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 52 毫秒
1.
The interaction of Np(VI), Pu(VI), Np(V), Np(IV), Pu(IV), Nd(III), and Am(III) with Al(III) in solutions at pH 0–4 was studied by the spectrophotometric method. It was shown that, in the range of pH 3–4, the hydrolyzed forms of neptunyl and plutonyl react with the hydrolyzed forms of aluminium. In the case of Pu(VI), the mixed hydroxoaqua complexes (H2O)3PuO2(-OH)2Al(OH)(H2O)3 2+ or (H2O)4PuO2OAl(OH)(H2O)4 2+ are formed at the first stage of hydrolysis. Np(VI) also forms similar hydroxoaqua complexes with Al(III). The formation of the mixed hydroxoaqua complexes was also observed when Np(IV) or Pu(IV) was simultaneously hydrolyzed with Al(III) at pH 1.5–2.5. The Np(IV) complex with Al(III) has, most likely, the formula (H2O) n (OH)Np(-OH)2Al(OH)(H2O)3 3+. At pH from 2 to 4.1 (when aluminium hydroxide precipitates), the Np(V) or Nd(III) ions exist in solutions with or without Al(III) in similar forms. When pH is increased to 5–5.5, these ions are almost not captured by the aluminium hydroxide precipitate.  相似文献   

2.
Summary The kinetics of the OsVIII-catalysed oxidation of glycols by alkaline hexacyanoferrate(III) ion exhibits zerothorder dependence in [Fe(CN) 6 3– ] and first-order dependence in [OsO4]. The order with respect to glycols is less than unity, whereas the rate dependence on [OH] is a combination of two rate constants; one independent of and the other first-order in [OH]. These observations are commensurate with a mechanism in which two complexes, [OsO4(H2O)G] and [OsO4(OH)G]2–, are formed either from [OsO4(H2O)(OH)] or [OsO4(OH)2]2– and the glycol GH, or by [OsO4(H2O)2] and [OsO4(H2O)(OH)] and the glycolate ion, G, which is in equilibrium with the glycol GH through the reaction between GH and OH. Hence there is an ambiguity about the true path for the formation of the two OsVIII-glycol complexes. A reversal in the reactivity order of glycols in the two rate-determining steps, despite the common attack of OH ion on the two species of OsVIII-complexes, indicates that the two complexes are structurally different because S changes from the negative (corresponding to k11) to positive (related to k2).  相似文献   

3.
PuO2(am) solubility was investigated as a function of time, for pH from 0.5 to 11, and in the presence of 0.001 M FeCl2 or 0.00052 M hydroquinone to determine the effect of environmentally important reducing agents on PuO2(am) solubilization under geological conditions. Equilibrium was reached in <4 days. The observed PuO2(am) solubilities were many orders of magnitude higher than the Pu(IV) concentrations predicted from thermodynamic data. Spectroscopic, solvent extraction, and thermodynamic analyses of data showed that Pu(III) was the dominant aqueous oxidation state. The experimental pH, pe, and Pu(III) concentrations from both the Fe(II) and hydroquinone systems provided a log K 0 value of 15.5 ± 0.7 for [PuO2(am) + 4H+ + e Pu3+ + 2H2O]. The data show that reduction reactions involving Fe(II) and hydroquinone are relatively rapid and that reductive dissolution of PuO2(am), hitherto ignored, may play an important role in controlling Pu behavior under reducing environmental conditions.  相似文献   

4.
The stability constans, 1, of each monochloride complex of Eu(III) have been determined in the methanol and water mixed system with 1.0 mol·dm–3 ionic strength using a solvent extraction technique. The values of 1 increase with an increase in the mole fraction of methanol (X S ) in the mixed solvent system when 0X S 0.40. The, distance of Eu3+–Cl in the mixed solvent system was calculated using the Born-type equation and the Gibbs' free energy derived from 1. Calculation of the Eu3+–Cl distance and the preferential solvation, of Eu3+ by water proposed the variation of the outersphere complex of EuCl2+ as follows: (1) [Eu(H2O)9]3+Cl, [Eu(H2O)8]3+Cl and [Eu(H2O)7(CH3OH)3+Cl inX S0.014, (2) [Eu(H2O)8]3–Cl and [Eu(H2O)7(CH3OH)]3+Cl in 0.014<X S <0.25 and (3) [Eu(H2O)7(CH3OH)]3–Cl and [Eu(H2O)6(CH3OH)[2 3+Cl in 0.25<X S 0.40.  相似文献   

5.
Summary The kinetics of the anation reaction of [Co(NH3)5H2O]3+ by H3PO3/H2PO 3 , to give [CoH2PO3(NH3)5]2+, have been studied at 60, 70 and 80°C, in the acidity range [H+](M)=1.5 · 10–1 –2.0 · 10–3. Only H2PO3 is found to be reactive. The rate data is consistent with an Id mechanism. The mean value of outer sphere association of [Co(NH3)H2O]3+ with H2PO 3 is 1.5 M–1. Values of the interchange constants are: 1044ki(s–1)= 0.29, 1.47, 5.13, at 60, 70 and 80 °C respectively (H= 1.4 · 102KJmol–1, S=8.3 · 10 JK–1 mol–1). The first acidity constant of H3PO3 at I=1.0 has also been determined: 102Ka(M)=4.8, 5.2 and 5.5, at 25, 40 and 50 °C respectively.  相似文献   

6.
Raman spectroscopic measurements were performed at ambient temperature onaqueous silica-bearing solutions (0.005 < m Si < 0.02; 0 < pH < 14). The spectraare consistent with the formation of monomeric Si(OH)o 4, SiO(OH) 3 andSiO2(OH)2– 2 species at acid to neutral, basic, and strongly basic pH, respectively.Raman spectra of aqueous Al-bearing solutions at basic pH confirm thepredominance of the Al(OH) 4 species in a wide concentration range (0.01 < m Al < 0.1).Raman spectra of basic solutions (12.4 < pH < 14.3), containing both Al andSi, exhibit a strong decrease in intensities of SiO(OH) 3, SiO2(OH)2– 2, andAl(OH) 4 bands in comparison with Al-free Si-bearing and Si-free Al-bearingsolutions of the same metal concentration and pH, suggesting the formation ofsoluble Al—Si complexes. The amounts of complexed Al and Si derived fromthe measurements of the Al and Si band intensities in strongly basic solutions(pH 14) are consistent with the formation, between Al(OH) 4 andSiO2(OH)2– 2, of the single Al—Si dimer SiAlO3(OH)3– 4 according to the reactionSiO2(OH)2– 2 + Al(OH) 4 SiAlO3(OH)3– 4 + H2OAt lower pH ( 12.5) the changes in band intensities are consistent with theformation of several, likely more polymerized, Al—Si complexes.  相似文献   

7.
Ru(PPh3)3Cl2 reacts with N(1)-alkyl-2-(arylazo)imidazoles, p-RC6H4N=NC3H2N2X, [RaaiX, R = H(a), Me(b), Cl(c); X = Me(1), Et(2), Bz(3)] under refluxing conditions in EtOH to give [Ru(RaaiX)2(PPh3)2](ClO4)2 · H2O complexes (4–6). RaaiX is a bidentate chelator (N, N) with N(imidazole), N and N(azo), N donor centres. Three isomers are present in the mixture in which the pairs of PPh3, N and N occupy cis–cis–trans, cis–trans–cis and cis–cis–cis, positions respectively. The isomers were identified by 1H-n.m.r. spectra. Four signals are observed in the aliphatic zone for N(1)-X; two are of equal intensity at higher and the other two signals at lower in the ratio 1:0.3:0.2 suggesting the presence of cis–cis–cis, cis–trans–cis and cis–cis–trans-geometry. The complexes display the allowed t 2(Ru) *(RaaiX) transition. Cyclic voltammetry indicates two consecutive RuIII/II couples along with azo reductions.  相似文献   

8.
The methods of optical, ESR, and IR spectroscopy were used to obtain data on the structure and mechanism for the formation of the products in the reaction of dioxasilirane groups (Si–O)2Si 2 (DOSG) stabilized on the silica surface. Depending on the regime of the reaction (temperature and methane pressure), the process is accompanied by the formation of various products: methoxy (–O–CH3) and ethoxy (–O–C2H5) groups. The process mechanism is elucidated: this is a free-radical reaction in which paramagnetic sites are generated in the reaction between DOSG and methane molecules. The formation of final products is due to the reactions >Si(O)(OCH3) + CH4 >Si(OH)(OCH3) + CH3 and >Si(O–CH2)(OH) + CH3 >Si(OH)(OC2H5). The ratio of the rate constants of methyl radical addition to (Si–O)2Si: and (Si–O)2Si 2 at room temperature was determined experimentally (4.6 ± 1.0).  相似文献   

9.
The synthesis, reduction, optical and e.p.r. spectral properties of a series of new binuclear copper(II) complexes, containing bridging moieties (OH, MeCO2 , NO2 , and N3 ), with new proline-based binuclear pentadentate Mannich base ligands is described. The ligands are: 2,6-bis[(prolin-1-yl)methyl]4-bromophenol [H3L1], 2,6-bis[(prolin-1-yl)methyl]4-t-butylphenol [H3L2] and 2,6-bis[(prolin-1-yl)methyl]4-methoxyphenol [H3L3]. The exogenous bridging complexes thus prepared were hydroxo: [Cu2L1(OH)(H2O)2] · H2O (1a), [Cu2L2(OH)(H2O)2] · H2O (1b), [Cu2L3(OH)(H2O)2] · H2O (1c), acetato [Cu2L1(OAc)] · H2O (2a), [Cu2L2(OAc)] · H2O (2b), [Cu2L3(OAc)] · H2O (2c), nitrito [Cu2L1(NO2)(H2O)2] · H2O (3a), [Cu2L2(NO2)(H2O)2] · H2O (3b), [Cu2L3(NO2)(H2O)2] · H2O (3c) and azido [Cu2L1(N3)(H2O)2] · H2O (4a), [Cu2L2(N3)(H2O)2] · H2O (4b) and [Cu2L3(N3)(H2O)2] · H2O (4c). The complexes were characterized by elemental analysis and by spectroscopy. They exhibit resolved copper hyperfine e.p.r. spectra at room temperature, indicating the presence of weak antiferromagnetic coupling between the copper atoms. The strength of the antiferromagnetic coupling lies in the order: NO2 N3 OH OAc. Cyclic voltammetry revealed the presence of two redox couples CuIICuII CuIICuI CuICuI. The conproportionality constant K con for the mixed valent CuIICuI species for all the complexes have been determined electrochemically.  相似文献   

10.
Summary The pentadentate macrocycle 1,4,7,10,13-penta-azacyclo-hexadecane [16]aneN5=(3)=L} has been prepared and a variety of copper(II), nickel(II) and cobalt(III) complexes of the ligand characterised. The copper complex [CuL](ClO4)2, on the basis of its d-d spectrum, appears to be square pyramidal, while [NiL(H2O)](ClO4)2 is octahedral. The copper(II) and nickel(II) complexes dissociate readily in acidic solution and these reactions have been studied kinetically. For the copper(II) complex, rate=kH[complex][H+]2 with kH =4.8 dm6 mol–2s–1 at 25 °C and I=1.0 mol dm–3 (NaClO4) with H=43 kJ mol–1 and S 298 =–89 JK–1 mol–1. Dissociation rates of the copper(II) complexes increase with ring size in the order: [15]aneN5 < [16]aneN5 < [17]aneN5. For the dissociation of the nickel(II) complex, rate=kH[Complex][H+] with kH=9.4×10–3 dm3mol–1 s–1 at 25 °C and I =1.0 mol dm–3 (NaClO4) with H=71 kJ mol–1 and S 298 =–47 JK–1mol–1.The cobalt(III) complexes, [CoLCl](ClO4)2, [CoL(H2O)]-(ClO4)3, [CoL(NO2)](ClO4)2, [CoL(DMF)](ClO4)3 (DMF=dimethylformamide) and [CoL(O2CH)](ClO4)2 have been characterised. The chloropentamine [CoCl([16]aneN5)]2+ undergoes rapid base hydrolysis with kOH=1.1× 105dm3 mol–1s–1 at 25°C and I=0.1 mol dm–3 (H=73 kJ mol–1 and S 298 =98 JK–1 mol–1). Rapid base hydrolysis of [CoL(NO2)]2+ is also observed and the origins of these effects are considered in detail.  相似文献   

11.
Summary The kinetics of oxygen-transfer from [MoO2(Et-L-cys)2] to PPh3 and the reaction between [Mo2O3(Et-L-cys)4] and O2 in benzene solution have been investigated using spectrophotometric techniques between 25 and 40°. The rate laws-d[Mo6+]/dt = k1[Mo6+][PPh3] with k1 (at 35°) = 2.95×10–4dm3mol–1s–1 and -d[Mo5+]/dt = 2k3[Mo5+][O2] with k3 (at 35°) = 6.3×10–2 dm3mol–1s–1 account for the kinetic data obtained with activation parameters (at 35°) of H = 46 kJ mol–1, S = –153 JK–1mol–1, and H = 50.8 kJ mol–1, S = –95 JK–1 mol–1 respectively.  相似文献   

12.
Summary Kinetics of formation of ternary complexes from diaquo-nitrilotriacetatonickelate(II), [Ni(nta)(H2O)2], and diaquoanthranilato-N, N-diacetatonickelate(II), [Ni(ada)-(H2O2] and amino acids have been studied by a pH indicator method using stopped-flow spectrophotontetry. The results conform to 1/kobs=1/k+[H+]/kK·TL, where K is the equilibrium constant for the formation of [Ni(A)(-L)(H2O)]2–(A=nta3– or ada3–) and k is the specific rate constant for the subsequent rate-determining ring closure leading to [Ni(A) (=L)]2–. For the different amino acids, the k values decrease in the sequence: glycine>-alanine>L-phenylalanine>L-Valine>L-methionine>-alanine>sarcosine>N,N-dimethylglycine, and areca. 1000 times smaller than the k values for complexation of [Ni)(nta)(H2O)2] with monodentate ligands, such as NH3 and imidazole. The spread of k values is much less than the pKa values of the amino acids, and can be accounted for on the basis of the proposed mechanism. The relative rates are enthalpy controlled and high negative S values are commensurate with ring closure as the rate-determining step.  相似文献   

13.
Summary The species, UO2H3L, UO2H2L2–, UO2HL3–, UO2L4–, UO2(OH)L5– and UO2(OH)2L6– are found in the equilibria between uranyl ions and 3,3-bis[N,N-di(carboxymethyl)-aminomethyl]-o-cresolsulphonphthalein (H6L; xylenol orange; dcac) in aqueous solution. The equilibria have been studied by the potentiometric method at 25° and at an ionic strength of 0.1M (KNO3). New algebraic equations have been employed to evaluate the equilibrium constants.  相似文献   

14.
The hydrolysis of the [Pt(dien)H2O]2+ and [Pd(dien)H2O]2+ complexes has been investigated by potentiometry at 298 K, in 0.1 mol dm–3 aqueous NaClO4. Least-squares treatment of the data obtained indicates the formation of mononuclear and -hydroxo-bridged dinuclear complexes with stability constants: log 11 = –6.94 for [Pt(dien)OH]+, log 11 = –7.16 for [Pd(dien)OH]+, and also log 22 = –9.37 for [Pt2(dien)2(OH)2]2+ and log 22 = –10.56 for [Pd2(dien)2(OH)2]2+. At pH values > 5.5, formation of the dimer becomes significant for the PtII complex, and at pH > 6.5 for the PdII complex. These results have been analyzed in relation to the antitumor activity of PtII complexes.  相似文献   

15.
A platinum-lined, flowing autoclave facility is used to investigate the solubility behavior of Cr2O3 and FeCr2O4 in alkaline sodium phosphate, sodium hydroxide, and ammonium hydroxide solutions between 21 and 288°C. Baseline Cr(III) ion solubilities were found to be on the order of 0.1 nmolal, which were enhanced by the formation of anionic hydroxo and phosphato complexes. At temperatures below 51°C, the activity of Cr(III) ions in aqueous solution is controlled by a Cr(OH)3·3H2O solid phase rather than Cr2O3; above 51°C the saturating solid phase is -CrOOH. Measured chromium solubilities were interpreted via a Cr(III) ion hydrolysis/complexing model and thermodynamic functions for the hydrolysis/complexing reaction equilibria were obtained from least-squares analyses of the data. The existence of four new Cr(III) ion complexes is reported: Cr(OH)3(H2PO4), Cr(OH)3(HPO4)2–, Cr(OH)3(PO4)3–, and Cr(OH)4(HPO4)-(H2PO4)4–. The last species is the dominant Cr(III) ion complex in concentrated, alkaline phosphate solutions at elevated temperatures.  相似文献   

16.
The solubility of CaSO3·1/2H2O(c) was studied under alkaline conditions (pH>8.2), in deaerated and deoxygenated Na2SO3 solutions ranging in concentration from 0.0002 to 0.4M and in CaCl2 solutions ranging in concentration from 0.0002 to 0.01M, for equilibration periods ranging from 1 to 7 days. Equilibrium was approached from both the over- and the under-saturation directions. In all cases, equilibrium was reached in <1 days. The aqueous Ca2+–SO 3 2– ion interactions can be satisfactorily modeled using either ion-association or ion-interaction aqueous thermodynamic models. In the ion-association model, the log K°=2.62±0.07 for Ca2++SO 3 2– CaSO 3 0 . In the Pitzer ion-interaction model, the binary parameters (0) and (1) for Ca2+–SO 4 2– were used, and the value of (2) was determined from the experimental data. As expected given the strong association constant, the value of (0) was quite small (about –134). We feel a combination of the two models is most useful. The logarithm of the thermodynamic equilibrium constant (K°) of the CaSO3·1/2H2O(c) solubility reaction (CaSO3·1/2H2O(c)Ca2++SO 3 2+ +0.5H2O) was found to be –6.64±0.07.  相似文献   

17.
Summary The kinetics and mechanism of the system: [FeL(OH)]2–n + 5 CN [Fe(CN)5(OH)]3– + Ln–, where L=DTPA or HEDTA, have been investigated at pH= 10.5±0.2, I=0.25 M and t=25±0.1 C.As in the reaction of [FeEDTA(OH)]2–, the formation of [Fe(CN)5(OH)]3– through the formation of mixed ligand complex intermediates of the type [FeL(OH)(CN)x]2–n–x, is proposed. The reactions were found to consist of three observable stages. The first involves the formation of [Fe(CN)5(OH)]3–, the second is the conversion of [Fe(CN)5(OH)]3– into [Fe(CN)6]3– and the third is the reduction of [Fe(CN)6]3– to [Fe(CN)6]4– by oxidation of Ln– The first reaction exhibits a variable order dependence on the concentration of cyanide, ranging from one at high cyanide concentration to three at low concentration. The transition between [FeL(OH)]2–n and [Fe(CN)5(OH)]3– is kinetically controlled by the presence of four cyanide ions around the central iron atom in the rate determining step. The second reaction shows first order dependence on the concentration of [Fe(CN)5(OH)]3– as well as on cyanide, while the third reaction follows overall second order kinetics; first order each in [Fe(CN)6]3– and Ln–, released in the reaction. The reaction rate is highly dependent on hydroxide ion concentration.The reverse reaction between [Fe(CN)5(OH)]3– and Ln– showed an inverse first order dependence on cyanide concentration along with first order dependence each on [Fe(CN)5– (OH)]3– and Ln–. A five step mechanism is proposed for the first stage of the above two systems.  相似文献   

18.
Two clathrate modifications of the title host with 4-methylpyridine (4-CH3C5H4N) as a guest have been determined at –50°C. [Mg(4-CH3C5H4N)4(NCS)2] · 2/3(4-CH3C5H4N) · 1/3H2O is trigonal, space group , witha=27.630(7),c=11.219(3) ÅV=7417(4) Å3,Z=9,D calc=1.171 g cm–3,(CuK )=18.506 cm–1, finalR=0.064. [Mg(4-CH3C5H4N)4(NCS)2] · (4-CH3C5H4N) is tetragonal, space group I4l/a, witha=16.944(7),c=23.552(9)Å,V=6762(5) Å,Z=8,D calc=1.191 g cm–3, (CuK )=18.200 cm–1, finalR=0.071.The structures consist of molecular packings of the same host complex units and the guest species. The Mg(II) cation is octahedrally coordinated to theN-atoms of four 4-methylpyridine and twotrans-coordinated isothiocyanato ligands in the host molecule. The conformations of the molecule are considerably different both in symmetry and in geometry in these two structures. The guest 4-methylpyridine molecules are disordered into channels which have different topology in these two clathrates resulting in different thermal stability.  相似文献   

19.
The interaction of thymidine, a nucleoside, with hydroxopentaaquarhodium(III), [Rh(H2O)5(OH)]2+ ion in aqueous medium is reported and the possible mode of binding is discussed. The kinetics of interaction between thymidine and [Rh(H2O)5OH]2+ has been studied spectrophotometrically as a function of [Rh(H2O)5OH2+], [thymidine], pH and temperature. The reaction has been monitored at 298 nm, the max of the substituted complex, and where the spectral difference between the reactant and product is a maximum. The reaction rate increases with [thymidine] and reaches a limiting value at a higher ligand concentration. From the experimental findings an associative interchange mechanism for the substitution process is suggested. The activation parameters (H=47.8 ± 5.7 kJ mol–1, S=–173 ± 17 J K–1 mol–1) supports our proposition. The negative G0 (–13.8 kJ mol–1) for the first equilibrium step also supports the spontaneous formation of the outer sphere association complex.  相似文献   

20.
The complexes formed by photosubstitution of pyrazine (Pz) in octacyanomolybdate(IV) and -tungstate(IV) with 8-hydroxyquinoline have been assigned the formulae [Mo(CN)2(OH)2(Pz)2(OX)] and [W(CN)2(OH)2(Pz)2(OX)·1.5H2O]. Coordination of Pz as an unidentate ligand by donating a lone pair of electron from nitrogen is shown by an absorption peak between 8–11 µ. Mechanism for the thermal decomposition of the complexes has been given. The formation of tungsten metal as residue in case of II has been confirmed by XRD analysis. The kinetic and thermodynamic parameters like activation energy (E a), pre-exponential factor (A) and entropy of activation (S #) were calculated employing different integral methods of Doyle, Coats and Redfern and Arrhenius.H for each stage of decomposition was obtained from DSC.This revised version was published online in November 2005 with corrections to the Cover Date.  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