首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Ion substitution and micromorphology control are two efficient strategies to ameliorate the electrochemical performance of supercapacitors electrode materials. Here, Al3+ doped α-Ni(OH)2 with peony-like morphology and porous structure has been successfully synthesized through a facile one-pot hydrothermal process. The Al3+ doped α-Ni(OH)2 electrode shows an ultrahigh specific capacitance of 1750 F g−1 at 1 A g−1, and an outstanding electrochemical stability of 72% after running 2000 cycles. In addition, the Al3+ doped α-Ni(OH)2 electrode demonstrates an excellent rate capability (92% retention at 10 A g−1). Furthermore, by using this unique Al3+ doped α-Ni(OH)2 as the positive electrode and a hierarchical porous carbon (HPC) as the negative electrode, the assembled asymmetric supercapacitor can demonstrate a high energy/power density (49.6 W h kg−1 and 14 kW kg−1). This work proves that synthesizing an Al3+ doped structure is an effective means to improve the electrochemical properties of α-Ni(OH)2. This scheme could be extended to other transition metal hydroxides to enhance their electrochemical performance.

Ion substitution and micromorphology control are two efficient strategies to ameliorate the electrochemical performance of supercapacitors electrode materials.  相似文献   

2.
Donor–acceptor π-conjugated random copolymers based on regioregular poly(3-hexylthiophene), rr-P3HT, with unsymmetrical monothienoisoindigo moieties were obtained by direct arylation polycondensation of 2-bromo-3-hexylthiophene with unsymmetrical monothienoisoindigo motifs under the optimized conditions [palladium-immobilized on thiol-modified silica gel with chloride counter anions, PITS-Cl (2.5 mol%), PivOH (1.0 equiv.), K2CO3 (3.0 equiv.), DMAc, 100 °C, 24 h]. Incorporation of unsymmetrical monothienoisoindigo electron-acceptor units into the polymers tuned their highest occupied and lowest unoccupied molecular orbital levels, which were close to those of the hole transport material (PEDOT) and electron transport material (PCBM), respectively, in thin-film organic solar cells. Alkyl chains of the unsymmetrical monothienoisoindigo units in the polymers tuned their macrostructural order, resulting in the observation of crystalline patterns and specific absorption peaks in thin films. An organic solar cell containing the most crystalline random copolymer showed an efficiency of 1.91%.

Donor–acceptor π-conjugated random copolymers based on regioregular poly(3-hexylthiophene) with unsymmetrical monothienoisoindigo moieties were obtained by direct arylation polycondensation.  相似文献   

3.
β-Ni(OH)2 nanoplatelets are prepared by a hydrothermal procedure and characterized by scanning and transmission electron microscopy, X-ray diffraction analysis, Raman spectroscopy, and X-ray photoelectron spectroscopy. The material is demonstrated to be an efficient electrocatalyst for oxygen reduction, oxygen evolution, and hydrogen evolution reactions in alkaline media. β-Ni(OH)2 shows an overpotential of 498 mV to reach 10 mA cm−2 towards oxygen evolution, with a Tafel slope of 149 mV dec−1 (decreasing to 99 mV dec−1 at 75 °C), along with superior stability as evidenced by chronoamperometric measurements. Similarly, a low overpotential of −333 mV to reach 10 mA cm−2 (decreasing to only −65 mV at 75 °C) toward hydrogen evolution with a Tafel slope of −230 mV dec−1 is observed. Finally, β-Ni(OH)2 exhibits a noteworthy performance for the ORR, as evidenced by a low Tafel slope of −78 mV dec−1 and a number of exchanged electrons of 4.01 (indicating direct 4e-oxygen reduction), whereas there are only a few previous reports on modest ORR activity of pure Ni(OH)2.

β-Ni(OH)2 nanoplatelets produced via a hydrothermal method exhibit good performance as trifunctional electrocatalysts for the ORR, OER, and HER in alkaline media along with excellent stability under cathodic/anodic polarisation conditions.  相似文献   

4.
Electrochemical oxidative degradation (EOD) is a particularly promising technique for removing organic pollutants from wastewater. However, due to the high overpotential of EOD in conventional anode materials, the energy cost of EOD is usually very high, which greatly promotes the search for highly active, stable, and energy-efficient anodic catalysts. Herein, we demonstrated that nickel-foam-supported (NF-supported) β-Ni(OH)2 (NF/β-Ni(OH)2) prepared via a facile hydrothermal method could be used as an energy efficient anode for EOD. The as-prepared 3D porous NF/β-Ni(OH)2 exhibited high activity toward the electrochemical oxidation of methyl orange (MO) in the low potential region (<1.07 V vs. SCE). This property differs greatly from those of the conventional anode materials that require a high positive potential to keep them active for EOD, making NF/β-Ni(OH)2 an energy-efficient and active anode material for EOD. With an oxidation current density of 0.25 mA cm−2, the decolorization of MO was completed within 30 min, and the COD removal after 3h of reaction was 63.0%. The normalized energy consumption for the 3 h degradation of MO was 22.2 kW h (kg COD)−1, which is only a fraction of (or even one tenth of) the values reported in the literature. Moreover, NF/β-Ni(OH)2 had a good stability and recyclability for EOD. No activity decay was observed during 10 h of EOD and the COD removal remained almost unchanged after four consecutive reaction cycles. We demonstrated experimentally that the NF/β-Ni(OH)2 anode could generate large amounts of hydroxyl radicals and that the oxidation of MO by hydroxyl radicals was the main mechanism during EOD. We believe that this work opens a new avenue for developing highly active and energy-efficient anode materials that can work in the low potential region for EOD.

A novel NF/β-Ni(OH)2 catalyst for energy efficient electrochemical degradation of methyl orange was fabricated via a facile hydrothermal method.  相似文献   

5.
We investigated the addition of alkali metal elements (namely Na+, K+, Rb+, and Cs+) to Cu-modified CH3NH3PbI3−δClδ photovoltaic devices and their effects on the photovoltaic properties and electronic structure. The open-circuit voltage was increased by CuBr2 addition to the CH3NH3PbI3−δClδ precursor solution. The series resistance was decreased by simultaneous addition of CuBr2 and RbI, which increased the external quantum efficiencies in the range of 300–500 nm, and the short-circuit current density. The energy gap of the perovskite crystal increased through CuBr2 addition, which we also confirmed by first-principles calculations. Charge carrier generation was observed in the range of 300–500 nm as an increase of the external quantum efficiency, owing to the partial density of states contributed by alkali metal elements. Calculations suggested that the Gibbs energies were decreased by incorporation of alkali metal elements into the perovskite crystals. The conversion efficiency was maintained for 7 weeks for devices with added CuBr2 and RbI.

Addition of alkali metal elements (Na+, K+, Rb+, and Cs+) to Cu-modified CH3NH3PbI3−δClδ devices improved the photovoltaic properties.  相似文献   

6.
The triosmium cluster [Os3(CO)10(μ-OH)(μ-H)] containing bridging hydride and hydroxyl groups at a common Os–Os edge was obtained in good yield (ca. 75%) from the hydrolysis of the labile triosmium cluster [Os3(CO)10(NCMe)2] in THF at 67 °C. [Os3(CO)10(μ-OH)(μ-H)] reacts with dppm at 68 °C to afford the isomeric clusters 1 and 2 with the general formula [Os3(CO)8(μ-OH)(μ-H)(μ-dppm)] that differ by the disposition of bridging dppm ligand. Cluster 1 is produced exclusively from the reaction of [Os3(CO)10(μ-OH)(μ-H)] with dppm in CH2Cl2 at room temperature in the presence of added Me3NO. Heating cluster 1 at 81 °C furnishes 2 in a process that likely proceeds by the release of one arm of the dppm ligand, followed by ligand reorganization about the cluster polyhedron and ring closure of the pendent dppm ligand. The oxo-capped [Os3(CO)73-CO)(μ3-O)(μ-dppm)] (3) has been isolated starting from the thermolysis of either 1 or 2 at 139 °C. Reactions of [Os3(CO)10(μ-dppm)] with ROH (R = Me, Et) in the presence of Me3NO at 80 °C furnish [Os3(CO)8(μ-OH)(μ,η11-OCOR)(μ-dppm)] (4, R = Me; 5, R = Et). Clusters 1–5 have been characterized by a combination of analytical and spectroscopic studies, and the molecular structure of each product has been established by X-ray crystallography. The bonding in these products has been examined by electronic structure calculations, and cluster 1 is confirmed as the kinetic product of substitution, while cluster 2 represents the thermodynamically favored isomer.

The cluster [Os3(CO)10(μ-OH)(μ-H)] was obtained in 75% from the hydrolysis of [Os3(CO)10(NCMe)2].  相似文献   

7.
The low utilization rate of active materials has been a critical obstacle for the industrialization of ultracapacitors. In this study, a thin layer of cross-structured ultrathin α-Ni(OH)2 nanosheets was successfully grown in situ on the surface of a nickel foam as a high-conductivity framework by a vibratory water bath route under a low temperature (80 °C) and mild conditions. Combining the ultrathin α-Ni(OH)2 nanosheets and ultrashort electron transport, the strategy of a perfect intercalation structure of α-Ni(OH)2 and a thin layer of active material on a continuous conductive framework resulted in a high utilization rate of active material, which further achieved high specific capacitance of 213.55 F g−1 at 1 A g−1 in a two-electrode system and high capacitance retention from three to two electrode system (753.79 F g−1 at 1 A g−1 in the three-electrode system). Meanwhile, the device also achieved high energy density of 74.94 W h kg−1 at power density of 197.4 W kg−1 and still retained 24.87 W h kg−1 at power density of 3642 W kg−1.

The low utilization rate of active materials has been a critical obstacle for the industrialization of ultracapacitors.  相似文献   

8.
In(OTf)3-catalyzed intramolecular hydroarylation of α-phenylallyl β-ketosulfones provides sulfonyl 1-benzosuberones and 1-tetralones in moderate to good yields in refluxing (CH2Cl)2 under open-vessel and easy-operation reaction conditions. A plausible mechanism is proposed and discussed. This highly regioselective protocol provides an atom-economic ring-closure route.

In(OTf)3-catalyzed intramolecular hydroarylation of α-phenylallyl β-ketosulfones provides sulfonyl 1-benzosuberones and 1-tetralones in moderate to good yields in refluxing (CH2Cl)2 under open-vessel and easy-operation reaction conditions.  相似文献   

9.
The synthesis of γG, γA, γM, β1C1A, C′1 esterase inhibitor, ceruloplasmin, transferrin, hemopexin, haptoglobin, fibrinogen, α1-antitrypsin, orosomucoid, β-lipoprotein, α2-macroglobulin, and prealbumin was studied in 15 normal human embryos and fetuses of 29 days to 18 wk gestation and in the yolk sacs of four embryos from 5.5 to 11.5 wk gestation using tissue culture in 14C-labeled amino acids followed by radioimmunoelectrophoresis. The human embryo as early as 29 day gestation synthesized β1C1A, C′1 esterase inhibitor, transferrin, hemopexin, α1-antitrypsin, β-lipoprotein, α2-macroglobulin, and prealbumin in culture. At 32 days gestation ceruloplasmin and orosomucoid were also synthesized, but synthesis of fibrinogen was not observed before 5.5 wk. Synthesis of γM occurred as early as 10.5 wk gestation, and γG synthesis was found in cultures as early as 12 wk gestation; γA synthesis was not detected in any of the tissue cultures. With the exception of the γ-globulins, each of the proteins studied was synthesized by the liver, but additional sites of synthesis for some of these proteins were also found. Synthesis of γG and γM occurred primarily in the spleen, but other sites of synthesis were noted as well.  相似文献   

10.
Two Dy(iii) single-ion magnets, formulated as [Dy(Phen)(Cl-tcpb)3] (Cl-1) and [Dy(Phen)(CH3O-tmpd)3] (CH3O-2) were obtained through β-diketonate ligands (Cl-tcpb = 1-(4-chlorophenyl)-4,4,4-trifluoro-1,3-butanedione and CH3O-tmpd = 4,4,4-trifluoro-1-(4-methoxyphenyl)-1,3-butanedione) with different substituent groups (–Cl/–OCH3) and auxiliary ligand, 1,10-phenanthroline (Phen). The Dy(iii) ions in Cl-1 and CH3O-2 are eight-coordinate, with an approximately square antiprismatic (SAP, D4d) and trigonal dodecahedron (D2d) N2O6 coordination environment, respectively, in the first coordination sphere. Under zero direct-current (dc) field, magnetic investigations demonstrate that both Cl-1 and CH3O-2 display dynamic magnetic relaxation of single-molecule magnet (SMM) behavior with different effective barriers (Ueff) of 105.4 cm−1 (151.1 K) for Cl-1 and 132.5 cm−1 (190.7 K) for CH3O-2, respectively. As noted, compound CH3O-2 possesses a higher effective barrier than Cl-1. From ab initio calculations, the energies of the first excited state (KD1) are indeed close to the experimental Ueff as 126.7 cm−1vs. 105.4 cm−1 for Cl-1 and 152.8 cm−1vs. 132.5 cm−1 for CH3O-2. The order of the calculated energies of KD1 is same as that of the experimental Ueff. The superior SIM properties of CH3O-2 could have originated from the larger axial electrostatic potential (ESP(ax)) felt by the central Dy(iii) ion when compared with Cl-1. The larger ESP(ax) of CH3O-2 arises from synergic effects of the more negative charge and shorter Dy–O distances of the axial O atoms of the first sphere. These charges and distances could be influenced by functional groups outside the first sphere, e.g., –Cl and –OCH3.

Further studies from the viewpoint of electrostatic potential demonstrate that the larger axial electrostatic potential (ESP) felt by the central Dy (iii) ion of CH3O-2 is responsible for its better SIM property when compared with Cl-1.  相似文献   

11.
To understand the dynamics of the cation in layered perovskite-type (CH3CH2NH3)2CuCl4, the temperature-dependent chemical shifts and spin–lattice relaxation times T in the rotating frame have been measured using 1H magic angle spinning nuclear magnetic resonance (MAS NMR) and 13C cross-polarization (CP)/MAS NMR techniques. Each proton and carbon in the (CH3CH2NH3)+ cation is distinguished in MAS NMR spectra. The Bloembergen–Purcell–Pound (BPP) curves for 1H T in CH3CH2 and NH3, and for the 13C T in CH3 and CH2 are revealed to have minima at low temperatures. This implies that the curves represent the CH3 and NH3+ rotational motions. The amplitude of the cationic motion is enhanced at the C-end, that is, the N-end of the organic cation is fixed to the inorganic layer through N–H⋯Cl hydrogen bonds, and T becomes short with larger-amplitude molecular motions.

To understand the dynamics of the cation in layered perovskite-type (CH3CH2NH3)2CuCl4, the temperature-dependent chemical shifts and spin–lattice relaxation times T have been measured using 1H MAS NMR and 13C CP/MAS NMR techniques.  相似文献   

12.
The site preference for ligand substitution in the benzothiazolate-bridged cluster HOs3(CO)10(μ-1,2-N,C-η11-C7H4NS) (1) has been investigated using PPh3. 1 reacts with PPh3 in the presence of Me3NO to afford the mono- and bisphosphine substituted clusters HOs3(CO)9(PPh3)(μ-1,2-N,C-η11-C7H4NS) (2) and HOs3(CO)8(PPh3)2(μ-1,2-N,C-η11-C7H4NS) (3), respectively. 2 exists as a pair of non-interconverting isomers where the PPh3 ligand is situated at one of the equatorial sites syn to the edge-bridging hydride that shares a common Os–Os bond with the metalated heterocycle. The solid-state structure of the major isomer establishes the PPh3 regiochemistry at the N-substituted osmium center. DFT calculations confirm the thermodynamic preference for this particular isomer relative to the minor isomer whose phosphine ligand is located at the adjacent C-metalated osmium center. 2 also reacts with PPh3 to give 3. The locus of the second substitution occurs at one of the two equatorial sites at the Os(CO)4 moiety in 2 and gives rise to a pair of fluxional stereoisomers where the new phosphine ligand is scrambled between the two equatorial sites at the Os(CO)3P moiety. The molecular structure of the major isomer has been determined by X-ray diffraction analysis and found to represent the lowest energy structure of the different stereoisomers computed for HOs3(CO)8(PPh3)2(μ-1,2-N,C-η11-C7H4NS). The fluxional behavior displayed by 3 has been examined by VT NMR spectroscopy, and DFT calculations provide evidence for stereoselective tripodal rotation at the Os(CO)3P moiety that serves to equilibrate the second phosphine between the two available equatorial sites.

The site preference for PPh3 substitution in HOs3(CO)10(PPh3)2(μ-1,2-N,C-η11-C7H4NS) (1) has been investigated.  相似文献   

13.
The β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by a bifunctional chiral tertiary amine has been developed, which provides an efficient access to optically active β-position functionalized pyrrolidin-2-one derivatives in both high yield and enantioselectivity (up to 78% yield and 95 : 5 er). This is the first catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach.

The asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by (DHQD)2AQN has been developed, which provides an access to β-position functionalized pyrrolidin-2-one derivatives in high levels yield and enantioselectivity.

Metal-free organocatalytic asymmetric transformations have successfully captured considerable enthusiasm of chemists as powerful methods for the synthesis of various kinds of useful chiral compounds ranging from the preparation of biologically important molecules through to novel materials.1 Chiral pyrrolidin-2-ones have been recognized as important structural motifs that are frequently encountered in a variety of biologically active natural and synthetic compounds.2 In particular, the β-position functionalized pyrrolidin-2-one backbones, which can serve as key synthetic precursors for inhibitory neurotransmitters γ-aminobutyric acids (GABA),3 selective GABAB receptor agonists4 as well as antidepressant rolipram analogues,5 have attracted a great deal of attention. Therefore, the development of highly efficient, environmentally friendly and convenient asymmetric synthetic methods to access these versatile frameworks is particularly appealing.As a direct precursor to pyrrolidin-2-one derivatives, recently, α,β-unsaturated γ-butyrolactam has emerged as the most attractive reactant in asymmetric organometallic or organocatalytic reactions for the synthesis of chiral γ-position functionalized pyrrolidin-2-ones (Scheme 1). These elegant developments have been achieved in the research area of catalytic asymmetric vinylogous aldol,6 Mannich,7 Michael8 and annulation reactions9 in the presence of either metal catalysts or organocatalysts (a, Scheme 1). These well-developed catalytic asymmetric methods have been related to the γ-functionalized α,β-unsaturated γ-butyrolactam to date. However, in sharp contrast, the approaches toward introducing C-3 chirality at the β-position of butyrolactam through a direct catalytic manner are underdeveloped (b, Scheme 1)10 in spite of the fact that β-selective chiral functionalization of butyrolactam can directly build up α,β-functionalized pyrrolidin-2-one frameworks.Open in a separate windowScheme 1Different reactive position of α,β-unsaturated γ-butyrolactam in catalytic asymmetric reactions.So far, only a few metal-catalytic enantioselective β-selective functionalized reactions have been reported. For examples, a rhodium/diene complex catalyzed efficient asymmetric β-selective arylation10a and alkenylation10b have been reported by Lin group (a, Scheme 2). Procter and co-workers reported an efficient Cu(i)–NHC-catalyzed asymmetric silylation of unsaturated lactams (b, Scheme 2).10c Despite these creative works, considerable challenges still exist in the catalytic asymmetric β-selective functionalization of γ-butyrolactam. First, the scope of nucleophiles is limited to arylboronic acids, potassium alkenyltrifluoroborates and PhMe2SiBpin reagents. Second, the catalytic system and activation mode is restricted to metal/chiral ligands. To our knowledge, an efficient catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach has not yet been established. Therefore, the development of organocatalytic asymmetric β-selective functionalization of γ-butyrolactam are highly desirable. In conjunction with our continuing efforts in building upon chiral precedents by using chiral tertiary amine catalytic system,11 we rationalized that the activated α,β-unsaturated γ-butyrolactam might serve as a β-position electron-deficient electrophile. This γ-butyrolactam may react with a properly designed electron-rich nucleophile to conduct an expected β-selective functionalized reaction of γ-butyrolactam under a bifunctional organocatalytic fashion, while avoiding the direct γ-selective vinylogous addition reaction or β,γ-selective annulation as outlined in Scheme 2. Herein we report the β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters12 catalyzed by a bifunctional chiral tertiary amine, which provides an efficient and facile access to optically active β-position functionalized pyrrolidin-2-one derivatives with both high diastereoselectivity and enantioselectivity.Open in a separate windowScheme 2β-Selective functionalization of γ-butyrolactam via metal- (previous work) or organo- (this work) catalytic approach.To begin our initial investigation, several bifunctional organocatalysts13 were firstly screened to evaluate their ability to promote the β-selective asymmetric addition of γ-butyrolactam 2a with cyclic imino ester 3a in the presence of 15 mol% of catalyst loading at room temperature in CH2Cl2 (entries 1–6,
EntryCat.SolventYieldeerf
11aCH2Cl270%40 : 60
21bCH2Cl2<5%57 : 43
31cCH2Cl270%65 : 35
41dCH2Cl268%70 : 30
51eCH2Cl258%63 : 47
61fCH2Cl271%77 : 23
71fDCE72%80 : 20
81fCHCl370%80 : 20
91fMTBE68%79 : 21
101fToluene63%78 : 22
111fTHF45%76 : 24
121fMeOH32%62 : 38
13b1fDCE : MTBE75%87 : 13
14c1fDCE : MTBE72%87 : 13
15d1fDCE : MTBE70%85 : 15
Open in a separate windowaReaction conditions: unless specified, a mixture of 2a (0.2 mmol), 3a (0.3 mmol) and a catalyst (15 mmol%) in a solvent (2.0 mL) was stirred at rt. for 48 h.bThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1).cThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) for 24 h.dThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) and 10 mol% of catalyst was used.eIsolated yields.fDetermined by chiral HPLC, the product was observed with >99 : 1 dr by 1H NMR and HPLC. Configuration was assigned by X-ray crystal data of 4a.The results of experiments under the optimized conditions that probed the scope of the reaction are summarized in Scheme 3. The catalytic β-selective asymmetric addition of γ-butyrolactam 2a with cyclic imino esters 3a in the presence of 15 mol% (DHQD)2AQN 1f was performed. A variety of phenyl-substituted cyclic imino esters including those bearing electron-withdrawing and electron-donating substituents on the aryl ring, heterocyclic were also examined. The electron-neutral, electron-rich, or electron-deficient groups on the para-position of phenyl ring of the cyclic imino esters afforded the products 4a–4m in 57–75% yields and 82 : 18 to 95 : 5 er values. It appears that either an electron-withdrawing or an electron-donating at the meta- or ortho-position of the aromatic ring had little influence on the yield and stereoselectivity. Similar results on the yield and enantioselectivities were obtained with 3,5-dimethoxyl substituted cyclic imino ester (71% yield and 91 : 9 er). It was notable that the system also demonstrated a good tolerance to naphthyl substituted imino ester (78% yield and 92 : 8 er value). The 2-thienyl substituted cyclic imino ester proceeded smoothly under standard conditions as well, which gave the desired product 4p in good enantioselectivity (88 : 12 er), although yield was slightly lower. However, attempts to extend this methodology to aliphatic-substituted product proved unsuccessful due to the low reactivity of the substrate 3q. It is worth noting that the replacement of Boc group with 9-fluorenylmethyl, tosyl or benzyl group as the protection, no reaction occurred. The absolute and relative configurations of the products were unambiguously determined by X-ray crystallography (4a, see the ESI).Open in a separate windowScheme 3Substrate scope of the asymmetric reaction of α,β-unsaturated γ-butyrolactam 2 to cyclic imino esters 3.a aReaction conditions: unless specified, a mixture of 2 (0.2 mmol), 3 (0.3 mmol) and 1f (15.0 mmol%) in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) was stirred at rt. bIsolated yields. cDetermined by chiral HPLC, all products were observed with >99 : 1 dr by 1H NMR and HPLC. Configuration was assigned by comparison of HPLC data and X-ray crystal data of 4a.We then examined the substrate scope of the imide derivatives (Scheme 4). Investigations with maleimides 4r–4u gave 48–61% yield of corresponding products as lower er and dr values than most of γ-butyrolactams. As for methyl substituted maleimides, the reaction failed to give any product.Open in a separate windowScheme 4Substrate scope of the asymmetric reaction of maleimides to cyclic imino esters.a aReaction conditions: unless specified, a mixture of 2 (0.2 mmol), 3 (0.3 mmol) and 1f (15.0 mmol%) in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) was stirred at rt. bIsolated yields. cDetermined by 1H NMR and chiral HPLC.The chloride product 4a ((R)-tert-butyl 4-((R)-3-((E)-(4-chlorobenzylidene)amino)-2-oxotetra hydrofuran-3-yl)-2-oxopyrrolidine-1-carboxylate) was recrystallized and the corresponding single crystal was subjected to X-ray analysis to determine the absolute structure. Based on this result and our previous work, a plausible catalytic mechanism involving multisite interactions was assumed to explain the high stereoselectivity of this process (Fig. 1). Similar to the conformation reported for the dihydroxylation and the asymmetric direct aldol reaction, the transition state structure of the substrate/catalyst complexes might be presumably in the open conformation. The acidic α-carbon atom of cyclic imino ester 3a could be activated by interaction between the tertiary amine moiety of the catalyst and the enol of 3avia a hydrogen bonding. Moreover, the enolate of 3a in the transition state might be in part stabilized through the π–π stacking between the phenyl ring of 3a and the quinoline moiety. Consequently, the Re-face of the enolate is blocked by the left half of the quinidine moiety. The steric hindrance between the Boc group of 2a and the right half of the quinidine moiety make the Re-face of 2a face to the enolate of 3a. Subsequently, the attack of the incoming nucleophiles forms the Si-face of enolate of 3a to Re-face of 2a takes place, which is consistent with the experimental results.Open in a separate windowFig. 1Proposed transition state for the reaction.In conclusion, we have disclosed the β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by a bifunctional chiral tertiary amine, which provides an efficient and facile access to optically active β-position functionalized pyrrolidin-2-one derivatives with high diastereoselectivity and enantioselectivity. To our knowledge, this is the first catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach. Current efforts are in progress to apply this new methodology to synthesize biologically active products.  相似文献   

14.
Three new metal coordination polymers of bifunctional imidazolate/tetrazolate bridges: the only example of a three-dimensional framework based on rare [Co4(μ3-OH)2(μ2-Cl)2]4+ mixed oxo-chloro-clusters     
Lili Yang  Jian Zhou  Hua-Hong Zou  Qiuling Tang 《RSC advances》2019,9(23):13082
Three new metal coordination polymers [Ni(μ2-L)2(H2O)2]n(1, HL = 1-tetrazole-4-imidazole-benzene), [Co(μ2-L)2]n (2), and [Co43-OH)22-Cl)25-L)22-L)2]n·7nH2O (3) were hydrothermally synthesized and structurally characterized. 1 displays a neutral [Ni(μ2-L)2(H2O)2]n chain built up from the Ni2+ ions bridged by deprotonated L ligands, while 2 shows another rare neutral [Co(μ2-L)2]n chain based on Co2+ ions connected by two different coordination modes of the L ligand. 3 exhibits a rare [Co43-OH)22-Cl)2]4+ mixed oxo-chloro-cluster-based three-dimensional framework with large elliptical channels, which are filled by unprecedented chilopod [(H2O)7]n chains. Both 1 and 2 show antiferromagnetic behavior, while 3 exhibits unusual spin-canting.

The interconnection of rare [Co43-OH)22-Cl)2]4+ mixed oxo-chloro-clusters and bifunctional bridges gives a new three-dimensional framework with unusual spin-canting behavior.  相似文献   

15.
Self-assembled membrane manufactured by metal–organic framework (UiO-66) coated γ-Al2O3 for cleaning oily seawater     
Cunlong Li  Yuqing Zhang  Ming Yong  Wei Liu  Jiaqi Wang 《RSC advances》2019,9(19):10702
In order to effectively clean oily seawater with anionic polyacrylamide (APAM), UiO-66 coated γ-Al2O3 (UA) composites were firstly synthesized using γ-Al2O3 as a template to induce the growth of high hydrophilic UiO-66 on its surface to form a uniform UA self-assembled membrane. The UA composites and self-assembled membrane were characterized and analyzed. Also, the membrane performance was investigated. The results show that the hydrophilicity of particles is enhanced with the water contact angle decreasing from 39.8° (γ-Al2O3 particles) to 26.2° (UA composites) by introducing the UiO-66 coating. Moreover, the UA self-assembled membrane performs attractive water yield and separation performance. The oil concentration in the permeate treated by the first class of UA self-assembled membrane declines apparently from 91.22 to 18.90 mg L−1, while the water yield is as high as 657.89 L m−2 h−1. The reclaimed separation experiments show that the membrane materials could be recycled by calcination at 200 °C and hydraulic cleaning, which gives the material potential application in cleaning oily seawater.

The self-assembled membrane manufactured by UA composites exhibits excellent separation performance and water yield in the treatment of oily seawater.  相似文献   

16.
The Common Cytokine Receptor γ Chain Controls Survival of γ/δ T Cells          下载免费PDF全文
Marie Malissen  Pablo Pereira  David J. Gerber  Bernard Malissen    James P. DiSanto 《The Journal of experimental medicine》1997,186(8):1277-1285
We have investigated the role of common γ chain (γc)-signaling pathways for the development of T cell receptor for antigen (TCR)-γ/δ T cells. TCR-γ/δ–bearing cells were absent from the adult thymus, spleen, and skin of γc-deficient (γc) mice, whereas small numbers of thymocytes expressing low levels of TCR-γ/δ were detected during fetal life. Recent reports have suggested that signaling via interleukin (IL)-7 plays a major role in facilitating TCR-γ/δ development through induction of V-J (variable-joining) rearrangements at the TCR-γ locus. In contrast, we detected clearly TCR-γ rearrangements in fetal thymi from γc mice (which fail to signal in response to IL-7) and reduced TCR-γ rearrangements in adult γc thymi. No gross defects in TCR-δ or TCR-β rearrangements were observed in γc mice of any age. Introduction of productively rearranged TCR Vγ1 or TCR Vγ1/Vδ6 transgenes onto mice bearing the γc mutation did not restore TCR-γ/δ development to normal levels suggesting that γc-dependent pathways provide additional signals to developing γ/δ T cells other than for the recombination process. Bcl-2 levels in transgenic thymocytes from γc mice were dramatically reduced compared to γc+ transgenic littermates. We favor the concept that γc-dependent receptors are required for the maintenance of TCR-γ/δ cells and contribute to the completion of TCR-γ rearrangements primarily by promoting survival of cells committed to the TCR-γ/δ lineage.  相似文献   

17.
Correction: Global diabatic potential energy surfaces for the BeH2+ system and dynamics studies on the Be+(2P) + H2(X1Σg+) → BeH+(X1Σ+) + H(2S) reaction     
Zijiang Yang  Jiuchuang Yuan  Shufen Wang  Maodu Chen 《RSC advances》2018,8(44):25181
  相似文献   

18.
Synthesis of 1-(β-coumarinyl)-1-(β-indolyl)trifluoroethanols through regioselective Friedel–Crafts alkylation of indoles with β-(trifluoroacetyl)coumarins catalyzed by Sc(OTf)3     
Lijun Shi  Ying Liu  Caixia Wang  Xinxin Yuan  Xiaobiao Liu  Lulu Wu  Zhenliang Pan  Qicheng Yu  Cuilian Xu  Guoyu Yang 《RSC advances》2020,10(24):13929
A highly efficient Friedel–Crafts alkylation of indole derivatives with β-(trifluoroacetyl)coumarins using Sc(OTf)3 as a catalyst has been developed, which gives regioselective 1,2-adducts to afford 1-(β-coumarinyl)-1-(β-indolyl)trifluoroethanols. A series of tertiary trifluoroethanols containing different indole and coumarin groups were synthesized in moderate to excellent yields (up to 95%) in the presence of 5 mol% catalyst in a short time (only 2 minutes at least). A mechanism of the reaction, in which the trace amount of water plays the role of proton transfer in catalyzing circulation was proposed and confirmed.

A Friedel–Crafts alkylation of indoles with β-(trifluoroacetyl)coumarins catalyzed by Sc(OTf)3 to afford 1-(β-coumarinyl)-1-(β-indolyl)trifluoroethanols in a short time and high yield was developed.  相似文献   

19.
A model study on the photodecarbonyl reaction of (η5-C5H5)M(CO)2 (M = Co,Rh, Ir)     
Zheng-Feng Zhang  Ming-Der Su 《RSC advances》2018,8(43):24641
The group 9 organometallic complexes η5-CpM(CO)2 (M = Co, Rh, and Ir) and Si(CH3)3(H) have been considered as a model system to study their photochemical decarbonyl reactions as well as the Si–H bond activation reactions using the CASSCF and MP2-CAS computational methods. For the cobalt complex, three kinds of reaction pathways, which result in the same oxidative addition product, are investigated. Our theoretical finding demonstrated that after the photoirradiation, η5-CpCo(CO)2 loses one CO ligand without any difficulty to form either the triplet ([η5-CpCo(CO)]3) or singlet ([η5-CpCo(CO)]1) species. The former plays a decisive role in the formation of the final oxidative addition product. On the other hand, the latter plays no role in the production of the final product molecule, but its singlet cobalt center interacts weakly with solvent molecules ((Me3)SiH) to produce an alkyl-solvated organometallic complex, which is experimentally detectable. The present works reveal that both η5-CpRh(CO)2 and η5-CpIr(CO)2 should adopt the conical intersection mechanism after they are irradiated by light. Moreover, our theoretical examinations strongly suggest that for the 16-electron monocarbonyl η5-CpM(CO) (M = Rh and Ir) species, the insertion into a Si–H bond by the Ir system is much more facile and more exothermic than that for the Rh counterpart.

Theoretical observations show that the triplet cobalt photoproduct [η5-CpCo(CO)] activates Si–H bonds much more readily than its singlet counterpart. The conical intersection mechanism may play a decisive role in photodecarbonyl reactions of (η5-C5H5)M(CO)2 (M = Rh and Ir).  相似文献   

20.
Monoallylation and benzylation of dicarbonyl compounds with alcohols catalysed by a cationic cobalt(iii) compound     
Mohan Chandra Sau  Smita Mandal  Manish Bhattacharjee 《RSC advances》2021,11(16):9235
Monoallylation and monoalkylation of diketones and β-keto esters with allylic and benzylic alcohols catalysed by [Cp*Co(CH3CN)3][SbF6]2 (I) are reported. The method does not require any additive and affords regioselective products. The mechanistic investigations were done by in situ1H NMR spectroscopy as well as control experiments. It has been shown that reactions proceed via η3-allyl complex formation or ally ether intermediate. The alkylation takes place via only ether intermediate. The resulting allylated and alkylated products have been used for the synthesis of eleven new trisubstituted pyrazoles and one pyrazolone.

Monoallylation and monoalkylation of diketones and β-keto esters with allylic and benzylic alcohols catalysed by [Cp*Co(CH3CN)3][SbF6]2 (I) are reported.  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