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.
The reaction of bis(dibenzylideneacetone)palladium(0) with white phosphorus was studied using the methods of NMR, UV spectroscopy, and X-ray powder diffraction. The products of the reaction are shown to be palladium phosphides, their composition depending on the ratio of the reagents. The mechanism of the formation of the palladium-enriched phosphides is suggested, which includes the formation of palladium diphosphide PdP2 that subsequently reacts with the excess of bis(dibenzylideneacetone)palladium(0) leading to palladium phosphides Pd5P2, Pd3P0.8, Pd4.8P, and free dibenzylideneacetone.  相似文献   

2.
The applicability of elemental phosphorus as a modifier of palladium catalysts for hydrogenation was demonstrated, and the conditions for the synthesis of nanoparticles that are highly efficient in hydrogenation catalysis were optimized. The modifying effect of elemental phosphorus depends on the P/Pd ratio; it is associated with changes in the catalyst dispersity and the nature of the formed nanoparticles containing various palladium phosphides (PdP2, Pd5P2, and Pd6P) and Pd(0) clusters. The main stages of the formation of palladium catalysts for hydrogenation were determined, and a model of an active catalyst, in which the Pd6P phosphide is the core of a nanoparticle and Pd(0) clusters form a shell, was proposed.  相似文献   

3.
The catalytic properties and nature of the nanoparticles forming in the system based on Pd(dba)2 and white phosphorus are reported. A schematic mechanism is suggested for the formation of nanosized palladium-based hydrogenation catalysts. The mechanism includes the formation of palladium nanoclusters via the interaction of Pd(dba)2 with the solvent (N,N-dimethylformamide) and substrate and the formation of palladium phosphide nanoparticles. The inhibiting effect exerted by elemental phosphorus on the catalytic process is due to the conversion of part of the Pd(0) into palladium phosphides, which are inactive in hydrogenation under mild conditions, and the formation of mainly segregated palladium nanoclusters and palladium phosphide nanoparticles. By investigating the interaction between Pd(dba)2 and white phosphorus in benzene, it has been established that the formation of palladium phosphides under mild conditions consists of the following consecutive steps: Pd(0) → PdP2 → Pd5P2 → Pd3P. It is explained why white phosphorus can produce diametrically opposite effects of on the catalytic properties of nanosized palladium-based hydrogenation catalysts, depending on the nature of the palladium precursor.  相似文献   

4.
Phase composition and surface layer state of the Pd–P hydrogenation catalyst formed at various P/Pd ratios from Pd(acac)2 and white phosphorus in a hydrogen atmosphere were determined. Palladium on the catalyst surface is mainly in two chemical states: as Pd(0) clusters and as palladium phosphides. As the P/Pd ratio increases, the fraction and size of palladium clusters decrease, and also the phase composition of formed palladium phosphides changes: Pd3P0.8 → Pd5P2 → PdP2. The causes of the modifying action of phosphorus on the properties of palladium catalysts for hydrogenation of unsaturated compounds were considered.  相似文献   

5.
The nature and catalytic properties of a hydrogenation catalyst based on Pd(acac)2 and PH3 are considered. As demonstrated by a variety of physicochemical methods (IR and UV spectroscopy, 31P and 1H NMR, electron microscopy, and X-ray powder diffraction), nanoparticles consisting of various palladium phosphides (Pd6P, Pd4.8P, and Pd5P2) and Pd(0) clusters form under the action of dihydrogen during catalyst preparation. The promoting effect of phosphine at low PH3: Pd(acac)2 ratios is mainly due to the ability of phosphine to increase the extent of dispersion of the catalyst.  相似文献   

6.
7.
The reactivity of palladium complexes of bidentate diaryl phosphane ligands (P2) was studied in the reaction of nitrobenzene with CO in methanol. Careful analysis of the reaction mixtures revealed that, besides the frequently reported reduction products of nitrobenzene [methyl phenyl carbamate (MPC), N,N′‐diphenylurea (DPU), aniline, azobenzene (Azo) and azoxybenzene (Azoxy)], large quantities of oxidation products of methanol were co‐produced (dimethyl carbonate (DMC), dimethyl oxalate (DMO), methyl formate (MF), H2O, and CO). From these observations, it is concluded that several catalytic processes operate simultaneously, and are coupled via common catalytic intermediates. Starting from a P2Pd0 compound formed in situ, oxidation to a palladium imido compound P2PdII?NPh, can be achieved by de‐oxygenation of nitrobenzene 1) with two molecules of CO, 2) with two molecules of CO and the acidic protons of two methanol molecules, or 3) with all four hydrogen atoms of one methanol molecule. Reduction of P2PdII?NPh to P2Pd0 makes the overall process catalytic, while at the same time forming Azo(xy), MPC, DPU and aniline. It is proposed that the Pd–imido species is the central key intermediate that can link together all reduction products of nitrobenzene and all oxidation products of methanol in one unified mechanistic scheme. The relative occurrence of the various catalytic processes is shown to be dependent on the characteristics of the catalysts, as imposed by the ligand structure.  相似文献   

8.
Three new palladium complexes containing a difunctional P,N‐chelate, namely tris­(chloro­{[1‐methyl‐1‐(6‐methyl‐2‐pyridyl)ethoxy]diphenylphospine‐κ2N,P}methyl­palladium(II)chloro­form solvate, 3[Pd(CH3)Cl(C21H22NOP)]·CHCl3, (III), dichloro­[2‐(2,6‐dimethyl­phen­yl)‐6‐(diphenyl­phosphinometh­yl)­pyridine‐κ2N,P]palladium(II), [PdCl2(C26H24NP)], (IV), and chloro­[2‐(2,6‐dimethyl­phen­yl)‐6‐(diphenyl­phos­phino­meth­yl)pyridine‐κ2N,P]methyl­palladium(II), [Pd(CH3)Cl(C26H24NP)], (V), are reported. Geometric data and the conformations of the ligands around the metal centers, as well as slight distortions of the Pd coordination environments from idealized square‐planar geometry, are discussed and compared with the situations in related compounds. Non‐conventional hydrogen‐bond inter­actions (C—H⋯Cl) have been found in all three complexes. Compound (III) is the first six‐membered chloro–meth­yl–phosphinite P,N‐type PdII complex to be structurally characterized.  相似文献   

9.
《Comptes Rendus Chimie》2015,18(7):785-789
The reaction of a dinuclear palladium complex [Pd2(CH3CN)6][X]2 (X = BF4, PF6) with excess furan afforded the sandwich-type bis-furan dinuclear palladium(I) complex [Pd2(μ-furan)2(CH3CN)2][X]2. The substitutionally labile bridging furan ligands in the dipalladium sandwich complex can be readily replaced with toluene to give the bis-toluene dipalladium(I) sandwich complex [Pd2(μ-toluene)2(CH3CN)2][X]2.  相似文献   

10.
Binuclear nitrosopalladium complexes Pd2(μ-COOR)22-CH2C6H4NO)2 (R = Me, CF3, or Ph) were studied by ESR spectroscopy. Analysis of parameters of ESR spectra of the polycrystalline samples and their toluene solutions suggests partial izomerization of the nitroso ligands to the nitroxide form to result in the oxidation of palladium(II) to palladium(III). __________ Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1746–1751, August, 2005.  相似文献   

11.
Evidence is presented that the dimeric π-allylic species [(η3-allyl)PdCl]2 is not intermediate in the Li2Pd2Cl6-catalysed allylic H/D exchange in alkenes. Neither H/D exchange in α-methylstyrene, nor enrichment of [(η3-2-PhC3H4)PdCl]2, was observed when the latter complex was incubated at 100°C in D3CCOOD either in the presence or in the absence of PhC(CH3)?CH2, respectively. The kinetics of H/D exchange in α-methylstyrene catalysed by Li2Pd2Cl6 were studied in some detail. The exchange proceeds at highest rates when reduction of palladium(II) takes place and is much slower in the presence of 1,4-benzoquinone as a palladium reoxidant. The exchange rate is directly proportional to the alkene and catalyst concentrations and independent of the reoxidant concentration. It is suggested that the palladium(II)-catalysed exchange involves an intermediate hydrid-allyl species where palladium has a formal oxidation state of IV.  相似文献   

12.
Rodman DL  Carrington NA  Xue ZL 《Talanta》2006,70(2):426-431
The advanced oxidation process (AOP) for the pretreatment of model palladium catalysts has been studied. Most standard metal analysis techniques are for metal ions free of organic ligands. Spent palladium catalysts contain organic ligands that need to be removed prior to analysis. AOP uses a combination of hydrogen peroxide and UV light to generate radicals that decompose such ligands, freeing up metals for further analysis. Palladium acetate Pd(OAc)2, palladium acetylacetonate Pd(acac)2, and tris(dibenzylideneacetone)dipalladium (Pd2(dba)3) were chosen as model precious metal catalysts for investigation. AOP was found to decompose ligands in Pd(OAc)2, Pd(acac)2 and give accurate Pd(II) quantification, while ligand decomposition and oxidation of Pd(0) to Pd(II) were demonstrated in treatments involving Pd2(dba)3. The effects of solubility of the palladium complexes, continuous addition of H2O2 during AOP treatments, sample pH, concentration of H2O2, and length of UV irradiation are reported.  相似文献   

13.
Palladium-based catalysts are widely used in pharmaceutical industries, which can sometimes cause palladium contamination in pharmaceutical drug manufacture. It is important to separately quantify the different oxidation states of palladium (Pd0 and Pd2+) in pharmaceuticals as they react with scavengers differently. Although palladium sensors have been under intense investigation, oxidation state differentiators are very rare. Here, we report a simple porphyrin–coumarin conjugate, PPIX-L2, that can selectively discriminate between the oxidation states of palladium. The reaction of PPIX-L2 with Pd0 showed a 24-fold fluorescence increase of the coumarin emission, meanwhile, the presence of Pd2+ led to a 98% quenching of the porphyrin emission. Fluorescent responses of PPIX-L2 towards Pd0 and Pd2+ are specific, and its sensitivity towards both palladium species is significantly increased with a detection limit of 75 nM and 382 nM for Pd0 and Pd2+ respectively.

A simple porphyrin–coumarin conjugate PPIX-L2 was developed for the discrimination of different oxidation states of palladium (Pd0 and Pd2+), and with a significantly improved sensitivity.  相似文献   

14.
The interaction of Ph3PPD(OAc)22 with molecular H2 yields a binuclear complex of zero-valent palladium, (Ph3P)2Pd2. This complex interacts reversibly with H2 in CH2Cl2, yielding (Ph3P)2Pd2H2. In argon atmosphere (Ph3P)2Pd2 reacts with [Ph3PPd(OAc)22 to form a binuclear complex of PdI with a metal—metal bond. These data, as well as the results of kinetic studies of the reactions between [Ph3PPd(OAc)22 and H2, are in agreement with an autocatalytic mechanism for the process, including catalysis of the reduction of PdII complexes by the Pd0 compounds. It has been established that the synthesized compound of PdII, PdI and Pd0 with the ratio P/Pd?1, are inactive in the hydrogenation of unsaturated compounds. The catalytically active complex (PPh)2Pd5 is formed when palladium acetate reacts with (Ph3P)2Pd2 in the presence of H2. The same compound is formed when a solution of (Ph3P)2Pd2 is treated with a mixture of H2 and O2 (or H2O2 in an atmosphere of H2). (PPh)2Pd5 is an effective catalyst for the hydrogenation of olefins, dienes, acetylenes, aldehydes, organic peroxides, quinones, O2, Schiff bases, and nitro, nitroso, and azo compounds.  相似文献   

15.
[(PPh3)3(PPh2)2Pd3Cl] Cl, benzene and aniline hydrochloride were isolated as products of the reactions of (PPh3)2PdCl2]2 or [(PPh3)PdCl2]2 with H2 in organic amines (Am). Similar products were obtained when (Ph3P)2Pd(Ph)Br was treated with H23 Both in amines and aromatic solvents. The reaction between H2 and [(PBu3)PdCl2]2 resulted in the formation of [(PBu3(PBu2)PdCl2 ·. 2 Am The kinetic data for H2 absorption by solutions of palladium(II) complexes are consistent with the heterolytic mechanism of cleavage fo hte HH bond in the coordination sphere of palladium(II); the function of the H+ acceptor being performed by the bases (e.g. Am or Ph). The reaction between the palladium complexes and H2 is autocatalytic. Reduction of the initial PdII complexes leads to lower oxidation state palladium complexes, which catalyse the reduction of PdII complexes. In the coordination sphere of the lower oxidation state palladium complexes, the oxidative addition of PR3 to Pd takes place with formation of compounds containing a Pd-R bond. It is the reaction between these complexes and H2 that yields palladium compounds with PR2 ligands.  相似文献   

16.
The size-dependent effects of the heterogeneous reaction PdCl42− + 2e = Pd0 + 4Cl were studied in hydrochloric acid solutions of H2PdCl4 at 40–70°C. Changes in the structural characteristics of palladium black were analyzed by transmission electron microscopy and X-ray diffraction. The temperature dependence of the redox potential of the PdCl42−/Pd0 pair was used to determine the thermodynamic characteristics of aggregation of fine-dispersed palladium. The heat effect of the reaction was in satisfactory agreement with direct differential scanning calorimetry measurements.  相似文献   

17.
Laser ablation of palladium was studied and velocity (energy) distributions of palladium ions evaporated by an Kr-F laser in a vacuum were obtained. The optimum values of energy fluence (fluence rate) of laser radiation for doping tin dioxide films, at which neither multiply charged PdN+ ions nor ionized clusters Pd N + , occur in a plasma, were determined. From time-of-flight probe measurement data, Pd+ implantation depths in SnO2 films were calculated, which qualitatively agree with the results obtained by secondary neutral mass spectrometry. Electric conductivity measurements on the obtained films in a gas phase showed that introduction of palladium into polycrystalline SnO2 films by laser ablation significantly enhanced their gas sensitivity to hydrogen.  相似文献   

18.
The formation of low-valence palladium species Pd n 2+ (n ≥ 2) in the course of oxidation of aliphatic C1-C4 alcohols with oxygen in the presence of palladium(II) tetraaqua complex in aqueous solution was proved UV-spectrophotometrically by the absorption band with a maximum at 312 or 316 nm depending on particular alcohol.  相似文献   

19.
Er3Pd7P4 — Crystal Structure Determination and Extended Hückel Calculations Er3Pd7P4 was prepared by heating the elements (1050°C) and investigated by means of single-crystal X-ray methods. The compound crystallizes in a new structure (C2/m; a = 15.180(3) Å, b = 3.955(1) Å, c = 9.320(1) Å, β = 125,65(1)°; Z = 2) with a three-dimensional framework of Pd and P atoms and with Er atoms in the holes. The Pd atoms are surrounded tetrahedrally, trigonally or linearly by P atoms, which are coordinated by nine metal atoms in the form of a tricapped trigonal prism. Therefore the atomic arrangement of Er3Pd7P4 is related to the structures of ternary transition metal phosphides with a metal: phosphorus ratio of 2:1. Band calculations using the Extended Hückel method show strong covalent Pd? P bonds and weak bonding interactions between Pd atoms with Pd? Pd distances shorter than 2.9 Å.  相似文献   

20.
The new dinuclear palladium complex Pd2(-S,N-SC7H5N2)4 with a Chinese-lantern structure was synthesized by the reaction of K2PdCl4 with 2-mercaptobenzimidazole and structurally characterized by X-ray diffraction analysis.  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