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.
《Electroanalysis》2004,16(11):949-954
The preparation and the electrochemical study of Disperse Blue 1‐chemically modified electrodes (DB1‐CME), as well as their efficiency for the electrocatalytic oxidation of NADH is described. The proposed mediator was immobilized by physical adsorption onto graphite electrodes. The electrochemical behavior of DB1‐CME was studied with cyclic voltammetry. The electrochemical redox reaction of DB1 was found to be reversible, revealing two well‐shaped pair of peaks with formal potentials 152 and ?42 mV, respectively, (vs. Ag/AgCl/3M KCl) at pH 6.5. The current Ip has a linear relationship with the scan rate up to 800 mV s?1, which is indicative for a fast electron transfer kinetics. The dissociation constants of the immobilized DB1 redox couple were calculated pK1=4 and pK2=5. The electrochemical rate constants of the immobilized DB1 were calculated k1°=18 s?1 and k2°=23 s?1 (Γ=2.36 nmol cm?2). The modified electrodes were mounted in a flow injection manifold, poised at +150 mV (vs. Ag/AgCl/3M KCl) and a catalytic current due to the oxidation of NADH was measured. The reproducibility was 1.4% RSD (n=11 for 30 μM NADH) The behavior of the sensor towards different reducing compounds was investigated. The sensor exhibited good operational and storage stability.  相似文献   

2.
A rapid, accurate and sensitive thin‐layer chromatography (TLC) method with densitometric detection has been developed and validated for the determination of cefepime in pharmaceuticals. Chromatographic separation was achieved on a silica gel TLC F254 plates with a mobile phase consisting of ethanol–2‐propanol–glacial acetic acid 99.5%–water (4:4:1:3, v/v). Densitometric detection was carried out at wavelength of 266 nm in reflectance/absorbance mode. The validation of the method was found to be satisfactory with high accuracy (from 99.24 to 101.37%) and precision (RSD from 0.06 to 0.36%). Additionally, the stability of cefepime in solution was investigated, including the effect of pH, temperature and incubation time. Favorable retention parameters (Rf, Rs, α) were obtained under the developed conditions, which guaranteed good separation of the studied components. The degradation process of cefepime hydrochloride was described by kinetic and thermodynamic parameters (k, t0.1, t0.5 and Ea). Moreover, the chemical properties of degradation products were characterized by the Rf values, absorption spectra, HPLC‐MS/MS and TLC‐densitometry analysis. As the method could effectively separate the active substance from its main degradation product (1‐methylpyrrolidine), it can be employed as a method to indicate the stability of this drug. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

3.
Photophysical properties in dilute MeCN solution are reported for seven RuII complexes containing two 2,2′‐bipyridine (bpy) ligands and different third ligands, six of which contain a variety of 4,4′‐carboxamide‐disubstituted 2,2′‐bipyridines, for one complex containing no 2,2′‐bipyridine, but 2 of these different ligands, for three multinuclear RuII complexes containing 2 or 4 [Ru(bpy)2] moieties and also coordinated via 4,4′‐carboxamide‐disubstituted 2,2′‐bipyridine ligands, and for the complex [(Ru(bpy)2(L)]2+ where L is N,N′‐([2,2′‐bipyridine]‐4,4′‐diyl)bis[3‐methoxypropanamide]. Absorption maxima are red‐shifted with respect to [Ru(bpy)3]2+, as are phosphorescence maxima which vary from 622 to 656 nm. The lifetimes of the lowest excited triplet metal‐to‐ligand charge transfer states 3MLCT in de‐aerated MeCN are equal to or longer than for [Ru(bpy)3]2+ and vary considerably, i.e., from 0.86 to 1.71 μs. Rate constants kq for quenching by O2 of the 3MLCT states were measured and found to be well below diffusion‐controlled, ranging from 1.2 to 2.0⋅109 dm3 mol−1 s−1. The efficiencies f of singlet‐oxygen formation during oxygen quenching of these 3MLCT states are relatively high, namely 0.53 – 0.89. The product of kq and f gives the net rate constant k for quenching due to energy transfer to produce singlet oxygen, and kqk equals k, the net rate constant for quenching due to energy dissipation of the excited 3MLCT states without energy transfer. The quenching rate constants were both found to correlate with ΔGCT, the free‐energy change for charge transfer from the excited Ru complex to oxygen, and the relative and absolute values of these rate constants are discussed.  相似文献   

4.
A new assay based on ultrafiltration, liquid chromatography and mass spectrometry was developed for the rapid screening and identification of the ligands for α‐glucosidase from the extract of Panax japonicus. Six saponins were identified as α‐glucosidase inhibitors. Subsequently, the specific binding ligands, namely, notoginsenoside R1, ginsenoside Rb1, chikusetsusaponin V, chikusetsusaponin IV, chikusetsusaponin IVa, and ginsenoside Rd (the purities were 94.18, 95.43, 96.09, 93.26, 94.50, 93.86%, respectively) were separated by counter‐current chromatography using two‐phase solvent systems composed of tert‐butyl methyl ether, acetonitrile, 0.1% aqueous formic acid (3.8:1.0:4.4, v/v/v) and the solvent system composed of methylene chloride, isopropanol, methanol, 0.1% aqueous formic acid (5.8:1.0:6.0:2.2, v/v/v). The results demonstrate that ultrafiltration, liquid chromatography and mass spectrometry combined with high‐speed counter‐current chromatography might provide not only a powerful tool for screening and isolating α‐glucosidase inhibitors in complex samples but also a useful platform for discovering bioactive compounds for the prevention and treatment of diabetes mellitus.  相似文献   

5.
Guest‐binding affinities of water‐soluble cyclophane heptadecamer (1) and pentamer (2) with immobilized guests such as 1‐pyrenylmethylamine (PMA) and 2‐(1‐ naphthyl)ethylamine (NEA) were investigated by surface plasmon resonance (SPR) measurements. As a typical example, the binding constants (K) for 1 and 2 with the immobilized PMA as a guest were evaluated to be 2.5 × 107 and 2.7 × 106 M?1, respectively, and were much larger than that of a monocyclic reference cyclophane (K, 2.5 × 104 M?1). Interestingly, in the complexation of 1 and 2 with the immobilized guests, more favorable association and dissociation rate constant values (ka and kd, respectively) were observed in comparison with those for the monocyclic cyclophane, reflecting multivalent effects in macrocycles. The multivalent effects in macrocycles as well as molecular recognition abilities of the cyclophane oligomers were confirmed even when the guest molecules were immobilized on SPR sensor chip surfaces. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
In this work, a sensitive and efficient method was established and validated for qualitative and quantitative analysis of major bioactive constituents in Dazhu Hongjingtian capsule by liquid chromatography tandem mass spectrometry. A total of 32 compounds were tentatively identified using ultra‐performance liquid chromatography coupled with quadrupole time‐of‐flight mass spectrometry. Furthermore, 12 constituents, namely gallic acid, 3,4‐dihydroxybenzoic acid, salidroside, p‐ coumaric acid‐4‐O β ‐d ‐glucopyranoside, bergeninum, 4‐hydroxybenzoic acid, 4‐hydroxyphenylacetic acid, syringate, 6′′‐O ‐galloylsalidroside, rhodiosin, rhodionin and kaempferol‐7‐O α ‐l ‐rhamnoside, were simultaneously quantified by the developed ultra‐performance liquid chromatography coupled with a triple quadrupole mass spectrometry method in 9 min. All of them were analyzed on an Agilent ZorBax SB‐C18 column (3.0 × 100 mm, 1.8 μm) with linear gradient elution of methanol–0.1% formic acid water. The proposed method was applied to analyze three batches of samples with acceptable linearity (R , 0.9979–0.9997), precision (RSD, 1.3–4.7%), repeatability (RSD, 1.7–4.9%), stability (RSD, 2.2–4.9%) and recovery (RSD, 0.6–4.4%) of the 12 compounds. As a result, the analytical method possessing high throughput and sensitivity is suitable for the quality control of Dazhu Hongjingtian capsule.  相似文献   

7.
A sensitive and high‐throughput inhibition screening liquid chromatography–mass spectrometry (LC‐MS/MS) method was developed and validated for the simultaneous quantification of five probe metabolites (7‐hydroxycoumarin, CYP2A6; 4‐hydroxytolbutamide, CYP2C9; 4′‐hydroxymephenytoin, CYP2C19; α‐hydroxymetoprolol, CYP2D6; and 1‐hydroxymidazolam, CYP3A4) for in vitro cytochrome P450 activity determination in human liver microsome and recombinant. All the metabolites and the internal standard, tramadol, were separated on a Waters 2695 series liquid chromatograph with a Phenomenex Luna C18 column (150 × 2.0 mm, 5 µm). Quality control samples and a positive control CYP inhibitor were included in the method. The IC50 values determined for typical CYP inhibitors were reproducible and in agreement with the literature. The method was selective and showed good accuracy (99.13–103.37%), and inter‐day (RSD < 6.20%) and intra‐day (RSD < 6.13%) precision. Also, the incubation extracts of the sample were stable at room temperature (20 °C) for 48 h and for 96 h in the autosampler (4 °C). The presented method is the first HPLC‐MS/MS method of this combination for simultaneous detection of the five metabolites 7‐hydroxycoumarin, 4‐hydroxytolbutamide, 4′‐hydroxymephenytoin, α‐hydroxymetoprolol and 1‐hydroxymidazolam in a single‐run process. It is possible that the high‐quality and ‐throughput cocktail provides suitable information in drug discovery and screening for new drug entities. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

8.
Films of linear and branched oligomer wires of Fe(tpy)2 (tpy=2,2′:6′,2′′‐terpyridine) were constructed on a gold‐electrode surface by the interfacial stepwise coordination method, in which a surface‐anchoring ligand, (tpy? C6H4N?NC6H4? S)2 ( 1 ), two bridging ligands, 1,4‐(tpy)2C6H4 ( 3 ) and 1,3,5‐(C?C? tpy)3C6H3 ( 4 ), and metal ions were used. The quantitative complexation of the ligands and FeII ions was monitored by electrochemical measurements in up to eight complexation cycles for linear oligomers of 3 and in up to four cycles for branched oligomers of 4 . STM observation of branched oligomers at low surface coverage showed an even distribution of nanodots of uniform size and shape, which suggests the quantitative formation of dendritic structures. The electron‐transport mechanism and kinetics for the redox reaction of the films of linear and branched oligomer wires were analyzed by potential‐step chronoamperometry (PSCA). The unique current‐versus‐time behavior observed under all conditions indicates that electron conduction occurs not by diffusional motion but by successive electron hopping between neighboring redox sites within a molecular wire. Redox conduction in a single molecular wire in a redox‐polymer film has not been reported previously. The analysis provided the rate constant for electron transfer between the electrode and the nearest redox‐complex moiety, k1 (s?1), as well as that for intrawire electron transfer between neighboring redox‐complex moieties, k2 (cm2 mol?1 s?1). The strong effect of the electrolyte concentration on both k1 and k2 indicates that the counterion motion limits the electron‐hopping rate at lower electrolyte concentrations. Analysis of the dependence of k1 and k2 on the potential gave intrinsic kinetic parameters without overpotential effects: k10=110 s?1, k20=2.6×1012 cm2 mol?1 s?1 for [n Fe 3 ], and k10=100 s?1, k20=4.1×1011 cm2 mol?1 s?1 for [n Fe 4 ] (n=number of complexation cycles).  相似文献   

9.
Aromatic sulfonyl groups have attracted increasing interest due to their unique electronic features. In this article, a series of IrIII complexes bearing fluorinated phenylsulfonyl groups were evaluated by density functional theory and time‐dependent density functional theory methods. To explore their phosphorescence efficiencies, factors that determine the radiative decay rate constant, kr, and the nonradiative decay rate constant, knr, were computed. As demonstrated by the results, complex 4 , which has fluorinated phenylsulfonyl groups at the 5‐positions of the phenyl rings for all three C^N ligands, was found to have the highest phosphorescence efficiencies with the largest kr and smallest knr values among these complexes. Moreover, it was found to exhibit significantly blueshifted behavior relative to complex 1 and emits in the blue region, and thus, it can serve as a highly efficient blue emitter for application in organic light‐emitting diodes. These findings successfully illustrated the structure–properties relationship and provided valuable information for the development of future highly efficient blue‐emitting phosphors.  相似文献   

10.
The mechanisms for the hydrolysis of organopalladium complexes [Pd(CNN)R]BF4 (R=P(OPh)3, PPh3, and SC4H8) were investigated at 25 °C by using UV/Vis absorbance measurements in 10 % v/v ethanol/water mixtures containing different sulphuric acid concentrations in the 1.3–11.7 M range. In all cases, a biphasic behavior was observed with rate constants k1obs, which corresponds to the initial step of the hydrolysis reaction, and k2obs, where k1obs>k2obs. The plots of k1obs and k2obs versus sulfuric acid concentration suggest a change in the reaction mechanism. The change with respect to the k1obs value corresponds to 35 %, 2 %, and 99 % of the protonated complexes for R=PPh3, P(OPh)3, and SC4H8, respectively. Regarding k2obs, the change occurred in all cases at about 6.5 M H2SO4 and matched up with the results reported for the hydrolysis of the 2‐acetylpyridinephenylhydrazone (CNN) ligand. By using the excess acidity method, the mechanisms were elucidated by carefully looking at the variation of ki,obs (i=1,2) versus ${c_{{\rm{H}}^ + } }$ . The rate‐determining constants, k0,A‐1, k0,A‐2, and k0,A‐SE2 were evaluated in all cases. The R=P(OPh)3 complex was most reactive due to its π‐acid character, which favors the rupture of the trans nitrogen–palladium bond in the A‐2 mechanism and also that of the pyridine nitrogen–palladium bond in the A‐1 mechanism. The organometallic bond exerts no effect on the relative basicity of the complexes, which are strongly reliant on the substituent.  相似文献   

11.
The potential of diorganotin compounds, in particular alkoxides and phenoxides, to function as neutral and non‐corrosive catalysts in the methanolysis of tripalmitin (the main triglyceride in palm oil) to methyl palmitate has been investigated. The compounds reveal a strong dependence of catalytic activity on the nature of the organic moiety on tin, the ring‐borne substituent on the phenoxyl group and the chain length of the alkoxyl fragment, as well as the ring size in cyclic alkoxides derived from bifunctional ligands such as diethanolamine. Kinetic studies, based on detailed compositional analysis of the reaction mixture by gas chromatography, were performed typically at 70.0 ± 0.1 °C in mixed methanol–tetrahydrofuran (3:2, v/v) medium and at 1.0 mol% catalyst concentration with respect to tripalmitin. The catalysts used for the kinetic studies were dibutyl bis( p‐chlorophenoxyl)tin 1, dibutyl bis(phenoxyl)tin 2, 1,1‐dibutyl‐5‐aza‐2,8‐dioxo‐1‐stannacyclo‐octane 3, 2,2‐dibutyl‐2‐stanna‐1,3‐benzdioxane 4 and dioctyltin oxide 5. The methanolysis was shown to proceed by a consecutive reaction pathway. Numerical analysis of the rate data yielded values of the three rate constants k1, k2 and k3 corresponding to the respective conversions, tripalmitin → dipalmitin → monopalmitin → glycerol. Based on t values ranging from 7.2 to 22.3 h−1, the following order of catalytic activity was established: 1 > 2 > 3 ≫ 5 ≫ 4; for catalyst 4 the t value was close to that of the uncatalysed reaction. A six‐fold increase in rate was observed when the catalyst concentration was raised from 1.0 to 3.0 mol% for 3. 119Sn NMR analysis of the chloroform extracts of the pot residue following solvent removal at the end of 24 h of the transesterification reaction revealed that the catalysts 1 and 3 essentially retained their chemical integrity. Copyright © 2000 John Wiley & Sons, Ltd.  相似文献   

12.
In this study, an efficient affinity purification protocol for an alkaline metalloprotease from marine bacterium was developed using immobilized metal affinity chromatography. After screening and optimization of the affinity ligands and spacer arm lengths, Cu‐iminmodiacetic acid was chosen as the optimal affinity ligand, which was coupled to Sepharose 6B via a 14‐atom spacer arm. The absorption analysis of this medium revealed a desorption constant Kd of 21.5 μg/mL and a theoretical maximum absorption Qmax of 24.9 mg/g. Thanks to this affinity medium, the enzyme could be purified by only one affinity purification step with a purity of approximately 95% pure when analyzed by high‐performance liquid chromatography and reducing sodium dodecyl sulfate polyacrylamide gel electrophoresis. The recovery of the protease activity reached 74.6%, which is much higher than the value obtained by traditional protocols (8.9%). These results contribute to the industrial purifications and contribute a significant reference for the purification of other metalloproteases.  相似文献   

13.
The aim of this work was to devise a one‐step purification procedure for monoclonal antibodies (MAbs) of IgG class by immobilized metal affinity chromatography (IMAC). Therefore, several stationary phases were prepared containing immobilized metal chelates in order to study the chromatographic behaviour of MAbs against wild‐type amidase from Pseudomonas aeruginosa. Such MAbs adsorbed to Cu(II), Ni(II), Zn(II) and Co(II)–IDA agarose columns. The increase in ligand concentration and the use of longer spacer arms and higher pH values resulted in higher adsorption of MAbs into immobilized metal chelates. The dynamic binding capacity and the maximum binding capacity were 1.33 ± 0.015 and 3.214 ± 0.021 mg IgG/mL of sedimented commercial matrix, respectively. A KD of 4.53 × 10−7 m was obtained from batch isotherm measurements. The combination of tailor‐made stationary phases of IMAC and the correct selection of adsorption conditions permitted a one‐step purification procedure to be devised for MAbs of IgG class. Culture supernatants containing MAbs were purified by IMAC on commercial‐Zn(II) and EPI‐30–IDA–Zn(II) Sepharose 6B columns and by affinity chromatography on Protein A‐Sepharose CL‐4B. This MAb preparation revealed on SDS–PAGE two protein bands with Mr of 50 and 22 kDa corresponding to the heavy and light chains, respectively. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

14.
The kinetics of the polycondensation and copolycondensation reactions of bis(3‐hydroxypropyl) terephthalate (BHPT) and bis(4‐hydroxybutyl) terephthalate (BHBT) as monomers were investigated at 270 °C in the presence of titanium tetrabutoxide as a catalyst. BHPT was prepared by the ester interchange reaction of dimethyl terephthalate and 1,3‐propanediol (1,3‐PD). Through the same method adopted for BHPT synthesis, BHBT was prepared with 1,4‐butanediol instead of 1,3‐PD. With second‐order kinetics applied for polycondensation, the rate constants of the polycondensation of BHPT and BHBT, k11 and k22, were calculated to be 4.08 and 4.18 min?1, respectively. The rate constants of the cross reactions in the copolycondensation of BHPT and BHBT, k12 and k21, were calculated with results obtained from proton nuclear magnetic resonance spectroscopy analysis. The rate constants during the copolycondensation of BHPT and BHBT at 270 °C decreased in the order k12 > k22 > k11 > k21, indicating that the reactivity of BHBT was larger than that of BHPT at 270 °C. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 2435–2441, 2002  相似文献   

15.
In this study, high‐performance liquid chromatography coupled with tandem mass spectrometry was used to assess the results of bioactive compound screening from natural products using immobilized enzyme magnetic beads. We compared three commercial magnetic beads with modified amino, carboxy, and N‐hydroxysuccinimide groups, respectively. Amino magnetic beads performed best for immobilization and were selected for further experiments. Xanthine oxidase was immobilized on amino magnetic beads and applied to screen potential inhibitors in fresh Zingiber officinale Roscoe, extracts of Scutellaria baicalensis Georgi, and Pueraria lobata Ohwi. In total, 12 potential xanthine oxidase ligands were identified from fresh Zingiber root and Scutellaria root extracts, of which eight were characterized and the concentration required for 50% inhibition was determined. Preliminary structure–function relationships were discussed based on these results. A convenient and effective method was therefore developed for the identification of active compounds from complex natural product mixtures.  相似文献   

16.
17.
The immobilized artificial membrane phosphatidylcholine (IAMPC) chromatography was evaluated for the predictability of oral absorption potential of 40 structurally unrelated drugs. The chromatographic capacity factors (kIAM) were determined as a function the pH and composition of the mobile phase, and were corrected for the molar volume of the solutes (kIAM/MWn). The correlation between kIAM/MWn and the human fraction of intestinal absorption (Fa) was highest when measured at 20% acetonitrile (pH 5.5) with the power function n = 2.5. The best-fit equation for the sigmoid relationship between kIAM/MWn and Fa was obtained: Fa (%) = 94.3 × {1-exp[-17.9 × (kIAM/MW2.5) × 106]}^2.1 (r = 0.925). This in vitro prediction method may be useful in a rapid screening of drug candidates with high oral absorption potential in humans.  相似文献   

18.
The most selective purification method for proteins and other biomolecules is affinity chromatography. This method is based on the unique biological‐based specificity of the biomolecule–ligand interaction and commonly uses biological ligands. However, these ligands may present some drawbacks, mainly because of their cost and lability. Dye‐affinity chromatography overcomes the limitations of biological ligands and is widely used owing to the low cost of synthetic dyes and to their resistance to biological and chemical degradation. In this work, immobilized aminosquarylium cyanine dyes are used in order to exploit affinity interactions with standard proteins such as lysozyme, α‐chymotrypsin and trypsin. These studies evaluate the affinity interactions occurring between the immobilized ligand and the different proteins, as a reflection of the sum of several molecular interactions, namely ionic, hydrophobic and van der Waals, spread throughout the structure, in a defined spatial manner. The results show the possibility of using an aminosquarylium cyanine dye bearing a N‐hexyl pendant chain, with a ligand density of 1.8 × 10?2 mmol of dye/g of chromatographic support, to isolate lysozyme, α‐chymotrypsin and trypsin from a mixture. The application of a decreasing ammonium sulfate gradient resulted in the recovery of lysozyme in the flowthrough. On the other hand, α‐chymotrypsin and trypsin were retained, involving different interactions with the ligand. In conclusion, this study demonstrates the potential applicability of ligands such as aminosquarylium cyanine dyes for the separation and purification of proteins by affinity chromatography. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

19.
Synthetic routes to aluminium ethyl complexes supported by chiral tetradentate phenoxyamine (salan‐type) ligands [Al(OC6H2(R‐6‐R‐4)CH2)2{CH3N(C6H10)NCH3}‐C2H5] ( 4 , 7 : R=H; 5 , 8 : R=Cl; 6 , 9 : R=CH3) are reported. Enantiomerically pure salan ligands 1–3 with (R,R) configurations at their cyclohexane rings afforded the complexes 4 , 5 , and 6 as mixtures of two diastereoisomers ( a and b ). Each diastereoisomer a was, as determined by X‐ray analysis, monomeric with a five‐coordinated aluminium central core in the solid state, adopting a cis‐(O,O) and cis‐(Me,Me) ligand geometry. From the results of variable‐temperature (VT) 1H NMR in the temperature range of 220–335 K, 1H–1H NOESY at 220 K, and diffusion‐ordered spectroscopy (DOSY), it is concluded that each diastereoisomer b is also monomeric with a five‐coordinated aluminium central core. The geometry is intermediate between square pyramidal with a cis‐(O,O), trans‐(Me,Me) ligand disposition and trigonal bipyramidal with a trans‐(O,O) and trans‐(Me,Me) disposition. A slow exchange between these two geometries at 220 K was indicated by 1H–1H NOESY NMR. In the presence of propan‐2‐ol as an initiator, enantiomerically pure (R,R) complexes 4 – 6 and their racemic mixtures 7 – 9 were efficient catalysts in the ring‐opening polymerization of lactide (LA). Polylactide materials ranging from isotactically biased (Pm up to 0.66) to medium heterotactic (Pr up to 0.73) were obtained from rac‐lactide, and syndiotactically biased polylactide (Pr up to 0.70) from meso‐lactide. Kinetic studies revealed that the polymerization of (S,S)‐LA in the presence of 4 /propan‐2‐ol had a much higher polymerization rate than (R,R)‐LA polymerization (kSS/kRR=10.1).  相似文献   

20.
Sulfur‐substituted methylmercury compounds [Hg(CH2SR)2]( 1a, R = Me; 1b, R = Ph ) react with aluminium amalgam in refluxing toluene with transmetallation to give homoleptic tris(thiomethyl)aluminium complexes [Al(CH2SR)3]( 2a, R = Me; 2b, R = Ph ) (degree of conversion: >80%, isolated yields: 2a 63%, 2b 41%). Their identities were confirmed by NMR spectros‐copy (1H, 13C) and X‐ray crystal structure analyses. In crystals of compound 2a the aluminium atoms possess a trigonal‐bipyramidal arrangement with the coordination polyhedron defined by three carbon and two sulfur atoms. Two of the three CH2SMe ligands are bridging ligands (μ‐η2; 1kC:2kS), the third one is terminal bound (η1; kC). The structure is polymeric. Crystals are threaded by helical chains built up of six‐membered Al2C2S2 rings. Crystals of 2b are built up of centrosymmetrical dimers with six‐membered Al2C2S2 rings having bridging CH2SPh ligands (μ‐η2; 1kC:2kS). On each Al atom two terminal (η1; kC)CH2SPh ligands are bound. They exhibit quite different Al‐C‐S angles (116.7(4) and 106.5(3)?). Similar values (114.32115.7? and 109.52109.9?) were found in ab initio calculations of model compounds [{Al(CH2SR)3}2]( 3a, R=H; 3b, R=Me; 3c, R=CH=CH2 ). A conformational energy diagram for rotation of one of the terminal CH2SH ligand in the parent compound 3a around the Al‐C bond is discussed in terms of repulsive interactions of lone electron pairs of sulfur atoms.  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