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.
Methaqualone [2-methyl-3-o-tolyl-4(3H)-quinazolinone] is reduced at pH 1.5–5 in a single two-electron step. At pH 1.5–3, wave i1 appears and is gradually replaced by wave i2 which predomiantes in the solution between pH 4.5 and 6.5. At pH > 5, the height of wave i2 decreases in the shape of a dissociaton curve with increasing pH. For both waves, the diprotonated form of methaqualone (Ia) is reduced to 1,2,-dihydromethaqualone as the final product. In wave i1, the monoprotonated form Ib is further protonated at the electrode surface before it accepts the first electron; in wave i2, the free base form I accepts two protons at the electrode surface before the first electron uptake. Polarographic curves are complicated by the presence of three waves of catalytic hydrogen evolution. Wave i1, cata appears at pH < 5, waves i2,cata at ?1.5 V and i3,cata at ?1.7 V at pH > 5. Citrate buffers pH 1.5–3 or Britton-Robinson buffers at pH 2.6–3.6 are most suitable for quantitative work with eitehr d.c. polarography or differential pulse polarography.  相似文献   

2.
Coordination equilibria in the Co(II)–Ni(II)–2-aminopropanoic acid (HAla)–EDTA system have been studied spectrophotometrically at different molar ratios of the reagents in a wide pH range. It has been found that, when metal ions are in excess with respect to EDTA at pH 5–9, polyheteronuclear complexonates [(CoAla)Edta(NiAla)]2–, [(CoAla2)Edta(NiAla2)]4–, [(NiAla2)Edta(CoAla2)2]4–, [(CoAla2)Edta(NiAla2)2]4–, and [(NiAla2)2Edta(CoAla2)2]4– form in a solution. The equilibrium constants of formation of these complexes and their overall stability constants have been calculated. Possible structures of the polynuclear complexonates are discussed.  相似文献   

3.
The interaction of Me2Sn(IV)2+ and Me3Sn(IV)+ with a prodrug, sodium 2-mercaptoethane sulfonate (HSCH2CH2SO3Na, MESNA) abbreviated as (HL), has been studied potentiometrically in aqueous solution (I = 0.1 mol·L?1 KNO3, 298 K). The concentration distribution of various species formed in the solution was studied with changes in pH (~3–11). A strong coordination of MESNA with metal through the S atom of thiol group has been found. In the Me2Sn(IV)–HL system, the species [Me2Sn(L)]+ (53.1–75.6%) is predominant at acidic pH (3.73 ± 0.02) and the species [Me2Sn(L)2OH]? (29.4–38.5%) is predominant at basic pH (10.32 ± 0.08). In contrast, for the Me3Sn(IV)+ system, [Me3SnL] (37.0–57.4%) is the major species at pH 7.65 ± 0.03 and [Me3Sn(OH)] (49.9–67.2%) and [Me3Sn(L)(OH)]? (30.2–46.5%) are the major species at pH 11.05 ± 0.01. However, at physiological pH (7.01 ± 0.32), in both (1:1) and (1:2) Me2Sn(IV)–HL systems, the species [Me2Sn(L)(OH)] (67.2–89.9%) is predominant, whereas for Me3Sn(IV)–HL (1:1) and (1:2) systems, [Me3Sn(OH)] (53.5%) and [Me3SnL] (56.8%) are the respective predominant species. In order to characterize the possible geometry of the proposed complex species, multinuclear (1H, 13C and 119Sn) NMR studies were carried out at different pHs. No polymeric species were detected in the experimental pH range.  相似文献   

4.
Values of the pH for four solutions in a KHPh-HCl-KOH system, where Ph denoting C8O4H4, are measured at 25–70°C depending on pressure (1–1000 bar). The experimental results and the available literature data are processed on the base of Helgeson-Kirkham-Flowers (HKF) equation of state [1] and the GIBBS, OptimA, and OptimB programs from the HCh software package [2]. The obtained standard thermodynamic properties and HKF parameters of H2Phaq, HPh?, and Ph2? aqueous species, provides calculation of the pH of phthalate buffers in a wide range of temperatures (up to 200°C) and pressures (up to 5 kbar). The calculated values for the biphthalate buffer (0.05m KHPh) correspond to the IUPAC recommendations (at 0–50°C and 1 bar) with an accuracy of 0.005 pH units, and to the values of pH measured in this study at elevated Tp parameters (25–70°C and pressures of up to 1 kbar) within the limits of ±0.02 pH units.  相似文献   

5.
Oxidations of indigocarmine (IC) by chloramine-T (CAT) and aqueous chlorine (HOCl) in acidic buffer media, pH 2–6, have been kinetically studied at 30°C using spectrophotometry. The CAT reaction rate shows a first-order dependence on [IC]0 and an inverse fractional order on [p-toluenesulfonamide]. The effect of [CAT] on the rate is strongly pH dependent with a variable order of 1–2 on [CAT]0 in the pH range 6–2. The chlorine reaction rate follows first-order in [IC]0 and [HOCl]0 each in the pH range 6–2. Addition of halide ions and variation of ionic strength of the medium have no influence on the reaction rate. There is a negative effect of dielectric constant of the solvent. The kinetics of the IC oxidation with CAT at pH 6 and of that with HOCl at pHs 2–6 are similar suggesting similarity in their rate determining steps. A two-pathway mechanism for the CAT reaction and a one-pathway mechanism for the HOCl reaction, consistent with the kinetic data, have been proposed. Activation parameters have been calculated using the Arrhenius and Erying plots. © 1995 John Wiley & Sons, Inc.  相似文献   

6.
Dioctylarsinic acid (HDOAA) in chloroform solution has been investigated as a reagent for the extraction of Hg(II), Ag(I), Co(II), and Cd(II). Silver, cobalt and cadmium are not extracted below pH 7. An extraction coefficient of 1.1, constant over the pH range 1–6.5, was observed for Hg(II). With HCl concentrations of 1–8 M the extractability of mercury decreased slowly, reaching Ea0 = 0.05 at 8 M HCl. Silver formed a silver dioctylarsinate precipitate which collected at the interface. The extraction coefficients for Hg(II), Co(II) and Cd(II) increased above pH 7 to values of 20 (pH 9.1), 30 (pH 8.0), and 23 (pH 10), respectively. Reagent- and pH-dependence studies indicated that Co(II) and Cd(II) are extracted as M(DOAA)2 or M(DOAA)Cl through interaction of HDOAA with M(OH)2 or M(OH)+. Mercury was extracted from solutions of pH 1–6.5 as HgCl2 (HDOAA)2.5.  相似文献   

7.
The reactions of 2-methoxy-4,6-diamino-s-triazine (MT) with formaldehyde (F) were studied in the pH region 2–11. In an alkaline medium the rate constants were proportional to the concentration of hydroxyl ion in the pH range 9.90–10.95; in a neutral medium they were approximately constant and in an acidic medium they achieved the maximum value at pH 3.50 (half-neutralization point of MT). The hydroxymethylation mechanism of MT was similar to that of melamine. The hydroxymethylation rate constants of various 2-substituted 4,6-diamino-s-triazines(XT) were determined at pH 7.00, 10.26, and 4.00. Linear relationship was observed between the rate constants at pH 7.00 and pKb of XT, σm, and σ* but not at pH 10.26 and 4.00.  相似文献   

8.
H2O2 production by electroreduction of O2 is an attractive alternative to the current anthraquinone process, which is highly desirable for chemical industries and environmental remediation. However, it remains a great challenge to develop cost‐effective electrocatalysts for H2O2 synthesis. Here, hierarchically porous carbon (HPC) was proposed for the electrosynthesis of H2O2 from O2 reduction. It exhibited high activity for O2 reduction and good H2O2 selectivity (95.0–70.2 %, most of them >90.0 % at pH 1–4 and >80.0 % at pH 7). High‐yield H2O2 generation has been achieved on HPC with H2O2 concentrations of 222.6–62.0 mmol L?1 (2.5 h) and corresponding H2O2 production rates of 395.7–110.2 mmol h?1 g?1 at pH 1–7 and ?0.5 V. Moreover, HPC was energy‐efficient for H2O2 production with current efficiency of 81.8–70.8 %. The exceptional performance of HPC for electrosynthesis of H2O2 could be attributed to its high content of sp3‐C and defects, large surface area and fast mass transfer.  相似文献   

9.
《Polyhedron》2005,24(3):443-450
The proctolin (Arg–Tyr–Leu–Pro–Thr, RYLPT) analogues modified in fifth position of the peptide chain (RYLPP, RYLPI, RYLP–Dab, where Dab – 2,4-diamminobutyric acid) have been synthesised and their complexes with H+ and Cu2+ studied by potentiometry and spectroscopy (UV–Vis, CD and EPR) at 25 °C and I = 0.10 mol dm−3 (KNO3). The results obtained support the earlier suggestion on the specific role of a proline residue as a “break-point” in copper complex formation with peptides. The presence of a proline residue into the fourth position of the proctolin analogues (RYLPP, RYLPI) leads in wide pH range of the existence the CuL and CuH−1L complexes with expected stabilities. Spectroscopic studies confirm that these are 2N {NH2, N, CO} and 3N {NH2, 2N, CO} species, respectively. The amine group of the Dab residue of the RYLP–Dab proctolin analogue, in whole pH range (2.5–10.5) is coordinated to the copper(II) ions, and the deprotonation and coordination of the second amide nitrogen atom to the metal ion is prevented. In solution in wide pH range (5–10.5) the 3N {NH2, N, CO, NH2Dab} complex is present. Proctolin and its analogues modified in fifth position contain in the second position of the peptide sequence the Tyr residue and the CD results show that TyrO–Cu2+ bonding is present at pH above 8.  相似文献   

10.
The hydrogen peroxide decomposition kinetics were investigated for both “free” iron catalyst [Fe(II) and Fe(III)] and complexed iron catalyst [Fe(II) and Fe(III)] complexed with DTPA, EDTA, EGTA, and NTA as ligands (L). A kinetic model for free iron catalyst was derived assuming the formation of a reversible complex (Fe–HO2), followed by an irreversible decomposition and using the pseudo‐steady‐state hypothesis (PSSH). This resulted in a first‐order rate at low H2O2 concentrations and a zero order rate at high H2O2 concentrations. The rate constants were determined using the method of initial rates of hydrogen peroxide decomposition. Complexed iron catalysts extend the region of significant activity to pH 2–10 vs. 2–4 for Fenton's reagent (free iron catalyst). A rate expression for Fe(III) complexes was derived using a mechanism similar to that of free iron, except that a L–Fe–HO2 complex was reversibly formed, and subsequently decayed irreversibly into products. The pH plays a major role in the decomposition rate and was incorporated into the rate law by considering the metal complex specie, that is, EDTA–Fe–H, EDTA–Fe–(H2O), EDTA–Fe–(OH), or EDTA–Fe–(OH)2, as a separate complex with its unique kinetic coefficients. A model was then developed to describe the decomposition of H2O2 from pH 2–10 (initial rates = 1 × 10−4 to 1 × 10−7 M/s). In the neutral pH range (pH 6–9), the complexed iron catalyzed reactions still exhibited significant rates of reaction. At low pH, the Fe(II) was mostly uncomplexed and in the free form. The rate constants for the Fe(III)–L complexes are strongly dependent on the stability constant, KML, for the Fe(III)–L complex. The rates of reaction were in descending order NTA > EGTA > EDTA > DTPA, which are consistent with the respective log KMLs for the Fe(III) complexes. Because the method of initial rates was used, the mechanism does not include the subsequent reactions, which may occur. For the complexed iron systems, the peroxide also attacks the chelating agent and by‐product‐complexing reactions occur. Accordingly, the model is valid only in the initial stages of reaction for the complexed system. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 24–35, 2000  相似文献   

11.
The sorption of a complex of europium (III) with acetylacetone on silica gel chemically modified with hexadecyl groups (SiO2-C16) and hyper crosslinked polystyrene (HLPS) was studied. Maximum extraction was observed at pH 5–7 when SiO2-C16 was used as the sorbent and at pH 4–7 in the case of crosslinked polystyrene. The partition coefficients for HLPS and silica gel were calculated as 7 × 103 and 1 × 102 cm3/g, respectively. Quantitative extraction of the europium (III) complex was possible in dynamic conditions using a microcolumn (length, 10 mm; internal diameter, 3mm) packed with HLPS at pH 5 (10–50 mL sample volume). Desorption of europium using solutions of nitric acid at different concentrations was investigated. Quantitative desorption was achieved using 5 mL of 1 M HNO3. A linear range of detection was observed at an amount of europium from 5 to 25 μg in a 10-mL sample (650 nm).  相似文献   

12.
Nitrophenols have been detected in some Antarctic lakes, the water of which is basic and rich in nitrate, nitrite and other nutrients. Nitrate or nitrite photolysis could be a possible reaction to explain the presence of these compounds. This work presents evidence for the formation of 2-nitrophenol (2NP), 4-nitrophenol (4NP) and 4-nitrosophenol (4NOP) upon UV irradiation of phenol and nitrite in aerated basic solutions.

The pH dependence of the 2NP initial formation rate is different from those of 4NP and 4NOP. The dependence of the first mainly reflects the phenol/phenolate equilibrium, with phenol yielding 2NP at a higher rate than phenolate. In the case of 4NOP, the initial formation rate vs pH has a maximum at pH 9.5. The pH dependence of 4NOP formation rate suggests that three pathways are likely to operate: nitrosation of undissociated phenol by N2O3, prevailing at pH<8.7, nitrosation of phenolate by N2O3, prevailing in the pH interval 8.7–10.8, and reaction between phenoxyl radical and ?NO, prevailing at pH>10.8. Phenol nitrosation by N2O3 is favoured when phenol is negatively charged (phenolate), but it is also disfavoured at alkaline pH values, owing to the depletion of N2O3 (the nitrosating agent) by basic hydrolysis. Differently from 2NP, the initial formation rate vs pH of 4NP is very similar to that of 4NOP, suggesting that 4NP may originate from the oxidation of 4NOP. Moreover, while in neutral and acidic solutions the formation rate of 2NP is slightly higher than that of 4NP, in the pH interval 8–12 the formation of 4NP is much more rapid than that of 2NP. This indicates that the pH of natural waters influences the ratio of nitroisomers.  相似文献   

13.
Precipitation of radiotellurium, containing trace radioimpurities, has been carried out from sulfate media at different pH-values. The highest precipitation yield was achieved at the region of pH ~4–6. Quantitative uptake by the formed precipitate was noticed for (i) 54Mn, 110mAg and 125Sb over all the pH-range of study (pH 1.7–9.2), (ii) for 65Zn and 60Co in the regions of pH ~6–8 and pH 6–8.8, respectively, and (iii) for 134Cs in the region of pH 1.7–2.8, while its percent uptake fluctuated around 60.5 % in the region of pH 4.4–6.4. Further precipitation studies have been conducted for a mixture of 125I and radiotellurium from sulfate, nitrate and chloride media at pH-values of 6.0 and 7.5. The highest 125I recovery yield in the obtained supernatant was 95.0 ± 1.3 %, which was achieved with sulfate medium at pH 6.0 with percent uptake values of 5.0 ± 1.3, 98.9 ± 0.9 and 62.0 ± 4.6 % of 125I, 123mTe and 134Cs, respectively, and quantitative uptake of 54Mn, 110mAg, 125Sb, 60Co and 65Zn by the precipitated tellurium. Thereafter, the supernatant was further acidified with H2SO4 and boiled, after adding H2O2, for 3 h. >99 % of 125I was distilled off from the acidified supernatant. The distilled of 125I was received in 0.1 M NaOH + 1 % Na2S2O3 solution, with a radionuclidic purity of >99.99 %, radiochemical purity of >99.8 % as I? and pH ~13.  相似文献   

14.
The hydrolysis of six selected pesticides has been studied in aqueous solution. Four organophosphorus pesticides (disulfoton, isofenfos, isazofos and profenfos) and two N-methylcarbamate derivatives (oxamyl and ethiofencarb) were selected. Hydrolysis was performed in purified buffered water at different pH in the range 7.0–10.0 (ionic strength?=?2.5?mM, T?=?25°C). At pH?=?8.0, isofenfos and disulfoton (t 1/2?≈?4 years, t 1/2?≈?1 year, resp.) were found to be far more stable than isazofos (t 1/2?≈?5 months), ethiofencarb and profenofos (t 1/2<1 month), themselves more stable than oxamyl (t 1/2?≈?1 day). As expected, a strong dependence on pH was observed for all pesticides: the rate of degradation increased when the pH increased. Degradation products were identified by GC–MS and/or LC–MS. Possible structures are presented in the article.  相似文献   

15.
Acrylic acid (AA) and N-vinylpyrrolidone (NVP) were copolymerized in aqueous solution at 30°C in the pH range 4–9 and the monomer reactivity ratios (r1 for AA and r2 for NVP) were determined as a function of pH from the high conversion data by using both the differential (YBR) and the integrated copolymerization equations. The value of r1 decreased from 5.2 at pH 4 to a minimum of 1.3 at pH 5 and then increased to 8.1, 6.6, and 7.2 at pH 7, 8, and 9, respectively. Addition of 1M NaCl at pH 6.5 restored the r1 nearly to that at pH 4, the predominantly un-ionized acid. The r2 values for NVP were nearly zero at all pH values except at pH 5. The variation of the reactivity ratios with pH was examined in terms of the electrostatic interactions between the ionized monomer molecules, the rate of homopolymerization of acrylic acid, and the cation binding to the poly(acrylic acid) molecules. The r2 values for NVP compared favorably with the literature values reported in bulk and organic solvent systems, although r1 values are much higher.  相似文献   

16.

We report the synthesis of electrochemically active LiMn2O4 nanoparticles at varied temperature and pH values by sol–gel method using urea as a chelating and combusting agent. The effect of pH and annealing temperature on the structure, morphology and electrochemical performance was evaluated. The results obtained by XRD, SEM, TEM, and FTIR show that LiMn2O4 has uniform porous morphology and highly crystalline particles that can be obtained at pH 7.0 and 8.0 and at a relatively lower temperature of 600°C. Cyclic voltammetry measurements showed reversible redox reactions with fast kinetics corresponding to Li ions intercalation/deintercalation at 600°C at neutral pH 7.0. Charge/discharge studies carried out at a current rate of 40 mA g–1 reveal that LiMn2O4 synthesized at 600°C and pH 7.0 has the best structural stability and excellent cycling performance.

  相似文献   

17.
Acrolein was copolymerized by radical initiation in aqueous solutions with sodium p-styrenesulfonate and acrylic acid, respectively, in the pH range of 3–7. The reactivities were shown to be pH-dependent. For the acrolein (M1)–sodium p-styrenesulfonate (M2) pair, r1 = 0.33 ± 0.15 and r2 = 0.32 ± 0.05 at pH 3; r1 = 0.23 ± 0.12 and r2 = 0.05 ± 0.03 at pH 5; r1 = 0.26 ± 0.03 and r2 = 0.025 ± 0.025 at pH 7. For the acrolein (M1)–acrylic acid (M2) pair, r1 = 0.50 ± 0.30 and r2 = 1.15 ± 0.2 at pH 3; r1 = 2.40 ± 0.50 and r2 = 0.05 ± 0.05 at pH 5; r1 = 6.70 ± 3.00 and r2 = 0.00 at pH 7. For acrolein, the new values of Q = 1.6 and e = 1.2 have been calculated. For sodium p-styrenesulfonate, the values Q = 0.76 and e = ?0.26 at pH 3, Q = 0.51 and e = ?0.87 at pH 5, Q = 0.39 and e = ?1.00 at pH 7 were obtained; and for acrylic acid, the values Q = 1.27 and e = 0.50 at pH 3, Q = 0.11 and e = ?0.22 at pH 5 were derived. The changes in reactivity are explained on the basis of inductive and resonance effects.  相似文献   

18.
Spectrophotometric methods were utilized for stability constant determinations of the Fe(III) interaction with glycinehydroxamic acid (GX) at I = 0.15 M NACl and T = 25°C. Program SQUAD II was used to assess the absorbance data in the wavelength range 300–520 nm. Four constants were determined for 1:1:1, 1:1:0, 2:1:1 or 3:1:3 and 2:1:0 complex species in the pH range 1.0–7.5. The kinetics of the interactions of Fe(III) with GX were also studied in the pH range 1.0–3.0 by the stopped flow method. The observed rate constant at a given pH was kobs = A + BTGX. The parameters A and B are functions of pH in the range 1.7–3.0 and only A is a function of pH in the range 1.0–1.7. The mechanism of complex formation was discussed in the light of the experimental results and the equilibrium study. It has been concluded that FeOH2+ is the reactive species in the complex formation of FeGXH3+ species while Fe(OH)2+ is the reactive species in the complex formation of FeGX2+ species.  相似文献   

19.
Hydrothermal microwave treatment of mixed solutions of potassium permanganate and hexamethylenetetramine within the pH range 0.5–6.9, resulted in various polymorphs of nanocrystalline manganese dioxide: α-MnO2 (cryptomelane), γ-MnO2 (nsutite), β-MnO2 (pyrolusite), and δ-MnO2 (birnessite). The pH values of the medium at which single-phase samples form were determined.  相似文献   

20.
4-(2-Pyridylazo)resorcinol (PAR) forms 2 red chelates with scandium in aqueous solution at pH ? 2 or 5: SCRH2+ and ScR2-. The former can be used to determine 0.05–2.4 μg Sc/ml at pH 3.5–4.5 with μ = 0.4; measurement is done at 530 mμ. Many ions interfere and separation is necessary.  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