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.
Conclusions The limiting and initiating stage in oxidation-reduction destruction of (Ru)Dipy 3 + with the hydroxyl anion, and in acidic reaction with a water molecule.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 6, pp. 1243–1247, June, 1988.  相似文献   

2.
Two solid complexes, fac–[Cr(gly)3] and [Cr(gly)2(OH)]2, (where gly is glycinato ligand) were prepared and their acid-catalysed aquation products were identified. The structure of [Cr(gly)3] was solved by X-ray diffraction, revealing a cationic 3D sublattice with perchlorate anions inside its cavities. Acid-catalysed aquation of [Cr(gly)3] and [Cr(gly)2(OH)]2 leads to the same inert product, [Cr(gly)2(H2O)2]+, in a two-stages process. At the first stage, intermediate complexes, [Cr(gly)2(O–glyH)(H2O)]+ and [Cr(gly)2(H2O)–OH–Cr(gly)2(H2O)]+, are formed respectively. Kinetics of the first aquation stage of [Cr(gly)3] were studied in HClO4 solutions. The dependencies of the pseudo first-order rate constants on [H+] are as follows: k obs1H = k 0 + k 1 K p1[H+], where k 0 and k 1 are rate constants for the chelate-ring opening via spontaneous and acid-catalysed reaction paths, respectively, and K p1 is the protonation constant. The proposed mechanism assumes formation of the reactive intermediate as a result of proton addition to the coordinated carboxylate group of the didentate ligand. Some kinetic studies on the second reaction stage, the one-end bonded glycine liberation, were also done. The obtained results were analogous to those for stage I. In this case, the proposed reactive species are intermediates, protonated at the carboxylate group of the monodentate glycine. Base hydrolysis of two complexes, [Cr(gly)2(O–gly)(OH)] and [Cr(gly)2(OH)2], was studied in 0.2–1.0 M NaOH. The pseudo first-order rate constants, k obsOH, were [OH] independent in the case of [Cr(gly)2(O–gly)(OH)], whereas those for [Cr(gly)2(OH)2] linearly depended on [OH]. The reaction mechanisms were proposed, where the OH -catalysed reaction path was rationalized in terms of formation of the reactive conjugate base, [Cr(gly)2(OH)(O)]2−, as a result of OH ligand deprotonation. Activation parameters were determined and discussed.  相似文献   

3.
Kinetics and Catalysis - The kinetics of glycolaldehyde and glyceraldehyde condensation with formaldehyde in a neutral aqueous medium in the presence of homogeneous phosphates and in a weakly...  相似文献   

4.
This work presents spectroscopic studies and electrochemical characterization of Cu(II)—dicyandiamide (DCDA) complex formation. The range of conditions leading to the precipitation of the complex is significantly larger than that presented in the literature. In all cases the stoichiometry of the compound is: [Cu(DCDA)2(SO4)(H2O)5). The spectroscopic data suggest that DCDA is a monodentate ligand forming a bond with Cu2+ via the nitryl nitrogen. Electroreduction of this complex is a two-step process occurring through a Cu(I)—DCDA intermediate.  相似文献   

5.
6.
Extraction of WO 4 2– and ReO 4 by Adogen-381, tricaprylmethylammonium chloride, Hyamine 10-X, trioctylphosphine oxide or dibenzylsulphoxide in xylene from HNO3, HCl or H2SO4 acid medium was investigated. Based on the separation factors obtained, the separation of ReO 4 from WO 4 2– was elucidated. ReO 4 was separated from WO 4 2– in high radiochemical purity: >99.9% by three successive extractions and strippings using Adogen-381 from HCl or HNO3 acid medium.  相似文献   

7.
The negative low-frequency capacitance that appears in interpretations of impedance of the iron electrode in weakly acidic solutions is shown to arise in the case of interaction of two consecutive nonequilibrium flows that constitute a two-stage anodic faradaic process of intermediate adsorption in the prepassivation potential range. The low-frequency capacitance is negative throughout a potential range where the logarithm of rate constant vs. potential (logk vs. E) dependence has the higher slope for the limiting stage. The low-frequency capacitance becomes positive at higher anodic potentials and for the other limiting stage.  相似文献   

8.
The synthesis, characterization, X-ray crystal structures, and reactivity in aqueous acidic solution of the Co(III) carbonate complexes [Co(tpa)(O2CO)]ClO4.H2O, [Co(Me-tpa)(O2CO)]ClO4.0.5H2O, [Co(Me2-tpa)(O2CO)]ClO4.0.5H2O, and [Co(Me3-tpa)(O2CO)]ClO4 are reported (tpa = tris(2-pyridylmethyl)amine; Me-tpa, Me2-tpa, and Me3-tpa are derivatives of tpa containing one, two, and three 6-methylpyridyl rings, respectively). The complexes display very different spectroscopic and 59Co NMR properties, consistent with the decreasing ligand field strength of the tripodal amine ligands in the order tpa > Me-tpa > Me(2)-tpa > Me3-tpa. X-ray structural data show an increase in the average Co-N bond distances as the number of methyl groups on the tripodal amine ligand increases, and this is the result of steric interactions between the methyl groups and the carbonate ligand and between the methyl groups themselves. Rate data for the acid hydrolysis of [Co(tpa)(O2CO)]+ (I = 1.0 M (NaClO4), 25.0 degrees C) over the [HClO4] range of 0.10-1.0 M are consistent with a previously proposed mechanism involving protonation of the carbonate ligand prior to ring-opening, but the equilibrium constant for protonation is smaller in this case than those obtained previously, as is the equilibrium constant for proton transfer from the exo to the endo O atoms. Comparative rate data ([HCl] = 6.0 M, 25.0 degrees C) for the four complexes show that those containing methylated ligands undergo acid hydrolysis between 25 and 90 times more slowly than [Co(tpa)(O2CO)]+ under the same conditions, and it is proposed that this rate difference is a result of steric factors. Inspection of space-filling diagrams shows that one of the endo oxygen atoms is significantly sterically hindered by the methyl groups of the tripodal amine ligands, thus inhibiting protonation at this site and leading to slower observed rates of hydrolysis. The results obtained in this study are consistent with the endo oxygen atoms being the mechanistically important site of protonation in the acid hydrolysis of metal complexes containing chelated carbonate.  相似文献   

9.
From UV-visible measurements and potentiometric titrations it follows that the lowest pK values (pK1) of imidazolinone herbicides correspond to the simultaneous protonation/dissociation equilibria of both the pyridinic (or quinolinic) nitrogen and the carboxyl group, the following pK (pK2) to the imminium nitrogen and the basic pK (pK3) to the dissociation of the imido nitrogen. Below pH 6 and down to pH c.a. 2.5, the dominant form of the herbicide is a double ion having both positive and negative charges, this being important in discussing the effect of pH in the natural dynamics of imidazolinone herbicides, especially in their soil sorption. Electrochemical studies of the reduction of the herbicides were made on mercury and carbon electrodes in strongly acidic media (0.1 to 2.7 M H2SO4) as well up to pH 7. The reduction signals were all attributed to the reduction of the imidazolinone ring except the second peak/wave that was found to have originated by the reduction of the pyridine/quinoline ring. A signal observed in strongly acidic media and at highly negative overpotentials was attributed to the reduction of the imidazolinone ring of the product of the previous reduction in a process consisting of two reversible electron transfers followed by a protonation reaction.  相似文献   

10.
The structure of copper complexes with phenanthroline in aqueous solutions has been studied by ESR in the X and Q bands at 300 and 77K. It has been shown that the Cu2+ ions are ordered by a cooperative Jahn-Teller exchange interaction when [phen][Cu2+] 2.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 22, No. 1, pp. 99–102, January–February, 1986.  相似文献   

11.
The kinetics of the oxidation of imipramine and desipramine using cerium(IV) complexes were studied in the presence of a large excess of azepine derivative (TCA) in acidic sulfate media using UV-Vis spectroscopy. The reaction proceeds via dibenzoazepine radical formation, identified by EPR measurements. The kinetics of the first degradation step were studied independently of the further slower degradation reactions. Linear dependences, with zero intercept, of the pseudo-first-order rate constants (k(obs)) on [TCA] were established for both dibenzoazepine radical formation processes. Rates of reactions decreased with increasing concentration of the H(+) ion indicating that cerium(IV) as well as both reductants exist in an equilibrium with their protolytic forms. The activation parameters for the degradation of dibenzoazepine derivatives in the first oxidation stage were as follows: ΔH(≠) = 39 ± 2 kJ mol(-1), ΔS(≠) = -28 ± 8 J K(-1) mol(-1) for imipramine and ΔH(≠) = 39 ± 2 kJ mol(-1), ΔS(≠) = -28 ± 6 J K(-1) mol(-1) for desipramine, respectively. Imipramine and desipramine radicals dimerized leading to an intermediate radical dimer, which decayed in a first-order consecutive decay process. These two further reactions proceed with rates which are characterized by non-linear dependences of the pseudo-first-order rate constants (k(obs)) on [TCA]. The degradation reaction of the intermediate radical dimer leads to an uncharged dimer as a final product. Mechanistic consequences of all the results are discussed.  相似文献   

12.
It is shown that stable metal complexes with ammonium pyrrolidine dithiocarbamate (APDC) are formed in strongly acidic (0.5–6 M) solutions and can be extracted into methyl isobutyl ketone (MIBK), although APDC is normally used for extractions from solutions at pH 2–12. Percentage extraction curves are presented for 24 elements (Ag, As, Au, Bi, Cd, Co, Cu, Fe, Ga, Ge, Hg, In, It, Ni, Os, Pb, Pd, Pt, Rh, Ru, Sb, Sn, Tl and Zn) from solutions of hydrochloric or nitric acid with and without addition of APDC. Some elements (e.g., Fe, Ga, Ge, In and Au) show identical extractions as their chloro complexes in hydrochloric acid with or without APDC. Others (e.g., Ni, Cu, Pd, As, Ag, Sb, It, Hg and Bi) are strongly extracted (Kd ? 20), from 2 M hydrochloric or nitric acid in the presence of APDC. Palladium (Kd = 8000), Sb (Kd = 10 000), and Bi (Kd = 3500) are particularly easily extracted. The potential of the extraction system was tested by extraction and quantification of palladium from the CANMET standard ore PTC-1; the mean value found was 12.55 μg g?1 (ppm) palladium with a relative standard deviation of 7.6% (n = 12) and a relative error of 1.2% from the recommended value of 12.70 μg g?1.  相似文献   

13.
The content of reactive groups such as OH, CO, and COOH was increased by modifying hydrolyzed lignin with sulfuric acid and sodium hydroxide. The increase was confirmed by IR spectral analysis. The sorptive capacity of the resulting hydrolyzed lignin derivatives was increased sharply by base activation. __________ Translated from Khimiya Prirodnykh Soedinenii, No. 1, pp. 70–72, January–February, 2006.  相似文献   

14.
Raman spectra of aqueous FeSO4 and (NH4)2SO4 solutions have been recorded over broad concentration and temperature ranges. Whereas the v1-SO 4 2- band profile is symmetrical in noncomplexing (NH4)2SO4 solutions, in FeSO4 solutions a shoulder appears on the high-frequency side, which increases in intensity with increasing concentration and temperature. The molar scattering coefficient of the v1-SO 4 2- band is the same for all forms of sulfate in (NH4)2SO4 and FeSO4 solutions and is independent of temperature up to 150‡C, the highest temperature studied. The high-frequency shoulder is attributed to the formation of a contact ion pair, Fe2+OSO3/2-, as is the splitting of the v3-SO 4 2- antisymmetric stretching mode which is observed in the FeSO4 solution. The bending modes v2-SO 4 2- and v4-SO 4 2- , normally forbidden in the isotropic spectrum, show a gain in intensity with increasing ion-pair formation. A polarized band has been assigned to the Fe2+-O ligand vibration. No higher associates or anionic complexes are required to interpret the spectroscopic data. No evidence of contact ion pairing between Fe2+ and HSO4 4 - could be detected at temperatures up to 303‡C in 1 molal solutions of FeSO4 with an excess of 2 molal H2SO4.  相似文献   

15.
The present investigation reports the first experimental measurements of the reorganization energy of unfolded metalloprotein in urea solution. Horse heart cytochrome c (cyt c) has been found to undergo reversible one-electron transfer reactions at pH 2 in the presence of 9 M urea. In contrast, the protein is electrochemically inactive at pH 2 under low-ionic strength conditions in the absence of urea. Urea is shown to induce ligation changes at the heme iron and lead to practically complete loss of the alpha-helical content of the protein. Despite being unfolded, the electron-transfer (ET) kinetics of cyt c on a 2-mercaptoethanol-modified Ag(111) electrode remain unusually fast and diffusion controlled. Acid titration of ferric cyt c in 9 M urea down to pH 2 is accompanied by protonation of one of the axial ligands, water binding to the heme iron (pK(a) = 5.2), and a sudden protein collapse (pH < 4). The formal redox potential of the urea-unfolded six-coordinate His18-Fe(III)-H(2)O/five-coordinate His18-Fe(II) couple at pH 2 is estimated to be -0.083 V vs NHE, about 130 mV more positive than seen for bis-His-ligated urea-denatured cyt c at pH 7. The unusually fast ET kinetics are assigned to low reorganization energy of acid/urea-unfolded cyt c at pH 2 (0.41 +/- 0.01 eV), which is actually lower than that of the native cyt c at pH 7 (0.6 +/- 0.02 eV), but closer to that of native bis-His-ligated cyt b(5) (0.44 +/- 0.02 eV). The roles of electronic coupling and heme-flattening on the rate of heterogeneous ET reactions are discussed.  相似文献   

16.
The paper presents a detailed experimental and theoretical study of the four mixed nickel-bisdithiolene complexes [Ni(Pr(i)(2)pipdt)(dmit)] (1b, Pr(i)(2)pipdt = 1,4-diisopropyl-piperazine-3,2-dithione; dmit = 1,3-dithiolo-2-tione-4,5-dithiolato), [Ni(R(2)pipdt)(mnt)] (2b", R = 2-ethylhexyl; mnt = maleonitriledithiolato), [Ni(Pr(i)(2)timdt)(dmit)] (3b, Pr(i)(2)timdt = 1,3-diisopropyl-imidazoline-2,4,5-trithione), and [Ni(Pr(i)(2)timdt)(mnt)] (4b), and their models. All the complexes, with common (C(2)S(2))Ni(C(2)S(2)) core and two different terminal groups, are uncharged and square-planar coordinated. Previous measurements of the first molecular hyperpolarizability indicated that some of the species are potential NLO chromophores due to the pi-delocalized character of two frontier levels (HOMO and LUMO) which is asymmetrically perturbed by the combination of one push (R(2)pipdt, R(2)timdt) with one pull ligand (dmit and mnt). The X-ray structure of complex 1b is presented and its geometry is compared with those available in the literature for the four types of complexes under study. The results of electrochemical and spectroscopic measurements (oxidation and reduction potentials, IR, dipole moment, molecular absorptivities, etc.) indicate rather different responses between the pairs of complexes 1-2 and 3-4. Hence, DFT calculations on the model compounds 1a-4a, where hydrogen atoms replace the alkyl groups of R(2)pipdt and R(2)timdt, have been carried out to correlate geometries and electronic structures. Moreover, the first molecular hyperpolarizabilities have been calculated and their components have been analyzed with the simplest two-level approximation. The derived picture highlights the different roles of the two push and pull ligands, but also the peculiar perturbation of the pi-electron density induced by the terminal CS(3) grouping of the ligand dmit.  相似文献   

17.
The interaction of aniline, pyridine and its homologues, phenol and benzamide with a series of neutral polar polymers and polyanions in aqueous solution has been studied by the equilibrium dialysis method. The interaction isotherms were generally sigmoidal. Significant association between polymer and organic molecule was only observed beyond some specific free organic molecule-polymer concentration ratio; i. e., after inter-chain attractive forces and solvation have been partially weakened by preliminary absorption of a surface layer of organic molecules. Benzamide, in fact, shows practically no interaction with neutral ‘coiled’ polymers. Maximum interaction is reached at up to thirty percent utilisation of ‘polar’ sites on the polymer chain. Klotz's method was modified to correct for these weak interactions at low organic molecule concentration. Equilibrium constants and free energies for the main association were then determined. Equilibrium constants are low;K=5–50 (g. moles per l.)?1 for interaction with neutral polymers andK=20–120 (g. moles per l.)?1 for interaction with polyions. The free energies are (?δG)=1–3RT and 3–5RT calories per g. mole respectively. Complexing is due to dipole-dipole and/or ion-dipole type interaction and is greatest for aniline, pyridine and phenol with fully extended polyion chains. Van der Waals' interaction between the aromatic ring and the organic section of the polymer backbone augments the dipole-dipole or ion-dipole binding with pyridine homologues, especially with acridine.  相似文献   

18.
This paper describes the use of an aluminum electrode covered by metallic palladium and modified by Prussian blue prepared by a simple and rapid electroless method for the electro-oxidation of morphine. Two different pathways for electro-oxidation of morphine at various pH ranges were suggested. Also, some thermodynamic and kinetic parameters such as the number of electrons involved in the rate determining step, n α , transfer coefficient α, and the total electrons (n) involved in morphine oxidation at the time scale of the cyclic voltammetric technique, the catalytic rate constant of the electrochemical process k, and diffusion coefficient of morphine D were determined. The mean values obtained are 0.5, 0.5, 1, 26.8 M-1 s-1 and 3.1 × 10−5 cm2 s−1, respectively.  相似文献   

19.
Laser flash photolysis of perfluoro-4-biphenyl azide andN-propyl-4-azido-2,3,5,6-tetrafluorobenzylamide in acetonitrile, water-acetonitrile mixtures, and HCl-containing solutions was studied. The absorption spectra of primary intermediates, singlet arylnitrenes(2a(S) and2b(S), respectively), were recorded. The absolute rate constants of their intersystem crossing in MeCN were measured, and the corresponding Arrhenius parameters were found from the temperature dependences of the rate constants of isomerization of singlet arylnitrenes2a(S) and2b(S) to azirines. Protonation of singlet arylnitrenes2a(S) and2b(S) was observed, the rate constants of their protonation were measured, and the transient absorption spectra of arylnitrenium ions were recorded. It was shown by quantum-chemical calculations (the hybrid B3LYP method) that the arylnitrenium ions that formed have the singlet ground state and the singlet-triplet gap is ∼20 kcal mol−1. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 49–54, January, 2000.  相似文献   

20.
Laser flash photolysis of perfluoro-4-biphenyl azide andN-propyl-4-azido-2,3,5,6-tetrafluorobenzylamide in acetonitrile, water-acetonitrile mixtures, and HCl-containing solutions was studied. The absorption spectra of primary intermediates, singlet arylnitrenes(2a(S) and2b(S), respectively), were recorded. The absolute rate constants of their intersystem crossing in MeCN were measured, and the corresponding Arrhenius parameters were found from the temperature dependences of the rate constants of isomerization of singlet arylnitrenes2a(S) and2b(S) to azirines. Protonation of singlet arylnitrenes2a(S) and2b(S) was observed, the rate constants of their protonation were measured, and the transient absorption spectra of arylnitrenium ions were recorded. It was shown by quantum-chemical calculations (the hybrid B3LYP method) that the arylnitrenium ions that formed have the singlet ground state and the singlet-triplet gap is ∼20 kcal mol−1. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 49–54, January, 2000.  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