共查询到20条相似文献,搜索用时 15 毫秒
1.
A variety of palladium(II)–acetylacetonate complexes bearing α-diimine ligands were synthesized by the reaction of [Pd(acac)(MeCN)2]BF4 with N⁀N ligands. When activated with BF3·OEt2, these complexes provide access to a new class of alkylating agent free Pd–diimine catalyst systems. Catalyst screening for the vinyl addition polymerization of norbornene showed their high activities of 102–104 kgpol molcat− 1 h− 1. 相似文献
2.
3.
Nonproteinogenic α- or β-amino acids have attracted tremendous attention, as they are widely utilized for biological, biochemical, pharmaceutical, and asymmetric chemical investigations. Recently, we developed a series of new strategies for preparing achiral and chiral nickel(ii) complexes for the synthesis of amino acids. We applied these new methods utilizing chiral nickel(ii) complexes for the asymmetric Mannich reaction to synthesize enantiopure α,β-diamino acids, the enantioselective tandem conjugate addition-elimination reaction to prepare glutamic acid derivatives, the Suzuki coupling reaction to yield β(2)-amino acid derivatives, the asymmetric Mannich reaction to synthesize 3-aminoaspartate, the asymmetric Michael addition reaction to give β-substituted-α,γ-diaminobutyric acid derivatives, the asymmetric alkylation reaction to prepare linear ω-trifluoromethyl containing amino acids, and the asymmetric Michael addition reaction to synthesize syn-β-substituted tryptophans. 相似文献
4.
《Inorganic chemistry communications》1999,2(10):472-475
Employing a variety of solvents and molar ratios, the reactions of Mn(NO3)2·6H2O with N,N′-dimethylurea (DMU) afforded the adduct [Mn(NO3)2(DMU)3] (1). X-ray analysis shows that the high-spin complex has a pentagonal bipyramidal geometry with two DMU oxygen atoms in axial positions and with two types of monodentate O-bonding for the DMU molecules, one of them being very unusual. The spectroscopic properties of the prepared complex are also discussed. 相似文献
5.
Enantiopure N,N-diamine has been obtained in an economical manner via demetalation of its corresponding dichloro Zn(II) complex, which is separated by fractional crystallization of two diastereomeric Zn(II) complexes. Polymerization of rac-lactide (rac-LA) initiated by diisopropoxide derivative proceeds rapidly at room temperature in a living fashion to yield heterotactic polylactide (PLA) with Pr = 0.90. 相似文献
6.
Phosphoribosylanthranilate (PRA) isomerase (TrpF) and tryptophan synthase α-subunit (TrpA) are (βα)(8)-barrel enzymes that are involved in the biosynthesis of tryptophan. They contain a conserved phosphate binding site, which indicates a common evolutionary origin. In order to experimentally back this hypothesis, we have established TrpF activity on the scaffold of TrpA from Salmonella typhimurium using protein engineering. Based on the superposition of crystal structures with bound ligands, two residues in the active site of TrpA were replaced with catalytic residues from TrpF using site-directed mutagenesis. This TrpA variant as well as wild-type TrpA were each subjected to random mutagenesis using error-prone PCR. The two resulting trpA gene libraries were used to transform an auxotrophic Escherichia coli trpF deletion strain, and TrpA variants with PRA isomerisation activity were isolated by in vivo complementation. The amino acid substitutions of the selected TrpA variants were recombined by DNA shuffling, again followed by complementation in vivo. Several TrpA variants were produced in E. coli and purified, and their catalytic TrpF activities were determined in vitro by steady-state enzyme kinetics. Our results support that TrpA and TrpF have evolved by gene duplication and diversification from each other or a common predecessor, and provide insights into the minimum requirements for the catalysis of PRA isomerisation. 相似文献
7.
《Inorganic chemistry communications》1999,2(7):288-291
Ruthenium(II) complexes containing both σ-alkenyl and η2-carboxylate ligands Ru(O2CR2)(CHCHR1)(CO)(PPh3)2 (R1tBu, Ph; R2=CHCHCHCHCH3) react with phenylacetylene in two stages. Their reaction with an equivalent amount of the alkyne led to the formation of a σ-alkynyl-containing ruthenium(II) complex Ru(O2CR2)(CCPh)(CO)(PPh3)2 (R2=CHCHCHCHCH3), the molecular structure of which was established by X-ray diffraction. This σ-alkynyl complex then reacts with phenylacetylene to form a σ-butenynyl-containing compound Ru(O2CR2)(C(CCPh)CPhH)(CO)(PPh3)2 (R2=CHCHCHCHCH3). Both reactions support the key role of alkynyl ligands in the dimerization of alkynes. 相似文献
8.
Synthesis, characterization and anti-Mycobacterium tuberculosis assays of new platinum(II)/dppf/N,N-disubstituted-N′-acyl thiourea complexes with general formulae [Pt(dppf)(L)]PF6, [dppf = 1,1′-bis(diphenylphosphino)ferrocene; L = N,N-disubstituted-N′-acyl thioureas] is reported. The complexes were characterized by elemental analysis, molar conductivity, IR, NMR (1H, 13C and 31P{1H}) spectroscopy. The spectroscopic data are consistent with the complexes containing one dppf and one O, S chelated ligand. The crystal structures of complexes with N,N-diphenyl-N′-benzoylthiourea (L4), N,N-diethyl-N′-furoylthiourea (L5) and N,N-diphenyl-N′-(thiophene-2-carbonyl)thiourea (L8) were determined by X-ray crystallography, confirming the coordination of the ligands with the metal through sulfur and oxygen atoms, forming distorted square-planar structures. The complexes were screened with respect to their anti-M. tuberculosis activity (H37Rv ATCC 27294). 相似文献
9.
Tuneo Suzuki Hirofusa Shirai Nobumasa Hojo 《Journal of Inorganic and Organometallic Polymers and Materials》1994,4(3):251-260
The kinetics of the metal exchange reaction between Cu(II)-poly(vinyl alcohol) [Cu(II)-PVA] and Zn(II)-ethylenediamine-N,N,N′,N′-tetraacetic acid [Zn(II)-EDTA] has been studied by mixing both solutions in a spectrophotometer at pH 10.0 to 11.0, ionic strength μ=0.10(KNO3), and 15 to 35°C. The reaction is initiated by the formation of unstable Cu(II)-H-PVA through attack of H+ ion on the Cu(II)-PVA complex, and both reactions, ligand exchange and metal exchange, proceed simultaneously. The metal exchange step may be rate determining. The rate equation and rate constants of this reaction were determined as follows: ?d[Cu(II)-PVA]/dt=k 0(H)[PVA?][Cu(II)-PVA] [Zn(II)-EDTA], wherek 0(H)=k 1+(k′2+k′3)[H+],k 1=5.98±1.64M ?1 s?1, andk 2+k 3=k′2 K Cu(II)-H-PVA ?H +k′3 K Zn(II)-EDTA H =(5.91±0.89)×107 M ?2 s?1. 相似文献
10.
Complex [Cu2{(py)2CNO}2{(py)2CNOH}2](ClO4)2 (2) has been isolated and proven to be an intermediate of the reaction that leads to the known, neutral dimer [Cu2{(py)2CNO}4]·2H2O (1), where (py)2CNOH = di-2-pyridyl ketoxime; the different topologies of the double oximato bridge in the two complexes lead to different magnetic responses. 相似文献
11.
Abstract N,N,N′,N′-tetra(2-ethylhexyl)diglycolamide (TEHDGA) was found to be a promising extractant for actinide partitioning from high-level waste (HLW) (Part I). In order to evaluate the applicability of TEHDGA to the HLW partitioning process, investigations on its radiolytic stability were carried out. The present work deals with the studies on the uptake of americium by γ-irradiated 0.2 M TEHDGA/n-dodecane in the absence and presence of phase modifiers—di(n-hexyl)octanamide (DHOA), isodecanol and n-decanol—against the absorbed dose up to 1 × 106 Gy in the presence of nitric acid at varying concentrations. The addition of phase modifiers suppressed the radiolysis of TEHDGA in n-dodecane and DHOA was found to be the most effective radiolysis suppressor. Investigations were also carried out on the degradation of neat TEHDGA by γ-irradiation, and it was attempted to isolate and identify its degradation products by instrumental analysis. The radiolysis study showed that the degradation products were formed by the cleavage of the –C-N bond, to eliminate an ethylhexyl group, and the bond adjacent to the ether bond. The results obtained for TEHDGA radiolysis were compared with that of its straight-chain isomer TODGA, and TEHDGA was observed to be more resistant to radiation than TODGA. The changes in the physico-chemical properties of γ-irradiated TEHDGA against the absorbed dose were also investigated. 相似文献
12.
Ru(II)–CO complexes of thioarylazoimidazoles: Syntheses,structures, spectroscopy and DFT calculation
T.K. Mondal P. Raghavaiah A.K. Patra C. Sinha 《Inorganic chemistry communications》2010,13(2):273-277
The complexes, cis-(CO)-trans-(Cl)-[Ru(SRaaiNR)(CO)2Cl2] (2) and trans-(Cl)-[Ru(SRaaiNR)(CO)Cl2] (3) (SRaaiNR = 1-alkyl-2-{(o-thioalkyl)phenylazo}imidazoles; R = Me (1a) and Et (1b)) have been synthesized and characterized. The structural confirmation is achieved by single crystal X-ray structure determinations. The complexes show Ru(III)/Ru(II) couple and ligand reductions. Electronic structure and spectral properties of the complexes have been explained with the DFT and TDDFT calculation. 相似文献
13.
The coordination chemistry of the ligand precursor 1-benzoyl-4,5-dihydro-3,5-bis(trifluoromethyl)-1H-pyrazol-5-ol (1a) to iron(II) acetate was studied. In dependence of the added co-ligand different complex geometries were observed. In case of 4-dimethylaminopyridine (DMAP) as co-ligand an octahedral iron(II) complex was found with an O,N,O′-coordination of the tridentate ligand (1a-2H), in which the ligand is planar, the oxygen donors are trans to each other, and the nitrogen donor is in a cis position. The other coordination sites on the iron center are occupied by DMAP ligands. In contrast to that, with triphenylphosphane as the co-ligand an oxidation process took place, which revealed an octahedral iron(III) complex with a comparable geometry for the tridentate ligand (O,N,O′-coordination, 1a-2H) demonstrating the usefulness of 1a-2H to stabilize different oxidation states. The additional coordination sites are occupied by one triphenylphosphane oxide and ethoxide ligands. Interestingly, the ethoxide ligands act as bridging ligands to form a bimetallic complex. 相似文献
14.
Tuneo Suzuki Hirofusa Shirai Nobumasa Hojo 《Journal of Inorganic and Organometallic Polymers》1994,4(3):251-260
The kinetics of the metal exchange reaction between Cu(II)-poly(vinyl alcohol) [Cu(II)-PVA] and Zn(II)-ethylenediamine-N,N,N,N-tetraacetic acid [Zn(II)-EDTA] has been studied by mixing both solutions in a spectrophotometer at pH 10.0 to 11.0, ionic strength =0.10(KNO3), and 15 to 35°C. The reaction is initiated by the formation of unstable Cu(II)-H-PVA through attack of H+ ion on the Cu(II)-PVA complex, and both reactions, ligand exchange and metal exchange, proceed simultaneously. The metal exchange step may be rate determining. The rate equation and rate constants of this reaction were determined as follows: –d[Cu(II)-PVA]/dt=k
0(H)[PVA–][Cu(II)-PVA] [Zn(II)-EDTA], wherek
0(H)=k
1+(k2+k3)[H+],k
1=5.98±1.64M
–1 s–1, andk
2+k
3=k2
K
Cu(II)-H-PVA
–H
+k3
K
Zn(II)-EDTA
H
=(5.91±0.89)×107
M
–2 s–1. 相似文献
15.
Xiaojuan Huang Zaiwen Yang Xiao-Juan Yang Qilong Zhao Yana Xia Biao Wu 《Inorganic chemistry communications》2010,13(9):1103-1107
Two zinc(II) complexes of a urea-functionalized pyridyl ligand, [Zn(SO4)L3]·CH3OH (1) and {[Zn(μ2-SO4)L2]·0.5CH3OH}n (2) (L = N-(1-naphthyl)-N′-(3-pyridyl)urea), have been synthesized by the reaction of L with ZnSO4·7H2O under different conditions. Complex 1 is a hydrogen-bonded 2D grid structure in which the sulfate anion not only coordinates to the ZnII ion as a monodentate ligand but also forms multiple hydrogen bonds with the urea groups of adjacent molecules. In complex 2 the sulfate ion serves as a μ2-bridging ligand, connecting the ZnII ions to form an infinite 1D organic–inorganic hybrid framework. Meanwhile, the sulfate anion is also bound by neighboring ligands through hydrogen bonds between urea groups and the non-coordinated oxygen atoms. The solid-state emission spectra of 1 and 2 show a red shift of the fluorescence emission of ligand L. 相似文献
16.
Suresh Chavda Kulbir Singh D. Gerrard Marangoni Vinod K. Aswal Pratap Bahadur 《Journal of surfactants and detergents》2012,15(3):317-325
The solution behavior of a typical cationic surfactant, tetradecyltrimethylammonium bromide, in mixed solvent systems composed
of water and varying concentrations of α,ω-alkanediols; 1,2-ethanediol (ED), 1,4-butanediol (BD), 1,6-hexanediol (HD) and
1,8-octanediol (OD) was examined via electrical conductance measurements, 13C-NMR spectroscopy and small angle neutron scattering (SANS) measurements. The critical micelle concentration (CMC) values
and degree of counterion dissociation (α) indicate that both ED and BD oppose micellization, whereas HD and OD enhance micelle
formation. Changes in the 13C-NMR chemical shifts (∆δ values) reveal that the short chain diols reside almost exclusively in the bulk phase and hence,
affect the formation of micelles by altering the solvent properties in the bulk of the solution, whereas HD and OD partition
between the pseudomicellar phase and the bulk phase. SANS studies indicated that both the micellar size and aggregation number
(N
agg) decrease in the presence of all diols. ED and BD behave like cosolvents and increase the α and CMC values and decrease N
agg. We note that the effect of HD and OD on the properties of the micelles is concentration dependent; at low concentrations,
these diols interact with the micelles and behave as cosurfactants (as evidenced by the trends in the micellar properties),
while at higher concentrations, they enhance the surfactant solubility and behave as a cosolvent. 相似文献
17.
Jens Kiesewetter 《Polymer》2006,47(10):3302-3314
Four cationic palladium(II) α-diimine complexes, [{ArNC(R)-C(R)NAr}Pd(Me)(CH3CN)] (1, R=H, Ar=2,6-Me2C6H3, =B[3,5-C6H3(CF3)2]4; 2, R=CH3, Ar=2,6-iPr2C6H3; 3, R=CH3, Ar=2-tBuC6H4; 4, R,R=An, Ar=2,6-iPr2C6H3) were used for the copolymerization of ethene with norbornene. The copolymerization behavior of the catalysts and the influence of the polymerization temperature were investigated. The copolymers were characterized using 13C NMR spectroscopy, differential scanning calorimetry, and gel permeation chromatography techniques. Sterically demanding ortho -N-aryl substituents and rigid bulky bridge units increase the copolymer molar masses, while the incorporation level of norbornene is decreased. Microstructures with isolated norbornene units and alternating sequences are predominant. Less bulky substituted catalysts yield copolymers with higher norbornene contents and lower molar masses. Norbornene diblock sequences are dominant which are exclusively racemic connected, indicating that the insertion proceeds under chain end control. Optimal polymerization results are achieved at temperatures between 10 and 30 °C, while temperatures below 0 °C result in lower polymerization rates and molar masses. Above 30 °C, activities, molar masses, and norbornene incorporation decreases due to catalyst decomposition. 相似文献
18.
Hyeong-Tak Jeon Kyung-Eun Lee Yong-Joo Kim Soon W. Lee 《Inorganic chemistry communications》2009,12(12):1234-1237
Pd(II)– and Pt(II)–azido complexes, [M(N3)(PMe3)2(C–L)] {LH = 2-(2′)-thienyl pyridine; M = Pd (1), Pt(2)}, which contain σ-bonded heterocycles (L), were treated with aryl isothiocyanate (Me2C6H3–NCS) to afford the corresponding Pd(II) and Pt(II) tetrazole–thiolato complexes, trans-{M[SCN4(2,6-Me2C6H3)](PMe3)2(C–L)} {M = Pd (3), Pt (4)}. Complexes 3 and 4 have a 1-D helical network formed by the intermolecular M?S van der Waals contacts. 相似文献
19.
Highly efficient carbocationic polymerization of styrene was achieved under environmentally advantageous conditions in benzotrifluoride, BTF (α,α,α-trifluorotoluene, TFT), an environmentally benign solvent at room temperature, that is without any energy consumption for cooling or heating, with even better yields than that in the usually applied volatile and harmful, widely used chlorinated solvent, dichloromethane (DCM). The polymerization was initiated by 1-phenylethyl chloride in conjunction with the TiCl4/TMEDA (N,N,N′,N′-tetramethylethylenediamine) catalyst (coinitiator) system. Within a very short reaction time (5 min), higher conversion values were obtained in BTF (89%) than in DCM (76%), that is the TiCl4/TMEDA combination proved to be a powerful catalyst for the carbocationic polymerization of styrene even in BTF. The molecular weight distributions of the synthesized polymers were relatively narrow in both solvents (Mw/Mn = 1.29–1.65). The effect of the increasing reaction temperature (up to room temperature) was also investigated. With increasing reaction temperature, the polydispersity decreased and Mn close to the theoretical one was obtained in BTF at room temperature. Structural analysis with 1H NMR revealed that the major chain breaking reaction in this polymerization is indanyl ring formation between the penultimate monomer unit and the propagating carbocation. These results indicate that BTF can be utilized as a unique, inert, non-volatile, environment friendly solvent with medium polarity for cationic polymerization of styrene, a nonfluorous monomer, and based on these results, presumably it may be also applied effectively as a quite universal solvent for a large array of various polymerizations and copolymerizations for not only fluorinated, but also for nonflourous monomers, and other chemical reactions as well. 相似文献
20.
Effect of Alkanediyl-α,ω-Type Cationic Dimeric (Gemini) Surfactants on the Reaction Rate of Ninhydrin with [Cu(II)-Gly-Tyr]+ Complex 下载免费PDF全文
Dileep Kumar Kian-Eang Neo Malik Abdul Rub 《Journal of surfactants and detergents》2016,19(1):101-109
The effect of gemini (16‐s‐16, s = 4, 5, 6) surfactants on the reaction rate of ninhydrin with [Cu(II)‐Gly‐Tyr]+ complex was determined using a spectrophotometric technique. The ninhydrin concentration was kept in excess in order to maintain pseudo‐first‐order conditions. The reaction followed irreversible first‐ and fractional‐order kinetics with respect to [Cu(II)‐Gly‐Tyr]+ and [ninhydrin], respectively. It is found that gemini surfactants effectively catalyze the reaction. The rate constants (kψ) first increase and then become relatively constant with increasing gemini surfactant concentration similar to conventional cetyltrimethylammonium bromide. At higher gemini surfactant concentration a third region of increasing kψ is observed. The unusual third region is ascribed to changes in micellar morphology. The kinetic data has been analyzed using a micellar pseudo‐phase model. 相似文献