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.
Summary Volumetric measurements of ethylene and simple EDTA titration of copper(I) and copper(II) ions confirm that [CuL]+ and [CuL2]+ are formed when an aqueous solution of copper(II) is reduced by copper metal in the presence of ethylene, (L). The formation constants,K 1=[CuL+]2[Cu2+]–1[L]–2 andK 2=[CuL 2 + ]–1[L]–1, have been estimated. The formation of [CuL]+ is accompanied by an enthalpy change, H, of –25 kJ mol–1, and a positive entropy change, S, of 13 J mol–1 K–1.  相似文献   

2.
The oxido-pincer ligand pydotH2 (2,6-bis(1-hydroxy-1-o-tolyl-ethyl-η2O,O′)pyridine) forms two different CuII containing complexes when prepared from anhydrous CuCl2. A combination of EPR spectroscopy and EXAFS allowed to structurally characterise the light-green dimer of the formula [(pydotH2)CuCl(μ-Cl)2ClCu(pydotH2)] and the penta-coordinate olive-green monomer [(pydotH2)CuCl2]. The molecular entities imply that the ligand remains protonated upon coordination. When dissolved in DMF both compounds form monomeric species [(pydotH2)CuCl2(DMF)] which could be characterised in detail by EPR, UV-Vis/NIR spectroscopy and electrochemical measurements. The assignments were supported by comparison with CuII complexes of the related ligands 2,6-bis(hydroxymethyl)pyridine (pydimH2) and 2,6-bis(1-hydroxy-1-methyl)pyridine (pydipH2).  相似文献   

3.
Knowledge of the thermodynamic properties of aqueous copper(II) chloride complexes is important for understanding and quantitatively modeling trace copper behavior in hydrometallurgical extraction processing. In this paper, UV–Vis spectra data of Cu(II) chloride solutions with various salinities (NaCl, 0–5.57 mol·kg?1) are collected at 25 °C. The concentration distribution of Cu–Cl species is in good agreement with those calculated by a reaction model (RM). The simple hydrated ion, Cu2+, is dominant at low concentration, whereas [CuCl]+, [CuCl2]0 and [CuCl3]? become increasingly important as the chloride concentration rises. Moreover, the RM calculation suggests the present of a small amount of [CuCl4]2?. The de-convoluted molar spectrum of each species is in excellent agreement with our previous theoretical results predicted by time-dependent density functional theory treatment of aqueous Cu-containing systems. The formation constants for these copper chloride complexes have been reported and are to be preferred, except log10 K 2 ([CuCl2]0).  相似文献   

4.
The stepwise decomposition of CoBr2py2(s) has been investigated on a thermobalance by the “modified entrainment” method yielding Δ1H=88.6 kJ mol1, Δ1S=156.6 JK?1 mol?1 and Δ2H=119.0 kJ mol?1, Δ2S=211.8 JK?1 mol?1 for the dissociation of the first and second pyridine. The evaporation of CoBr2py2(l) and the association of gaseous pyridine to CoBr2py(l) forming CoBr2py2(g) has been studied by vis spectroscopy at 250?420°C. By combining the new results with literature values, a complete thermodynamic cycle for the solid-liquid-gas equilibria in the CoBr2-pyridine system could be established. It shows that in solution the formation of CoBr2py2 is not determined by the cobalt-pyridine bond energy but by the solvation energy of the rectants.  相似文献   

5.
Syntheses and Structures of the Lithiumtitanates(III)/(IV) (py)2Li[(py)2Ti(OPh)4] and (py)2Li[(py)Ti(OPh)5] The new lithiumtitanates (py)2Li[(py)2Ti(OPh)4] ( 1 ) and (py)2Li[(py)Ti(OPh)5] ( 2 ) have been obtained from the reaction of titaniumtrichloride (respectively titaniumtetrachloride 2 ) with LiOPh in the presence of the base pyridine (py). The crystal structures of both compounds show that the titanium atoms are in the centres of distorted octahedral coordination figures. In compound 1 , four oxygen and two nitrogen atoms (in cis orientation) are bonded to titanium, whereas in 2 , five oxygen and one nitrogen atom form the coordination polyeder around titanium. In both compounds, the lithium atoms are attached through phenolate bridges to the octahedra. The titanate (py)2Li[(py)2Ti(OPh)4] ( 1 ) has a single absorption band in the visible region of the UV‐spectrum showing a shoulder shifted to the bathochromic region, due to the Jahn‐Teller‐effect for d1‐systems.  相似文献   

6.
Density functional theory calculations, with an effective core potential for the copper ion, and large polarized basis set functions have been used to construct the potential energy surface of the Cu+·(CO)n (n = 1–3) complexes. A linear configuration is obtained for the global minimum of the Cu+·CO and Cu+·(CO)2 complexes with a bond dissociation energy (BDE) of 35.9 and 40.0 kcal mol-1, respectively. For the Cu+·(CO)3 complex, a trigonal planar geometry is obtained for the global minimum with a BDE of 16.5 kcal mol?1. C-coordinated copper ion complexes exhibit stronger binding energy than O-coordinated complexes as a result of Clp → 4s σ-donation. The computed sequential BDEs of Cu+·(CO)n (n = 1–4) complexes agree well with experimental findings, in which the electrostatic energy and σ-donation play an important role in the observed trend.  相似文献   

7.
The addition of [(L)4Ca(I)Mes] (Lewis base L=thf, Et2O) to mesityl copper(I) and the transmetalation reaction of mesityl copper(I) with activated calcium are suitable pathways for the synthesis of dimesityl cuprates(I) of calcium. However, the structures of the calcium cuprates(I) depend on the preparative procedure. The transmetalation reaction leads to the formation of [Mes‐Cu‐Mes]? anions whereas the addition yields dinuclear [(Mes‐Cu)2(μ‐Mes)]? anions. The solvent‐separated counterions are [Ca(thf)6]2+ and [(thf)5CaI]+, respectively. In contrast to these findings, the addition of [(L)4Ca(I)Mes] to mesityl copper(I) in an Et2O/toluene mixture led to formation of tetrameric solvent‐free iodocalcium dimesityl cuprate(I) [ICa(μ‐η16‐Mes2Cu)]4, representing a rare example of a heavy Normant‐type organocuprate.  相似文献   

8.
A dependence of the activation energy upon the extent of conversion has been discovered for the thermal decomposition of Cu4OCl6L4 with piperidine (1), morpholine (2), and triphenylphosphine oxide (3) as the ligand (L). Within the interval of conversions 0–0.3 the process shows a decrease in the activation energy (230–130 (1), 130–50 (2), and 200–100 (3) kJ mol?1). The processes considered show an isokinetic relationship with Tiso = 255 ± 15 K which corresponds to a vibrational frequency of viso = 177 ± 10 cm?1. This value accords well with IR absorption bands assigned to the stretching vibration in the trigonal CuCl3 chromophore as predicted by theory. Based on this, an assumption about the CuCl3-group as a central site of the reaction can be made. The IR- and X-ray data are presented to support the assumption made. © 1995 John Wiley & Sons, Inc.  相似文献   

9.
Binuclear and tetranuclear copper(II) complexes are of interest because of their structural, magnetic and photoluminescence properties. Of the several important configurations of tetranuclear copper(II) complexes, there are limited reports on the crystal structures and solid‐state photoluminescence properties of `stepped' tetranuclear copper(II) complexes. A new CuII complex, namely bis{μ3‐3‐[(4‐methoxy‐2‐oxidobenzylidene)amino]propanolato}bis{μ2‐3‐[(4‐methoxy‐2‐oxidobenzylidene)amino]propanolato}tetracopper(II), [Cu4(C11H13NO3)4], has been synthesized and characterized using elemental analysis, FT–IR, solid‐state UV–Vis spectroscopy and single‐crystal X‐ray diffraction. The crystal structure determination shows that the complex is a stepped tetranuclear structure consisting of two dinuclear [Cu2(L )2] units {L is 3‐[(4‐methoxy‐2‐oxidobenzylidene)amino]propanolate}. The two terminal CuII atoms are four‐coordinated in square‐planar environments, while the two central CuII atoms are five‐coordinated in square‐pyramidal environments. The solid‐state photoluminescence properties of both the complex and 3‐[(2‐hydroxy‐4‐methoxybenzylidene)amino]propanol (H2L ) have been investigated at room temperature in the visible region. When the complex and H2L are excited under UV light at 349 nm, the complex displays a strong blue emission at 469 nm and H2L displays a green emission at 515 nm.  相似文献   

10.
The electrochemical behaviour of the copper-substituted Keggin-type and sandwich-type polyoxotungstate anions of the compounds α-[(C4H9)4N]4H[PW11CuIIO39] and α-B-[(C4H9)4N]7H3[CuII4(H2O)2(PW9O34)2] was studied by cyclic voltammetry in acetonitrile. In both cases two copper 1-electron reduction waves were detected in the cathodic scan. The first one was due to the reduction of one CuII to CuI in the polyoxoanion and the second one to the consecutive reduction of the preformed CuI to Cu0, with the consequent deposition/adsorption of the ejected metal atom at the glassy carbon electrode surface. In the anodic scan, Cu0 was re-oxidised with regeneration of the initial copper(II) complexes, via a CuI intermediate. The observed two-step reduction of copper(II) to copper(0) and the formation of intermediate species containing copper(I) is here reported for the first time for copper substituted polyoxotungstates. The co-ordination of the acetonitrile molecules to the copper ions must play a role in the formation of the copper(I) species, which are not detected in aqueous solution.  相似文献   

11.
The solutions containing one of the copper salts (CuCl2, Cu(ClO4)2, Cu(NO3)2, and CuSO4) and one of the non-steroidal anti-inflammatory drugs (NSAIDs, ibuprofen, ketoprofen or naproxen) were analyzed by electrospray ionization mass spectrometry. Three of the salts, namely CuCl2, Cu(ClO4)2 and Cu(NO3)2, yielded binuclear complexes of drug:metal stoichiometry 1:2. Existence of the complexes of such stoichiometry has not been earlier observed. For copper(II) chloride the complexes (ions of the type [M-HCOOH+Cu2Cl]+ and [M+Cu2Cl]+, M stands for the drug molecule) were formed in the gas phase. When copper(II) perchlorate or copper(II) nitrate was used, the observed binuclear copper complexes (ions of the type [M-H+Cu2(ClO4)2+CH3OH]+, [M-H+Cu2(ClO4)2]+ and [M-H+Cu2(NO3)2+CH3OH]+, [M-H+Cu2(NO3)2]+) were observed at low cone voltage, thus these complexes must have already existed in the solution analysed. Therefore, such complexes may also exist under physiological conditions.   相似文献   

12.
A study of the strong N?X????O?N+ (X=I, Br) halogen bonding interactions reports 2×27 donor×acceptor complexes of N‐halosaccharins and pyridine N‐oxides (PyNO). DFT calculations were used to investigate the X???O halogen bond (XB) interaction energies in 54 complexes. A simplified computationally fast electrostatic model was developed for predicting the X???O XBs. The XB interaction energies vary from ?47.5 to ?120.3 kJ mol?1; the strongest N?I????O?N+ XBs approaching those of 3‐center‐4‐electron [N?I?N]+ halogen‐bonded systems (ca. 160 kJ mol?1). 1H NMR association constants (KXB) determined in CDCl3 and [D6]acetone vary from 2.0×100 to >108 m ?1 and correlate well with the calculated donor×acceptor complexation enthalpies found between ?38.4 and ?77.5 kJ mol?1. In X‐ray crystal structures, the N‐iodosaccharin‐PyNO complexes manifest short interaction ratios (RXB) between 0.65–0.67 for the N?I????O?N+ halogen bond.  相似文献   

13.
The reaction of copper(II) perchlorate with the hydrochloride salt of 3,6,9,15-tetra-azabicyclo[9.3.1]penta-deca-1,11,13-triene (L1) in acetonitrile forms two macrocyclic complexes that can be characterized: [L1CuIICl][ClO4] (1) and [L1CuIICl]2[CuCl4] (2). The structural, electronic, and redox properties of these complexes were studied using spectroscopy (EPR and UV–visible) and electrochemistry. In addition, the solid-state structure of 1 was obtained using X-ray diffraction. The copper(II) is five-coordinate ligated by four N-atoms of the macrocycle and a chloride atom. EPR studies of 1 both in DMF and aqueous solution indicate the presence of a single copper(II) species. In contrast, EPR studies of 2 performed in frozen DMF and in the solid-state reveal the presence of two spectroscopically distinct copper(II) complexes assigned as [L1CuIICl]+ and [CuIICl4]2?. Lastly, electrochemical studies demonstrate that both [L1CuIICl]+ and [CuIICl4]2? are redox active. Specifically, the [L1CuIICl]+ undergoes a quasi-reversible Cu(II)/(I) redox reaction in the absence of excess chloride. In the presence of chloride, however, the chemical irreversibility of this couple becomes evident at concentrations of chloride that exceed 50 mM. As a result, the presence of chloride from the chemical equilibrium of this latter species impedes the reversibility of the reduction of [L1CuIICl]+ to [L1CuICl]0.  相似文献   

14.
Summary The aquation ofcis-[(en)2Co(CO2H)2]+ tocis-[(en)2Co(OH2)(CO2H)]2+ is catalysed by Cu2+ and the rate equation, –d[complex]t/dt=(kCu[Cu2+]+kH [H+]) [complex)T is valid at [Cu2+]T=0.01–0.1, I=0.5 and [HClO4]=0.005 mol dm–3. The rate measurements are reported at 30, 35, 40 and 45°C and the rate and activation parameters for the Cu2+ and H+-catalysed paths are: kH(35°C)=(2.44±0.09)×10–2 dm3 mol–1 s–1, H=83±13 kJ mol–1, S=–8±42 JK–1 mol–1, k Cu (35°C)=(3.30±0.09)×10–3 dm3 mol–1 s–1, H=73.2±6.1 kJ mol–1, S=–55±20 JK–1 mol–1. The formate-bridged innersphere binuclear complex,cis-[(en)2Co{(O2CH)2Cu}]3+ may be involved as the catalytically active intermediate in the copper(II)-catalysed path, just as the corresponding H+-bridged species presumed to be present in the acidcatalysed path.  相似文献   

15.
The structures of five compounds consisting of (prop‐2‐en‐1‐yl)bis[(pyridin‐2‐yl)methylidene]amine complexed with copper in both the CuI and CuII oxidation states are presented, namely chlorido{(prop‐2‐en‐1‐yl)bis[(pyridin‐2‐yl)methylidene]amine‐κ3N,N′,N′′}copper(I) 0.18‐hydrate, [CuCl(C15H17N3)]·0.18H2O, (1), catena‐poly[[copper(I)‐μ2‐(prop‐2‐en‐1‐yl)bis[(pyridin‐2‐yl)methylidene]amine‐κ5N,N′,N′′:C2,C3] perchlorate acetonitrile monosolvate], {[Cu(C15H17N3)]ClO4·CH3CN}n, (2), dichlorido{(prop‐2‐en‐1‐yl)bis[(pyridin‐2‐yl)methylidene]amine‐κ3N,N′,N′′}copper(II) dichloromethane monosolvate, [CuCl2(C15H17N3)]·CH2Cl2, (3), chlorido{(prop‐2‐en‐1‐yl)bis[(pyridin‐2‐yl)methylidene]amine‐κ3N,N′,N′′}copper(II) perchlorate, [CuCl(C15H17N3)]ClO4, (4), and di‐μ‐chlorido‐bis({(prop‐2‐en‐1‐yl)bis[(pyridin‐2‐yl)methylidene]amine‐κ3N,N′,N′′}copper(II)) bis(tetraphenylborate), [Cu2Cl2(C15H17N3)2][(C6H5)4B]2, (5). Systematic variation of the anion from a coordinating chloride to a noncoordinating perchlorate for two CuI complexes results in either a discrete molecular species, as in (1), or a one‐dimensional chain structure, as in (2). In complex (1), there are two crystallographically independent molecules in the asymmetric unit. Complex (2) consists of the CuI atom coordinated by the amine and pyridyl N atoms of one ligand and by the vinyl moiety of another unit related by the crystallographic screw axis, yielding a one‐dimensional chain parallel to the crystallographic b axis. Three complexes with CuII show that varying the anion composition from two chlorides, to a chloride and a perchlorate to a chloride and a tetraphenylborate results in discrete molecular species, as in (3) and (4), or a bridged bis‐μ‐chlorido complex, as in (5). Complex (3) shows two strongly bound Cl atoms, while complex (4) has one strongly bound Cl atom and a weaker coordination by one perchlorate O atom. The large noncoordinating tetraphenylborate anion in complex (5) results in the core‐bridged Cu2Cl2 moiety.  相似文献   

16.
A rare, low‐spin FeIV imide complex [(pyrr2py)Fe?NAd] (pyrr2py2?=bis(pyrrolyl)pyridine; Ad=1‐adamantyl) confined to a cis‐divacant octahedral geometry, was prepared by reduction of N3Ad by the FeII precursor [(pyrr2py)Fe(OEt2)]. The imide complex is low‐spin with temperature‐independent paramagnetism. In comparison to an authentic FeIII complex, such as [(pyrr2py)FeCl], the pyrr2py2? ligand is virtually redox innocent.  相似文献   

17.
The title compound, (C10H10N2S)[CuCl4], was obtained by the reaction of cupric chloride with pyridine‐4‐thiol in a mixture of aceto­nitrile and tetra­hydro­furan, suggesting that the desulfurization and coupling reactions of pyridine‐4‐thiol occurred in the presence of the Cu2+ ion. X‐ray diffraction analysis reveals the presence of one 4,4′‐thio­dipyridinium cation, H2bps2+, and one [CuCl4]2− anion. The cations interact with the anions via N—H⋯Cl hydrogen‐bonding interactions to form a closed `chair' conformation.  相似文献   

18.
A dinuclear copper(II) complex, [CuII2(L)2] is afforded by the reaction of CuCl2 · 2H2O with a triazenido ligand, 1-[(2-carboxymethyl) benzene]-3-[2-carboxybenzene] triazene (H2L). Structural investigation shows that the copper-copper distance [2.3985(7) Å] is significantly shorter than the sum of the van der Waals radii of Cu (1.40 Å), suggesting that there are metal-metal bonds in [CuII2(L)2]. In solid, there is a strong antiferromagnetic interaction between copper(II) ions (J = –135.6 cm–1). In homogeneous environment, [CuII2(L)2] shows electrocatalytic activities for hydrogen generation both from acetic acid with a turnover frequency (TOF) of 32 mol of hydrogen per mole of catalyst per hour [mol(H2) · mol–1(catalyst) · h–1] at an overpotential (OP) of 941.6 mV, and neutral buffer with a TOF of 512 mol(H2) · mol–1(catalyst) · h–1 at an OP of 836.7 mV.  相似文献   

19.
A stopped-flow investigation of the reversible addition of Ph3P to [(C8H11)Co(C5H5)]+ indicates the rate law, kobs = k1[Ph3P] + k?1. The low Δ2 of 21.0 ± 1.2 kJ mol?1 and the negative ΔS2 of ?114 ± 5 J K?1 mol?1 are consistent with rapid addition to the enyl ligand. The higher Δ2 of 86.2 ± 5.1 kJ mol?1 and the positive ΔS2 of +60 ± 17 J K?1 mol?1are as expected for the reverse dissociation. Preliminary studies show that the related complex [(C7H9)Co(C5H5)]+ is at least 65 times more electrophilic towards Ph3P.  相似文献   

20.
A new dicarboxylic acid, LH2 , derived from 2-[(2-hydroxy-3,5-dimethylphenyl) (phenyl)methyl]-4,6-dimethylphenol (1) was prepared by reacting it with methyl bromoacetate followed by alkaline hydrolysis. Two five-coordinate mononuclear complexes of L with zinc (+2) and copper (+2), [Zn(L)(py)2(H2O)]?·?H2O?·?py (IV), and [Cu(L)(py)2(H2O)] (V), were prepared (py?=?pyridine) and characterized. The packing patterns of these two complexes are different and the H-bond interactions in their lattices are controlled by the presence or absence of water molecules. This difference arises from subtle change in the orientation of carbonyl groups of the carboxylates.  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