms or vesicles, as judged by oscillatory rheology, dynamic light scattering, small-angle X-ray scattering and TEM studies. Finally, a variable temperature phase diagram is constructed for 15% w/w aqueous dispersions of eight PEG113–PHBA200–700 diblock copolymers. Notably, PEG113–PHBA350 can switch reversibly from spheres to worms to vesicles to lamellae during a thermal cycle.

RAFT aqueous dispersion polymerization of 4-hydroxybutyl acrylate (HBA) affords shape-shifting thermoresponsive diblock copolymer nano-objects. 1H NMR studies suggest that such behavior involves uniform plasticization of the PHBA block.  相似文献   

17.
Solvent coordination to palladium can invert the selectivity of oxidative addition     
Emily K. Elias  Steven M. Rehbein  Sharon R. Neufeldt 《Chemical science》2022,13(6):1618
Reaction solvent was previously shown to influence the selectivity of Pd/PtBu3-catalyzed Suzuki–Miyaura cross-couplings of chloroaryl triflates. The role of solvents has been hypothesized to relate to their polarity, whereby polar solvents stabilize anionic transition states involving [Pd(PtBu3)(X)] (X = anionic ligand) and nonpolar solvents do not. However, here we report detailed studies that reveal a more complicated mechanistic picture. In particular, these results suggest that the selectivity change observed in certain solvents is primarily due to solvent coordination to palladium. Polar coordinating and polar noncoordinating solvents lead to dramatically different selectivity. In coordinating solvents, preferential reaction at triflate is likely catalyzed by Pd(PtBu3)(solv), whereas noncoordinating solvents lead to reaction at chloride through monoligated Pd(PtBu3). The role of solvent coordination is supported by stoichiometric oxidative addition experiments, density functional theory (DFT) calculations, and catalytic cross-coupling studies. Additional results suggest that anionic [Pd(PtBu3)(X)] is also relevant to triflate selectivity in certain scenarios, particularly when halide anions are available in high concentrations.

In the presence of the bulky monophosphine PtBu3, palladium usually prefers to react with Ar–Cl over Ar–OTf bonds. However, strongly coordinating solvents can bind to palladium, inducing a reversal of selectivity.

Oxidative addition is a key elementary step in diverse transformations catalyzed by transition metals.1 For instance, this step is common to traditional cross-coupling reactions, which are among the most widely used methods for small molecule synthesis. During the oxidative addition step of cross-coupling reactions, a low valent metal [usually Pd(0)] inserts into a C–X bond with concomitant oxidation of the metal by two electrons. The “X” group of the C–X bond is commonly a halogen or triflate. Despite a wealth of research into this step,2–5 uncertainties remain about its mechanistic nuances. The mechanistic details are especially pertinent to issues of selectivity that arise when substrates contain more than one potentially reactive C–X bond.6One of the best-studied examples of divergent selectivity at the oxidative addition step is the case of Pd-catalyzed Suzuki couplings of chloroaryl triflates. In 2000, Fu reported that a combination of Pd(0) and PtBu3 in tetrahydrofuran (THF) effects selective coupling of 1 with o-tolylB(OH)2via C–Cl cleavage, resulting in retention of the triflate substituent in the final product 2a (Scheme 1A).7 In contrast, the use of PCy3 (ref. 7) or most other phosphines8 provides complementary selectivity (product 2b) under similar conditions. The unique selectivity imparted by PtBu3 was later attributed to this ligand''s ability to promote a monoligated oxidative addition transition state on account of its bulkiness.5,8 Smaller ligands, on the other hand, favor bisligated palladium, which prefers to react at triflate. The relationship between palladium''s ligation state and chemoselectivity has been rationalized by Schoenebeck and Houk through a distortion/interaction analysis.5 In brief, the selectivity preference of PdL2 is dominated by a strong interaction between the electron-rich Pd and the more electrophilic site (C–OTf). On the other hand, PdL is less electron-rich and its selectivity preference mainly relates to minimizing unfavorable distortion energy by reacting at the more easily-distorted C–Cl bond.Open in a separate windowScheme 1Seminal reports on the effects of (A) ligands and (B) solvents on the selectivity of cross-coupling of a chloroaryl triflate.5,7,9Proutiere and Schoenebeck later discovered that replacing THF with dimethylformamide (DMF, Scheme 1B, entry 1) or acetonitrile caused a change in selectivity for the Pd/PtBu3 system.9,10 In these two polar solvents, preferential reaction at triflate was observed, and PtBu3 no longer displayed its unique chloride selectivity. The possibility of solvent coordination to Pd was considered, as bisligated Pd(PtBu3)(solv) would be expected to favor reaction at triflate. However, solvent coordination was ruled out on the basis of two intriguing studies. First, DFT calculations using the functional B3LYP suggested that solvent-coordinated transition states are prohibitively high in free energy (about 16 kcal mol−1 higher than the lowest-energy monoligated transition structure). Second, the same solvent effect was not observed in a Pd/PtBu3-catalyzed base-free Stille coupling in DMF (Scheme 1B, entry 2). Instead, the Stille coupling was reported to favor reaction at chloride despite the use of a polar solvent. This result appears inconsistent with the possibility that solvent coordination induces triflate-selectivity, as coordination of DMF to Pd should be possible in both the Stille and Suzuki conditions, if it happens at all. Instead, it was proposed that the key difference between the Suzuki and Stille conditions was the absence of coordinating anions in the latter (unlike traditional Suzuki couplings, Stille couplings do not necessarily require basic additives such as KF to promote transmetalation). Indeed, when KF or CsF was added to the Stille reaction in DMF, selectivity shifted to favor reaction at triflate (Scheme 1B, entry 3), thereby displaying the same behavior as the Suzuki coupling in this solvent. On the basis of this and the DFT studies, it was proposed that polar solvents induce a switch in chemoselectivity if coordinating anions like fluoride are available by stabilizing anionic bisligated transition structures (Scheme 1B, right).However, our recent extended solvent effect studies produced confounding results.11 In a Pd/PtBu3-catalyzed Suzuki cross-coupling of chloroaryl triflate 1, we observed no correlation between solvent polarity and chemoselectivity (Scheme 2). Although some polar solvents such as MeCN, DMF, and dimethylsulfoxide (DMSO) favor reaction at triflate, a number of other polar solvents provide the same results as nonpolar solvents by favoring reaction at chloride. For example, cross-coupling primarily takes place through C–Cl cleavage when the reaction is conducted in highly polar solvents like methanol, water, acetone, and propylene carbonate. In fact, the only solvents that promote reaction at triflate are ones that are commonly thought of as “coordinating” in the context of late transition metal chemistry.12 These are solvents containing nitrogen, sulfur, or electron-rich oxygen lone pairs (nitriles, DMSO, and amides). The observed solvent effects were upheld for a variety of chloroaryl triflates and aryl boronic acids.11Open in a separate windowScheme 2Expanded solvent effect studies in the Pd/PtBu3-catalyzed Suzuki coupling.11We have sought to reconcile these observations with the earlier evidence9 against solvent coordination. Herein we report detailed mechanistic studies indicating that coordinating solvents alone are sufficient to induce the observed selectivity switch. In solvents like DMF and MeCN, stoichiometric oxidative addition is favored at C–OTf even in the absence of anionic additives. The apparent contradiction between our observations and the previously-reported DFT calculations and base-free Stille couplings is reconciled by a reevaluation of those studies. In particular, when dispersion is considered in DFT calculations, neutral solvent-coordinated transition structures involving Pd(PtBu3)(solv) become energetically feasible. Furthermore, we find that the selectivity analysis in the Stille couplings is convoluted by low yields, the formation of side products, and temperature effects. When these factors are disentangled, the Stille coupling in DMF displays selectivity similar to the Suzuki coupling in the same coordinating solvent. In light of these new results, anionic bisligated [Pd(PtBu3)(X)] does not appear to be the dominant active catalyst in nonpolar or polar solvents unless special measures are taken to increase the concentration of free halide, such as adding tetraalkylammonium halide salts or crown ethers.  相似文献   

18.
The Janus face of high trans-effect carbenes in olefin metathesis: gateway to both productivity and decomposition     
Giovanni Occhipinti  Daniel L. Nascimento  Marco Foscato  Deryn E. Fogg  Vidar R. Jensen 《Chemical science》2022,13(18):5107
Ruthenium–cyclic(alkyl)(amino)carbene (CAAC) catalysts, used at ppm levels, can enable dramatically higher productivities in olefin metathesis than their N-heterocyclic carbene (NHC) predecessors. A key reason is the reduced susceptibility of the metallacyclobutane (MCB) intermediate to decomposition via β-H elimination. The factors responsible for promoting or inhibiting β-H elimination are explored via density functional theory (DFT) calculations, in metathesis of ethylene or styrene (a representative 1-olefin) by Ru–CAAC and Ru–NHC catalysts. Natural bond orbital analysis of the frontier orbitals confirms the greater strength of the orbital interactions for the CAAC species, and the consequent increase in the carbene trans influence and trans effect. The higher trans effect of the CAAC ligands inhibits β-H elimination by destabilizing the transition state (TS) for decomposition, in which an agostic MCB Cβ–H bond is positioned trans to the carbene. Unproductive cycling with ethylene is also curbed, because ethylene is trans to the carbene ligand in the square pyramidal TS for ethylene metathesis. In contrast, metathesis of styrene proceeds via a ‘late’ TS with approximately trigonal bipyramidal geometry, in which carbene trans effects are reduced. Importantly, however, the positive impact of a strong trans-effect ligand in limiting β-H elimination is offset by its potent accelerating effect on bimolecular coupling, a major competing means of catalyst decomposition. These two decomposition pathways, known for decades to limit productivity in olefin metathesis, are revealed as distinct, antinomic, responses to a single underlying phenomenon. Reconciling these opposing effects emerges as a clear priority for design of robust, high-performing catalysts.

In ruthenium catalysts for olefin metathesis, carbene ligands of high trans influence/effect suppress decomposition via β-H elimination, but increase susceptibility to bimolecular decomposition.  相似文献   

19.
Guest-mediated phase transitions in a flexible pillared-layered metal–organic framework under high-pressure     
Gemma F. Turner  Scott C. McKellar  David R. Allan  Anthony K. Cheetham  Sebastian Henke  Stephen A. Moggach 《Chemical science》2021,12(41):13793
The guest-dependent flexibility of the pillared-layered metal–organic framework (MOF), Zn2bdc2dabco·X(guest), where guest = EtOH, DMF or benzene, has been examined by high-pressure single crystal X-ray diffraction. A pressure-induced structural phase transition is found for the EtOH- and DMF-included frameworks during compression in a hydrostatic medium of the guest species, which is dependent upon the nature and quantity of the guest in the channels. The EtOH-included material undergoes a phase transition from P4/mmm to C2/m at 0.69 GPa, which is accompanied by a change in the pore shape from square to rhombus via super-filling of the pores. The DMF-included material undergoes a guest-mediated phase transition from I4/mcm to P4/mmm at 0.33 GPa via disordering of the DMF guest. In contrast, the benzene-included framework features a structure with rhombus-shaped channels at ambient pressure and shows direct compression under hydrostatic pressure. These results demonstrate the large influence of guest molecules on the high-pressure phase behavior of flexible MOFs. Guest-mediated framework flexibility is useful for engineering MOFs with bespoke pore shapes and compressibility.

The guest-dependent flexibility of the pillared-layered metal–organic framework (MOF), Zn2bdc2dabco·X(guest), where guest = EtOH, DMF or benzene, has been examined by high-pressure single crystal X-ray diffraction.  相似文献   

20.
Rh(iii)-Catalyzed [5 + 1] annulation of 2-alkenylanilides and 2-alkenylphenols with allenyl acetates     
Anurag Singh  Rahul K. Shukla  Chandra M. R. Volla 《Chemical science》2022,13(7):2043
Herein, we report a mild and highly regioselective Rh(iii)-catalyzed non-oxidative [5 + 1] vinylic C–H annulation of 2-alkenylanilides with allenyl acetates, which has been elusive so far. The reaction proceeds via vinylic C–H activation, regioselective 2,3-migratory insertion, β-oxy elimination followed by nucleophilic cyclization to get direct access to 1,2-dihydroquinoline derivatives. The strategy was also successfully extended to C–H activation of 2-alkenylphenols for constructing chromene derivatives. In the overall [5 + 1] annulation, the allene serves as a one carbon unit. The acetate group on the allene is found to be crucial both for controlling the regio- and chemoselectivity of the reaction and also for facilitating β-oxy elimination. The methodology was scalable and also further extended towards late stage functionalization of various natural products.

A highly regioselective Rh(iii)-catalyzed non-oxidative [5 + 1] vinylic C–H annulation of 2-alkenylanilides and 2-alkenylphenols with allenyl acetates was described for accessing dihyroquinoline and chromene derivatives.  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