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 46 毫秒
1.
The electrodeposition of three different heteropolyoxometalates:(HPOM)(NH4)3[RhMo6O18(OH)6]· 7 H2O, HPOM (1), Cu(II)(NH4)[RhMo6O18(OH)6]· 7 H2O, HPOM (2) and Cu(II)NH4[CoMo6O18(OH)6]· 7 H2O HPOM (3) was studied by in situ atomic force microscopy (AFM) and cyclic voltammetry. It was found that the voltammetric response of compounds 2 and 3 show the deposition of the copper countercation as well as the simultaneous reduction of the corresponding heteropolyanion (HPA). AFM was used to monitor the in situ film formation of the electroreduced species on the working electrode surface. The AFM images show important differences in the film texture when copper is present in the complex.  相似文献   

2.
The reaction of the Cu(II) bis N,O‐chelate‐complexes of L‐2,4‐diaminobutyric acid, L‐ornithine and L‐lysine {Cu[H2N–CH(COO)(CH2)nNH3]2}2+(Cl)2 (n = 2–4) with terephthaloyl dichloride or isophthaloyl dichloride gives the polymeric complexes {‐OC–C6H4–CO–NH–(CH2)n–CH(nh2)(COO)Cu(OOC)(NH2)CH–CH2)n–NH‐}x 1 – 5 . From these the metal can be removed by precipitation of Cu(II) with H2S. The liberated ω,ω′‐N,N′‐diterephthaloyl (or iso‐phthaloyl)‐diaminoacids 6 – 10 react with [Ru(cymene)Cl2]2, [Ru(C6Me6)Cl2]2, [Cp*RhCl2]2 or [Cp*IrCl2]2 to the ligand bridged bis‐amino acidate complexes [Ln(Cl)M–(OOC)(NH2)CH–(CH2)nNH–CO]2–C6H4 11 – 14 .  相似文献   

3.
The synthesis and characterisation of a novel [(η2-dppf)(η5-C5H5)Ru(CC)-1,4-(C6H4)PPh2–Au–CC-bipy({[Ti](μ-σ,π-CCSiMe3)2}Cu)]PF6 (dppf = 1,1′-bis(diphenylphosphino)ferrocene) is reported in which five different transition metals (Fe–Ru–Au–Cu–Ti) are linked by carbon-rich organic bridging units.  相似文献   

4.
Three dinuclear coordination complexes generated from 1-n-butyl-2-((5-methyl-1H-pyrazole-3-yl)methyl)-1H-benzimidazole ( L ), have been synthesized and characterized spectroscopically and structurally by single crystal X-ray diffraction analysis. Reaction with iron(II) chloride and then copper(II) nitrate led to a co-crystal containing 78 % of [Cu(NO3)(μ-Cl)( L’ )]2 ( C1 ) and 22 % of [Cu(NO3)(μ-NO3)( L’ )]2 ( C2 ), where L was oxidized to a new ligand L . A mechanism is provided. Reaction with copper chloride led to the dinuclear complex [Cu(Cl)(μ-Cl)( L) ]2 ( C3 ). The presence of N−H⋅⋅⋅O and C−H⋅⋅⋅O intermolecular interactions in the crystal structure of C1 and C2 , and C−H⋅⋅⋅N and C−H⋅⋅⋅Cl hydrogen bonding in the crystal structure of C3 led to supramolecular structures that were confirmed by Hirshfeld surface analysis. The ligands and their complexes were tested for free radical scavenging activity and ferric reducing antioxidant power. The complex C1 / C2 shows remarkable antioxidant activities as compared to the ligand L and reference compounds.  相似文献   

5.
Cu–SAPO-34 and CuY–SAPO-34 catalysts for NH3-SCR were prepared by the wet-impregnation method. XRD, UV–vis DRS, ESR and NH3-TPD results showed that the introduction of Y effectively improved the dispersion of copper species, increased the amount of isolated copper ions and enhanced the acid density. In addition, the activity test, NH3-TPD and TGA results reflected that the CuY–SAPO-34 catalyst showed better C3H6 oxidation activity, lower dropping degree of acid sites after C3H6/O2 treatment and less adsorption of C3H6/O2 than Cu–SAPO-34 catalyst. Therefore, the addition of Y promoted the NH3-SCR performance and the hydrocarbon (HC) resistance of Cu–SAPO-34 catalyst.  相似文献   

6.
Schiff base bis (2— quinolidene)- diamine gives a purple-red complex with copper ions with a λmax at 530 mμ. The optimum pH for production of this complex is approximately 9.5. None of the metallic ions of the analytical groups II and III give rise to red complexes with the above Schiff base. Manganous (Mn2+) and stannous (Sn2+)ions as well as reducing agents such as hydrazine or hydroxylamine exert a catalytic effect on the rate of formation of these copper complexes. The thermal stability of the copper complex formed from the Schiff bases (2-quinoline-aldehyde with NH2-(CH2)n-NH2) is greater for diamine for which n = 4–6 than for n = 2–3.  相似文献   

7.
Mononuclear and dinuclear palladacycles derived from 1,3-bis(N,N-dimethylaminomethyl)benzenes, [{Pd(Cl)}2,6-(Me2NCH2)2C6H3] (1) and [1-{Pd(H2O)(Py)}-5-{Pd(OTf)(Py)-2,4-(Me2NCH2)2C6H2]-(OTf) (2), were synthesized and their structures were fully characterized. Complex 1 is a pincer complex with η3-mer NCN phenyl backbone, complex 2 is a bispalladium(II) complex with 1,2- and 4,5- two C,N-ortho phenyl backbone. Whereas the pincer complex 1 acted as a poor catalyst on methanolysis of fenitrothion, complex 2 demonstrated high catalytic activity in the same reactions, but there is no synergetic effect between two palladium ions. The results clearly indicate that a dissociable co-ligand in the palladacycle compounds significantly promotes the catalytic methanolysis.  相似文献   

8.
The reaction of copper salts with AlPO4-5 or Vv-VAPO-5 under acidic (CuCl2, pH adjusted to 2) but especially basic conditions (Cu(NH3) 4 2+ , pH adjusted to 9) gives ion incorporations greater than expected by a simple ion exchange mechanism (both AlPO4-5 and Vv-VAPO-5 could be expected to have no cation exchange capacity). Ion incorporation is proposed to occur initially at defect sites, and examination of the ESR spectrum of a dehydrated, evacuated CuCl2-exchanged AlPO4-5 shows that these defect sites give rise to a number of unique environments upon CuII incorporation. The CuCl2-exchanged VAPO-5 retains a significant toluene accessibility to the Vv sites in the VAPO-5. However, the toluene accessibility in the Cu(NH3) 4 2+ -exchanged VAPO-5 is significantly reduced and we propose this is due to a combination of the presence of crystalline CuO and structural collapse from reaction with base (NH4OH). The ability of treatment with base (NH4OH, pH 13) to restrict access of toluene to the Vv sites of the original VAPO-5 was verified in a separate experiment.  相似文献   

9.
The kinetics of the ligand exchange reaction of the Cu(II)-ammine complex with poly(vinyl alcohol) (PVA) has been studied by a stopped-flow method at pH 9–10, at μ=0.1 (NH4Cl) and at 25°C. The reaction is initiated by the formation of unstable [Cu(NH3)3]2+ by the attack of H+ on Cu(II)-ammine complex, and proceeds through the mixed complex {[Cu(NH3)3(O?PVA)]2+}. This step may be rate-determining, followed by a rapid reaction. Finally, the Cu(II) ion is taken up by PVA. The rate is given by d[Cu(II)?PVA]/dt=k[H+]{[Cu(NH3)4]2+}[PVA]/[NH4Cl], where k=k1 + k2[H+], k1=4.25× 10s?1 and k2=5.20× 1011l mol?1s?1.  相似文献   

10.
The complexes, tetra-μ-[2-(phenylamino)benzoato](O,O′)-bis[(ethanol)copper(II)] (1) and di-μ-[2-(phenylamino)benzoato](O,O′)-bis[(hydroxo)copper(II)] (2), were synthesized by the reaction of N-phenylanthranilic acid and CuCl2·2H2O in an ethanol water mixture. In complex 1, each Cu(II) atom, which is in a slightly distorted square pyramidal environment, is coordinated equatorially by four N-phenylanthranilate O-atoms and axially by the ethanol O-atoms. In complex 2, [Cu2(C6H5NHC6H4COO)2(OH)2], each Cu(II) atom, which is in tetrahedral environment, is coordinated by two N-phenylanthranilate O-atoms and hydroxo ligands. The crystal structure (monoclinic, P21/c space group) of complex 1 comprises a dinuclear [Cu2(C6H5NHC6H4COO)4(CH3CH2OH)2] species and the dimer is located on a crystallographic inversion centre. The Cu(II) ions, 2.591(2) Å apart, are bridged by the carboxylate groups of four N-phenylanthranilate ligands. The complex molecules show three-dimensional supramolecular networks by O–H···O, C–H···O and C–H···π interactions.  相似文献   

11.

Aluminum citrate is used as a conformance control agent to improve oil production and excess water production. This paper discusses the formation of mono and polynuclear aluminum species from the synthesis of aluminum citrate and evaluates these compounds as crosslinkers in hydrogels for conformance control. The products obtained from the synthesis were characterized by Fourier-transform infrared spectrometry (FTIR), elemental analysis (CHN), scanning electron microscopy (SEM), and inductively coupled plasma–optical emission spectrometry (ICP-OES). The FTIR analyses indicated the presence of mononuclear aluminum citrate complexes at pH 3 and polynuclear species starting at pH 4. These results were corroborated by CHN and ICP-OES techniques, which revealed the variation of carbon, oxygen, hydrogen, and alumina precipitate levels as functions of pH variation. The focus of the study was to assess how these crosslinking agents perform in hydrogel formation under reservoir conditions. Rheological analysis showed that the values of tan (delta) of the hydrogel synthesized with aluminum citrate at pH 6 were lower than 0.1, indicating strong gels, while at pH 9, the values were above 0.1, indicating weak gels. These results are in agreement with those obtained by FTIR, which showed that at pH 6, the structures of the aluminum citrate complex were probably in the form [Al3(C6H5O7)3(OH)4(H2O)]4−. This structure appears to allow easier access to the aluminum orbital for the crosslinking process compared to the gel composed of aluminum citrate synthesized at pH 9 [Al3(C6H6O7)3(OH)4(H2O)5]4−.

  相似文献   

12.
In situ high‐pressure NMR spectroscopy of the hydrogenation of benzene to give cyclohexane, catalysed by the cluster cation [(η6‐C6H6) (η6‐C6Me6)2Ru33‐O)(μ2‐OH)(μ2‐H)2]+ 2 , supports a mechanism involving a supramolecular host‐guest complex of the substrate molecule in the hydrophobic pocket of the intact cluster molecule.  相似文献   

13.
The reaction of the PCP-type complex Pd(Me){2,6-(iPr2PCH2)2C6H3}( 3 ) with phenyl iodide results in the formation of Pd(I){2,6-(iPr2PCH2)2C6H3} ( 5 ), methyl iodide, toluene, and biphenyl. Formation of Pd(Ph){2,6-(iPr2PCH2)2C6H3}( 4 ) is observed during the reaction by 31P NMR. Reaction of 4 with aryl iodides results in the formation of 5 and Ph–Ph, Ph–Ar, and Ar–Ar, products indicative of a radical reaction. Under pseudo-first-order conditions, the rates of the reactions follow the order p-OMe > p-Me > H > p-NO2 > m-Cl. The reaction is likely to involve electron transfer from 4 to the aryl iodide followed by fast decomposition of a postulated radical cation [Pd(Ph){2,6-(iPr2PCH2)2C6H3}]+. ( 4 +.) to give a phenyl radical and [Pd{2,6-(iPr2PCH2)2C6H3}]+ ( 6 +). Facile decomposition of the aryl iodide radical anion generates an aryl radical and I. Recombination of aryl radicals gives rise to mixed biaryls, and 6 + combines with I to give 5 .  相似文献   

14.
The hemiketal complex [Cu{NH2C(Me)2C(Ph)(OMe)O}2] (1) was generated by the reaction of [Cu(NCMe)4](BF4) and 4 equivs of 2,2-dimethyl-3-phenyl-2H-azirine in the presence of NCNMe2 in wet MeOH and isolated in 92% yield. In 1, the in situ formed hemiketal NH2C(Me)2C(Ph)(OMe)OH in its deprotonated form is stabilized due to chelation to copper(II). In the X-ray structure of 14MeOH, the intermolecular hydrogen bonding was detected between O and N atoms of the organic ligand and the HO group of solvated MeOH as well as between two molecules of the solvated MeOH. Three types of hydrogen bonds in the obtained structure were studied by the DFT calculations (M06/6-31 ++G** level of theory, MDF10 pseudopotentials on the Cu atoms) and topological analysis of the electron density distribution within the formalism of Bader's theory (QTAIM method). Estimated strength of these non-covalent interactions is 3–9 kcal/mol.  相似文献   

15.
Viscosity measurements under Newtonian flow conditions had been performed on cetyltrimethylammonium bromide (CTAB) aqueous solutions in the combined presence of sodium salts of aromatic acids (sodium salicylate, NaSal; sodium benzoate, NaBen; sodium anthranilate, NaAn) and organic additives (1-hexanol, C6OH; n-hexylamine, C6NH2) at 30°C. On addition of C6OH or C6NH2, the viscosity of 25 mM CTAB solution remained nearly constant without salt as well as with a lower salt concentration. This is due to low CTAB concentration which is not sufficient to produce structural changes in this concentration range of salts. However, as the salt concentration was increased further, the effect of C6OH/C6NH2 addition was different with different salts: The viscosity first increased; then a decrease was observed with the former while with C6NH2 a decrease followed by constancy appeared in plots of relative viscosities (η r ) vs. organic additive concentrations. At further higher salt concentration, the magnitude of η r was much higher. The viscosity increase is explained in terms of micellar growth and the decrease in terms of swollen micelle formation (due to interior solubilization of organic additive) or micellar disintegration (due to formation of water + additive pseudophase).  相似文献   

16.
It has been shown that one-pot reaction of copper powder, ammonium tris-oxalatoferrate(III) (NH4)3[Fe(C2O4)3]·3H2O as a source of metalloligand, and ethylenediamine (en) leads to the formation of heterobimetallic complex (NH4)[Cu(en)2Fe(C2O4)3]?2dmso (1). This complex has been characterised by elemental analysis, IR spectroscopy and single crystal X-ray crystallography. Crystal structure analysis reveals that 1 consists of [Cu(en)2Fe(C2O4)3]? anionic chains with regularly alternated [Cu(en)2]2+ and [Fe(C2O4)3]3? moieties. Magnetic susceptibility measurements indicate absence of a significant exchange interaction between the metal units. The polycrystalline X-band EPR spectrum of 1 exhibits broad line characteristic of Fe(III) centers, whereas the signals observed in EPR spectrum of frozen dmf solution of 1 are distinctly associated with zero field splitting in the spin states of rhombically distorted Fe(III) centers, non interacting with Cu(II) centers.  相似文献   

17.
The decavanadate with a novel glycine–glycinato complex of copper (II) in the cationic part, (NH4)2[Cu2(H2O)4(NH3CH2COO)2(NH2CH2COO)2]H2V10O28·6H2O (1), has been prepared and characterized by elemental analysis, infrared and EPR spectroscopies. The triplet X band EPR spectrum of powdered sample evidenced magnetic interaction between the Cu(II) atoms in the dimeric unit which is probably realized through the bridging water molecules in this part of the complex. A single crystal X-ray diffraction study revealed that the structure of 1 contains cationic copper complex with a rare Cu(H2O)2Cu double bridge. In this cation, a simultaneous bidentate N, O– and monodentate O– coordination of glycine to the same central atom was observed for the first time.  相似文献   

18.
The activation of CO2 by chemical, electrochemical, and photochemical means is discussed. Binuclear transition metal complexes mediate oxygen atom transfers from CO2 by three distinct chemical pathways: (i) deoxygenation of CO2, (ii) multiple bond metathesis, and (iii) disproportionation. The complex Ir2 (μ-CNR)2(CNR)2(dmpm),(dmpm = bis(dimethylphosphino)methane) undergoes double cycloaddition of CO2 to its μ-CNR ligands. A subsequent reaction produces the bis(carbamoyl) complex [Ir2(μ-CO)(μ-H)(CONHR)2(CNR)2(dmpm)2]Cl. Isotope labelling studies show that the μ-CO ligand results from net deoxygenation of CO2. In contrast, the binuclear nickel complex Ni2(μ-CNMe)(CNMe)2(dppm)2 (dppm = bis-(diphenylphosphino)methane) reacts with liquid CO2 to give the tricarbonyl complex Ni2(μ-CO)(CO)2(dppm)2. Isotope labelling indicates that the carbonyl ligands are not derived from CO2 deoxygenation, but from C/CO triple bond metathesis. The reaction of CO2 with the Ir(0) complex Ir2(CO)3(dmpm)2 leads to CO2 disproportionation by formation of the carbonate, Ir2(CO3)(CO)2(dmpm)2, and tetracarbonyl, Ir2(CO)4(dmpm)2, complexes. The complex Ir2(CO3)(CO)2 (dmpm)2 undergoes reversible O-atom transfers from its carbonate ligand. The electrochemical activation of CO2 by the binuclear Ni2(μ-CNMe)(CNMe)2(dppm)2 and trinuclear [Ni3(μ-CNMe)(μ-I)(dppm)3][PF6] species is described. The triangular nickel complex [Ni33-CNMe)(μ3-I)(dppm)3][PF6] is an electrocatalyst for the reduction of CO2. The cluster exhibits a reversible single electron reduction at E0(+/0) = −1.09 V vs. Ag/AgCl. In the presence of CO2, the cluster reduces CO2 by an EC' electrochemical mechanism. The reduction products correspond to the disproportionation and H-atom abstraction products of CO2*−, with a partitioning ratio of 10:1. Isotope labelling studies with 13CO2 indicate that 13CO2*− disproportionation produces 13CO and 13CO32−. Studies of the photochemical activation of CO2 by Ni2(μ-CNMe)(CNMe)2(dppm)2 are described. The bimolecular photochemical addition of CO2 to the complex was examined by laser transient absorbance spectroscopy. Photolysis at 355nm in the presence of CO2 (1 atm) leads to cycloaddition of CO2 to the μ-CNMe ligand and the complex Ni2(μ-CN(Me)C(O)O)(CNMe)2(dppm)2 with Φ355=0.05. The triplet excited state of Ni,(μ-CNMe)(CNMe)2(dppm)2 was determined to react with CO2 with the bimolecular reaction rate constant k = 1 × 104 M−1 s−1. Bridging ligand substituent effects and solvent dependence of the lowest energy electronic absorption spectral bands of the series of complexes, Ni2(μ-L)(CNMe)2(dppm)2, L = CNMe, CNC6H5, CN-p-C6H4Cl. and CN-p-C6H4Me, confirm the assignment of di-metal to bridging ligand charge transfer (M2→μ-LCT). This assignment is supported by results of extended Hückel calculations which indicate a LUMO of predominantly μ-isocyanide π* character. A systematic study of the nature of the lowest excited states of related d10–d10 binuclear complexes of the type Ni2(μ-L)(CNMe)2(dppm)2, where L = CNMe(Ph)+, CNMe2+, CNMe(C5H11)+, CNMe(H)+, CNMe(CH2C6H5)+, and NO+ reveals dramatic differences in the lowest excited states of the three classes of complexes: μ-isocyanide, μ-aminocarbyne, and μ-nitrosyl. Spectroscopic and extended Hückel MO studies confirm that the μ-isocyanide complexes are characterized by di-metal to bridging ligand charge transfer (M2 → μ-LCT) excited states. However, the μ-aminocarbyne and μ-nitrosyl complexes exhibit bridging ligand to metal charge transfer (μ-L→M2) and intraligand (IL) lowest excited states, respectively.  相似文献   

19.
The deposition of diamond on a molybdenum substrate was studied in an Ar–H2 plasma jet adding phenol (C6H5OH) as a carbon source and was compared with adding benzene (C6H6) to examine the effect of OH species on diamond deposition. Better faceted and larger crystal size diamond was deposited from the Ar–C6H5OH–H2 plasma jet than from the Ar–C6H6–H2 plasma jet. Furthermore, the amount of co-deposited graphite and/or amorphous carbon in the deposit from the Ar–C6H5OH–H2 plasma jet was smaller than that in the deposit from the Ar–C6H6–H2 plasma jet. At the beginning of the exposure to the Ar–C6H5OH–H2 plasma jet, in which OH radicals were identified, the surface of the substrate was slightly oxidized. Oxide formation on the surface of the substrate by reaction with OH radicals would contribute to the purification and increase of crystal size in the diamond deposition with an Ar–C6H5OH–H2 plasma jet.  相似文献   

20.
The effect of the nature of the anion on the performance of ionic rhodium catalysts has received little attention. Herein it is shown that the use of highly fluorous tetraphenylborate anions can enhance catalyst activity in both conventional and fluorous media. For hydrogenation catalysts of the type [Rh(COD)(dppb)][X] {COD=1,5‐cis,cis‐cyclooctadiene; dppb=1,4‐bis(diphenylphosphino)butane; X=BF4 ( 1a ), [BPh4] ( 1b ), [B{C6H4(SiMe3)‐4}4] ( 1c ), [B{C6H3(CF3)2‐3,5}4] ( 1d ), [B{C6H4(SiMe2CH2CH2C6F13)‐4}4] ( 1e ), [B{C6H4(C6F13)‐4}4] ( 1f ) and [B{C6H3(C6F13)2‐3,5}4] ( 1 g )} the activity towards the hydrogenation of 1‐octene in acetone increased in the order 1c < 1b < 1e < 1a < 1d ~ 1f < 1g with 1g being twice as active as the commonly applied 1a . Despite the fluorophilic character introduced by the substituted tetraarylborate anions, the presence of some perfluoroalkyl‐substituents in the cation was still required for achieving high partition coefficients. Therefore, [Rh(COD)(Ar2PCH2CH2PAr2)][X] {Ar=C6H4(SiMe2CH2CH2C6F13)‐4, X=[B{C6H3(C6F13)2‐3,5}4] ( 3f ); Ar=C6H4(SiMe(CH2CH2C6F13)2)‐4 and X=[B{C6H4(C6F13)‐4}4] ( 2g )} were prepared, which were active in the hydrogenation of 1‐octene, 2g even more so than 3f . Both these highly fluorous catalysts could be recycled with 99% efficiency through fluorous biphasic separation, whereas the corresponding BF4 complex of 2g ( 2a ) did not show any affinity for the fluorous phase.  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