首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 234 毫秒
1.
焦燕妮  侯万国 《中国化学》2007,25(6):756-764
The interface electrochemical properties of clay were theoretically analyzed to obtain some relationships among point of zero net charge (PZNC), point of zero net proton charge (PZNPC), intrinsic surface reaction equilibrium constants (K in 1-pK model, Ka1^int and Ka2^int in 2-pK model, *KNa^int and *KNO3^int in inert electrolyte chemical binding model) and structural negative charge density (σst) of clay, and some interface electrochemical parameters of kaolinite were measured. The following main conclusions were obtained. For clay possessing structural negative charges, the PZNC independent of electrolyte concentration (c) should exist just as amphoteric solid without structural charges such as oxides or hydroxides. A common intersection point (CIP) should appear among the potentiometric (or acid-base) titration curves at different c and the pH at the CIP should be PHPZNC. A CIP among potentiometric titration curves at different c for kaolinite was observed, and the value of PHPZNC of kaolinite was 2.16. The values of pHPZNPC were decreased with increasing c, which arises from the presence of structural negative charges of kaolinite. In addition, it was observed that a good linear relationship existed between pHPZNPC and 1g c. According to the values of PHPZNC and σst measured, the intrinsic surface reaction equilibrium constants, pK and pKa1^int and pKa2^int of 1-pK and 2-pK models could be directly calculated for clay, and the values of pK, pKa1^int and pKa2^int for kaolinite were 2.93, 1.90 and 3.97, respectively. These experimental values of pKa1^int and pKa2^int for kaolinite are obviously lower than those optimized with fitting programs in literatures, which maybe arises from the introduction of a type of permanent negatively charged sites in the models of literatures. An interesting result obtained in this study is that the inert electrolyte chemical binding does not exist for kaolinite, which also arises from the presence of structural negative charges.  相似文献   

2.
The relationship among intrinsic surface reaction constant (K) in 1-pK model, point of zero net charge (PZNC) and structural charge density (σst) for amphoteric solid with structural charges was established in order to investigate the effect of σst on pK. The theoretical analysis based on 1-pK model indicates that the independent PZNC of electrolyte concentration (c) exists for amphoteric solid with structural charges. A common intersection point (CIP) should appear on the acid-base titration curves at different c, and the pH at the CIP is pHPZNC. The pK can be expressed as pK=-pHPZNC log[(1 2αPZNC)/(1-2αPZNC)], where αPZNC≡σst/eNANs, in which e is the elementary charge, NA the Avogadro‘s constant and Ns the total density of surface sites. For solids without structural charges, pK=-pHPZNC. The pK values of hydrotalcite-like compounds (HTlc) with general formula of [Mg1-xFex(OH)2](Cl,OH)x were evaluated. With increasing x, the pK increases, which can be explained based on the affinity of metal cations for H^- or OH^- and the electrostatic interaction between charging surface and H^- or OH^-.  相似文献   

3.
Immobilization of hydrous ferric oxide(HFO) particles inside solid hosts of porous structure is an important approach to improve their applicability in advanced water treatment such as arsenic and heavy metal removal. Here, we fabricated three polystyrene(PS)-based nano-HFOs and explored the effect of host pore structure on the surface chemistry of the immobilized HFOs. Potentiometric titration of the hybrids and surface complexation modeling of their adsorption towards arsenite and arsenate were performed to evaluate the surface chemistry variation of the loaded HFOs. Polymer hosts of higher surface area and narrower pore size would result in smaller particle size of HFOs and lower the value of the point of zero charge. Also, the site density(normalized by Fe mass) and the deprotonation constants of the loaded HFOs increased with the decreasing host pore size. Arsenite adsorption did not change the surface charge of the loaded HFOs, whereas arsenate adsorption accompanied more of the negative surface charges. Adsorption affinity of both arsenic species with three HFO hybrids were compared in terms of the intrinsic surface complexation constants optimized based on the adsorption edges. HFO loaded in polystyrene host of smaller pore size exhibits stronger affinity with arsenic species.  相似文献   

4.
The photodegradation reaction rate of CHCl3 in TiO2 particulate suspension was imperoved significantly by HCl-treatment.The effect of HCl-treatment on the photocatalytic activity of TiO2 was strdied in a PEC cell by using Intensity-Modulated Photocurrent Spectroscopy(IMPS).The magntude of photocurrent response and the characteristic frequencies of the upper and lower semicircles in the complex plane of IMPS response were analyzed,and the ccathodic and anodic reaction processes of photogenerated holes and electrons were discussed.The increases in the cathodic and anodic photocurrent response and in the time constants of both cathodic and anodic reaction processes of photogenerated holes and electrons were discussed.The increases in the cathodic and anodic photocurrent response and in the time constants of both cathodic and anodic reaction processes indicate that HCl-treatment leads to the improvement of the photocatalytic activity of TiO2 and a change of the photocatalytic kinetic mechanism.  相似文献   

5.
Three-dimensionally ordered macro-porous (3DOM) Pt/TiO2 catalysts were prepared by template and impregnation methods, and the resultant samples were characterized by using TG-DTA, XRD, SEM, TEM, and TPR techniques. The catalytic performance for water-gas shift (WGS) reaction was tested, and the influences of some conditions, such as reduction temperature of catalysts, the amount of Pt loadings and space velocity on catalytic performance were investigated. It was shown that Pt particles were homogeneously dispersed on 3DOM TiO2. The reduction of TiO2 surface was important for the catalytic performance. The activity test results showed that the 3DOM Pt/TiO2 catalysts exhibited very good catalytic performance for WGS reaction even at high space velocity, which was owing to the better mass transfer of 3DOM porous structure besides the high intrinsic activity of Pt/TiO2.  相似文献   

6.
邻位吡啶自由基多通道分解反应的动力学和反应机理研究   总被引:1,自引:0,他引:1  
Utilizing Gaussian94 program package, all species involved in decomposition reactions of o-pyridyl radical were optimized fully at B3LYP/6-311++G^** level. Intrinsic reaction coordinate calculations were employed to confirm the connections of the transition states and products, and transition states were ascertained by the number of imaginary frequency (0 or 1). The reaction mechanism was elucidated by the vibrational mode analysis and electronic population analysis, and the reaction rate constants were calculated with transition state theory.  相似文献   

7.
Ag2S‐graphene/TiO2 composites were synthesized by a facile sonochemical method.The products were characterized by X‐ray diffraction,scanning electron microscopy,energy dispersive X‐ray spectroscopy,transmission electron microscopy,and UV‐Vis diffuse reflectance spectrophotometry.During the synthesis reaction,the reduction of graphene oxide and loading of Ag2S and TiO2 particles were achieved.The Ag2S‐graphene/TiO2 composites possessed a large adsorption capacity for dyes,an extended light absorption range,and efficient charge separation properties.Hence,in the photodegradation of rhodamine B,a significant enhancement in the reaction rate was observed with the Ag2S‐graphene/TiO2 composites as compared to pure TiO2.The generation of reactive oxygen species was detected by the oxidation of 1,5‐diphenyl carbazide to 1,5‐diphenyl carbazone.The high activity was attributed to the synergetic effects of high charge mobility and the red shift in the absorption edge of the Ag2S‐graphene/TiO2 composites.  相似文献   

8.
The ab initio calculations were performed by the intrinsic reaction coordinated(IRC) method for the reaction paths of the dehydrogenations of methylenimine. We determined the geometries and energies of the transition states and obtained the activation energies, activation entropies and the statistical A factors on the RHF/4 -31G singlet potential energy surface. The mode selective study reveals the intrinsic reaction coordi-nates(URC) of 1 ,1 and 1 ,2 dehydrogenations of methylenimine to be related with the H2C= group symmetrical stretch mode and the =NH group stretch mode, respectively. The vibration coupling between the normal coordinates occurs in either of these two reaction paths.  相似文献   

9.
MENG Zeda  OH Wonchun 《催化学报》2012,(9):1495-1501
CoS2,CoS 2-C60 /TiO2,CoS2-CNT/TiO2,and CoS2-Graphene/TiO2 were prepared.The TiO2 products had the anatase phase structure and interesting surface compositions.X-ray diffraction patterns of the CoS2-carbon/TiO2 composites showed a single and clear anatase phase and the CoS2 structure.Scanning electron microscopy characterization of the texture on the CoS 2-carbon/TiO2 composites showed a homogenous composition.Energy-dispersive X-ray spectra for elemental identification showed the presence of C and Ti with strong Co and S peaks from the CoS2-carbon/TiO2 composites.The composites obtained were also characterized by transmission electron microscopy and UV-Vis spectroscopy.CoS2-carbon/TiO2 composites showed excellent photocatalytic activity for the degradation of methylene blue under visible light irradiation.This was attributed to both photocatalysis on the TiO2 support and charge transfer by the carbon nanomaterial,and the introduction of CoS2 to enhance transfer of photogenerated electrons.  相似文献   

10.
The kinetics theory of the substrate reaction during modification of enzyme activity previously described by Tsou has been applied to a study on the kinetics of the course of inactivation of aminoacylase I by DPDS and PCMB.From the results obtained we have found that the inactivation reaction of aminoacylase I by DPDS is noncomplexing inhibition,and PCMB reaction is complexing inhibition.The microscopic constants for the reaction of the inactivator with free enzyme and the enzyme-substrate complex were determined.  相似文献   

11.
徐洁  侯万国  周维芝  台培东  王文兴 《化学学报》2007,65(13):1191-1196
分别采用电势滴定(Potentiometric titration)法, 质量滴定(Mass titration)法和惰性电解质滴定(Inert electrolyte titration)法测定了荷结构负电荷的东北草甸棕壤的零净电荷点(PZNC), 研究表明荷结构负电荷的土壤依然存在与电解质浓度无关的零净电荷点, 三种方法的实验结果分别为2.9, 5.0和3.3. 相比较而言, 电势滴定法和惰性电解质滴定法的结果相近, 而质量滴定法的结果偏高. 对文献中推算表面质子活性位密度(Ns)的方法进行了改进, 得到东北草甸棕壤样品的Ns为2.5 mmol•g-1. 根据实验测定的pHPZNC, Ns和结构电荷密度(σst)值直接计算得出东北草甸棕壤样品的界面反应特征平衡常数即1-pK模型中的pK, 2-pK模型中pKa1int和pKa2int, 分别为3.37, 2.42和4.32, 与文献报道的沉积物和蒙脱土的值相吻合.  相似文献   

12.
Hou  Wan-Guo  Jiang  Peng  Han  Shu-Hua  Hu  Ji-Fan  Li  Dong-Qing 《Colloid and polymer science》2003,281(8):738-744
The zero point of charge (ZPC) and the intrinsic ionization constant ( % MathType!MTEF!2!1!+- % feaaeaart1ev0aaatCvAUfKttLearuavTnhis1MBaeXatLxBI9gBae % bbnrfifHhDYfgasaacH8srps0lbbf9q8WrFfeuY-Hhbbf9v8qqaqFr % 0xc9pk0xbba9q8WqFfea0-yr0RYxir-Jbba9q8aq0-yq-He9q8qqQ8 % frFve9Fve9Ff0dmeaabaqaciGacaGaaeqabaWaaeaadaaakeaacqWG % lbWsdaqhaaWcbaGaeeyyaegabaGaeeyAaKMaeeOBa4MaeeiDaqhaaa % aa!304C! Kaint K_{\rm a}^{{\rm int}} ) of Zn-Mg-Al hydrotalcite-like compounds (HTlc) with the general formula [(Zn,Mg) 1- x Al x (OH) 2] x + [(OH,Cl) x] x - were determined by potentiometric titration (PT). The variation of the ZPC and % MathType!MTEF!2!1!+- % feaaeaart1ev0aaatCvAUfKttLearuavTnhis1MBaeXatLxBI9gBae % bbnrfifHhDYfgasaacH8srps0lbbf9q8WrFfeuY-Hhbbf9v8qqaqFr % 0xc9pk0xbba9q8WqFfea0-yr0RYxir-Jbba9q8aq0-yq-He9q8qqQ8 % frFve9Fve9Ff0dmeaabaqaciGacaGaaeqabaWaaeaadaaakeaacqWG % lbWsdaqhaaWcbaGaeeyyaegabaGaeeyAaKMaeeOBa4MaeeiDaqhaaa % aa!304C! Kaint K_{\rm a}^{{\rm int}} with xwas investigated. For the colloidal particles possessing permanent charges, the ZPC determined by the PT method is the zero point of net charge (ZPNC). The ZPNC ( pH ZPNC) values were 9.63 {[Zn 0.27Mg 0.36Al 0.37 (OH) 2]Cl 0.13 (OH) 0.24 }, 9.68 {[Zn 0.13Mg 0.58Al 0.29 (OH) 2]Cl 0.12 (OH) 0.17 }, 9.67 {[Zn 0.17Mg 0.54Al 0.29 (OH) 2]Cl 0.16 (OH) 0.12 },10.16 {[Zn 0.08Mg 0.67Al 0.25 (OH) 2]Cl 0.17 (OH) 0.08}, 10.33 {[Zn 0.16Mg 0.60Al 0.24 (OH) 2] Cl 0.16 (OH) 0.08} and 10.60 {[Zn 0.19Mg 0.60Al 0.21 (OH) 2] Cl 0.15 (OH) 0.06}; the intrinsic ionization constants % MathType!MTEF!2!1!+- % feaaeaart1ev0aaatCvAUfKttLearuavTnhis1MBaeXatLxBI9gBae % bbnrfifHhDYfgasaacH8srps0lbbf9q8WrFfeuY-Hhbbf9v8qqaqFr % 0xc9pk0xbba9q8WqFfea0-yr0RYxir-Jbba9q8aq0-yq-He9q8qqQ8 % frFve9Fve9Ff0dmeaabaqaciGacaGaaeqabaWaaeaadaaakeaacqqG % WbaCcqWGlbWsdaqhaaWcbaGaeeyyaeMaeeOmaidabaGaeeyAaKMaee % OBa4MaeeiDaqhaaaaa!329E! pKa2int {\rm p}K_{{\rm a2}}^{{\rm int}} of the same HTlc samples were 10.31, 10.44, 10.44, 11.02, 11.19 and 11.54. With decreasing x, ZPNC and % MathType!MTEF!2!1!+- % feaaeaart1ev0aaatCvAUfKttLearuavTnhis1MBaeXatLxBI9gBae % bbnrfifHhDYfgasaacH8srps0lbbf9q8WrFfeuY-Hhbbf9v8qqaqFr % 0xc9pk0xbba9q8WqFfea0-yr0RYxir-Jbba9q8aq0-yq-He9q8qqQ8 % frFve9Fve9Ff0dmeaabaqaciGacaGaaeqabaWaaeaadaaakeaacqWG % lbWsdaqhaaWcbaGaeeyyaeMaeeOmaidabaGaeeyAaKMaeeOBa4Maee % iDaqhaaaaa!3137! Ka2int K_{{\rm a2}}^{{\rm int}} of HTlc increased, and the acidity of the HTlc samples decreased.  相似文献   

13.
The effect of pH and neutral electrolyte on the interaction between humic acid/humate and γ-AlOOH (boehmite) was investigated. The quantitative characterization of surface charging for both partners was performed by means of potentiometric acid–base titration. The intrinsic equilibrium constants for surface charge formation were logK a,1 int=6.7±0.2 and logK a,2 int = 10.6±0.2 and the point of zero charge was 8.7±0.1 for aluminium oxide. The pH-dependent solubility and the speciation of dissolved aluminium was calculated (MINTEQA2). The fitted (FITEQL) pK values for dissociation of acidic groups of humic acid were pK 1 = 3.7±0.1 and pK 2 = 6.6±0.1 and the total acidity was 4.56 mmol g−1. The pH range for the adsorption study was limited to between pH 5 and 10, where the amount of the aluminium species in the aqueous phase is negligible (less than 10−5 mol dm−3) and the complicating side equilibria can be neglected. Adsorption isotherms were determined at pH ∼ 5.5, ∼8.5 and ∼9.5, where the surface of adsorbent is positive, neutral and negative, respectively, and at 0.001, 0.1, 0.25 and 0.50 mol dm−3 NaNO3. The isotherms are of the Langmuir type, except that measured at pH ∼ 5.5 in the presence of 0.25 and 0.5 mol dm−3 salt. The interaction between humic acid/humate and aluminium oxide is mainly a ligand-exchange reaction with humic macroions with changing conformation under the influence of the charged interface. With increasing ionic strength the surface complexation takes place with more and more compressed humic macroions. The contribution of Coulombic interaction of oppositely charged partners is significant at acidic pH. We suppose heterocoagulation of humic acid and aluminium oxide particles at pH ∼ 5.5 and higher salt content to explain the unusual increase in the apparent amount of humic acid adsorbed. Received: 20 July 1999 /Accepted in revised form: 20 October 1999  相似文献   

14.
We have obtained pKa values of p-nitrophenol–TiO2 by measuring the adsorption equilibrium constants of p-nitrophenol (PNP) on the TiO2 surface at different pH values. These values have been obtained from Langmuir isotherms and from a plot of 1/rate vs. 1/[PNP]o obtained during TiO2 catalyzed solar light photo-degradation of PNP. Two limit equilibrium constants are readily obtained depending on the solution pH: at pH 5 at which the TiO2 surface is mainly positively charged and at pH 8 when it is negatively charged. With these and other adsorption equilibrium constants and the PNP pKa value in solution, thermodynamic cycles are established in order to obtain the PNP pKa when it is adsorbed on positively charged, neutral and negatively charged TiO2 surfaces. From these pKa values useful information on the PNP–TiO2 interaction is readily obtained. For instance, the PNP nitro group interacts with the TiO2 surface via a hydrogen bond, arising from the complex of water molecules with the Ti4+ ions on its surface. The weaker the hydrogen bond donor, the stronger the oxygen nitro group basicity. Therefore, pKa changes on the phenolic hydroxyl group result from these interactions. Linear free energy correlations, maximum PNP adsorption capacity values (QL) and FTIR ATR, spectrum support this proposal. A kobs vs. pH degradation profile of p-nitrophenol is also provided.  相似文献   

15.
Rate constants are reported for collisional quenching of K* (4p2p and K* (5p2p) by H2O, CF4 and CH4. The K* (4p2p or K*(5p2p) is produced by photodissociation of K1 vapor at 2450 Å or 1925 Å, respectively. As for Na* (3p2p, H2O, CF4, and CH4 are very inefficient at quenching K* (4p2p); however, they are very efficient at quenching K* (5p2p).  相似文献   

16.
Alkaline hydrolyses of p-nitrophenyl-3-indoleacetate (p-NPIA) and N-(indole-3-acryloyl)imidazole (IAI) were studied in the presence of poly(4-vinyl-N-propylpyridinium bromide) (C3PVP), poly(4-vinyl-N-benzylpyridinium chloride) (BzPVP), and copolymers of 4-vinyl-N-benzylpyridinium chloride and 4-vinyl-N-cetylpyridinium bromide (C16BzPVP). The hydrolyses were enhanced by the addition of these cationic polyelectrolytes. The magnitudes of the enhancement were in the order C16BzPVP > BzPVP > C3PVP, which is explainable in terms of the hydrophobicity of the polymers. The result and activation parameters obtained indicated that the substrates bound to polymers were more reactive than free substrates. The association constants obtained from the kinetic measurements (K) increased in the order BzPVP < C3PVP < C16BzPVP, which may suggest an important contribution of charge transfer interaction, in addition to electrostatic and hydrophobic interactions, between the polymers and the substrates. The association constants between the polymers and IAI were also obtained independently from the spectrophotometric measurements (K*) with the charge transfer absorption bands. The lack of a satisfactory agreement between K and K* is discussed.  相似文献   

17.
Acidobasic properties of purine and pyrimidine bases (adenine, cytosine) and relevant nucleosides (adenosine, cytidine) were studied by means of glass-electrode potentiometry and the respective dissociation constants were determined under given experimental conditions (I = 0.1 M (NaCl), t = (25.0 ± 0.1) °C): adenine (pK HL = 9.65 ± 0.04, pK H2L = 4.18 ± 0.04), adenosine (pK H2L = 3.59 ± 0.05), cytosine (pK H2L = 4.56 ± 0.01), cytidine (pK H2L = 4.16 ± 0.02). In addition, thermodynamic parameters for bases: adenine (ΔH 0 = (−17 ± 4) kJ mol−1, ΔS 0 = (23 ± 13) J K−1 mol−1), cytosine (ΔH 0 = (−22 ± 1) kJ mol−1, ΔS 0 = (13 ± 5) J K−1 mol−1) were calculated. Acidobasic behavior of oligonucleotides (5′CAC-CAC-CAC3′ = (CAC)3, 5′AAA-CCC-CCC3′ = A3C6, 5′CCC-AAA-CCC3′ = C3A3C3) was studied under the same experimental conditions by molecular absorption spectroscopy. pH-dependent spectral datasets were analyzed by means of advanced chemometric techniques (EFA, MCR-ALS) and the presence of hemiprotonated species concerning (C+-C) a non-canonical pair (i-motif) in titled oligonucleotides was proposed in order to explain experimental data obtained according to literature.  相似文献   

18.
The determination of equilibrium constants is difficult when several chemical species are simultaneously present in solution. In this investigation, optical emission spectroscopic determinations of chromium(III) concentration in a 10−4 mol⋅dm−3 solution, prepared from K2Cr2O7 reduced in HNO3 or HCl media, were used to construct the pCr(aq)–pC H diagram. This diagram was used to calculate the pC H borderline of precipitation, to estimate the solubility product (log10Ksp,Cr(OH)3*)(\log_{10}K_{\mathrm{sp,Cr(OH)}_{3}}^{*}), and the hydrolysis constants (log10bCr,H*,log10bCr,2H*(\log_{10}\beta_{\mathrm{Cr,H}}^{*},\log_{10}\beta_{\mathrm{Cr,2H}}^{*}, and log10bCr,3H*)\log_{10}\beta_{\mathrm{Cr,3H}}^{*}) of Cr(III). The hydrolysis constants were also calculated using the SQUAD and SUPERQUAD software, along with the average ligand number method. UV-Vis absorption data and associated variables were used in SQUAD, SUPERQUAD, and the average ligand calculations. Results are: 9.00±0.04 for the pC H at the onset of precipitation, 12.40 for log10Ksp,Cr(OH)3*\log_{10}K_{\mathrm{sp,Cr(OH)}_{3}}^{*}, −3.52±0.02 for log10bCr,H*\log_{10}\beta_{\mathrm{Cr,H}}^{*}, −9.30±0.87 for log10bCr,2H*\log_{10}\beta_{\mathrm{Cr,2H}}^{*} and −17.18±0.16 for log10bCr,3H*\log_{10}\beta_{\mathrm{Cr,3H}}^{*}, respectively. All methods produced essentially the same values for the hydrolysis constants of Cr(III).  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