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.
Nuclear magnetic resonance line-widths data have been used to determine the rate of solvent exchange from the first coordination sphere of ferro-and ferriprotoporphyrin(IX) dimethylester (Fe-PPD) in pyridine/chloroform. The average values of kinetic parameters for pyridine (PY) exchange indicate an SN2 mechanism tor Fe(III)-PPD(ΔH&;#; = 36 kJ · mol−1 ; ΔS&;#; = −53 J·mol−1K−1; TM(298 K) = 0.07 msec) and an SNI mechanism for Fe(II)-PPD (ΔH&;#; = 67 kJ·mol−1; ΔS&;#; = 42 J · mol−1K−1; TM(298 K) = 0.06 msec). Parallel to the accelerated ligand exchange rate at rising temperatures a redistribution of the electrons causing a transition of the metal porphyrin from the low-spin state to the high-spin state is observed. Enthalpy and entropy of the thermodynamic equilibrium between low- and high-spin Fe-PPD have been determined from experimental values of the average magnetic moment. A mean lifetime of low-spin Fe(III)-PPD was estimated from line. widths changes (TL→H(298 K)≈ 20 msec) and the corresponding activation parameters have been obtained (ΔH&;#;L→H(298 K) = 26 kJ · mol−1; ΔS&;#;L→H(298K) = −125 J · mol−1K−1).  相似文献   

2.
《Inorganica chimica acta》1987,128(2):169-173
The axial adduct formation of the iron(II) complex of 2,3,9,10-tetraphenyl-l,4,8,11-tetraaza-1,3,8,10-cyclotetradecatetraene (L) with imidazole in dimethyl sulfoxide has been investigated spectrophotometrically at various temperatures and pressures. In the presence of a large excess of imidazole the reaction with the two phases has been observed. The first faster reaction is the formation of the monoimidazole complex of FeL2+, and the second slower reaction corresponds to the formation of the bisimidazole complex. Activation parameters are as follows: for the first step with k1 (25.0°C) = (6.8 ±0.2)×105 mol−1 kg s−1, ΔH31 = 47.5 ± 4.9 kJ mol−1, ΔS31 = 26±16 J K−1 mol−1, and ΔV31 (30.0°C) = 27.2±1.5 cm3 mol−1; for the second step with k2 (25.0°C) = 26.8±0.8 mol−1 kg s−1, ΔH32 = 91.6± 0.8 kJ mol−1, ΔS32 = 90±3 J K−1 mol−1, and ΔV32 (35.0°C) = 21.8±0.9 cm3 mol−1. The large positive activation volumes strongly indicate a dissociative character of the activation process.  相似文献   

3.
The outer sphere reductions of Co(NH3)5B3+ by Fe(CN)5A3− have been studied. The observed pseudo first order rate constants (Co complex in excess) obey the dependence kobs=Kosket[Co]/(1 +Kos[Co]), as expected for outer sphere electron transfer reactions. Values of the fundamental electron transfer rate constants ket have been determined, along with the equilibrium constant Kos for a range of reactions in which A and B are pyridyl ligands of different sizes. The first order electron transfer rate constants vary in a manner that is consistcnt with adiabatic electron transfer. The outer sphere ion pairing equilibrium constants Kos have been calculated: Kos=8.6 ± 0.1 × 102 M−1 when A and B=pyridine; Kos=1.07 ± 0.09 × 103 M−1 where A=pyridine, B=1-phenyl-3-(4-pyridyl)propane; Kos=1.86 ± 0.11 × 103 M−1 when A=4,4′-bipyridine, B=pyridine; Kos=1.27 ± 0.08 × 103 M−1 when A=4,4′-bipyridine, B=4-phenylpyridine. Distances of closest approach between the metal centers in the reactive ion pairs are compared, and it is concluded that there is a common mechanism, in which the ammonia side of the cobalt complex approaches the cyano side of the iron complex in each reactive ion pair.The distance of closest approach between the two metal centers (a) was calculated from the experimental values for the ion pairing equilibrium constant Kos at 25 °C: 5.2 Å when A=4,4′-bipyridine, B=pyridine; 5.4 Å when A=4,4′-bipyridine, B=4-phenylpyridine; 5.5 Å when A=pyridine, B=1-phenyl-3-(4-pyridyl)propane; 5.7 Å when A=B=pyridine. These relatively short metal-metal distances, when compared to the X-ray structure of the compound [Co(NH3)5(4-phenylpyridine)]2[S2O6]3· 4H2O, do not support an ion pair orientation in which the two substituted pyridine ligands A and B are oriented toward each other. [P21/c,a=7.399(3), b=22.355(10), c=13.776(4) Å, β=92.02(3)°, R=0.070.] The crystallographic results show that if the two pseudo-octahedral coordination spheres are oriented in the reactive ion pair so that an ammonia face of the cobalt complex is at hydrogen bonding distance from a cyano face on the iron complex, the metal-metal distance is 5.3 Å, a distance which is in agreement with the kinetic results.  相似文献   

4.
《Inorganica chimica acta》1988,149(1):151-154
The extraction equilibrium of the hydronium-uranium(VI)-dicyclohexano-24-crown-8 complex was carried out in the crown ether1,2-dichloroethaneHCl aqueous solution system at different temperatures. The extraction complex has the overall composition (L)2·(H3O+·χH2O)2·UO2Cl42− (L = dicyclohexano-24-crown-8). The values of the extraction equilibrium constants (Kex) increase steadily with a decrease in temperature: 13.5 (298 K), 7.96 (301 K), 4.20 (303 K) and 2.07 (305 K). A plot of log Kex against 1/T shows a straight line. The value of the enthalpy change, ΔH°, was calculated from the slope and equals −212 kJ mol−1. The value of the entropy change, ΔS°, was calculated from ΔH° and Kex and equals −690 J K−1 mol−1, whereas ΔG° = −6.45 kJ mol−1. Comparing these thermodynamic parameters with those of the dicyclohexano-18-crown-6 isomer A [1] (ΔS° = −314 J K−1 mol−1, ΔH° = −101 kJ mol−1 and ΔG° = −8.37 kJ mol−1), it can be seen that ΔH° and ΔS° are more negative for the former than for the latter, and both are enthalpy-stabilized complexes. The molecular structure of the complex has the feature that there are two H5O2+ ions in it, in contrast to the H3O+ ions in the dicyclohexano-18-crown-6 isomer A complex [1]. Each of the H5O2+ ions is held in the crown ether cavity by four hydrogen bonds. The H5O2+ ion has a central bond. The uranium atom forms UO2Cl42− as a counterion away from the crown ether. The formation of this complex is in good agreement with more negative entropy change and less negative free energy change, as mentioned above.  相似文献   

5.
《Inorganica chimica acta》1988,145(2):211-217
The hydrolysis of the ester 2,4-dinitrophenyl- ethyl methylphosphonate has been examined by both stop-flow spectrophotometric and pH-stat techniques. These reactions have been carried out in the presence of several nucleophiles including simple non-labile (w.r.t. substitution) mono-aquo metal ion complexes. Comparison of reaction rates of the metal complexes with sterically hindered organic nucleophiles has led to the conclusion that the metal ions function predominantly as general base catalysts in dilute aqueous solution. Reaction rates for the various nucleophiles studied are tabulated together with solvolysis constants for hydroxide ion and water at 35 °C and I=0.1 mol dm−3 (KNO3). These later two values are respectively 32.7 mol−1 dm3 s−1 and 1.37 x 10−4 s−1. A Brönsted β value of 0.52 for the phosphonate ester studied has also been derived.  相似文献   

6.
The kinetics and mechanism of the oxidation of L- ascorbic acid by trisoxalatocobaltate(III) were studied as a function of pH, ascorbate concentration, ionic strength and temperature in a weakly basic aqueous solution. The pH dependence of the process can be ascribed to the oxidation of the doubly deprotonated ascorbate ion for which k = 20 M−1 s−1 at 25 °C, ΔH# = 34 ± 2 kJ mol−1 and ΔS# = −108 ± 7 J K−1 mol−1. The results are discussed in reference to literature data for this reaction in weakly acidic medium and for the oxidation by a series of other oxidants.  相似文献   

7.
《Inorganica chimica acta》1986,120(2):131-134
The equilibrium, kinetics and mechanism of the reaction of chromium(III) with pentane-2,4-dione (Hpd) have been investigated in aqueous solution at 55°C and ionic strength 0.5 mol dm−3 NaClO4. The equilibrium constant (log β1) is 10.08(±0.01) while the pK of Hpd is 8.69(±0.01). The kinetics are consistent with a mechanism in which [Cr(H20)6]3+ and [Cr(H20)5(OH)]2+ react with the enol tautomer of Hpd with rate constants of 1.05(±0.26) × 10−2 and 2.78(±0.08) × 10−1 dm3 mol−1 s−1 respectively. These rate constants are considerably more rapid than those predicted by the Eigen-Wilkins mechanism. These data are compared with literature values.  相似文献   

8.
The enthalpies of the hexokinase-catalyzed phosphorylation or glucose, mannose, and fructose by ATP to the respective hexose 6-phosphates have been measured calorimetrically in TRIS/TRIS HCl buffer at 25.0, 28.5, and 32.0°C. The effects on the measured enthalpy of the glucose/hexokinase reaction due to variation of pH (over the range 6.7 to 9.0) and ionic strength (over the range 0.02 to 0.25) have been examined. Correction for enthalpy of buffer protonation leads to δHo and δCpo values for the processes: eq-D-hexose + ATP4− = eq-D-hexose 6-phosphate2− + ADP3−+ H+. Results are δHo = −23.8 ± 0.7 kJ · mol−1 and δCpo = −156 ± 280 J·mol−1·K−1 for glucose. δHo = −21.9 ± 0.7 kJ·mol−1 and δCpo = 10 ± 140 J·mol−1·K−1 for mannose, and δHo = −15.0 ± 0.9 kJ·mol−1 and δCpo = −41 ± 160 J·mol−1·K−1 for fructose. Combination of these measured enthalpies with Gibbs energy data for hydrolysis of ATP4− and that for the hexose 6-phosphates lead to δSo values for the above hexokinase-catalyzed reactions.  相似文献   

9.
The kinetics of malonate replacement in bis- (malonato)oxovanadate(IV), [VO(mal)2H2O]2−(hereafter water molecule will be omitted), by oxalate has been studied by the stopped-flow method. The reaction was found to consist of two consecutive steps (k1 and k2: first-order rate constants) passing through a mixed ligand complex, [VO(mal)(ox)]2−. The rates for each step depended linearly on the concentrations of free oxalate species, Hox and ox2−. The second-order rate constants for the replacement by ox2− were much larger in the k1 step than in the k2 step and the activation parameters were determined as follows: ΔH= 43.5 ± 5.6 kJ mol−1, ΔS±-53 ± 19 J K−1 mol−1 and ΔH≠= 43.6 ± 0.5 kJ mol−1, δS≠ = -62 ± 2 J K−l mol−1 for the k1 and k2 steps, respectively. The volume of activation was determined to be -0.65 ± 0.75 cm3 mol−1 at 20.2 °C by the high-pressure stopped-flow method for the apparent rate constants.  相似文献   

10.
The reactions of PtCl2en or cis-Pt(NH3)2Cl2 and their aqua species with adenine and adenosine were studied by means of ion-pair HPLC. From the chromatograms, it was found that the first binding site of Pt(II) was the N(7) site of adenine under both acidic and neutral conditions. The rates of Pt(II) binding at the (N7) site of adenosine and deoxyadenosine were measured. The rate constants, k1, were obtained for the reactions of PtCl2en or cis-Pt(NH3)2Cl2 with adenosine and deoxyadenosine at pH 3 and 7 over the temperature range 9–25 °C. The k1 values were 6.8–7.7 × 10−4 dm3 mol−1 s−1 at 25 °C. For the aqua species, the rate of [cis-Pt(NH3)2ClH2O]+ with adenosine N(7) was measured. The rate constants, k2 which were found to be smaller than those of hydrolysis, kh, were calculated at pH 3 over the temperature range 25–40 °C. The k2 value obtained at 25 °C was 1.1 × 10−2 dm3 mol−1 s−1, 15 time larger than k1. The activation parameters were also calculated.  相似文献   

11.
《Inorganica chimica acta》1987,130(2):183-184
cis,cis,trans-[PtIV(NH3)2Cl2(OH)2] reacts reversibly with ascorbic acid to give dehydroascorbic acid and mainly cis-[PtII(NH2Pri)2Cl2]. The parameters for the forward reaction are: kf = 0.584 M s at 37.0 °C, ΔHf = 108.6 −+ 6.4 kJ mol−1 andΔSf = 101 −+ 22 J K−1 mol−1.  相似文献   

12.
《Inorganica chimica acta》1988,154(2):209-214
The diastereoisomeric complex Δ-(+)-tris(cyclicO,O′, 1 (R), 2(R)(−)dimethylethylene dithiophosphato)chromium(III), was synthesized stereoselectively in tetrahydrofuran (THF) solution. The complex proves optically labile, [α]D=+106, in CHCl3, changing quickly to [α]D=+211. The CD spectra in THF enable us to characterize the complex and show a configuration inversion which gives the diastereoisomeric equilibrium Λ⇌Δ with an excess of the Λ-(R,R)(R,R)(R,R) diastereoisomeric form. The equilibrium constant K=0.86 at 25 °C is indicative of a different thermodynamic stability between the two diastereoisomers in THF solution, Λ-(R,R)> Δ-(R,R), δΔH°=1.5 kJ mol−1, δΔG°=0.3 kJ mol−1, δΔS°=4 J mol−1 K−1. The kinetic diastereoisomer Δ-(R,R)(R,R)(R,R) is stabilized in CHCl3, CH2Cl2, EtOH solvents where it is highly soluble and optically stable with a maximum negative chirality factor, g=−5×10−3, in CHCl3.  相似文献   

13.
《Inorganica chimica acta》1988,148(2):233-240
The complexes CodptX3 and [Codpt(H2O)X2]ClO4 (X = Cl, Br; dpt = dipropylenetriamine = NH(CH2CH2CH2NH2)2) have been prepared and characterized. Rate constants (s−1) for aqueous solution at 25 °C and μ = 0.5 M (NaClO4), for the acid-independent sequential ractions.
have been measured spectrophotometrically. For X = Cl: k1 ⋍ 2 × 10−2, k2 = 1.7 × 10−4 and k3 = 4.8 × 10−6, and for X = Br: k1 ⋍ 2 × 10−2, k2 = 5.25 × 10−4 and k3 = 2.5 × 10−5 The primary equation was found to be acid independent, while the secondary and tertiary aquations were acid-inhibited reactions. For the second step, the rate of the reaction was given by the rate equation
where Ct is the complex concentration in the aqua-and hydroxodihalo species, k2 is the rate constant for the acid-dependent pathway and Ka is the equilibrium constant between the hydroxo and aqua complex ions. The activation parameters were evaluated, for X = Cl: ΔH2 = 106.3 ± 0.4 kJ mol−1 and ΔS2 = 40.2 ± 1.7 J K−1 mol, and for X = Br: ΔH2 = 91.6 ± 0.4 kJ mol−1 and ΔS2 = 0.4 ± 1.7 J K−1 mol−1. The results are discussed and detailed comparisons of the reactivities of these complexes with other haloaminecobalt(III) species are presented.  相似文献   

14.
《Inorganica chimica acta》1986,115(1):95-100
Racemic mer-[CoCl(en)(NH2CH2CHNCH2CH2NH2)]ZnCl4, which contains no dissymetric chelate rings, no asymmetric carbon centers and no asymmetric nitrogen centers, has been resolved using sodium arsenic(III)—(+)-tartrate. The chirality arises by virtue of the coordinated unsymmetric tridentate ligand and the less soluble diastereoisomeride is associated with the (+)-cation. The rate of base hydrolysis was measured spectrophotometrically using tris buffers. Kinetic parameters (25 °C, \3m; ∼ 0.04 M) are kOH = 1.28 X 103 M−3 s−1, Ea = 87.0 ± 0.7 kJ mol−1 and ΔS#; = +98 ± 1.4 J K−1 mol−1. Complete racemisation accompanies the base hydrolysis reaction and the rate of loss of optical activity is 0.5 times that of base hydrolysis. These data are interpreted in terms of the formation of a symmetrical trigonal bipyramid intermediate generated from the conjugate base.  相似文献   

15.
Binding of [3H]aflatoxin B1 to rat plasma was investigated in vivo and vn vitro. Column chromatographic and polyacrylamide gel electrophoretic analysis clearly demonstrated that aflatoxin B1 bound primarily plasma albumin. Very little binding activity was shown by other plasma proteins. Spectrofluorimetric studies were undertaken to gain some insight into the nature of the aflatoxin-albumin interaction. Quenching of the lone tryptophan fluorescence intensity upon aflatoxin binding was due, at least in part, to a ligand-induced conformational change in the albumin molecule. Aflatoxin B1 binds an apolar site with an association constant of 30 mM−1 at pH 7.4 and 20°C. Neither charcoal treatment of rat albumin nor the presence of 0.15 M NaCl had many significant effect on the interaction. The association constant was pH-dependent, increasing about 1.7-fold as the pH increased from 6.1 to 8.4. This pH dependence is ascribed to a pH-induced conformational change in the albumin molecule. Thermodynamic studies indicated that the aflatoxin-albumin interaction was exothermic (ΔH = −29.3 kJ·mol), with a ΔS value of −13.8 J·mol−1·K−1.  相似文献   

16.
A detailed investigation on the oxidation of aqueous sulfite and aqueous potassium hexacyanoferrate(II) by the title complex ion has been carried out using the stopped-flow technique over the ranges, 0.01≤[S(IV)]T≤0.05 mol dm−3, 4.47≤pH≤5.12, and 24.9≤θ≤37.6 °C and at ionic strength 1.0 mol dm−3 (NaNO3) for aqueous sulfite and 0.01≤[Fe(CN)6 4−]≤0.11 mol dm−3, 4.54≤pH≤5.63, and 25.0≤θ≤35.3 °C and at ionic strength 1.0 or 3.0 mol dm−3 (NaNO3) for the hexacyanoferrate(II) ion. Both redox processes are dependent on pH and reductant concentration in a complex manner, that is, for the reaction with aqueous sulfite, kobs={(k1K1K2K3+k2K1K4[H+])[S(IV)]T]/([H+]2+K1[H+]+K1K2) and for the hexacyanoferrate(II) ion, kobs={(k1K3K4K5+k2K3K6[H+])[Fe(CN)6 4−]T)/([H+]2+K3[H+]+K3K4). At 25.0 °C, the value of k1′ (the composite of k1K3) is 0.77±0.07 mol−1 dm3 s−1, while the value of k2′ (the composite of k2K4) is (3.78±0.17)×10−2 mol−1 dm3 s−1 for aqueous sulfite. For the hexacyanoferrate(II) ion, k1′ (the composite of k1K5) is 1.13±0.01 mol−1 dm3 s−1, while the value of k2′ (the composite of k2K6) is 2.36±0.05 mol−1 dm3 s−1 at 25.0 °C. In both cases there was reduction of the cobalt(III) centre to cobalt(II), but there was no reduction of the molybdenum(VI) centre. k22, the self-exchange rate constant, for aqueous sulfite (as SO3 2−) was calculated to be 5.37×10−12 mol−1 dm3 s−1, while for Fe(CN)6 4−, it was calculated to be 1.10×109 mol−1 dm3 s−1 from the Marcus equations.  相似文献   

17.
《Bioorganic chemistry》1987,15(1):31-42
The use of NAD+ analogs lacking a carbonyl function at position C-3 of the pyridinium moiety allowed the manipulation of the kinetic mechanism of calf spleen NAD+ glycohydrolase so as to render the cleavage of the pyridinium-ribose bond rate limiting. The analogs used in this study are relatively poor substrates of the enzyme. They present an affinity for the active site which is independent of the nature of their substituent (Ki = 10 ± 2 μm), suggesting that the specificity of the NAD+ glycohydrolase reflects the dynamic steps occurring after the formation of the Michaelis complex. The maximal rates of hydrolysis of the NAD+ analogs are very sensitive to the pKa of the departing pyridine; a Brönsted plot (r = 0.99) gave a βtg = −0.90 (at 37°C). From this plot we could estimate that for NAD+, the specific interaction of the 3-carboxamide group with the active site contributed to the catalysis by decreasing the energy barrier by about 2 kcal mol−1. We have also studied the nonenzymatic hydrolysis of NAD+ and its analogs under conditions (pH-independent hydrolysis) which favor a unimolecular mechanism. In this case a linear Brönsted plot was also found (r = 0.99) with βtg = −1.11 (at 37°C). Our data indicate that NAD+ glycohydrolase catalyzes the chemical cleavage of the pyridinium-ribose bond, over a 103 rate difference, according to a single mechanism involving a late transition state in which the scissile bond is broken. The present study strongly supports our previous hypothesis (F. Schuber, P. Travo, and M. Pascal (1979) Bioorg. Chem. 8, 83) according to which NAD+ glycohydrolase catalyses unimolecular decomposition of its substrates with generation of an ADP-ribosyl oxocarbonium ion intermediate which must be stabilized by the active site of the enzyme.  相似文献   

18.
《Bioorganic chemistry》1987,15(2):100-108
Nonenzymatic rates of hydrolytic deamination of adenosine and cytidine by acids and bases analogous to side chains of naturally occurring amino acids are compared with the rates of uncatalyzed deamination in water and with the rates of the hydroxide- and hydrogen ion-catalyzed reactions. For adenosine, hydroxide ion is an effective catalyst, with a second-order rate constant of 7.5 × 10−6 m−1 s−1 at 85°C and an energy of activation of 19.9 kcal/mol. Acid-catalyzed deamination of adenine proceeds with a second-order rate constant of 1.5 × 10−6 m−1 s−1 at 85°C. At concentrations of 1 m and at pH values corresponding to their respective pKa values, dimethylamine, acetate, selenide, imidazole, phosphate, and zinc(II) do not enhance the rate of deamination of adenosine beyond that observed in water, and 2-mercaptoethanol produces only a modest rate enhancement. The uncatalyzed rate of adenosine deamination in water is 8.6 × 10−9 s−1 at 85°C: extrapolation to 37°C and comparison with kcat for rat hepatoma adenosine deaminase yield a rate enhancement by the enzyme of approximately 2 × 1012-fold. 1,6-Dimethyladenosine, the conjugate acid of which has a pKa value much higher than that of adenosine, is not readily deaminated, suggesting that the uncatalyzed deamination of adenosine does not proceed by hydroxide ion attack on the rare protonated form of adenosine, but rather by attack on the neutral species. Deamination of cytidine is catalyzed most effectively by hydroxide ion, with a second-order rate constant of 4.5 × 10−4 m−1 s−1 at 85°C and an energy of activation of 28.5 kcal/mol. The uncatalyzed rate of deamination of cytidine in water, which also exhibits an energy of activation of 28.5 kcal/mol, is 8.8 × 10−8 s−1 at 85°C. Comparison of the rate extrapolated to 25°C with kcat for bacterial cytidine deaminase gives a rate enhancement for the enzyme of 4 × 1011-fold. The C-5 proton of the pyrimidine ring of cytidine does not exchange with solvent during alkaline hydrolysis, suggesting that deamination under these conditions does not involve prior addition of water across the 5,6 double bond.  相似文献   

19.
Aldehyde dehydrogenases (ALDHs) are a diverse family of enzymes that catalyze the NAD(P)+-dependent detoxification of toxic aldehyde compounds. ALDHs are also involved in non-enzymatic ligand binding to endobiotics and xenobiotics. Here, the enzyme crucial non-canonical and non-catalytic interaction with kolaflavanone, a component of kolaviron, and a major bioflavonoid isolated from Garcinia kola (Bitter kola) was characterized by various spectroscopic and in silico approaches under simulated physiological condition. Kolaflavanone quenched the intrinsic fluorescence of ALDH in a concentration dependent manner with an effective quenching constant (Ksv) of 1.14 × 103 L.mol−1 at 25 °C. The enzyme has one binding site for kolaflavanone with a binding constant (Ka) of 2.57 × 104 L.mol−1 and effective Forster resonance energy transfer (FRET) of 4.87 nm. The bonding process was enthalpically driven. The reaction was not spontaneous and was predominantly characterized by Van der Waals forces and hydrogen bond. The flavonoid bonding slightly perturbed the secondary and tertiary structures of ALDH that was ‘tryptophan-gated’. The interaction was regulated by both diffusion and ionic strength. Molecular docking showed the binding of kolaflavanone was at the active site of ALDH and the participation of some amino acid residues in the complex formation with −9.6 kcal mol−1 binding energy. The profiles of atomic fluctuations indicated the rigidity of the ligand-binding site during the simulation. With these, ALDH as a subtle nano-particle determinant of kolaviron bioavailability and efficacy is hereby proposed.  相似文献   

20.
The kinetics of the base hydrolysis of the complex ion chloropentaamminecobalt(III) have been studied by conventional spectrophotometry at 25.0 °C in water and in the presence of the anionic surfactant sodium dodecyl sulfate (SDS) over the SDS concen- tration range from 1.0 × 10−3 to 7.5 × 10−2 mol dm−3. The hydrolysis rate is strongly inhibited by the surfactant, fitting a model in which the cobalt- (III) complex is distributed between water and the micellar pseudo-phase with a binding constant equal to 3.7 × 103 dm3 mol−1. The effects of different added electrolytes on the critical micelle concentra- tion of the surfactant and on the hydrolysis rate have also been investigated and discussed.  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