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 78 毫秒
1.
The reactions of Fe(CN)5dpa3? and Ru(NH3)5dpa2+ (dpa = 4,4′-dipyridylamine) with Co(edta)? have been investigated kinetically. For Fe(CN)5dpa3? complex, a linear relationship was observed between the pseudo-First-order rate constants and the concentrations of Co(edta) which leads to a specific rate 0.876 ± 0.006 M?1S?1 at T = 25°C., μ = 0.10 M and pH = 8.0. For the Ru(NH3)5dpa2+ system, the plots kobs vs [Co(edta)?] become nonlinear at concentrations of Co(edta) greater than 0.01 M and the reaction is interpreted on the basis of a mechanism involving the formation of an ion pair between Ru(NH3)5dpa2+ and Co(edta)? followed by electron transfer from Ru(II) to Co(III). The nonlinear least squares fit of the kinetic results shows that Qip = 10.6 ± 0.7 M?1 and ket = 93.9 ± 0.7 s?1 at pH = 8.0,μ = 0.10 M and T = 25°C.  相似文献   

2.
The isomerization of the complex trans-meso-CH3Co(H2O)L2+ (L = 5,7,7,12,14,14-hexamethyl-1,4,8,11-tetraazacyclotetradeca-4,11-diene) to trans-primary, rac-CH3Co(H2O)L2+ has been investigated from pH range 7.11 to 8.09 in aqueous solution. The reaction rate law has been determined as: -d[meso-CH3Co(H2O)L2+]/dt = kOH [OH?][meso-CH3Co(H2O)L2+], where kOH = 600 ± 10 M?1s?1 at 25 °C and μ = 0.5 M. The activation parameters of the reaction were also studied with ΔH± = 19.1 ± 0.9 Kcal mol?1 and ΔS± = 18.0 ± 0.8 cal K?1mol?1. A mechanism that involves a secondary NH inversion is proposed.  相似文献   

3.
The anilinepentacyanoferrate (II) complex has been characterized in aqueous solution. The complex exhibits a predominant ligand field transition at λmax = 415 nm with ?max = 494 M?1 cm?1. The corresponding Fe(III) complex displays a strong absorption at λmax = 700nm(?max = 1.61×104 M?1 sec?1) which can be assigned as a ligand to metal charge transfer transition. The rate constants of formation and dissociation for the Fc(II) complex are (3.14±0.18)×102 M?1W?1 and 0.985±0.005 sec?1, respectively, at μ = 0.10 M LiClO4, pH = 8 and T = 25°C. The cyclic voltammetry of the complex shows that a reversible redox process is observed with E1/2 value of 0.51±0.01 V vs. NHE at μ = 0.10 M LiClO4, pH = 8 and T = 25°C. The kinetic study of the oxidation of the Fe(II) complex by ferricyanide ion yielded the rate constant of the reaction ket = (1.43±0.04)x10 M sec?1 at μ = 0.10 M LiClO4, pH = 8 and T = 25°C.  相似文献   

4.
Kinetic measurements for the forward reaction Fe(CN)54-AmPy3? + Co(edta)? ? Fe(CN)5s4-AmPy2? + Co(edta)2? have been carried out; the rate constant is 2.72 ± 0.07 M?1s?1, at pH = 8, μ = 0.10 M LiClO4, and T = 25°C. The activation parameters of the reaction were also studied with and . The mechanism of the reaction is discussed in the context of the Marcus cross relation for an outer-sphere process.  相似文献   

5.
J.G. Leipoldt  H. Meyer 《Polyhedron》1985,4(9):1527-1531
The reaction of Cl?, Br?, I?, Co(CN)63? and NCS? with meso-tetrakis (p-trimethylammoniumphenyl)porphinatodiaquorhodate(III), [RhTAPP(H2O)2]5+, has been studied at 15, 25 and 35°C in 0.1 M [H+] with μ = 1.00 M (NaNO3). The value of the acidity constant, Kal, at 25°C is 4.39 × 10?9 M. The reactions are first order in anion concentration up to 0.9 M. The values of the stability constants, K1, and the second order rate constants, k1, for the reaction with Cl?, Br?, I?, Co(CN)63? and NCS? are respectively 0.23 M?1 and 2.5 × 10?3 M?1 s?1, 1.1 M?1 and 6.92 × 10?3 M?1 s?1, 40.0 M?1 and 17.0 × 10?3 M?1 s?1, 550 M?1 and 20.0 × 10?3 M?1 s?1, 3400 M?1 and 20.9 × 10?3 M?1 s?1. The porphine greatly labilizes the Rh(III). There has been about a 500-fold increase in the rate constant for substitution compared to that of [Rh(NH3)5H2O]3+. The substitution rates are however about the same as for [Rh(TPPS)(H2O)2]3?, indicating that the overall charge on the complex plays only a minor role. The kinetic results indicate that dissociative activation is occurring in these reactions.  相似文献   

6.
The metallothermic reduction of praseodymium tribromide, PrBr3, with lithium metal (molar ratio 1:1) in sealed tantalum containers at 850°C yields bronze lustrous rods of Pr2Br5. The crystal structure (monoclinic, P21/m, Z = 2, a = 774.38(4), b = 415.33(3), c = 1327.06(9) pm, β = 90.816(6)°, Vm = 128.52(2) cm3 mol?1) contains seven- and eight-coordinate trivalent praseodymium in mono- and bicapped trigonal prisms. Pr2Br5 should therefore be formulated as (Pr3+)2(Br?)5(e?). It is isostructural with Pr2I5 and is believed to be identical with PrBr2,38 reported for the Pr/PrBr3 system.  相似文献   

7.
The equilibrium constants of the reactions MBr2(s) + Al2Br6(sln) ? MAl2Br8(sln) M = Cr, Mn, Co, Ni, Zn, Cd have been measured at 298 K in toluene. Ni: 0.017 ± 0.0024, Co: 0.54 ± 0.07, Zn: 1.5 ± 0.2, Mn: 2.1 ± 0, 7, Cr: 2.2 ± 1, Cd: 7 ± 5. They are compared with literature values of the equilibrium constants of analogous reactions in the gas phase MX2(s) + Al2X6(g) ? MAl2X8(g), X = Cl, Br. For CoAl2Br8(sln) the temperature dependence of the equilibrium constant yielded ΔfH = ?9.4 ± 1 kJ mol?1 and ΔfS = ?39.5 ± 3 J mol?1 K?1 while literature values for CoAl2Br8(g) are ΔfH = 42.4 ± 2 kJ mol?1 and ΔfS = 42.9 ± 2 J mol?1 K?1. The solubility of Al2Br6 in toluene as well as its enthalpy of dissolution have been measured in order to evaluate ΔH° and ΔS° of the solvation of Al2Br6(g) and CoAl2Br8(g) in toluene by a thermodynamic cycle. Solvation of Al2Br6(g): ΔH = ?72.7 ± 1 kJ mol?1, ΔS = ?139.6 ± 4 J mol?1 K?1, solvation of CoAl2Br8(g): ΔH = ?124.5 ± 4kJ mol?1, ΔS = ?222 ± 9J mol?1 K?1. Thus, CoAl2Br8 interacts more strongly with the solvent toluene than Al2Br6 does.  相似文献   

8.
Redox reactions of Co(edta)? with Ru(NH3)5L2+ (L = 3- and 4-aminopyridine (AmPy)) were found to follow an outer-sphere electron transfer mechanism. The specific rate constants are (3.26 ± 0.03) × 102 and (3.07 ± 0.04) × 103 M?1S?1, for L = 3- and 4-AmPy, respectively, at μ, = 0.10 M LiClO4, pH = 8.0 (tris) and T = 25 °C. The rate constants of oxidations for a series of Ru(NH3)5L2+ complexes are higher than those of the corresponding Fe(CN)5L3- complexes by factors of 4 to 15 even after corrections for differences in reduction potentials and in charges of the complexes. Nonadiabaticity in the reactions of Fe(CN)5L3 complexes may account for the difference in the relative reactivities.  相似文献   

9.
High-pressure 1H-NMR. has been used to determine volumes of activation (ΔV#) for solvent exchange with [M(S)6]3+ ion (M = Al(III), Ga(III); S = dimethylsulfoxide (DMSO) and N,N-dimethylformamide (DMF)) in [2H]3-nitromethane solution. For Al(III),Δ V# = + 15.6 ± 1.4 (S = DMSO, 358.5 K) and ΔV# = + 13.7 ± 1.2 cm3mol?1 (S = DMF, 354.5 K), whilst for Ga(III), ΔV# = + 13.1 ± 1.0 (S = DMSO, 334.6 K) and ΔV# = +7.9 ± 1.6 cm3mol?1 (S= DMF, 313.8 K). Variable temperature studies over a temperature range of 107.2 K (Al(III)) and 101.1 K (Ga(III)) were carried out for solvent exchange with [M(DMF)6]3+ ions in [2H]3-nitromethane solution, using stopped-flow NMR, and conventional linebroadening, and gave ΔH# = 88.3 ± 0.9 and 85.1 ± 0.6 kJ+ mol?1, and ΔS# = 28.4 ± 2.7 and 45.1 ± 1.9 JK?1 mol?1 for Al(III) and Ga(III) ions respectively. All of these results are consistent with dissociative modes of activation.  相似文献   

10.
Properties indirectly determined, or alluded to, in previous publications on the titled isomers have been measured, and the results generally support the earlier conclusions. Thus, the common five‐coordinate intermediate generated in the OH?‐catalyzed hydrolysis of exo‐ and endo‐[Co(dien)(dapo)X]2+ (X=Cl, ONO2) has the same properties as that generated in the rapid spontaneous loss of OH? from exo‐ and endo‐[Co(dien)(dapo)OH]2+ (40±2% endo‐OH, 60±2% exo‐OH) and an unusually large capacity for capturing (R=[CoN3]/[CoOH][]=1.3; exo‐[CoN3]/endo‐[CoN3]=2.1±0.1). Solvent exchange for spontaneous loss of OH? from exo‐[Co(dien)(dapo)OH]2+ has been measured at 0.04 s?1 (k1, 0.50M NaClO4, 25°) from which similar loss from the endo‐OH isomer may be calculated as 0.24 s?1 (k2). The OH?‐catalyzed reactions of exo‐ and endo‐[Co(dien)(dapo)N3]2+ result in both hydrolysis of coordinated via an OH?‐limiting process =153 M ?1 s?1; =295 M ?1 s?1; KH=1.3±0.1 M ?1; 0.50M NaClO4, 25.0°) and direct epimerization between the two reactants =33 M ?1 s?1; =110 M ?1 s?1; 1.0M NaClO4, 25.0°). Comparisons are made with other rapidly reacting CoIII‐acido systems.  相似文献   

11.
The crystal structure of racemic [Co(NSSSN)Cl](ClO4)Cl was determined by X-ray diffraction methods. It crystallizes in the monoclinic system, space group P21/c, with cell constants of a = 9.795(3), b = 10.412(3) and c = 16.323(8) Å, and β = 93.87(4)°; V = 1661 Å3d (meas.; flotation) = 1.85 gm-cm?3, d (calc.; Z = 4 molecules/unit cell) = 1.88 gm-cm?3. The molecules, a racemic mixture, have the absolute configurations λλδλ or δδλδ at each of the four five-membered rings and resemble, in general, the so called αα conformer already described by Snow1 in his study of the Co(tetraen)Cl2+ cation. However, the torsional angles at C2, C3 and C8, C9 in the two terminal C-C-NH2 fragments are quite different in the two systems. For Co(tetraen)Cl2+ they are 44.7° and ?20.2° respectively, whereas for Co(NSSSN)C2+ the values –52.3° and –44.6° obtain. Also, the ring Co-S1-C3-C2-N1 does not have the classical, low energy conformation found in Co(tetraen)Cl2+. The presence of the larger Co-S bonds causes the two terminal -NH2 groups to be pushed toward each other, and to minimize steric hindrance between adjacent -NH2 hydrogens and ligand twists C2 down and staggers the terminal hydrogens. We visualize the propagation of these distortion effects in solution as being transferred from one side to the other across the entire ligand chain with concomittant effects on the activation of the precursor complex in electron transfer reactions, resulting in ~107 rate enhancement over the Co(tetraen)Cl2 system. Kinetic data for the reduction of Co(NSSSN)X2+ and Co(NSNSN)X2+ (X = Cl?, Br?) by Fe(II) is also presented and discussed.  相似文献   

12.
On dehydration of La[Co(CN)6]·5H2O, the color of the complex, changes from white to pale blue at around 230°C. Heating the pale blue specimen, the color changes to deep blue at around 290°C. This deep blue specimen is easily rehydrated to a pink one. As reported previously, in the pale blue specimen, Co3+ ions are situated in the center of the D4h crystal field formed by six CN- ions. The deep blue specimen is due to the presence of [Co(CN)4]2- ions in which Co2+ was situated in a Td coordination field formed by four CN- ions and the Co-C bond length is 1.67 Ĺ. The pink species corresponded to trans-[Co(CN)4(H2O)2]2- and the bond lengths of Co-C and Co-O are 1.89 and 1.85 Ĺ, respectively. The Raman spectra of the complex observed at 25°C displays two bands at 2157 and 2176 cm-1 associated with the vibration of C-N bond, and the band of 2157 cm-1 was split into two bands, 2150 and 2156 cm-1, at around 100°C. When the complex was heated to around 230°C, three new bands were observed at 2103, 2116 and 2141 cm-1. The bands of 2103 and 2116 cm-1 were assigned to the stretching vibration of C=N bonding to Co2+. The band of 2141 cm-1 was assigned to the stretching vibration of the inverted CN- as follows: Co-C=N-La→Co-N=C-La. The activation energy for the inversion of CN- was estimated as 67 kJ mol-1. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

13.
Kinetics of the Aquation [Co(NH3)5DMSO](ClO4)3 · 2 H2O The aquation rate constants of (dimethylsulfoxide)pentaamminecobalt(III)perchlorate in aqueous perchloric acid media has been determined spectrophotometrically under various conditions of acidity and complex concentrations at 25–50°C. The reaction proceeds by an first-order rate law presumably with D-mechanism, and is independently of the acidity. The values for the activation enthalpy and entropy has been calculated: ΔH≠ = 24,7 kcal mol?1; ΔS≠ = 4,5 cal K?1 mol?1.  相似文献   

14.
The equilibrium quotients for the formation of Co(NH3)5Cl2+ from Co(NH3)5OH23+ and Cl? were 3.74±0.25 M?1 and 6.07±0.54 M?1 at 45.0°C in 10:1 mole ratio water: dimethyl sulfoxide and in 25 w/w % aqueous ethanol, respectively, and those forthe formation of the ion pair Co(NH3)5OH23+ . Cl? were 1.21±0.20 M?1 and 1.58±0.17 M?1, respectively, in the same solvents. The aquation and anation rateconstants were determined at 45.0°C for these two solvents over the range of chloride-ion concentrations 0.0 ≤ [Cl?] ≤ 0.9 M. The aquation rate constant was essentially independent of chloride-ion concentration in each solvent over this range. The inverse of the pseudo-first-order anation rate constant was linearly dependent on the inverse of the chloride-ion concentration in each solvent. The least squares relationships between (1/kan) and (1/[Cl?]) gave intercepts and ratios of intercept to slope which were analyzed interms of Id and D mechanisms. It was concluded that the data were not satisfied by a D mechanism, but that they were consistent with an Id mechanism.  相似文献   

15.
The microwave spectra of 32SPHF2, 34SPHF2 and 32SPDF2 have been analyzed. The structural parameters obtained from this analysis are: d(S-P) = 1.867±0.005Å, d(P-F) = 1.551 ±0.005Å, (P-H) = 1.392±0.005Å, ∠SPF = 117.4 ± 0.2 °, ∠SPH = 119.2±0.2 °, ∠FPF = 98.6±0.2 °. Centrifugal distortion coefficients were obtained for 32SPHF2. The spectra of two vibrational excited states of 32SPHF2 were observed. The two sets of rotational constants (A) 8336.72, 3726.70, 2807.56 MHz and (B) 8344.88, 3727.73, 2798.75 MHz were associated with the vibrational states with measured infrared frequencies 419 cm?1 and 344 cm?1 respectively. An analysis of the infrared spectrum is included. Dipole moment measurements yielded μ = 1.87±0.03 D for 32SPHF2 and μ = 1.86±0.03 D for 32SPDF2  相似文献   

16.
The kinetics and mechanism of Hg2+‐catalyzed substitution of cyanide ion in an octahedral hexacyanoruthenate(II) complex by nitroso‐R‐salt have been studied spectrophotometrically at 525 nm (λmax of the purple‐red–colored complex). The reaction conditions were: temperature = 45.0 ± 0.1°C, pH = 7.00 ± 0.02, and ionic strength (I) = 0.1 M (KCl). The reaction exhibited a first‐order dependence on [nitroso‐R‐salt] and a variable order dependence on [Ru(CN)64?]. The initial rates were obtained from slopes of absorbance versus time plots. The rate of reaction was found to initially increase linearly with [nitroso‐R‐salt], and finally decrease at [nitroso‐R‐salt] = 3.50 × 10?4 M. The effects of variation of pH, ionic strength, concentration of catalyst, and temperature on the reaction rate were also studied and explained in detail. The values of k2 and activation parameters for catalyzed reaction were found to be 7.68 × 10?4 s?1 and Ea = 49.56 ± 0.091 kJ mol?1, ΔH = 46.91 ± 0.036 kJ mol?1, ΔS = ?234.13 ± 1.12 J K?1 mol?1, respectively. These activation parameters along with other experimental observations supported the solvent assisted interchange dissociative (Id) mechanism for the reaction. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 215–226, 2009  相似文献   

17.
At bromide concentrations higher than 0.1 M, a second term must be added to the classical rate law of the bromate–bromide reaction that becomes ?d[BrO3?]/dt = [BrO3?][H+]2(k1[Br?] + k2[Br?]2). In perchloric solutions at 25°C, k1 = 2.18 dm3 mol?3 s?1 and k2 = 0.65 dm4 mol?4 s?1 at 1 M ionic strength and k1 = 2.60 dm3 mol3 s?1and k2 = 1.05 dm4 mol?4 s?1 at 2 M ionic strength. A mechanism explaining this rate law, with Br2O2 as key intermediate species, is proposed. Errors that may occur when using the Guggenheim method are discussed. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 39: 17–21, 2007  相似文献   

18.
Cobalt (II) phthalocyanine tetracarboxylate [Co (II)Pc-COOH] has been prepared and used in aqueous solutions as a novel chromogenic reagent for the spectrophotometric determination of cyanide ion. The method is based on measuring the increase in the intensity of the monomer peak in the reagent absorbance at 682 nm due to the formation of a 1 : 2 [Co (II)Pc-COOH] : [CN] complex. The complex exhibits a molar absorptivity (ε) of 7.7 × 104 L mol?1 cm?1 and a formation constant (Kf ) of 5.4 ± 0.01 × 106 at 25 ± 0.1°C. Beer's law is obeyed over the concentration range 0.15–15 µg mL?1 (5.8 × 10?6–5.8 × 10?4 M) of cyanide ion, the detection limit is 20 ng mL?1 (7.7 × 10?7 M) the relative standard deviation is ±0.7% (n = 6) and the method accuracy is 98.6 ± 0.9%. Interference by most common ions is negligible, except that by sulphite. The proposed method is used for determining cyanide concentration in gold, silver and chromium electroplating wastewater bath solutions after a prior distillation with 1 : 1 H2SO4 and collection of the volatile cyanide in 1 M NaOH solution containing lead carbonate as recommended by ASTM, USEPA, ISO and APAHE separation procedures. The results agree fairly well with potentiometric data obtained using the solid state cyanide ion selective electrode.  相似文献   

19.
The reductions of Co(terpy)23+ and Co(edta)? complexes by ascorbic acid have been subjected to a detailed kinetic study in the range of pH =1–10.9. For each complex the rate law of the reaction is interpreted as a rate determining reaction between Co(III) complex and the ascorbic acid in the form of HA? (k1) and A2? (k2), depending on the pH of the solution, followed by a rapid scavenge of the ascorbic acid radicals by Co(III) complex. With given Ka1 and Ka2, the rate constants are k1 = 0.25 and 9.87 × 10?5 M?1s?1, k2 = 1.28 × 106 and 18.7 M?1s?1 for Co(terpy)23+and Co(edta)? complexes, respectively, at T = 25 °C and μ = 0.50M (terpy)and 1.0 M (edta) HClO4/LiClO4. The mechanism of the reaction is discussed on the basis of Marcus theory for outer sphere electron transfer process. Spin change and charge effect, duly considered, account for the non‐adiabatic behavior in the reduction of Co(edta)? complex.  相似文献   

20.
IR spectra of 2-haloethanols (CH2XCH2OH, X = Cl, Br, and I) in carbon disulfide were measured at 25°C up to 2.5 kbar. The volume changes accompanying the transformation to the Gg conformer of the compounds were ?1.2, +0.5, and +1.3 cm3 mol?1 for X = Cl, Br, and I, respectively.  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