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 31 毫秒
1.
A new asymmetric oligobipyridine ligand, 1- (5’-methyl-2, 2’-bipyridin-5-y1)-2- (6’-methyl-2, 2’-bipyridin-6-yl)ethane (L), in which the bipyridine units are bridged by CH2CH2 at 5,6’-position has been synthesized. The ligand L reacts with Cu(I) and Cu(I1) ions giving double-stranded helical complexes [Cu 2 1 L2](C104)2.Et20 (1) and [Cu 2 II L2,(OH)(H20) ] [ClO4]3(2), respectively. Complexes 1 and 2 were characterized by X-ray diffraction analyses, ES-MS, ESR and cyclic voltammetry, etc. Differing from the oligobipyridine ligands bridged by CH2CH2 at 6,6’-or 5,5’-position, the ligand L not only forms a double-stranded helicate with Cu(1) ion, but also gives a double-stranded helicate with Cu(I1) ion. The results show that the linkage mode of the spacer group to the bipyridine units exerts a great impact on the formation of helix. Project supported by the National Natural Science Foundation of China (Grant NO. 29601003).  相似文献   

2.
采用基于密度泛函理论的第一性原理方法和平板模型研究了CH3SH分子在Cu(111)表面的吸附反应.系统地计算了S原子在不同位置以不同方式吸附的一系列构型, 第一次得到未解离的CH3SH分子在Cu(111)表面顶位上的稳定吸附构型,该构型吸附属于弱的化学吸附, 吸附能为0.39 eV. 计算同时发现在热力学上解离结构比未解离结构更加稳定. 解离的CH3S吸附在桥位和中空位之间, 吸附能为0.75-0.77 eV. 计算分析了未解离吸附到解离吸附的两条反应路径, 最小能量路径的能垒为0.57 eV. 计算结果还表明S―H键断裂后的H原子并不是以H2分子的形式从表面解吸附而是以与表面成键的形式存在. 通过比较S原子在独立的CH3SH分子和吸附状态下的局域态密度, 发现S―H键断裂后S原子和表面的键合强于未断裂时S原子和表面的键合.  相似文献   

3.
The kinetics of isothermal decomposition of Cu(CH2CH2COO)2 were studied at 483–503 K. The end-product was identified as CuO by X-ray diffraction and chemical analysis. The kinetics follow the Prout-Tompkins equation with an activation energy of 191 ± 10 kJ/moIe. The activation energies and the order of reaction were also evaluated from analysis of the DTG, DTA and TG curves of the sample.  相似文献   

4.
Reaction of (NBu4)2[Mo2O7] with [Cu(CH3CN)4](PF6) in acetonitrile results in isolation of the orange β-octamolybdate [Cu(CH3CN)4]2[Mo8O26Cu2(CH3CN)4] (1) along with the colourless α-octamolybdate [Cu(CH3CN)4]4[Mo8O26]·2CH3CN (2). Both products decompose rapidly upon removal from their mother liquors, forming an insoluble, dark brown coloured phase with the composition Cu4[Mo8O26]·0.6CH3CN·16H2O (3). The copper(I) acetonitrile derivatised isopolyanion in 1 thus represents an intermediate structure between the simple, underivatised octamolybdate 2 and fully condensed, polymeric phase 3.  相似文献   

5.
Oxidized copper surfaces have attracted significant attention in recent years due to their unique catalytic properties, including their enhanced hydrocarbon selectivity during the electrochemical reduction of CO2. Although oxygen plasma has been used to create highly active copper oxide electrodes for CO2RR, how such treatment alters the copper surface is still poorly understood. Here, we study the oxidation of Cu(100) and Cu(111) surfaces by sequential exposure to a low-pressure oxygen plasma at room temperature. We used scanning tunnelling microscopy (STM), low energy electron microscopy (LEEM), X-ray photoelectron spectroscopy (XPS), near edge X-ray absorption fine structure spectroscopy (NEXAFS) and low energy electron diffraction (LEED) for the comprehensive characterization of the resulting oxide films. O2-plasma exposure initially induces the growth of 3-dimensional oxide islands surrounded by an O-covered Cu surface. With ongoing plasma exposure, the islands coalesce and form a closed oxide film. Utilizing spectroscopy, we traced the evolution of metallic Cu, Cu2O and CuO species upon oxygen plasma exposure and found a dependence of the surface structure and chemical state on the substrate''s orientation. On Cu(100) the oxide islands grow with a lower rate than on the (111) surface. Furthermore, while on Cu(100) only Cu2O is formed during the initial growth phase, both Cu2O and CuO species are simultaneously generated on Cu(111). Finally, prolonged oxygen plasma exposure results in a sandwiched film structure with CuO at the surface and Cu2O at the interface to the metallic support. A stable CuO(111) surface orientation is identified in both cases, aligned to the Cu(111) support, but with two coexisting rotational domains on Cu(100). These findings illustrate the possibility of tailoring the oxidation state, structure and morphology of metallic surfaces for a wide range of applications through oxygen plasma treatments.

A low-pressure oxygen plasma oxidized Cu(100) and Cu(111) surfaces at room temperature. The time-dependent evolution of surface structure and chemical composition is reported in detail for a range of exposure times up to 30 min.  相似文献   

6.
Copper(II) complexes of the general composition Cu(ligand)2X2 (where X = Cl, Br, NO3, ClO4, and 12SO4) and Cu(ligand)(CH3COO)2 have been synthesised with benzymethylketonesemicarbazone. All the complexes prepared have been characterised by elemental analysis, magnetic moment, conductance, IR, electronic and electron spin resonance spectral studies. The complexes Cu(ligand)2X2 (X = Cl, Br, NO3) and Cu(ligand)(CH3COO)2 may have tetragonal symmetry while the Cu(ligand)2X2 (ClO4 and 12SO4) may be five-coordinate trigonal bipyramidal in structure.  相似文献   

7.
Structure and Thermal Decomposition of Bis(triethanolamine)copper(II) Acetate [Cu{N(CH2CH2OH)3}2](CH3COO)2 Bis(triethanolamine)copper(II) acetate [Cu{N · (CH2CH2OH)3}2](CH3COO)2 was prepared using the basic components; the structure was determined by single crystal X-ray diffraction. The complex crystallizes in the monoclinic space group P21/c with a = 9.101 Å, b = 13.136 Å, c = 9.819 Å, β = 111.63°. Details of the synthesis, X-ray data, and the thermal decomposition are reported.  相似文献   

8.
The kinetics of the direct synthesis reaction (Si + 2CH3Cl → (CH3)2SiCl2) were measured on a Cu3Si alloy containing 1.2 atom % Zn. Reaction was carried out in a differential reactor (520–595 K, 1 atm) attached to an ultrahigh vacuum (UHV) system. Auger spectroscopy was used to characterize the surface before and after reaction. Zinc does not significantly change the overall rate of reaction, but it changes selectivity to dimethyldichlorosilane (the desired product), surface composition, activation energies, and induction times. The rate of silicon diffusion to the surface is not limiting in the presence of zinc. Zinc is found to be a promoter for improved selectivity only in low concentrations, and only a fraction of the surface appears to be active for reaction. The kinetics appear relatively insensitive to the surface composition or the form of surface carbon. A Cu3Si surface with Zn is shown to be a good model catalyst for the direct synthesis reaction.  相似文献   

9.
The coordination capability of the new tetrapyridyl pendant-armed azamacrocyclic ligand L, towards Co(II), Ni(II), Cu(II) and Cd(II) ions was studied. The ligand and the complexes were characterized by microanalysis, LSI mass spectrometry, IR, UV-Vis and NMR spectroscopy, magnetic studies and conductivity measurements. Crystal structures of [Co2L(CH3CN)2](ClO4)4·2CH3CN and [Cd2L(NO3)2](NO3)2·2H2O complexes have been determined. The X-ray studies show the presence of dinuclear endomacrocyclic complexes with the metal ion in a similar distorted octahedral environment, coordinated by one pyridyl bridgehead group, two amine nitrogen atoms and two pyridyl pendant-arms. The sixth coordination position around the metal ion is completed by one acetonitrile molecule in [Co2L(CH3CN)2](ClO4)4·2CH3CN and by one monodentate nitrate anion in [Cd2L(NO3)2](NO3)2·2H2O. Different sort of intramolecular non-classical hydrogen bonds were found in the crystal lattice of both structures.  相似文献   

10.
A series of bis(trifluoromethyl)carbinyl acrylate monomers [Y-C(CF3)2 O? CO? CH?CH2] in which Y is CH3, CH3CH2, CH3CH2CH2, CH3CH2CH2CH2, C6H5, H, F, CF3, N3, CN, and CH3OCH2CH2O, was prepared. Polymers were easily prepared from all of these monomers except where Y = CN, wherein a variety of initiation methods failed to produce high molecular polymer. Wettabilities of the polymer films were examined by means of contact angle measurements by using n-alkane test liquids and water. Values of the dispersion force contribution (γsd) and the polar force contribution (γsp) to the solid surface energy were calculated by employing both geometric and harmonic mean approximations. Values of γsd calculated by either method agreed well with γc (critical surface tension) values determined graphically from contact angle data employing n-alkane test liquids, confirming the suggestion that γc is an approximate measure of the dispersion force contribution to solid surface energy. Values of γsd ranged from 15 dyne/cm (Y = F or CF3) to 25 dyne/cm (Y = C6H5). Values of the polar force contribution to solid surface energy (γsp) varied from 0.6 dyne/cm (Y = CH3CH2CH2CH2) to 3.4 dyne/cm (Y = CH3OCH2CH2O) when calculated by the geometric mean equation. The values of γsp obtained from the harmonic mean equation followed the same trend upon varying substituents, but were larger in value, ranging from 2.9 dyne/cm (Y = CH3CH2CH2CH2) to 7.5 dyne/cm (Y γ CH3OCH2CH2O).  相似文献   

11.
In the presence of copper(II) ion, two asymmetrical tripodal tetraamine ligands N{(CH2)3NH2}{(CH2)2NH2}2 (pee), N{(CH2)3NH2}2{(CH2)2NH2} (ppe) and one symmetrical ligand, N{(CH2)3NH2}3 (tpt), were condensed with 2-acetylpyridine. In EtOH–H2O solutions the reaction stops after the first condensation stage, and complexes of acyclic pentadentate(N5) mono-Schiff-base ligands were obtained. With asymmetrical tetraamines there are two possible condensation sites: the primary amine of the propylene, or the ethylene chain. The X-ray structure analysis of one complex, [Cu(ppe-py)](ClO4)2, indicates that condensation with 2-acetylpyridine in this case occurs at the propylene chain and the geometry around the copper ion is trigonal-bipyramidal.  相似文献   

12.
Triethylphosphanimine Complexes of the Acetates of Copper(II) and Zinc. Crystal Structures of [Zn(O2C–CH3)2(HNPEt3)], [Cu5(O2C–CH3)10(HNPEt3)2], and [Cu(O2C–CH3)2(HNPEt3)2] The title compounds originate from the anhydrous acetates of zinc and copper(II) with trimethylsilyl-triethylphosphanimine, Me3SiNPEt3, in the presence of water in dichloromethane. They form colourless ( 1 ), bluish-green ( 2 ), and blue ( 3 ), respectively, single crystals, which were characterized by IR spectroscopy and by crystal structure analyses. [Zn(O2C–CH3)2(HNPEt3)] ( 1 ): Space group P 4 21c, Z = 8, lattice dimensions at –83 °C: a = b = 1709.6(2), c = 982.4(1) pm, R = 0.0551. 1 has a polymeric chain structure in which the zinc atoms are μ2-bridged via the oxygen atoms of one of the two acetato groups, while the second acetato group and the phosphanimine are bonded terminally. [Cu5(O2C–CH3)10(HNPEt3)2]( 2 · 4 CH2Cl2): Space group P21/c, Z = 8, lattice dimensions at –80 °C: a = 1761.18(13), b = 4074.5(2), c = 1733.34(15) pm, β = 91.383(10)°, R = 0.0413. 2 consists of the two structural units [Cu2(O2C–CH3)4] and [Cu3(O2C–CH3)6(HNPEt3)2], which are connected via two of the acetato groups of the Cu3-unit along the crystallographic a-axis to form three crystallographically independent polymeric strands. [Cu(O2C–CH3)2(HNPEt3)2] ( 3 ): Space group P21/n, Z = 2, lattice dimensions at 20 °C: a = 695.49(8), b = 1217.85(10), c = 1380.05(7) pm, β = 96.451(7)°, R = 0.0291. 3 forms monomeric, centrosymmetric molecules with a square planar environment at the Cu atoms.  相似文献   

13.
The structures of [Cu (S2CN (CH2)4)2] (1) and [Zn2(S2CN‐(CH2)4)4] (2) have been determined by X‐ray crystallography analysis. They are all isomorphous and triclinic, space group of P1?, with Z = 1. The lattice parameters of compound 1 is: a = 0.63483(2) nm, b = 0.74972(3) nm, c=0.78390(1) mn, α = 75.912(2)°, β = 78.634(2)° and γ = 86.845(2)°; compound 2: a = 0.78707(6) nm, b=0.79823(6) nm, c = 1.23246(9) nm, α = 74.813(2)°, β = 73.048(2)° and γ = 88.036(2)°. The copper atom is located on a crystallographic inversion center and zinc atom lies across centers of symmetry. The Cu(II) ion has a square‐planar geometry while Zn(II) has a distorted tetrahedral geometry. The thermal gravity (TG) data indicate that no structural transitions in the two compounds were abserved and the decomposition products can adsorb gas. Also they all have a high thermal stability.  相似文献   

14.
The mechanisms of methanol (CH3OH) oxidation on the PtPd(111) alloy surface were systematically investigated by using density functional theory calculations. The energies of all the involved species were analyzed. The results indicated that with the removal of H atoms from adsorbates on PtPd(111) surface, the adsorption energies of (i) CH3OH, CH2OH, CHOH, and COH increased linearly, while those of (ii) CH3OH, CH3O, CH2O, CHO, and CO exhibited odd‐even oscillation. On PtPd(111) surface, CH3OH underwent the preferred initial C H bond scission followed by successive dehydrogenation and then CHO oxidation, that is, CH3OH → CH2OH → CHOH → CHO → CHOOH → COOH → CO2. Importantly, the rate‐determining step of CH3OH oxidation was found to switch from CO → CO2 on Pt(111) to COOH → CO2 + H on PtPd(111) with a lower energy barrier of 0.96 eV. Moreover, water also decomposed into OH more easily on PtPd(111) than on Pt(111). The calculated results indicate that alloying Pt with Pd could efficiently improve its catalytic performance for CH3OH oxidation through altering the primary pathways from the CO path on pure Pt to the non‐CO path on PtPd(111).  相似文献   

15.
2-Methyl-2-propoxypropyl isocyanide [CNCH2C(OPr)(CH3)2] and its Cu(I) tetrafluoroborate complex [Cu{CNCH2C(OPr)(CH3)2}4]BF4 were prepared in good yields and purities. Both compounds were fully characterized by IR and NMR spectroscopy.  相似文献   

16.
Cu2O is an attractive catalyst for the selective reduction of CO2 to methanol. However, the mechanism of the reaction and the role of the Cu species in different oxidation states are not well understood yet. In this work, by first-principles calculations, we investigate the mechanism of the reaction on the Cu2O(110) surface, which is the most selective for methanol, in different degrees of reduction: ideal surface, slightly reduced surface (SRS), and partially reduced surface (PRS). The most favorable reaction pathways on the three surfaces were identified. We found that Cu(I) on the ideal surface is not capable of chemisorbing CO2, but surface oxygen serves as the active site which selectively converts CO2 to CH3OH with a limiting potential of −0.77 V. The Cu(0) on the SRS and PRS promotes the adsorption and reduction of CO2, while the removal of the residue O* becomes potential/rate limiting with a more negative limiting potential than the ideal surface. The SRS is selective to methanol while the PRS becomes selective to methane. The result suggests that the key to high methanol selectivity is to avoid the reduction of Cu(I), which provides a new strategy for the design of more efficient catalysts for selective CO2 reduction to methanol.  相似文献   

17.
N-(2-Hydroxybenzyl)aminopyridines (Li) react with Cu(II) and Pd(II) ions to form complexes in the compositions Cu(Li)2(CH3COO)2 · nH2O (n = 0, 2, 4), Pd(Li)2Cl2 · nC2H5OH (n = 0, 2) and Pd(L2)2Cl2 · 2H2O. In the complexes, the ligands are neutral and monodentate which coordinate through pyridinic nitrogen. Crystal data of the complexes obtained from 2-amino pyridine derivative have pointed such a coordinating route and comparison of the spectral data suggests the validity of similar complexation modes of other analog ligands. Cu(II) complex of N-(2-hydroxybenzyl)-2-aminopyridine (L1), [Cu(L1)2(CH3COO)2] has slightly distorted square planar cis-mononuclear structure which is built by two oxygen atoms of two monodentate carboxylic groups disposed in cis-position and two nitrogen atoms of two pyridine rings. The remaining two oxygen atoms of two carboxylic groups form two Cu and H bridges containing cycles which joint at same four coordinated copper(II) ion. IR and electronic spectral data and the magnetic moments as well as the thermogravimetric analyses also specify on mononuclear octahedric structure of complexes [Cu(L2)2(CH3COO)2 · 2H2O] and [Cu(L3)2(CH3COO)2 · 4H2O] where L2 and L3 are N-(2-hydroxybenzyl)-2- or 3-aminopyridines, respectively.  相似文献   

18.
The adsorption and decomposition of trimethylgallium (Ga(CH3)3, TMG) on Pd(111) and the effect of pre-covered H and O were studied by temperature programmed desorption spectroscopy and X-ray photoelectron spectroscopy. TMG adsorbs dissociatively at 140 K and the surface is covered by a mixture of Ga(CH3)x (x=1, 2 or 3) and CHx(a) (x=1, 2 or 3) species. During the heating process, the decomposition of Ga(CH3)3 on clean Pd(111) follows a progressive Ga-C bond cleavage process with CH4 and H2 as the desorption products. The desorption of Ga-containing molecules (probably GaCH3) is also identi ed in the temperature range of 275-325 K. At higher annealing temperature, carbon deposits and metallic Ga are left on the surface and start to di use into the bulk of the substrate. The presence of precovered H(a) and O(a) has a signi cant effect on the adsorption and decomposition behavior of TMG. When the surface is pre-covered by saturated H2, CH4, and H2 desorptions are mainly observed at 315 K, which is ascribed to the dissociation of GaCH3 intermediate. In the case of O-precovered surface, the dissociation mostly occurs at 258 K, of which a Pd-O-Ga(CH3)2 structure is assumed to be the precusor. The presented results may provide some insights into the mechanism of surface reaction during the lm deposition by using trimethylgallium as precursor.  相似文献   

19.
The complex of empirical formula CuI[(CH3)2(CH2)2P] has been found to be dimeric with two digonal copper(I) atoms bridged by two ligand molecules through coppercarbon σ-bonds, resulting in a centrosymmetric eight-membered heterocycle.  相似文献   

20.
The polymeric compounds [{Cu2I2(C6H5CN)2[cyclo‐(CH3AsO)4]} · C6H5CN] ( 1 ) and [Cu6Br6(C6H5CN)4{cyclo‐(CH3AsO)4}] ( 2 ) may be prepared by reaction of the copper(I) halide with methylcycloarsoxane (CH3AsO)n in benzonitrile at 100 °C. 1 contains four‐membered (CuI)2 rings, 2 tricyclic Cu6Br6 units, that are connected through bridging (CH3AsO)4 ligands into infinite chains. π–π Stacking of terminal C6H5CN ligands from parallel chains leads to the construction of porous frameworks, whose cavities are large enough in the case of 1 to accommodate guest C6H5CN molecules. In the presence of CsI, the self‐assembly reaction of CuI with (CH3AsO)4 in H2O–CH3OH–CH3CN (at 20 °C) or CH3CN (at 130 °C) affords [Cs(H2O)2][Cu3I4{cyclo‐(CH3AsO)4}2] · 0.5 CH3OH ( 3 ) and Cs[Cu3I4{cyclo‐(CH3AsO)4}2] ( 4 ), whose 1‐ and 2‐dimensional anionic coordination polymers are linked together through respectively [Cs{cyclo‐(CH3AsO)4‐κ4O}2]+ and [Cs{Cu3I4‐κ4I}{cyclo‐(CH3AsO)4‐κ4O}] sandwiches.  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