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 156 毫秒
1.
The system Pd(OAc)2/BQ/Co(acac)3 (BQ=benzoquinone), in combination with tetrabutylammonium bromide (TBAB) as a surfactant agent and a chelating ligand such as 2,9-dimethyl-1,10-phenanthroline (dmphen) or 2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline (dmdpphen), is an efficient catalyst for the oxidative carbonylation of phenol to diphenyl carbonate (DPC). The best results have been obtained using the system Pd(OAc)2/BQ/Co(acac)3/dmphen=1/30/8/5 (molar ratio) in which [Pd]=10−3 mol l−1 and TBAB/Pd=60/1. This system gives the maximum productivity of 700 mol DPC/mol Pd h at 135°C and under Ptot=60 atm (CO/O2=10/1 molar ratio). The role of each component of the catalytic system is discussed and a catalytic cycle is proposed.  相似文献   

2.
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   

3.
Pd/Ag films were electrolessly deposited onto p-silicon (100)-activated seed layers of Ag and Pd, respectively, in the solution of 0.005 mol l−1 AgNO3 + 0.005 mol l−1 PdCl2 + 4.5 mol l−1 NH3 + 0.16 mol l−1 Na2EDTA+0.1 mol l−1 NH2NH2 (pH 10.5) at room temperature. The morphology and composition of the films were studied comparatively by using atomic force microscopy (AFM) and X-ray photoelectron spectroscopy (XPS). Cathodic polarization curves for hydrogen evolution were recorded in 0.5-mol l−1 H2SO4 without illumination, in which the obtained films served as working electrodes. The experimental results show that the film obtained on the Ag seed layer was rather a pure Ag film and not a Pd/Ag film, and the Ag deposition rate on Pd sites was much faster than that on Ag sites.  相似文献   

4.
Metal Sulfur Nitrogen Compounds. 20. Reaction Products of PdCl2 and Pd(CN)2 with S7NH. Preparation and Structure of the Complexes [Ph6P2N][Pd(S3N)(S5)] and X[Pd(S3N)(CN)2] X = [Me4N]+, [Ph4P]+ With PdCl2 and [Ph6P2N]OH S7NH forms the complex salt [Ph6P2N][Pd(S3N)(S5)], which could be isolated in two modifications (α- and β-form). The α-form is triclinic, a = 9.347(4), b = 14.410(8), c = 15.440(11) Å, α = 76.27°(5), β = 77.06°(4), γ = 76.61α(4), Z = 2, space group P1 . The β-form is orthorhombic, a = 9.333(2), b = 17.659(4), c = 23.950(6) Å, Z = 4. The structure of the metal complex is the same in the two modifications. One S3N? and one S52? are coordinate as chelate ligands to Pd. From S7NH, Pd(CN)2, and XOH X = [(CH3)4N]+ and [(C6H5)4P]+ the salts X[Pd(S3N)(CN)2] were formed. The (CH3)4N-salt is isomorphous with the analogous Ni compound described earlier, the (C6H5)4P-salt is triclinic, a = 9.372(4), b = 10.202(5), c = 13.638(6) Å, α = 86.36α(4), β = 85.66°(4), γ = 88.71°(4), Z = 2, space group P1 . One S3N? chelate ligand and two CN? ions are bound to Pd. In all these complexes the coordination of Pd is nearly square planar.  相似文献   

5.
Crystal Structure and Data from Vibrational Spectra of cis-Na2[Pd(SO3)2en] · 4 H2O The compound cis-Na2[Pd(SO3)2en] · 4 H2O (en = 1,2-diaminoethane) crystallizes in the orthorhombic space group Pnma with a = 623.7(2), = 1070.9(10), c = 1989.5(30) pm and Z = 4. In the [Pd(SO3)2en]2? anions the trans-influence of the sulfite ligands manifests itself in long Pd? N bonds with short Pd? S distances. A set of Na+ ions is present in face-sharing octahedra Na(OH2)6+, forming rods [Na(OH2)6/2]+ parallel to [100]. A second set of Na+ ions is surrounded by two H2O molecules and four O atoms from SO3 ligands of two anions to form likewise octahedra with face-sharing, yielding rods [Na(OH2)2/2{(OSO2)2Pd en}2/2]? parallel to [100]. Comparatively low v(Pd? N)-frequencies reveal the trans-influence of the sulfite ligands also in the vibrational spectra.  相似文献   

6.
New palladium(0) complexes with a variety of coordinated olefins [Pd(olefin)(PMePh2)2] (II) (olefin = styrene, ethyl methacrylate, methyl methacrylate, methyl acrylate, methacrylonitrile, and dimethyl maleate), were prepared by the reactions of [PdEt2(PMePh2)2] (I) with corresponding olefins in toluene. These complexes were characterized by means of elemental analysis, IR and 1H NMR spectroscopy and the chemical reactions. The dissociation of the coordinated olefin from complex II in solution was confirmed by spectroscopic studies of [Pd(mma)(PMePh2)2] (mma = methyl methacrylate). From the variable temperature NMR study, kinetic parameters for the dissociation process were determined as Ea = 7 kcal/mol, and ΔS3 (293 K) = -30 cal/deg · mol. Some new hydrido complexes, [Pd(H)ClL2] (IV) (L = PMePh2, PEtPh2 and PEt2Ph), were prepared by the reactions of [Pd(olefin)L2] with dry HCl.  相似文献   

7.
Abstract

The reaction between 5,5-dimethyl-2-thioxoimidazolidin-4-one (H2L) and [PdCl4]2- has been studied in aqueous solution by potentiometric and spectrophotometric measurements. In the presence of the palladium salt, H2L is completely monodeprotonated (HL?); from spectrophotometric measurements, only two complexes having 1:1 and 1:2 Pd/ligand mol ratios have been identified. Potentiometric titrations, carried out on solutions with 1:1, 1:2, 1:3 and 1:4 metal/ligand mol ratios, show that these complexes must be formulated as Pd(HL)2 and [Pd2(HL)2(μ-H2O)(μ-OH)]+. Ionization constants of the pure ligand and formation constants of the complexes give pH distribution curves of the various species and the spectra of the two complexes. From MeOH, S-coordinated Pd(H2L)nCl2 (n = 2–4) complexes have been separated in the solid state; from water, two complexes of formula Pd(H2L)(HL)Cl and Pd(HL)Cl have been obtained with HL? N,S-coordinated to the metal.  相似文献   

8.
Metal Complexes of Biologically Important Ligands. CXXVI. Palladium(II) and Platinum(II) Complexes with the Antimalarial Drug Mefloquine as Ligand The coordination sites of the antimalarial drug mefloquine (L) were studied. Reactions of the chloro bridged complexes (allyl)Pd(μ‐Cl)2Pd(allyl) and (R3P)(Cl)M(μ‐Cl)2M(Cl)(PR3) (M = Pd, Pt) with racemic mefloquine give the compounds (allyl)(Cl)Pd(L) ( 1 ), Cl2(Et3P)Pt(L) ( 2 ) and Cl2(Et3P)Pd(L) ( 3 ) with coordination of the piperidine N atom of mefloquine. In the presence of NaOMe the N,O‐chelate complexes Cl(Et3P)Pt(L–H+) ( 4 ) and Cl(R3P)Pd(L–H+) ( 5 , 6 , R = Et, nBu) were obtained. Protection of the piperidine N atom of mefloquine by protonation allows the synthesis of the complexes Cl2(Et3P)Pt(L + H+) ( 7 ) in which mefloquine is coordinated via the quinoline N atom. The structures of 2 , 3 and 4 were determined by X‐ray diffraction analysis. In the crystal of 4 pairs of enantiomers are found which are linked by two hydrogen bridges between the amine group and the chloro ligand.  相似文献   

9.
Base hydrolysis reactions of [Cr(tmpa)(NCSe)]2O2+, [Cr(tmpa)(N3)]2O2+, [Cr2(tmpa)2(μ−O)(μ−PhPO4)]4+ and [Cr2(tmpa)2(μ−O)(μ−CO3)]2+ follow the pseudo‐first‐order relationship (excess OH): kobsd=ko+kbQp[OH]/(1+Qp[OH]). For the CO32− complex, kb(60°C)=(1.50±0.03)×10−2 s−1; ΔH‡=61±2 kJ/mol, ΔS‡=−99±7 J/mol K; Qp(60°C)=(3.8±0.3)×101 M−1; ΔH°=67±2 kJ/mol, ΔS°=230±7 J/mol K (I=1.0 M). An isokinetic relationship among kOH(=kbQp) activation parameters for five (tmpa)CrOCr(tmpa) complexes shows that all follow essentially the same pathway. Activated complex formation is thought to require nucleophilic attack of coordinated OH at the chromium‐leaving group bond in the kb step, accompanied by reattachment of a tmpa pyridyl arm displaced by OH in the Qp preequilibrium. Abstraction of both thiocyanate ligands was observed upon mixing [Cr(tmpa)(NCS)]2O2+ with [Pd(CH3CN)4]2+ in CH3CN solution. The proposed mechanism requires rapid complexation of both reactant thiocyanate ligands by Pd(II) (Kp(25°C)=(4.5±0.2)×108 M−2; ΔH°=−32±6 kJ/mol, ΔS°=59±19 J/mol K) prior to rate‐limiting Cr NCS bond‐breaking (k2(25°C)=(1.17±0.02)×10−3 s−1; ΔH‡=98±2 kJ/mol, ΔS‡=27±5 J/mol K). Pd(II)‐assisted NCS abstraction is not driven by weakening of the Cr( )NCS bond through ligation of the sulfur atom to palladium, but rather by a favorable ΔS‡ resulting from the release of Pd(NCS)+ fragments and weak solvation of the activated complex in CH3CN solution. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 351–356, 1999  相似文献   

10.
The reaction of isoprene with aniline, catalyzed by Pd (acac)2–(RO)3P‐CF3COOH, (1:4:4) (R = Me, Et, acac = (CH3CO)2CH‐) in MeCN solution, results in high (up to 89 mol.%) selectivity of N–(3‐methyl‐2‐buten‐1‐yl) aniline. The presence of telomeric products in the reaction mixture is observed at a P/Pd ratio of 1:2 and 1:1. The use of (1,1,1‐trifluoro, 4‐perfluorocyclo hexyl ‐2,4‐butanedionato) palladium as the catalyst gives rise to 92 mol% mol.selectivity of telomers by the favored tail‐to‐head and head to head coupling.  相似文献   

11.
A novel heterogeneous Al2O3–Pd catalyst has been prepared by the sol–gel method; bayberry tannin (BT) was used as stabilizer to prevent the migration and aggregation of Pd species during calcination. According to N2 adsorption/desorption determination, Al2O3–Pd has a mesoporous structure and its specific area is as high as 336.5 m2/g. Transmission electron microscopy observation indicated that the size of the Pd particles was greatly reduced by the presence of BT. On the basis of X-ray photoelectron spectra analysis, it was found that the most of Pd nanoparticles were dispersed in the pores, implying that BT can prevent migration of Pd particles from the pores to the outer surface of Al2O3 during calcination. For comparison, Al2O3–Pd* was prepared by the sol–gel method but without use of BT. In the hydrogenation of acrylic acid, Al2O3–Pd had high catalytic activity and excellent reusability compared with commercial and traditionally prepared heterogeneous Pd catalysts. The turnover number of Al2O3–Pd is as high as 11,328.0 mol/mol after recycling seven times, which is much higher than that of a commercial Pd–C catalyst (8048.0 mol/mol).  相似文献   

12.
The preparations of the binuclear hydrido-bridged cations [(terdentate ligand)Pd(μ-H)Pd(terdentate ligand)]+ from [(terdentate ligand)Pd(acetone)]+ and NaO2CH and [(terdentate ligand)Pd(μ-H)Pt(terdentate ligand)]+ from [(terdentate ligand)Pd(acetone)]+ and [(terdentate ligand)PtH] (terdentate ligand = 2,6-(Ph2PCH2)2C6H3) are reported. The preparation of the cation [(terdentate ligand)Pt(μ-H)Pt(terdentate ligand)]+ is also reported.  相似文献   

13.
A sensor based on graphite electrode modified with palladium‐platinum‐palladium film is proposed for phosphite determination by flow‐injection amperometry. The modified electrode was prepared by a sequential cathodic deposition of Pd, Pt and Pd on a graphite electrode from 0.5% m/v PdCl2+28% m/v NH4OH and 2% m/v H2PtCl6+10% v/v H2SO4 solutions. After suitable conditioning, the electrode showed catalytic activity for phosphite oxidation when 0.15 V was applied. The proposed system handles approximately 50 samples per hour (0.01–0.05 mol L?1 Na2HPO3; R2=0.9997), consuming ca. 70 μL of sample per determination. The limit of detection and amperometric sensibility were 5×10?4 mol L?1 and 1.5 mA L mol?1, respectively. The proposed method was applied to analysis of fertilizer samples without pre‐treatment. Results are in agreement with those obtained by spectrophotometry and titrimetry at 95% confidence level and good recoveries (96–109%) of spiked samples were found. Relative standard deviation (n= 12) of a 0.01 mol L?1 Na2HPO3 sample was 2%. The useful lifetime of modified electrode was around 220 determinations. For routine purposes it means that this electrode can be continuously used for 5 hours.  相似文献   

14.
The aim of this study is to clarify the effect of doped metal type on CO2 reduction characteristics of TiO2 with NH3 and H2O. Cu and Pd have been selected as dopants for TiO2. In addition, the impact of molar ratio of CO2 to reductants NH3 and H2O has been investigated. A TiO2 photocatalyst was prepared by a sol-gel and dip-coating process, and then doped with Cu or Pd fine particles by using the pulse arc plasma gun method. The prepared Cu/TiO2 film and Pd/TiO2 film were characterized by SEM, EPMA, TEM, STEM, EDX, EDS and EELS. This study also has investigated the performance of CO2 reduction under the illumination condition of Xe lamp with or without ultraviolet (UV) light. As a result, it is revealed that the CO2 reduction performance with Cu/TiO2 under the illumination condition of Xe lamp with UV light is the highest when the molar ratio of CO2/NH3/H2O = 1:1:1 while that without UV light is the highest when the molar ratio of CO2/NH3/H2O = 1:0.5:0.5. It is revealed that the CO2 reduction performance of Pd/TiO2 is the highest for the molar ratio of CO2/NH3/H2O = 1:1:1 no matter the used Xe lamp was with or without UV light. The molar quantity of CO per unit weight of photocatalyst for Cu/TiO2 produced under the illumination condition of Xe lamp with UV light was 10.2 μmol/g, while that for Pd/TiO2 was 5.5 μmol/g. Meanwhile, the molar quantity of CO per unit weight of photocatalyst for Cu/TiO2 produced under the illumination condition of Xe lamp without UV light was 2.5 μmol/g, while that for Pd/TiO2 was 3.5 μmol/g. This study has concluded that Cu/TiO2 is superior to Pd/TiO2 from the viewpoint of the molar quantity of CO per unit weight of photocatalyst as well as the quantum efficiency.  相似文献   

15.
Acrylonitrile (AN) displaces the ethyl ether ligand of the cationic complex [Pd(N-N)Me(Et2O)]+ (N-N = (2,6-(i-Pr)2C6H3)-NCMeCMeN-(2,6-(i-Pr)2C6H3)) to form the N-bonded AN complex [Pd(N-N)Me(AN)]+, which exists as two interconverting rotamers. On standing or heating, [Pd(N-N)Me(AN)]+ undergoes 2,1-insertion to give [Pd(N-N)(CH(CN)CH2Me)(AN)]+, which undergoes β-hydrogen elimination to give the intermediate hydride, [Pd(N-N)H(AN)]+, which in turn inserts AN to give the cyanoethyl complex [Pd(N-N)(CH(CN)Me)]+. Dimerization of the [Pd(N-N)(CH(CN)CH2CH3)]+ moiety via bridging nitrile groups also occurs, giving the dicationic species . Although [Pd(N-N)Me(AN)]+ does behave as a typical Brookhart ethylene polymerization catalyst, it does not catalyze AN polymerization and in fact added AN suppresses ethylene polymerization.  相似文献   

16.
We revisit the singlet–triplet energy gap (ΔEST) of silicon trimer and evaluate the gaps of its derivatives by attachment of a cation (H+, Li+, Na+, and K+) using the wavefunction‐based methods including the composite G4, coupled‐cluster theory CCSD(T)/CBS, CCSDT and CCSDTQ, and CASSCF/CASPT2 (for Si3) computations. Both 1A1 and 3 states of Si3 are determined to be degenerate. An intersystem crossing between both states appears to be possible at a point having an apex bond angle of around α = 68 ± 2° which is 16 ± 4 kJ/mol above the ground state. The proton, Li+ and Na+ cations tend to favor the low‐spin state, whereas the K+ cation favors the high‐spin state. However, they do not modify significantly the ΔEST. The proton affinity of silicon trimer is determined as PA(Si3) = 830 ± 4 kJ/mol at 298 K. The metal cation affinities are also predicted to be LiCA(Si3) = 108 ± 8 kJ/mol, NaCA(Si3) = 79 ± 8 kJ/mol and KCA(Si3) = 44 ± 8 kJ/mol. The chemical bonding is probed using the electron localization function, and ring current analyses show that the singlet three‐membered ring Si3 is, at most, nonaromatic. Attachment of the proton and Li+ cation renders it anti‐aromatic. © 2015 Wiley Periodicals, Inc.  相似文献   

17.
In catalytic two-step n-butene oxidation with dioxygen to methyl ethyl ketone, the first step is the oxidation of n-C4H8 with an aqueous solution of Mo-V-P heteropoly acid in the presence of Pd(II) complexes. The kinetics of n-butene oxidation with solutions of H7PV4Mo8O40 (HPA-4) in the presence of the Pd(II) dipicolinate complex (H2O)PdII(dipic) (I), where dipic2− is the tridentate ligand 2,6-NC5H3(COO)2, is studied. Calculation shows that, at the ratio dipic2−: Pd(II) = 1: 1, the ligand decreases the redox potential of the Pd(II)/Pdmet system from 0.92 to 0.73–0.77, due to which Pd(II) is stabilized in reduced solutions of HPA-4. The reaction is first-order with respect to n-C4H8. Its order with respect to Pd(II) is slightly below unity, and its order with respect to HPA-4 is relatively low (∼0.63). The activation energy of but-1-ene oxidation in the temperature range from 40 to 80°C is 49.0 kJ/mol, and that of the oxidation of but-2-ene is 55.6 kJ/mol. The mechanism of the reaction involving the cis-diaqua complex [(H2O)2PdII(Hdipic)]+, which forms reversibly from complex I, is proposed. The reaction rate is shown to increase with an increase in the HPA-4 concentration due to an increase in the acidity of the solution.  相似文献   

18.
The facile and tunable preparation of unique dinuclear [(L?)Pd?X?Pd(L?)] complexes (X=Cl or N3), bearing a ligand radical on each Pd, is disclosed, as well as their magnetochemistry in solution and solid state is reported. Chloride abstraction from [PdCl( NNOISQ )] ( NNOISQ =iminosemiquinonato) with TlPF6 results in an unusual monochlorido‐bridged dinuclear open‐shell diradical species, [{Pd( NNO ISQ)}2(μ‐Cl)]+, with an unusually small Pd‐Cl‐Pd angle (ca. 93°, determined by X‐ray). This suggests an intramolecular d8–d8 interaction, which is supported by DFT calculations. SQUID measurements indicate moderate antiferromagnetic spin exchange between the two ligand radicals and an overall singlet ground state in the solid state. VT EPR spectroscopy shows a transient signal corresponding to a triplet state between 20 and 60 K. Complex 2 reacts with PPh3 to generate [Pd(NNOISQ)(PPh3)]+ and one equivalent of [PdCl( NNOISQ )]. Reacting an 1:1 mixture of [PdCl( NNOISQ )] and [Pd(N3)( NNOI SQ)] furnishes the 1,1‐azido‐bridged dinuclear diradical [{Pd( NNO ISQ)}21‐N;μ‐N3]+, with a Pd‐N‐Pd angle close to 127° (X‐ray). Magnetic and EPR measurements indicate two independent S=1/2 spin carriers and no magnetic interaction in the solid state. The two diradical species both show no spin exchange in solution, likely because of unhindered rotation around the Pd?X?Pd core. This work demonstrates that a single bridging atom can induce subtle and tunable changes in structural and magnetic properties of novel dinuclear Pd complexes featuring two ligand‐based radicals.  相似文献   

19.
IR and NMR data showed that the ionic complex Pd2(CHCC6H5)2(C5H7O2)3(BF3)2BF4 isolated in the reaction Pd(Acac)2 + PA + 5BF3OEt2 (Acac is C5H7O2, PA is phenylacetylene) is an adduct of two complexes, namely, (Acac)PdBF4 and [(PA)2Pd(C3-Acac · BF3)]+(Acac · BF3) (coordinatively unsaturated). On dissolution in deuteroacetone or deuteromethanol, the [(Acac)PdF2BF2Pd(C3-Acac · BF3)(PA)2]+(Acac · BF3) adduct decomposed to Pd(Acac)2, 2BF3 · L (L = (CD3)2CO, CD3OD) and the [L(PA)2Pd(C3-Acac]+BF4 complex.  相似文献   

20.
Substitution reactions of five monofunctional Pd(II) complexes, [Pd(terpy)Cl]+ (terpy = 2,2′;6′,2″-terpyridine), [Pd(bpma)Cl]+ (bpma = bis(2-pyridylmethyl)amine), [Pd(dien)Cl]+ (dien = diethylenetriamine or 1,5-diamino-3-azapentane), [Pd(Me4dien)Cl]+ (Me4dien = 1,1,7,7-tetramethyldiethylenetriamine), and [Pd(Et4dien)Cl]+ (Et4dien = 1,1,7,7-tetraethyldiethylenetriamine), with unsaturated N-heterocycles such as 3-amino-4-iodo-pyrazole (pzI), 5-amino-4-bromo-3-methyl-pyrazole (pzBr), 1,2,4-triazole, pyrazole, pyrazine, and imidazole were investigated in aqueous 0.10 M NaClO4 in the presence of 10 mM NaCl using variable-temperature stopped-flow spectrophotometry. The second-order rate constants k2 indicate that the reactivity of the Pd(II) complexes decrease in the order [Pd(terpy)Cl]+ > [Pd(bpma)Cl]+ > [Pd(dien)Cl]+ > [Pd(Me4dien)Cl]+ > [Pd(Et4dien)Cl]+. The most reactive nucleophile of the heterocycles is pyrazine, while the slowest reactivity is with pyrazole. Activation parameters were determined for all reactions and negative entropies of activation, ΔS, supporting an associative mode of substitution. The reactions between [Pd(bpma)Cl]+ and 1,2,4-triazole, pzI, and pzBr were also investigated by 1H NMR to define the manner of coordination. These results could be useful for better explanation of structure-reactivity relationships of Pd(II) complexes as well as for the prediction of potential targets of Pd(II) complexes toward common N-heterocycles, constituents of biomolecules and different N-bonding pharmaceutical agents.  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