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.
The kinetics of the redox reaction between mandelic acid (MA) and ceric sulfate have been studied in aqueous sulfuric acid solutions and in H2SO4? MClO4 (M+ = H+, Li+, Na+) and H2SO4? MHSO4 (M+ = Li+, Na+, K+) mixtures under various experimental conditions of total electrolyte concentration (that is, ionic strength) and temperature. The oxidation reaction has been found to occur via two paths according to the following rate law: rate = k[MA] [Ce(IV)], where k = k1 + k2/(1 + a)2[HSO4?]2 = k1 + k2/(1 + 1/a)2[SO42?]2, a being a constant. The cations considered exhibit negative specific effects upon the overall oxidation rate following the order H+ ? Li+ < Na+ < K+. The observed negative cation effects on the rate constant k1 are in the order Na+ < Li+ < H+, whereas the order is in reverse for k2, namely, H+ ? Li+ < Na+. Lithium and hydrogen ions exhibit similar medium effects only when relatively small amounts of electrolytes are replaced. The type of the cation used does not affect significantly the activation parameters.  相似文献   

2.
The rate of the cerium (IV) oxidation of p-chloromandelic acid has been studied in perchlorate media at an ionic strength of 1.50 mol/dm3 by the stopped-flow technique and in H2SO4? MHSO4 (M+ = Li+, Na+, K+) and H2SO4? MClO4 (M+ = H+, Li+, Na+) mixtures at constant total electrolyte concentrations of 1.00 and 2.00 mol/dm3 using the conventional spectrophotometric method. In perchlorate media the kinetic data indicate the formation of two intermediate complexes between cerium (IV) and the organic substrate, but only one is significantly involved in the intramolecular electron-transfer process. The oxidation rate is markedly lower in sulfate media, where two reaction paths have been found to contribute to the overall redox reaction. The univalent cations examined exhibit negative specific effects upon the overall oxidation rate increasing in the order H+ < Li+ < Na+ < K+. Activation parameters have been also estimated.  相似文献   

3.
In aqueous H2SO4, Ce(IV) ion oxidizes rapidly Arnold's base((p-Me2NC6H4)2CH2, Ar2CH2) to the protonated species of Michler's hydrol((p-Me2NC6H4)2CHOH, Ar2CHOH) and Michler's hydrol blue((p-Me2NC6H4)2CH+, Ar2CH+). With Ar2CH2 in excess, the rate law of the Ce(IV)-Ar2CH2 reaction in 0.100 M H2SO4 is expressed -d[Ce(IV)]/dt = kapp[Ar2CH2]0[Ce(IV)] with kapp = 199 ± 8M?1s?1 at25°C. When the consumption of Ce(IV) ion is nearly complete, the characteristic blue color of Ar2CH+ ion starts to appear; later it fades relatively slowly. The electron transfer of this reaction takes place on the nitrogen atom rather than on the methylene carbon atom. The dissociation of the binuclear complex [Ce(III)ArCHAr-Ce(III)] is responsible for the appearance of the Ar2CH+ dye whereas the protonation reaction causes the dye to fade. In highly acidic solution, the rate law of the protonation reaction of Michler's hydrol blue is -d[Ar2CH+]/dt = kobs[Ar2CH+] where Kobs = ((ac + 1)[H*] + bc[H+]2)/(a + b[H+]) (in HClO4) and kobs= ((ac + 1 + e[HSO4?])[H+] + bc[H+]2 + d[HSO4?] + q[HSO4?]2/[H+])/(a + b[H+] + f[HSO4?] + g[HSO4?]/[H+]) (in H2SO4), and at 25°C and μ = 0.1 M, a = 0.0870 M s, b = 0.655 s, c = 0.202 M?1s?1, d = 0.110, e = 0.0070 M?1, f = 0.156 s, g = 0.156 s, and q = 0.124. In highly basic solution, the rate law of the hydroxylation reaction of Michler's hydrol blue is -d[Ar2CH+]/dt = kOH[OH?]0[Ar2CH+] with kOH = 174 ± 1 M?1s?1 at 25°C and μ = 0.1 M. The protonation reaction of Michler's hydrol blue takes place predominantly via hydrolysis whereas its hydroxylation occurs predominantly via the path of direct OH attack.  相似文献   

4.
The presence of ceric and bromide ions catalyzes the isomerization of maleic acid (MA) to fumaric acid (FA) in aqueous sulfuric acid. A kinetic study of this bromine-catalyzed reaction was carried out. The reaction between ceric ion and maleic acid is first order with respect to Ce(IV). For [Ce(IV)]0=5.0×10?4 M, [H2SO4]0=1.2 M, μ=2.0 M (adjusted by NaClO4), and [MA]0=(0.5–1.0)M, the observed pseudo-first-order rate constant (k03) at 25° is k03=7.622×10?5 [MA]0/(1+0.205[MA]0). The reaction between ceric and bromide ions is first order with respect to Ce(IV). For [Ce(IV)]0=5.0×10?4 M, [H2SO4]0=1.2 M, μ=2.0 M, and [Br?]0=(0.025–0.150)M, the pseudo-first-order rate constant (k02) at 25° is k02= (4.313±0.095)x10?2[Br?]2+(2.060±0.119)x10?3[Br?]. The reaction of Ce(IV) with maleic acid and bromide ion is also first order with respect to Ce(IV). For [Ce(IV)]0=5.0×10?4 M, [MA]0=0.75 M, [H2SO4]0=1.2 M, μ=2.0 M, and [Br?]0= (0.025–0.150)M, the pseudo-first-order rate constant (k03) at 25° is k03= (5.286±0.045)x10?2[Br?]2+(3.568±0.056)x10?3[Br?]. For [Ce(IV)]0=5.0 × 10?4 M, [Br?]0=0.050 M, [H2SO4]0=1.2 M, μ=2.0 M, and [MA]0=(0.15–1.0)M at 25°, k03=(2.108×10?4+2.127×10?4[MA]0)/(1+0.205[MA]0). A mechanism is proposed to rationalize the results. The effect of temperature on the reaction rate was also studied. The energy barrier of Ce(IV)—Br? reaction is much less than that of Ce(IV)—MA reaction. Maleic and fumaric acids have very different mass spectra. The mass spectrum of fumaric acid exhibits a strong metastable peak at m/e 66.5.  相似文献   

5.
The effect of the nature of the exchanged cation M z+ (M z+ = Li+, Na+, Rb+, Cs+, Mg2+, Ca2+, and Ba2+) of a Fiban K-1 fibrous sulfo cation exchanger on the degree of reduction of the immobilized complex cations [Pd(NH3)4]2+ to Pd0 was studied. A linear correlation was found between the degree of palladium reduction and the difference of the relative electronegativities of atoms that participate in the O–M z+ bond. The activity of the catalysts in the oxidation of H2 depends on the degree of palladium reduction.  相似文献   

6.
A comprehensive thermodynamic model based on the electrolyte NRTL (eNRTL) activity coefficient equation is developed for the NaCl + H2O binary, the Na2SO4 + H2O binary and the NaCl + Na2SO4 + H2O ternary. The NRTL binary parameters for pairs H2O-(Na+, Cl) and H2O-(Na+, SO42−), and the aqueous phase infinite dilution heat capacity parameters for ions Cl and SO42− are regressed from fitting experimental data on mean ionic activity coefficient, heat capacity, liquid enthalpy and dissolution enthalpy for the NaCl + H2O binary and the Na2SO4 + H2O binary with electrolyte concentrations up to saturation and temperature up to 473.15 K. The Gibbs energy of formation, enthalpy of formation and heat capacity parameters for solids NaCl(s), NaCl·2H2O(s), Na2SO4(s) and Na2SO4·10H2O(s) are obtained by fitting experimental data on solubilities of NaCl and Na2SO4 in water. The NRTL binary parameters for the (Na+, Cl)-(Na+, SO42−) pair are regressed from fitting experimental data on dissolution enthalpies and solubilities for the NaCl + Na2SO4 + H2O ternary.  相似文献   

7.
From extraction experiments and γ-activity measurements, the exchange extraction constants corresponding to the general equilibrium M+(aq)+NaL+(nb)⇔ML+(nb)+Na+(aq) taking place in the two-phase water-nitrobenzene system (M+ = H+, NH4+, Ag+, Tl+; L = tetramethyl p-tert-butylcalix[4]arene tetraketone; aq = aqueous phase, nb = nitrobenzene phase) were evaluated. Moreover, the stability constants of the ML+ complexes in water saturated nitrobenzene were calculated; they were found to increase in the order Tl+<NH4+<Ag+ <H+ <Na+.  相似文献   

8.
The kinetics of the cerium(IV) oxidation of glycolic acid have been studied in the medium HClO4? Na2SO4? NaClO4 at varying organic substrate (HL), hydrogen, and bisulfate ion concentrations at 25.0°C and ionic strength 2.0M. Under the experimental conditions used (0.03 ? [H+] ? 0.5M; 0.02 ? [HSO4?] ? 0.1M; 0.01 ? [HL] ? 0.1M) the observed pseudo-first-order rate constant kobs has been found to follow the complex expression where the values of the various constants have been estimated by a nonlinear least-squares method. According to this expression the oxidation process occurs significantly through three simultaneous pathways. Moreover three equilibria involving cerium(IV) and HSO4? (or SO42?) ions are important from a kinetic point of view, whereas only two equilibria involving the corresponding complexes with the organic substrate are predominant.  相似文献   

9.
When the sodium ion (Na+) concentration is increased above 0.5 mol-dm−3 (M), the concentrations of dissolved silica in aqueous sodium chloride (NaCl) and sodium nitrate (NaNO3) solutions decrease because of the salting out effect. On the other hand, the concentration of the dissolved silica in aqueous sodium sulfate (Na2SO4) solutions increases monotonously as the concentration of Na+ is increased above 0.5 M. The purpose of this study is to determine the reasons why the salting-out effect is not observed in Na2SO4 solutions. FAB-MS (Fast Atom Bombardment Mass Spectrometry) was used to sample directly the silica species dissolved in aqueous Na2SO4, NaCl, and NaNO3 solutions. In the FAB-MS spectra of these solutions, the peak intensity ratios of the linear tetramer to the cyclic tetramer largely increased for Na+ concentrations between (0.1 and 1) M. This shows that some characteristics of the Na2SO4 solutions are similar to those of the NaCl and NaNO3 solutions. In Na2SO4 solutions, however, when the concentration of Na+ is higher than 1 M, the peak intensity of the dimer is much higher than those of the other silicate complexes. In Na2SO4 solutions, the SO42− ion undergoes partial hydrolysis to form HSO4 and OH is produced. In particular, in the range where the concentration of SO42− is high, the pH of the solution increases slightly. This higher pH yields more dimers from the hydrolysis of silicate complexes. This increase in dimer production agrees with the observation that silica dissolves in sodium hydroxide (NaOH) solutions mainly as a dimer when the concentration of NaOH is less than 0.1 M. In Na2SO4 solutions at high concentrations, a salting-out effect is not observed for silica. This is due to the increase in the concentration of OH, which accelerates the hydrolysis of silica and results in dimer formation.  相似文献   

10.
The mass spectrometry of substituted benzenesulfonylhydrazides (X.C6H4.SO2NHNH2) has been studied, with X = p-CH3, H, p-CH3O and p-Br. The intensities of [X? C6H4SO2]+ and [X? C6H4]+ follow the Hammett relationship [In(Z/Z0) =ρδp+] with ρ of 0.375 and 2.37, respectively.  相似文献   

11.
Studies of the stoichiometry and kinetics of the reaction between hydroxylamine and iodine, previously studied in media below pH 3, have been extended to pH 5.5. The stoichiometry over the pH range 3.4–5.5 is 2NH2OH + 2I2 = N2O + 4I? + H2O + 4H+. Since the reaction is first-order in [I2] + [I3?], the specific rate law, k0, is k0 = (k1 + k2/[H+]) {[NH3OH+]0/(1 + Kp[H+])} {1/(1 + KI[I?])}, where [NH3OH+]0 is total initial hydroxylamine concentration, and k1, k2, Kp, and KI are (6.5 ± 0.6) × 105 M?1 s?1, (5.0 ± 0.5) s?1, 1 × 106 M?1, and 725 M?1, respectively. A mechanism taking into account unprotonated hydroxylamine (NH2OH) and molecular iodine (I2) as reactive species, with intermediates NH2OI2?, HNO, NH2O, and I2?, is proposed.  相似文献   

12.
The rate of oxidation of Ge(II) chloride by large excess of ClO2? ions in HCl, NaCl and Na2SO4 mixed solutions was polarographically observed at various H2O+ and Cl? ion concentrations. The observed rate constant, kobs, is expressed by ko=Kobs/(ClO3?)={k1,(H+)+k2K1(Cl?)2+ K3K2(SO42?)} (H+)/{(H+)1+K1(Cl-)2 +K2(SO42?)} for the following reaction processes, The values were obtained aa k1=1.5410-3liter2 mole2? sec-1, k2=5.00×10-2liter2 mole2? sec-2 and k2=4.30×10-3liter2 mole2? sec-2, K1=1.80× 10-2, K2= 2.43×10-2 mole liter-1 at constant ionic strength I=0.50 M at 30°C.  相似文献   

13.
    
The oxidation of lactic acid, mandelic acid and ten monosubstituted mandelic acids by hexamethylenetetramine-bromine (HABR) in glacial acetic acid, leads to the formation of the corresponding oxoacid. The reaction is first order with respect to each of the hydroxy acids and HABR. It is proposed that HABR itself is the reactive oxidizing species. The oxidation of α-deuteriomandelic acid exhibits the presence of a substantial kinetic isotope effect (k H /k D = 5.91 at 298 K). The rates of oxidation of the substituted mandelic acids show excellent correlation with Brown’s σ+ values. The reaction constants are negative. The oxidation exhibits an extensive cross conjugation between the electron-donating substituent and the reaction centre in the transition state. A mechanism involving transfer of a hydride ion from the acid to the oxidant is postulated.  相似文献   

14.
The reaction of N-nitro-O-(4-nitrophenyl)hydroxylamine (1) with conc. H2SO4 affords 4-nitropyrocatechol and that with conc. sulfonic acids (RSO3H where R = Me, CF3) affords 2-hydroxy-5-nitrophenyl-R-sulfonates in yields of 80?C85%. These reactions are assumed to proceed through an intermediate (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+, which eliminates the N2O molecule to form the aryloxenium ion [NO2C6H4O]+. The latter reacts with acid anions at the ortho-carbon atom of the phenyl ring. The thermodynamical parameters of the elementary reactions resulting in the formation of the (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+ and aryloxenium ion [NO2C6H4O]+ were calculated in the B3LYP/6?311+G(d) study of the combined molecular system (nitrohydroxylamine 1 + [H3SO4]+). The reaction of nitrohydroxylamine 1 with aqueous solutions of strong acids (??70% H2SO4, CF3SO3H) affords mainly 4-nitrophenol. It appears that the mechanism of this reaction does not involve the formation of the aryloxenium ion.  相似文献   

15.
Gas-phase ion–molecule reactions of transition metal ions, M+ (M+ = Ni+, Co+, Fe+ and Mn+), with six aromatic ring-containing nitriles were investigated in a modified fast atom bombardment (FAB) source. It is shown that the monoadduct, (Ph(CH2)nCN)–M+, is one of the most abundant ion–molecule reaction products. The main fragments in the FAB source are the [C7H7]+ and [C8H9]+ ions, and their formation is shown to involve metal ion insertion into the nitriles rather than direct bond cleavage from the ‘free’ or complexed nitriles after FAB ionization. An intramolecular oxidation–reduction reaction, giving [C7H7]+, is found in the metastable and collisionally induced dissociations of benzyl nitrile adducts accompanied by neutral MCN formation, but not seen for longer chain samples. An ortho effect is observed in the elimination of HCN from the 2-methylbenzyl nitrile adduct ions. This reaction dominates the metastable ion spectrum of the adduct of Mn+, whereas metal detachment is nearly the major process for the other complexes of Mn+. The different bond-insertion selectivities of the metal ions are also shown.  相似文献   

16.
The oxidation of glycolic, lactic, malic, and a few substituted mandelic acids by 2,2′‐bipyridinium chlorochromate (BPCC) in dimethylsulphoxide leads to the formation of corresponding oxoacids. The reaction is first order each in BPCC and the hydroxy acids. The reaction is catalyzed by the hydrogen ions. The hydrogen ion dependence has the form: kobs = a + b [H+]. The oxidation of α‐deuteriomandelic acid exhibited a substantial primary kinetic isotope effect (kH/kd = 5.29 at 303 K). Oxidation of p‐methylmandelic acid was studied in 19 different organic solvents. The solvent effect has been analyzed by using Kamlet's and Swain's multiparametric equations. A mechanism involving a hydride ion transfer via a chromate ester is proposed. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 248–254, 2002  相似文献   

17.
The kinetics of the Fe3+/Fe2+ reaction on a Pt rotating disc electrode was studied in solutions of 0.5 M H2SO4 and 0.5 M Na2SO4 (pH 2.2). Taking into account formation of sulphate complexes the conclusion was made that the main contribution to the reaction rate is due to FeSO4+ and FeSO4 complexes. Extended Tafel plots obtained by Randles analysis from experimental current-voltage curves were corrected for the 2 potential. The latter was evaluated according to the Gouy-Chapman theory by using the surface charge density values deduced from thermodynamic theory and measurements of other authors. Tafel plots were approximated by parabolas and the reorganization energy was calculated as 33 kJ mol?1 and 51 kJ mol?1 for Fe3+/Fe2+ in H2SO4 and Na2SO4, respectively. The comparison of these values with theoretically predicted ones was made. From the magnitude of the pre-exponential factor of the true rate constant it was concluded that the Fe3+/Fe2+ electron transfer reaction is non-adiabatic in nature.  相似文献   

18.
In this work, seven inorganic salts, KCl, Na2SO4, MgSO4-7H2O, ZnCl2, Na2CrO4, CuSO4-5H2O, and K3[Fe(CN)6], were used as catalysts to induce chemiluminescent luminol oxidation in alkaline aqueous media. It was observed that simple salts containing either Mg2+, Zn2+, Na+ and K+ cations or SO42– and Cl anions, are not active as catalysts. On the other hand, the relative order of activity detected for the active chemiluminescent salts containing Fe(III), Cu(II) and Cr(VI) cations is K3[Fe(CN)6] < CuSO4-5H2O < Na2CrO4. The intensity of the emitted light agrees with the standard reduction potentials of the corresponding redox couple and with the presence of paramagnetic species in the aqueous solutions. The inhibition effect of mannitol was also studied.  相似文献   

19.
The solid‐liquid equilibria in the quinary system Na+, K+//Cl?, SO2?4, B4O2?7‐H2O at 298 K had been studied experimentally using the method of isothermal solution saturation. Solubilities and densities of the solution of the quinary system were measured experimentally. Based on the experimental data, the dry‐salt phase diagram and water content diagram of the quinary system were constructed, respectively. In the equilibrium diagram of the quinary system Na+, K+//Cl?, SO2?4, B4O2?7‐H2O at 298 K, there are five invariant points F1, F2, F3, F4 and F5; eleven univariant curves E1F1, E2F2, E3F3, E4F5, E5F2, E6F4, E7F5, F1F4, F2F4 F1F3 and F3F5, and seven fields of crystallization saturated with Na2B4O7 corresponding to Na2SO4, Na2SO4·10H2O, Na2SO4·3K2SO4 (Gla), K2SO4, K2B4O7·4H2O, NaCl and KCl. The experimental results show that Na2SO4·3K2SO4 (Gla), K2SO4 and K2B4O7·4H2O have bigger crystallization fields than other salts in the quinary system Na+, K+//Cl?, SO2?4, B4O2?7‐H2O at 298 K.  相似文献   

20.
Summary The kinetics of oxidation of ,-unsaturated alcohols (UA's), such as prop-2-ene-1-ol, but-2-ene-1-ol and 3-phenyl-prop-2-ene-1-ol, by manganese(III) acetate in aqueous H2SO4 at constant ionic strength and different acidities has been studied. The reaction was found to proceed through an outer sphere mechanism. The reactions were first order with respect to [MnIII] and fractional order in [UA]. The reaction showed first order dependence in [H+], and the rate decreased on addition of [MnII]. Added salts, such as Na2SO4, had a negligible effect on the rate. The data suggested that disproportionation of the MnIII-UA complex into free radicals was the rate determining step in the presence of [MnII]. A mechanism consistent with the experimental data is proposed. The activation parameters have been evaluated for the temperature range 298–313 K.  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