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.
The removal of excess reagent extracted into an organic phase in the solvent extraction of a metal complex anion with a quaternary ammonium ion is discussed. With a given chelating ligand (HO—R—SO3H), the order of extractability is HO—R—SO3- > M(OR—SO3)n- > X- > -O—R—SO3- when an anion such as nitrate or halide is added. If suitable amounts of the anion are added, only the excess of reagent can be removed. The principle is applied to the extraction with zephiramine of the cobalt complex anion formed with 2-nitroso-1-naphthol-4-sulfonic acid. Micro amounts of cobalt in pure nickel salts were determined spectrophotometrically.  相似文献   

2.
The rates of formation of 1-(4′-chlorophenylazo-)-2-naphthol by SO3 of release from the π-complex [2-naphthol-1-sulfonate anion …? p-chlorobenzenediazonium cation] have been measured in chloroform and in methylene chloride; they are first-order with respect to the complex. They are catalysed by pyridine and co-catalysed by acetic acid. Acids alone, in stoichiometric or higher concentrations, inhibit the reaction. A mechanism is postulated involving proton transfer to the sulfonate group, followed by rearrangement to the σ-complex which, in the catalysed reaction, first loses a proton to the base and then releases SO3, but in the uncatalysed reaction loses SO3H. The function of the co-catalyst (pyridinium ion) is explained (see text). This reaction is - in contrast to electrophilic aromatic substitutions in which the leaving group is a protonan SN1-type re-aromatization of the σ-complex to the product.  相似文献   

3.
A new spectrophotometric method for the determination of gallium is described using 1-(2,4-dihydroxyphenylazo) -2-naphthol-4-sulfonic acid (DHPAN) as a reagent. The color reaction has a sensitivity of 0.013 μg Ga per cm2 for log 10/1=0.005 at 500 mμ and obeys Beer's law up to 2.8 p.p.m. The effects of pH, time, order of addition of the reagents, temperature and diverse ions were investigated. Gallium is separated from interfering ions by solvent extraction.  相似文献   

4.
The oxidation of D ‐mannitol by cerium(IV) has been studied spectrophotometrically in aqueous sulfuric acid medium at 25°C at constant ionic strength of 1.60 mol dm?3. A microamount of ruthenium(III) (10?6 mol dm?3) is sufficient to enhance the slow reaction between D ‐mannitol and cerium(IV). The oxidation products were identified by spot test, IR and GC‐MS spectra. The stoichiometry is 1:4, i.e., [D ‐mannitol]: [Ce(IV)] = 1:4. The reaction is first order in both cerium(IV) and ruthenium(III) concentrations. The order with respect to D ‐mannitol concentration varies from first order to zero order as the D ‐mannitol concentration increases. Increase in the sulfuric acid concentration decreases the reaction rate. The added sulfate and bisulfate decreases the rate of reaction. The active species of oxidant and catalyst are Ce(SO4)2 and [Ru(H2O)6]3+, respectively. A possible mechanism is proposed. The activation parameters are determined with respect to a slow step and reaction constants involved have been determined. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 440–452, 2010  相似文献   

5.
The rates of substitution of the group X in 1-X-2-naphthol-6-sulfonic acids (X = H, Cl, Br, and I) by p-chlorobenzenediazonium ions in aqueous solution have been measured. The rates of the halogenated naphthols relative to that of the parent compound (X = H) are 0.0070:0.0089:0.149 for X = Cl, Br, and I respectively. The reaction of 1-bromo-2-naphthol-6-sulfonic acid is catalysed by thiosulfate ions; the relative rate observed for this compound does not, therefore, represent the ipso factor. It is postulated that in its substitution the release of the electrofugal leaving group (Br) is rate-limiting.  相似文献   

6.
The reaction kinetics of autocatalytic oxidation of thiamine hydrochloride (vitamin B1) by the permanganate ion in aqueous sulfuric acid medium has been investigated spectrophotometrically under the pseudo–first‐order condition at 25°C. The observed stoichiometry is 6:5 in terms of the mole ratio of permanganate ions and thiamine hydrochloride. Formation of products was confirmed by UV–vis, IR, GC‐MS, and NMR spectral data. Usually in the permanganate oxidation–reduction reactions, one of the products, Mn2+ autocatalyzes the reaction, but in the present investigation the autocatalytic effect is due to the (4‐methyl–thiazol‐5‐yl) acetic acid, a product formed from the oxidation of vitamin B1, which is a rare case. The added Mn2+ does not have any significant effect on the rate of reaction. The reaction is first order with respect to both permanganate and thiamine hydrochloride concentrations. An increase in the sulfuric acid concentration decreases the rate of reaction. A composite scheme and rate laws were proposed. The activation parameters with respect to the slow step and reaction constants involved in the mechanism were determined and discussed.  相似文献   

7.
Three sulfonated 1-(2-pyridylazo)-2-naphthols and six sulfonated 2-(2-pyridylazo)-1-naphthols were synthesized, and their application to the spectrophotometric determination of metals was studied. The acidity constants of the reagents and the stability constants of the nickel chelates are reported, and the relationship between their properties and the position of the sulfonic acid group is discussed. 1-(2-Pyridylazo)-2-naphthol-6-sulfonic acid (PAN-6S) and 1-(2-pyridylazo)-2-naphthol-7-sulfonic acid (PAN-7S) are sensitive and selective reagents for nickel. The determination of nickel in the presence of cobalt with PAN-6S is described. Extraction of the chelate as the ion-pair with tetraphenylarsonium ions into chloroform is suitable for the determination of 1–10 μg Ni at 570 nm; the molar absorptivity is 56 000 l mol-1 cm-1, and interferences are easily avoided.  相似文献   

8.
The oxidation of fluorene by vanadium(V) in aqueous acetic acid containing sulfuric acid (1.0M) at 50°C produces fluorenone and 2-hydroxy diphenyl 2′-carboxaldehyde. The order with respect to each reactant is found to be 1. The order dependence on sulfuric acid concentration is 4, indicating that V(OH)23+ could be the active species. An increase in the acetic acid percentage in the solvent medium increases the rate of the reaction. The effect of solvent variation has been discussed in the light of the acidity function and the polarity of the medium. The effect of substituents on the rate has been studied for seven substituted fluorenes, and a linear relationship exists between log k versus σ values with the slope ρ = -3.2. A small isotope effect is observed for the oxidation of the parent compound (kH/kD = 1.2). The effect of temperature on the rate of the reaction has been studied, and the activation parameters are discussed. A mechanism involving the rate-limiting formation of a cation-radical intermediate is proposed.  相似文献   

9.
On treatment with HSO3F in SO2C1F at 0°, 3-hydroxy-2,2-dimethyl-3-phenyl-propionic acid ( 1a ) is transformed into 2-phenyl-3-methyl-2-butenoic acid ( 2a ) (isolated yield: 40–44%). Using monolabelled [3-13C]- 1a ( 1a *) and doubly labelled [1,3-13C2]- 1a ( 1a **), the migration of HOOC (or a mechanistically equivalent group) was proved; a cross experiment established the intramolecular character of the rearrangement. By following the reaction at low temperature in an NMR. spectrometer, the formation of intermediates and side products was demonstrated.  相似文献   

10.
The kinetics and mechanism of the reaction between nitrous acid and bromine are studied in dilute sulfuric acid medium, using both the stopped‐flow method and conventional spectrophotometry. The partial reaction order with respect to Br2 moderately differs from 1, showing a saturation at a higher concentration of bromine. The second order of the reaction towards nitrous acid has been observed. Hydrogen and bromide ions significantly suppress the rate of reaction. Despite the apparent simplicity, the mechanism is rather complex, with two reaction pathways proposed. The first one is represented by the reaction of bromine with the intermediate dinitrogen trioxide. A direct nucleophilic attack of NO2 ion towards the bromine molecule is suggested as the second pathway. The proposed mechanism accounts for the observed behavior; in almost all cases a satisfactory quantitative agreement with the experiments is obtained. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 279–285, 2000  相似文献   

11.
The reaction kinetics for the oxidation of l ‐histidine by permanganate ions have been investigated spectrophotometrically in sulfuric acid medium at constant ionic strength and temperature. The order with respect to permanganate ions was found to be unity and second in acid concentration, whereas a fractional order is observed with respect to histidine. The reaction was observed to proceed through formation of a 1:1 intermediate complex between oxidant and substrate. The effect of the acid concentration suggests that the reaction is acid catalyzed. Increasing the ionic strength has no significant effect on the rate. The influence of temperature on the rate of reaction was studied. The presence of metal ion catalysts was found to accelerate the oxidation rate, and the order of effectiveness of the ions was Cu2+ > Ni2+ > Zn2+. The final oxidation products were identified as aldehyde (2‐imidazole acetaldehyde), ammonium ion, manganese(II), and carbon dioxide. Based on the kinetic results, a plausible reaction mechanism is proposed. The activation parameters were determined and discussed with respect to a slow reaction step.  相似文献   

12.
Investigations on Aromatic Amino-Claisen Rearrangements The thermal and acid catalysed rearrangement of p-substituted N-(1′,1′-dimethylallyl)anilines (p-substituent=H (5) , CH3 (6) , iso-C3H7 (7) , Cl (8) , OCH3 (9) , CN (10) ), of N-(1′,1′-dimethylallyl)-2,6-dimethylaniline (11) , of o-substituted N-(1′-methylallyl)anilines (o-substituent=H (12) , CH3 (13) , t-C4H9 (14) , of (E)- and (Z)-N-(2′-butenyl)aniline ((E)- and (Z)- 16 ), of N-(3′-methyl-2′-butenylaniline (17) and of N-allyl- (1) and N-allyl-N-methylaniline (15) was investigated (cf. Scheme 3). The thermal transformations were normally conducted in 3-methyl-2-butanol (MBO), the acid catalysed rearrangements in 2N -0,1N sulfuric acid. - Thermal rearrangements. The N-(1′,1′-dimethylallyl)anilines rearrange in MBO at 200-260° with the exception of the p-cyano compound 10 in a clean reaction to give the corresponding 2-(3′-methyl-2′-butenyl)anilines 22–26 (Table 2 and 3). The amount of splitting into the anilines is <4% ( 10 gives ? 40% splitting). The secondary kinetic deuterium isotope effect (SKIDI) of the rearrangement of 5 and its 2′,3′,3′-d3-isomer 5 amounts to 0.89±0.09 at 260° (Table 4). This indicates that the partial formation of the new s?-bond C(2), C(3′) occurs already in the transition state, as is known from other established [3,3]-sigmatropic rearrangements. The rearrangement of the N-(1′-methylallyl)anilines 12–14 in MBO takes place at 290–310° to give (E)/(Z)-mixtures of the corresponding 2-(2′-Butenyl)anilines ((E)- and (Z)- 30,-31 , and -32 ) besides the parent anilines (5–23%). Since a dependence is observed between the (E)/(Z)-ratio and the bulkiness of the o-substituent (H: (E)- 30 /(Z)- 30 =4,9; t-C4H9: (E)- 32 /(Z)- 32 =35.5; cf. Table 6), it can be concluded, that the thermal amino-Claisen rearrangement occurs preferentially via a chair-like transition state (Scheme 22). Methyl substitution at C(3′) in the allyl chain hinders the thermal amino-Claisen-rearrangement almost completely, since heating of (E)-and (Z)- 16 , in MBO at 335° leads to the formation of the expected 2-(1′-methyl-allyl) aniline (33) to an extent of only 12 and 5%, respectively (Scheme 9). The main reaction (?60%) represents the splitting into aniline. This is the only observable reaction in the case of 17 . The inversion of the allyl chain in 16 - (E)- and (Z)- 30 cannot be detected - indicated that 33 is also formed in a [3, 3]-sigmatropic process. This is also true for the thermal transformation of N-allyl- (1) and N-allyl-N-methylaniline (15) into 2 and 34 , respectively, since the thermal rearrangement of 2′, 3′, 3′-d3- 1 yields 1′, 1′, 2′-d3- 2 exclusively (Table 8). These reaction are accompanied to an appreciable extent by homolysis of the N, C (1′) bond: compound 1 yields up to 40% of aniline and 15 even 60% of N-methylaniline ((Scheme 10 and 11). The activation parameters were determined for the thermal rearrangements of 1, 5, 12 and 15 in MBO (Table 22). All rearrangements show little solvent dependence (Table 5, 7 and 9). The observed ΔH values are in the range of 34-40 kcal/mol and the ΔS values very between -13 to -19 e.u. These values are only compatible with a cyclic six-membered transition state of little polarity. - Acid catalysed rearrangements. - The rearrangement of the N-(1′, 1′-dimethylallyl) anilines 5-10 occurs in 2N sulfuric acid already at 50-70° to give te 2-(3′-methyl-2′-butenyl)anilines 22-27 accompanied by their hydrated forms, i.e. the 2-(3′-hydroxy-3′-methylbutyl) anilines 35-40 (Tables 10 and 11). The latter are no more present when the rearrangement is conducted in 0.1 N sulfuric acid, whilst the rate of rearrangement is practically the same as in 2 N sulfuric acid (Table 12). The acid catalysed rearrangements take place with almost no splitting. The SKIDI of the rearrangement of 5 and 2′, 3′, 3′-d3- 5 is 0.84±0.08 (2 N H2SO4, 67, 5°, cf. Table 13) and thus in accordance with a [3,3]-sigmatropic process which occurs in the corresponding anilinium ions. Consequently, the rearrangement of a 1:1 mixture of 2′, 3′, 3′-d3- 5 and 3, 5-d2- 5 in 2 N sulfuric acid at 67, 5° occurs without the formation of cross-products (Scheme 13). In the acid catalysed rearrangement of the N-1′-methylallyl) anilines 12-14 at 105-125° in 2 N sulfuric acid the corresponding (E)- and (Z)-anilines are the only products formed (Table 14 and 15). Again no splitting is observed. Furthermore, a dependence of the observed (E)/(Z) ratio and the bulkiness of the o-substituent ( H : (E)/(Z)- 30 = 6.5; t- C 4 H 9: (E)- 32 /(Z)- 32 = 90; cf. Table 15) indicates that also in the ammonium-Claisen rearrangement a chair-like transition state is preferentially adopted. In contrast to the thermal rearrangement the acid catalysed transformation in 2 N-O, 1 N sulfuric acid (150-170°) of (E)- and (Z)- 16 as well as of 1 and 15 , occurs very cleanly to yield the corresponding 2-allylated anilines 33, 2 and 34 (Scheme 15 and 18). The amounts of the anilines formed by splitting are <2%. During longer reaction periods hydration of the allyl chain of the products occurs, and in the case of the rearrangement of (E)- and )Z)- 16 the indoline 45 is formed (Scheme 15 and 18). All transformations occur with inversion of the allyl chain. This holds also for the rearrangement of 1 , since 3′, 3′-d2- 1 gives only 1′, 1′-d2- 2 (Scheme 17). The activation parameters were determined for the acid catalysed rearrangement of 1, 5, 12 and 15 in 2 N sulfuric acid (Table 22). The ΔH values of 27-30 kcal-mol and the ΔS values of +9 to -12 e.u. are in agreement with a [3, 3]-sigmatropic process in the corresponding anilinium ions. The acceleration factors (kH+/kΔ) calculated from the activation parameters of the acid catalysed and thermal rearrangements of the anilines are in the order of 105 - 107. They demonstrate that the essential driving force of the ammonium-Claisen rearrangement is the ‘delocalisation of the positive charge’ in the transition state of these rearrangements (cf. Table 23). Solvation effects in the anilinium ions, which can be influenced sterically, also seem to play a role. This is impressively demonstrated by N-(1′, 1′-dimethylallyl)-2, 6-dimethylaniline (11) : its rearrangement into 4-(1′, 1′-dimethylallyl)-2, 6-dimethylaniline (43) cannot be achieved thermally, but occurs readily at 30° in 2 N sulfuric acid. From a preparative standpoint the acid catalysed rearrangement in 2 N-0, 1 N sulfuric acid of N-allylanilines into 2-allylanilines, or if the o-positions are occupied into 4-allylanilines, is without doubt a useful synthetic method (cf. also [17]).  相似文献   

13.
Summary The reaction between hydroxopentaaquochromium(III) and octacyanomolybdate(IV) was investigated spectrophotometrically and obeyed a 2:1 reactant stoichiometry with respect to formation of the [Cr(H2O)4OH]2 Mo(CN)8 complex. Kinetic studies reveal that the reaction is first order with respect to hydroxopentaaquochromium(III) in the presence of an excess of octacyanomolybdate(IV). The reaction rate increased with an increase in the ionic strength and temperature, and decreased with an increase in hydrogen ion concentration. A mechanism has been proposed based upon ion-pair formation. The results are best accounted for by the Eigen-Tamm mechanism. Anation of [Cr(H2O)5OH]2+ is discussed in terms of an associative interchange (I a) where bond breaking and bond making are equally important. The activation parameters were calculated using Arrhenius's equation.  相似文献   

14.
Kinetics of the oxidation of tris(2,2′-bipyridine)iron(II) sulfate by ceric sulfate was spectrophotometrically studied in an aqueous sulfuric acid medium. Different methods, including isolation, integration and half-life, were employed to determine the reaction order. The redox reaction was found to be first-order with respect to the reductant, tris(2,2′-bipyridine)iron(II) sulfate, and the oxidant, ceric sulfate. Complex kinetics was observed with an increase in the initial concentration of the oxidant. The influence of the dielectric constant, [H+] and [SO4 2-] on the rate was also investigated. The increase in the dielectric constant and H+ ion concentration of the medium retard the rate, while an increase in the SO4 2- ion concentration first accelerates the rate, and then retards the reaction. The effect of each factor, i.e., the dielectric constant, H+ ions and SO4 2- ions, suggests that Ce(SO4)3 2- is the active species of cerium(IV). A rate law consistent with the observed kinetic data and the proposed mechanism is suggested to be: {fx631-1  相似文献   

15.
In the presence of trifluoroacetic acid, N-phenyl-N′-phenoxyurea (1a) rearranges to N-(4′-hydroxy-2-biphenylyl)urea (2a) and N-carbamoyl-2-hydroxy-diphenylamine (3a). The rearrangement is an intramolecular reaction, and the transition state of the breakage of the N-O bond is deduced to be polarized in the form Nδ- --- Oδ+. The reaction is entirely new and constitutes a fundamental aromatic rearrangement.  相似文献   

16.
The reaction which occurs when lamiolgenin 2 is treated with an excess of NaBH4 is unusually complex owing to an intramolecular aldolic condensation which leads to a bicyclo[3.2.0]heptane derivative 4. The mechanism of this reduction has been investigated by using NaBD4. Spectral data (IR, 1H NMR, 13 C NMR) of the products are presented.  相似文献   

17.
The kinetics of the oxidation of sulfanilic acid (SAA) by sodium N-chloro-p-toluenesulfonamide (CAT) in the presence and absence of ruthenium(III) chloride have been investigated at 303 K in perchloric acid medium. The reaction shows a first-order dependence on [CAT]o and a non-linear dependence on both [SAA]o and [HClO4] for both the ruthenium(III)-catalyzed and uncatalyzed reactions. The order with respect to [RuIII] is unity. The effects of added p-toluenesulfonamide, halide, ionic strength, and dielectric constant have been studied. Activation parameters have been evaluated. The rate of the reaction increases in the D2O medium. The stoichiometry of the reaction was found to be 1:1 and the oxidation product of SAA was identified as N-hydroxyaminobenzene-4-sulfonic acid. The ruthenium(III)-catalyzed reactions are about four-fold faster than the uncatalyzed reactions. The protonated conjugate acid (CH3C6H4SO2NH2Cl+) is postulated as the reactive oxidizing species in both the cases.  相似文献   

18.
Two reagents of the 4-(phenylazo)-1-naphthol series were synthesized: 4-(4-sulfobenzeneazo)-1-naphthol-2-sulfonic acid (SBANS) and 4-(benzeneazo)-1-naphthol-2-sulfonic acid (BANS). These reagents contain the sulfo group at the ortho position with respect to the hydroxy group in the naphthol ring and differ by the presence or the absence of one more sulfo group at the para position in the benzene ring. The effect of ethanol and micelles of three types of nonionic surfactants (Proxamine-385, Brij-35, and Triton X-100) on the azo–quinonehydrazone tautomerism of BANS, SBANS, and their analogue Tropaeolin 000 was studied. It was demonstrated that an increase in the concentration of ethanol or the solubilization of the reagents in the micelles of nonionic surfactants shifts the tautomeric equilibrium to the less polar azo form. It was found that the sulfo group in the naphthol ring controls the lability of the tautomeric equilibrium. Tautomeric constants in aqueous–ethanolic solutions and nonionic surfactant micelles were calculated, and the effective dielectric constant of the medium at the position of localization of the reagents in the hydroxyethylene layer of the micelles was estimated using a model approach.  相似文献   

19.
Summary Quadrivalent selenium can be determined with fair accuracy by mixing with an excess of KMnO4 in the presence of 25–75 ml of 2% NaF solution and 4–7 ml of 9 N sulfuric acid. After leaving the reaction mixture for 10–30 minutes the excess KMnO4 is estimated by one of the following procedures: A) Titration of the excess KMnO4 with monovalent mercury, B) Adding an excess of Hg2 2+ solution to react with the excess KMnO4 followed by titrating the excess mercurous with KMnO4 solution.Part I: Issa, I. M., and M. Hamdy, Z. analyt. Chem. 172, 94 (1980).  相似文献   

20.
The iodous acid disproportionation is autocatalytic, and it is not easy to measure the rate constant of the step 2IO2H → IO3? + IOH + H+ separately. Hg(II) was used previously to suppress the autocatalytic pathway, but this method presents difficulties discussed in this work. A more effective method is the use of crotonic acid, an effective IOH scavenger. It suppresses side reactions, and a purely second‐order rate law is obtained. The rate constant decreases from 5 to 0.2 M?1 s?1 when the sulfuric acid concentration increases from 0.08 to 0.60 M. The observed decrease could be explained if IO2? reacts faster than IO2H. This may have consequences for the mechanism of the oscillating Bray–Liebhafsky reaction.  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