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 125 毫秒
1.
d-Block transition-metal-containing polymer blends which form coordination complexes are described in this treatise. The model compounds are zinc acetate dihydrate, copper acetate dihydrate, nickel acetate tetrahydrate, cobalt chloride hexahydrate, palladium chloride bis (acetonitrile), and the dimer of dichlorotricarbonylruthenium (II). Two classes of ligands are of interest. Poly (4-vinylpyridine), P4VP, and copolymers that contain 4-vinylpyridine repeat units form complexes with zinc, copper, nickel, cobalt, and ruthenium salts. Atactic 1,2-polybutadiene contains olefinic sidegroups that displace weakly bound acetonitrile ligands and coordinate to palladium chloride. Thermal analysis via differential scanning calorimetry suggests that the glass transition temperature of the polymeric ligand is enhanced by these low-molecular-weight transition-metal salts in binary and ternary blends. In some cases, d-block salts function as transition-metal compatibilizers for copolymers that would otherwise be immiscible. The isothermal ternary phase diagram for polybutadiene with palladium chloride highlights regions of gelation, precipitation, and transparent solutions during blend preparation in tetrahydrofuran. Fourier transform infrared spectroscopy provides molecular-level data that support the concept of polymeric coordination complexes. High-resolution carbon-13 solid-state NMR spectroscopy identifies (1) near-neighbor interactions between polymeric pyridine ligands and the ruthenium salt, and (2) a considerable reduction in the molecular mobility of the polybutadiene chain backbone when it forms a coordination complex with palladium chloride. The elastic modulus of polybutadiene increases by three orders of magnitude when the palladium salt concentration is 4 mol % in a solid-state glassy film. A thermodynamic interpretation of ligand field stabilization energies appropriate to tetrahedral cobalt and octahedral nickel complexes is employed to estimate the synergistic enhancement of the glass transition temperature, particularly when coordination crosslinks are present. ©1995 John Wiley & Sons, Inc.  相似文献   

2.
This study focuses on the thermal and mechanical properties of 1,2-polybutadiene and 3,4-polyisoprene with an inorganic salt, bis(acetonitrile)dichloropalladium (II). Upon mixing in THF, effective crosslinks are formed because the acetonitrile ligands of the palladium salt are displaced by olefinic pendant groups of the polymers. Using a simple nth-order irreversible kinetic rate model, the palladium-catalyzed Heck-like exothermic reaction in solid films was characterized via isothermal and nonisothermal DSC. Thermal energy and mass balances appropriate to a batch reactor are developed from first principles and applied to the isothermal DSC output curve to calculate the time dependence of reactant conversion. Relevant kinetic parameters, such as the order of the reaction, the characteristic time constant for the chemical reaction, and the activation energy, have been determined. The kinetic data suggest that the palladium-catalyzed crosslinking reactions are diffusion controlled in the solid state because the reaction order is very close to unity. Higher glass transition temperatures (Tg) are measured by dynamic mechanical thermal analysis (DMTA) and differential scanning calorimetry (DSC) when (i) palladium concentration, (ii) annealing (heat treatment) time, and (iii) annealing temperature increase. After 2 h of annealing at 80°C, which corresponds to a temperature below the first exothermic crosslinking reaction (≅ 115°C) during nonisothermal DSC kinetic studies, rubbery materials containing very low concentrations of PdCl2 (i.e., 0.5 mol %) exhibit reinforced ductile stress-strain response. When annealing is performed at the peak temperature of the first exothermic event, the rubbery materials are transformed into glasses. Transition-metal compatibilization of atactic 1,2-polybutadiene and 3,4-polyisoprene via PdCl2 is demonstrated by monitoring the glass transition obtained from dynamic mechanical tan δ profiles. The effect of annealing this ternary reactive “blend” produces a glassy material exhibiting an elevated Tg and synergistic mechanical properties. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
In the work, reactions of a partially deprotonated polypyrrole doped with hydroxide ions (PPyOH) in various PdCl2 aqueous solutions which differed in acidity were studied. Using X-ray photoelectron spectroscopy, X-ray diffraction and scanning electron microscopy it was established that in the PdCl2 solutions of lower acidity PPyOH was oxidatively doped and Pd0 and Pd2+ were incorporated into the polymer matrix. Pd2+ formed palladium(II) hydroxy-and/or aquochlorocomplex dopant anions and/or was coordinated by nitrogen atoms of the polymer (Pd-N bond). Additionally, deprotonation of PPyOH occurred in the PdCl2 solutions of lower acidity. It was proposed that deprotonation of PPyOH was caused by nucleophilic attack of [PdCl3(H2O)] on the positively charged, doped polymer chain. By comparison of the PPyOH and chloride-doped polypyrrole (PPyCl)-palladium systems prepared in similar PdCl2 solutions of lower acidity it was shown that the type of the counterion in the starting polymer has a decisive effect on the deprotonation process.PPyOH was less reactive towards palladium species in the PdCl2 solutions of higher acidity where [PdCl4]2− was the dominant complex. PPy-palladium systems containing exclusively Pd2+ were obtained in this case. It was proposed that incorporation of palladium species in these conditions proceeded via an acid-base reaction or coordination of palladium ions by the polymer chain (Pd-N bond formation).Results of the studies may serve as the basis for the preparation of a variety of polypyrrole-supported palladium catalysts.  相似文献   

4.
The mechanical properties of atactic 1,2-polybutadiene and 3, 4-polyisoprene can be modified significantly with the addition of bis(acetonitrile)dichloropalladium(II). These weak rubbery polymers are transformed into glassy materials when the salt concentration is ≅ 4 mol %, in the absence of high-temperature annealing. Stress-strain measurements and Fourier transform infrared (FTIR) spectra for blends of cis-polybutadiene and PdCl2, without high-temperature annealing, suggest that π-complexes form between palladium and the olefinic groups within the backbone of the polymer. These solid complexes cannot be dissolved in the original solvent (tetrahydrofuran), nor can they be disrupted by triphenylphosphine. Young's modulus of the cis-polymer is enhanced by a factor of 50 when the salt concentration is 4 mol %, and the fracture strain is approximately 300%. An exothermic process centered at ≅ 250°C accompanied by minimal weight loss suggests that PdCl2 could trigger high-temperature dimerization reactions of the carbon–carbon double bonds in the backbone of the cis-polymer. High-temperature annealing effects on the stress-strain response of cis-polybutadiene with 4 mol % PdCl2 are consistent with the data from calorimetry, suggesting that catalytically induced chemical crosslinking is operative at high temperatures. This latter claim is verified by infrared spectroscopy at ambient and elevated temperatures. Hence, bis(acetonitrile)dichloropalladium(II) coordinates to and catalyzes dimerization reactions of olefinic groups when they are present in the main chain or the sidegroup. This square-planar transition-metal salt also enhances the high-strain mechanical response of commercial styrene-butadiene-styrene triblock copolymers (KratonTM D series). Reactive blending and compatibilization with transition-metal salts are attractive strategies to modify the mechanical properties of commercially important diene-based polymers that contain unsaturation in the main chain or the sidegroup. © 1996 John Wiley & Sons, Inc.  相似文献   

5.
The interactions of palladium cation-anion compounds (C4H10NO)2[PdCl4], K2[PdCl4], and K2[PdBr4] with DNA in 0.005 M NaCl and 0.15 M NaCl solutions were studied by spectrophotometry, circular dichroism, viscosimetry, dynamic birefringence, and atomic force microscopy. The interactions are primarily effected by coordination of the donor atoms of DNA bases by palladium. The end products of interactions with palladium acido complexes are independent of the macromolecule and the nature of halogen X in [PdX4]2−. The significant changes in the conformation of DNA in palladium complexes resulted from both intra- and intermolecular cross-linkings induced by palladium.  相似文献   

6.
In this work, the static and dynamic light scattering measurements were used to investigate the solution properties and the aging effects on PVA/DMSO/water ternary system in dilute region at 25 °C. It was found that the phase separation and aggregate behavior occurs rapidly and obviously when DMSO mole fraction (X1) in the solvent mixture is between 0.2 and 0.33, especially at 0.25. In this solvent composition range, a broad peak which indicates phase separation and chain aggregation can be observed from static light scattering measurement. However, when DMSO mole fraction is increased to 0.37, no such peak is present. For this ternary system, the gelation mechanism and the relationship between the phase separation behavior and the gelation of the formed physical gels were also investigated through the gelation kinetic analyses in the dilute and semi-dilute region. It is concluded that the cononsolvency effect in the dilute solution is not the sole origin that affects the phase separation, aggregation, and gelation behavior for the ternary system in a higher polymer concentration range. The hydrodynamic factors such as the higher viscosity and slower polymer chain diffusion that are resulted from higher polymer concentration should be also considered.  相似文献   

7.
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   

8.
[(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   

9.
Reaction of PdCl2(CH3CN)2 with the sodium salt of 5‐mercapto‐1‐methyltetrazole (MetzSNa) in methanol solution affords an interesting dinuclear palladium complex [Pd2(MetzS)4 ] ( 1 ). However, treatment of PdCl2(CH3CN)2 with neutral MetzSH ligand in methanol solution produces a mononuclear palladium complex [Pd(MetzSH)4]Cl2 ( 2 ). Both complexes were characterized by IR, 1HNMR, UV‐Vis spectroscopy as well as X‐ray crystallography. Single‐crystal X‐ray diffraction analyses of two complexes lead to the elucidation of the structures and show that 1 possesses an asymmetric structure: one Pd atom is tetracoordinated by three sulfur atoms and one nitrogen atom to form PdS3N coordination sphere, the other Pd atom is tetracoordinated by three nitrogen atoms and one sulfur atom to form PdSN3 coordination sphere. The molecules of 1 are associated to 1‐D infinite linear chain by weak intermolecular Pd···S contacts in the crystal lattice. In 2 , the Pd atom lies on an inversion center and has a square‐planar coordination involving the S atoms from four MetzSH ligands. The two chloride ions are not involved in coordination, but are engaged in hydrogen bonding.  相似文献   

10.
Palladium(I) carbonyl chloride (PdCOCl) n has been synthesized by the effect of CO on PdCl2 in the presence of trace water. Anionic palladium(I) carbonyl chloride has been synthesized during treatment of ethanol-based or acetone-based solutions of H2PdCI4 containing small amounts of water; it has been isolated from the solution as the salt Cs2[Pd2(CO)2Cl4]. IR-spectra, X-ray diffraction patterns, and thermogravimetry data on synthesized palladium(I) carbonyl complexes are presented.  相似文献   

11.
The crystal structures of PdCl2[(?)-DIOP], PtCl2[(?)-DIOP] and of NiCl2-[(?)-DIOP] have been determined by X-ray analysis and refined by least-squares methods [(?)-DIOP=(?)-2,2-dimethyl-4,5-bis(diphenylphosphinomethyl)-1,3-dioxolane]. The coordination around the nickel atom is tetrahedral, the coordination around palladium and platinum is square planar. The unit cell of the palladium complex contains two non-equivalent molecules with different conformations of the seven-membered chelate ring involving the metal and the two phosphorus atoms. PtCl2[(?)-DIOP] is isostructural with the corresponding palladium complex.  相似文献   

12.
Palladium(II) extraction from hydrochloric acid solutions with a novel weakly basic complexing reagent, 4-[(hexylsulfanyl)methyl]-3,5-dimethyl-1H-pyrazole, dissolved in chloroform was studied. Palladium(II) was found to be highly efficiently extracted from 0.1–3 mol/L HCl solutions. A coordination mechanism of palladium(II) extraction with a protonated form of the reagent via fast interphase transfer of ion associates was proposed. The composition of the extracted compound, [PdCl2μ-L]n (n > 2), was found, and the way of coordination of the reagent to metal ions through N(2) nitrogen atom and thioether sulfur atom was determined. The reagent can be recommended for concentrating palladium(II) and selectively separating it from platinum(IV), copper(II), nickel(II), and iron(III).  相似文献   

13.
The effect of acidity and equilibrium chloride ion concentration on the interaction of PdCl4 2- with cystine (H2CySS) in hydrochloric acid solutions was studied. Pd(H4CySS)Cl3+ complex was found to form at [Cl] = 1.0, 0.5, or 0.25 mol/Linthe [H+] range from 0.10 to 1.00 mol/L; the relevant equilibrium constant was determined. Monodentate coordination of cystine to palladium(II) through the sulfur atom was proposed on the basis of analysis of conditional stability constants as functions of [Cl] and [H+].  相似文献   

14.
The electroless metallization of polymers needs an activation of their surface which consists of palladium chemisorption. In this study, the effect of surface treatments of polystyrene and polyamide substrates by reactive gas plasmas (O2, NH3, N2) has been followed by XPS analysis. According to the functional groups grafted on the surface, specific chemisorption reactions can occur. The latter have been highlighted through a comparative investigation of two activation processes, viz. a conventional way using successively SnCl2 and PdCl2 solutions and a new procedure, developed by the authors, using only a PdCl2 solution. This work shows that this simplified process can be extended to any polymer whose surface is grafted with nitrogenated functions.  相似文献   

15.
Triphenylantimony was used as an efficient agent for C-phenylation of methyl acrylate in the presence of ButOOH (1 to 2 mol) and a palladium salt (PdCl2, Li2PdCl4; 0.04 mol) in AcOH at 50 °C. The yield of methyl cinnamate is two moles per mole of the starting Ph3Sb.  相似文献   

16.
The effects of four d8 transition-metal complexes from Group 10 on the thermal, mechanical, optical, and spectroscopic properties of atactic 1,2-polybutadiene are compared, in addition to their ability to induce gelation. Olefin coordination and subsequent metal-catalyzed chemical crosslinking occur much more quickly, and to a greater extent, at ambient temperature with PdCl2(CH3CN)2 than with PtCl2(C6H5CN)2. Alkene side groups in the polymer attack the pseudo-square-planar metal center (i.e., Pd2+ or Pt2+) from above or below the plane of the coordinatively unsaturated low-molecular-weight organometallic complex and displace neutral acetonitrile or benzonitrile ligands via an associative mechanism. Gelation occurs much more quickly with Pd2+ than with Pt2+, and the ambient-temperature elastic modulus of solid polybutadiene/palladium complexes increases significantly, without high-temperature annealing, so that a weak rubbery polymer is transformed into a glass via 3 mol % Pd2+. Alkene functional groups in the side chain of the polymer do not coordinate to bis(dimethyl)glyoximatonickel(II) at ambient temperature because (1) it is difficult to displace anionic dimethylgloxime ligands that are bidentate; (2) these planar nickel complexes with C2h symmetry are stacked along the c axis via interlocking methyl groups on adjacent molecules; and (3) there is a lack of π back-bonding between dxy on Ni(II) and empty π* antibonding orbitals of CC, which typically stabilizes olefin complexes with pseudo-square-planar d8 metal centers. Pseudo-octahedral nickel(II) chloride hexahydrate does not form a complex with the polymer, in agreement with some macroscopic properties of these materials. The observed trend in the transition-metal-modified properties of atactic 1,2-polybutadiene in the solid state and in the gel state is Pd(II) > Pt(II) ≫ Ni(II). © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 2270–2285, 2004  相似文献   

17.
Wen Huang  Jie Tang 《Tetrahedron》2005,61(41):9783-9790
Palladium-benzimidazolium salt catalyst systems have been studied for the Suzuki coupling. A different substitutent effect has been uncovered with respect to nitrogen substituents in the benzimidazolium salts from the palladium-imidazolium salt analogs. A practical and highly active palladium catalyst system, PdCl2/N,N′-dibenzylbenzimidazolium chloride 2, has been identified for the Suzuki coupling of aromatic halides with arylboronic acids. The coupling of a wide array of aromatic halides with arylboronic acids with the PdCl2-2 catalyst system gave good to excellent yields. The effective palladium loading could be as low as 0.0001 mol% and 0.01-0.1 mol% for iodide and bromide substrates, respectively. The coupling of unactivated aromatic chlorides with arylboronic acids also gave good results using Cs2CO3 as base with a 2 mol% palladium loading. The electronic factors from aromatic halides exert a significant influence on the Suzuki coupling catalyzed by the PdCl2-2 system while the electronic effect from the arylboronic counterparts is negligible. The aromatic halides with modest steric hindrance could also couple smoothly with phenylboronic acids using the PdCl2-2 catalyst system.  相似文献   

18.
EXAFS spectroscopy was used to study the influence of various factors on the structure of PdCl2 complexes with organic sulfides in organic solvents. Absolute interatomic distances in the first coordination sphere of Pd were determined for the complexes [PdCl2·2(C6H13)2S] (I), [PdCl2·(C6H13)2S]2 (II), [PdCl2·2(C6H5)2S] (III), and [PdCl2·(C4H9)S(C4H7)] (IV) and for their solutions in some organic solvents. Our hypothesis that aromatic solvent molecules are coordinated to palladium atoms through weak π-bonds, which was proposed for complex (I) in benzene, is supported fror benzene and pseudocumene solutions of complexes (I), (II), and (III). It is shown that the characteristic features of the specific solvation of the complexes under study are determined by the electron properties and spatial structures of the molecules as well as by the donating abilities of the solvents. Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences. Translated fromZhurnal Strukturnoi Khimii, Vol. 36, No. 6, pp. 1030–1037, November–December, 1995. Translated by I. Izvekova  相似文献   

19.
Synthesis and Crystal Structure of the Complexes [(n‐Bu)4N]2[{(THF)Cl4Re≡N}2PdCl2], [Ph4P]2[(THF)Cl4Re≡N‐PdCl(μ‐Cl)]2 and [(n‐Bu)4N]2[Pd3Cl8] The threenuclear complex [(n‐Bu)4N]2[{(THF)Cl4Re≡N}2 PdCl2] ( 1 ) is obtained in THF by the reaction of PdCl2(NCC6H5)2 with [(n‐Bu)4N][ReNCl4] in the molar ration 1:2. It forms orange crystals with the composition 1· THF crystallizing in the monoclinic space group C2/c with a = 2973.3(2); b = 1486.63(7); c = 1662.67(8)pm; β = 120.036(5)° and Z = 4. If the reaction is carried out with PdCl2 instead of PdCl2(NCC6H5)2, orange crystals of hitherto unknown [(n‐Bu)4N]2[Pd3Cl8] ( 3 ) are obtained besides some crystals of 1· THF. 3 crystallizes with the space group P1¯ and a = 1141.50(8), b = 1401.2(1), c = 1665.9(1)pm, α = 67.529(8)°, β = 81.960(9)°, γ = 66.813(8)° and Z = 2. In the centrosymmetric complex anion [{(THF)Cl4Re≡N}2PdCl2]2— a linear PdCl2 moiety is connected in trans arrangement with two complex fragments [(THF)Cl4Re≡N] via asymmetric nitrido bridges Re≡N‐Pd. For Pd(II) thereby results a square‐planar coordination PdCl2N2. The linear nitrido bridges are characterized by distances Re‐N = 163.8(7)pm and Pd‐N = 194.1(7)pm. The crystal structure of 3 contains two symmetry independent, planar complexes [Pd3Cl8]2— with the symmetry 1¯, in which the Pd atoms are connected by slightly asymmetric chloro bridges. By the reaction of equimolar amounts of [Ph4P][ReNCl4] and PdCl2(NCC6H5)2 in THF brown crystals of the heterometallic complex, [Ph4P]2[(THF)Cl4Re≡N‐PdCl(μ‐Cl)]2 ( 2 ) result. 2 crystallizes in the monoclinic space group P21/n with a = 979.55(9); b = 2221.5(1); c = 1523.1(2)pm; β = 100.33(1)° and Z = 2. In the central unit ClPd(μ‐Cl)2PdCl of the centrosymmetric anionic complex [(THF)Cl4Re≡N‐PdCl(μ‐Cl)]22— the coordination of the Pd atoms is completed by two nitrido bridges Re≡N‐Pd to nitrido complex fragments [(THF)Cl4Re≡N] forming a square‐planar arrangement for Pd(II). The distances in the linear nitrido bridges are Re‐N = 163.8(9)pm and Pd‐N = 191.5(9)pm.  相似文献   

20.
Palladium colloids revealing narrow particle size distributions can be obtained by chemical reduction using tetra–alkylammonium hydrotriorganoborates. Combining the stabilizing agent [NR] with the reducing agent [BEt3H?] provides a high concentration of the protecting group at the reduction centre. Alternatively, NR4X (X = halogen) may be coupled to the metal salt prior to the reduction step: addition of N(octyl)4Br to Pd(ac)2 in THF, for example, evokes an active interaction between the stabilizing agent and the metal salt. Reduction of NR-stabilized palladium salts with simple reducing agents such as hydrogen at room temperature yields stable palladium organosols which may be isolated in the form of redispersible powders. The anion of the palladium salt is crucial for the success of the colloid synthesis. Electron microscopy shows that the mean particle size ranges between 1.8 and 4.0 nm. An X–ray–photoelectron spectrscopic examination demonstrated the presence of zerovalent palladium. These palladium colloids may serve as both homogeneous and heterogeneous hydrogenation catalysts. Adsorption of the colloids onto industrially important supports can be achieved without agglomeration of palladium particles. The standard activity of a charcoal catalyst containing 5% of colloidal palladium determined through the cinnamic acid standard test was found to exceed considerably the activity of the conventional technical catalysts. In addition, the lifespan of the catalyst containing a palladium colloid, isolated from the reduction of [N(octyl)4]2PdCl2Br2 with hydrogen, is superior to conventionally prepared palladium/charcoal (Pd/C) catalysts. For example, the activity of a conventional Pd/C catalyst is completely suppressed after 38×103 catalytic cycles per Pd atom, whereas the colloidal Pd/C catalyst shows activity even after 96times;103 catalytic cycles.  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