首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 546 毫秒
1.
采用等体积浸渍法制备了Cu-K-La/γ-Al2O3催化剂,考察了KCl对该催化剂催化HCl氧化制Cl2反应性能的影响. 当KCl的负载量为5 wt%时,Cu-K-La/γ-Al2O3催化剂表现出较好的催化活性和稳定性,可在较大的原料气空速变化范围内使用. 在0.1 MPa,360 ℃,空速450 L/(kg-cat·h)和HCl/O2摩尔比为2:1的反应条件下,Cu-K-La/γ-Al2O3催化剂上HCl转化率在100 h内保持85%以上. 表征结果表明,Cu,K和La物种均高度分散于γ-Al2O3载体表面;一定量KCl的加入可降低Cu2+ → Cu+的还原温度,从而提高Cu2+活性中心的催化活性.  相似文献   

2.
The activity coefficients of KCl in the KCl–CoCl2–H2O system were estimated at constant total ionic strengths of 0.5, 1.0, 2.0 and 3.0 mol-kg–1 and at 25°C. The data has been analyzed using the Harned, Scatchard, Reilly, Wood and Robinson and the Pitzer equations. The solubilities of KCl in these mixtures at 25°C were also analyzed.  相似文献   

3.
The liquid‐solid flotation separation behaviors of Cd2+ in ammonium sulfate‐potassium iodide‐cetylpyridine chloride‐water system and the conditions for the separation of Cd2+ from other metal ions were studied. The results showed that in the presence of 1.0 g (NH4)2SO4 and when the dosage of 0.1 M potassium iodide was 2.0 mL and 0.01 M cetylpyridine chloride (CPC) solution was 1.0 mL respectively, the formed water‐insoluble ternary association complex of KI‐CPC‐Cd floated above water phase and liquid‐solid phases were formed with clear interface. In this condition, Zn2+, Mn2+, Fe2+, Co2+, Ni2+ and Al3+ could not be floated and Cd2+ was floated quantitatively at pH 5.0. Therefore, the quantitative separation of Cd2+ from the above metal ions could be achieved. The quantitative flotation separation determination of Cd2+ in the sample of synthetic water and industrial waster water was performed, and the results agreed well with those by AAS method. The recoveries were 97.2%~102.4%, and the RSD was 1.8%.  相似文献   

4.
The reactions of .OH radicals with deoxyribose, DR, form five different DR. radicals, only one of which is transformed into malondialdehyde (MDA)‐like products. The radiolytic yield of the MDA‐like products increases with the increase in the DR concentration indicating that some of the initially formed “unproductive” radicals react with DR to form the “productive” radicals. The yield of the MDA‐like products also increases with the dose rate delivered to the solution suggesting that the formation of the MDA‐like products involves the reaction of the “productive” radicals with a radical. The addition of ascorbate, AH?, to the solution decreases the yield of the MDA‐like products as expected from the relative rates of the reaction of DR and AH? with .OH radicals. On the other hand the addition of the exogenous thiol, N‐acetylcysteine (NAC), to the solutions decreases the yield of the MDA‐like products considerably more than expected from the rate constants of the reaction with .OH radicals. The addition of the endogenous thiol, glutathione (GSH), to the solutions affects the yield of the MDA‐like products at low concentration less than expected and at “high” concentrations more than expected from the rate constant of the reaction. Addition of low concentration of AH? to solutions containing GSH increases considerably its antioxidant activity whereas addition of small concentrations of AH? to solutions containing NAC has no effect on its antioxidant activity. The results point out that the DR. radicals react differently with NAC and GSH and that the GS. and NAC. radicals react differently with DR, the GS. radical being considerably more active than the NAC. radical. Thus it has to be concluded that the relative activity of antioxidants depends also on the rate constants of many secondary reactions and on the concentrations of all the solutes present in the system.  相似文献   

5.
An empirical equation has been developed to model the electrolytic conductivity of potassium chloride (KCl) solutions over the temperature range 0 to 55 °C (ITS-90) and molality concentrations from 0 to 5 mol⋅kg−1 (H2O). The National Institute of Standards and Technology (NIST) Demal values are treated as the defining data and selected historical data provided further data. Changes in the definition of the ohm, liter and temperature scales have been identified and the historical data were converted to modern units. The accuracy of the equation is estimated to be equal to 0.03% as suggested by NIST as the accuracy of these data.  相似文献   

6.
The solvation and confinement of coumarin C153 within supramolecular host/guest complexes based on β‐cyclodextrin (β‐CD) and 6‐deoxy‐6‐thio‐β‐cyclodextrin (β‐CD‐SH) in water are studied by fluorescence spectroscopy. For β‐CD/C153, the 1:1 complex is proposed, and for β‐CD‐SH/C153 both the 1:1 and 2:1 complexes are believed to be formed. The 2:1 β‐CD‐SH/C153 complex has an association constant of 4.2×105 M ?1 and a C153 population of 82 %, which are interestingly high values, indicating that the proposed β‐CD‐SH dimers structure are connected by covalent disulfide bonds; this is supported by mass spectrometry. Solvation related to fast hydrogen‐bond rearrangement as a part of fluorescence relaxation is determined by the ultrafast components of time‐resolved spectroscopy to be 3 and 7 ps for the 1:1 β‐CD/C153 and 2:1 β‐CD‐SH/C153 complexes, respectively.  相似文献   

7.
A spectrophotometric flow-injection procedure for the determination of sulphite in aqueous media over the range 0.5–20 mg 1?1 is described. The reagent used was the organic disulphide 5,5′-dithiobis(2-nitrobenzoic acid). Results are presented for a laboratory-based method for sulphite in water and a potential on-line method for sulphite in high ionic strength potassium chloride brine. The general attractions of flow-injection-based monitors for the on-line analysis of liquid process streams are also discussed.  相似文献   

8.
We have measured the densities at temperatures T = (278.15 to 363.15) K and heat capacities at T = (278.15 to 393.15) K of aqueous solutions of 18-crown-6 and of (18-crown-6 + KCl) at molalities m = (0.02 to 0.3) mol · kg−1 and at the pressure 0.35 MPa. We have calculated apparent molar volumes V? and apparent molar heat capacities Cp,? for 18-crown-6(aq), and we have applied Young’s Rule and have accounted for chemical speciation and relaxation effects to resolve V? and Cp,? for the (18-crown-6: K+,Cl)(aq) complex in the mixture. We have also calculated estimates of the change in volume ΔrVm, the change in heat capacity ΔrCp,m, the change in enthalpy ΔrHm, and the equilibrium quotient log Q for formation of the complex at T = (278.15 to 393.15) K and m = (0 to 0.3) mol · kg−1.  相似文献   

9.
In this article, a synthetic concept for the preparation of polyamides with functional side groups is described. First, the synthesis of a bis(thiolactone) monomer is shown in a concise three‐step route from itaconic acid and DL‐homocysteine thiolactone. The reactivity of the resulting bis(thiolactone) toward hexyl amine is examined. Next, the bis(thiolactone) is reacted as A,A‐type monomer with different B,B‐type comonomers (1,12‐diaminododecane and 1,3‐bis(aminopropyl)tetramethyldisiloxane). Ring opening of the thiolactones by the diamines leads to polyamides with pendant thiol groups. Using two diamines in different ratios, the properties of the resulting polyamides are tuned (thermal properties are determined) and different molecular weights are acquired. Subsequently, the thiol groups are reacted with methyl acrylate via Michael addition to functionalize the polyamides. Functionalization of thiol‐functional polyamides using poly(ethylene glycol) monomethyl ether (mPEG) acrylates ( = 480 and 1700 g mol−1) results in water‐soluble amphiphilic poly­amides with molecular weights higher than 10 000 g mol−1.

  相似文献   


10.
The ionic liquid 1-N-butyl-3-methylimidazolium chloride ([C4mim]+Cl) was investigated as reaction media for the homogeneous acylation of cellulose with 2-furoyl chloride in the presence of pyridine. The preparation of cellulose furoate depending on the reaction conditions, the cellulose type and the pyridine content was studied. Cellulose furoates with a degree of substitution in the range from 0.46 to 3.0 were accessible, i.e., under mild conditions, with a low excess of reagent and in a short reaction time. The products were characterized by elemental analysis, perpropionylation, 1H- and 13C NMR spectroscopy and FTIR spectroscopy. Thomas Heinze is the member of the European Polysaccharide Network of Excellence (EPNOE), www.epnoe.eu  相似文献   

11.
Crystal Structure, Phase Transition, and Potassium Ion Conductivity of Potassium Trifluoromethanesulfonate According to the results of temperature dependent powder diffractometry (Guinier‐Simon‐technique) potassium trifluoromethanesulfonate is dimorphic. The phase transition occurs between –63 °C and –45 °C. The low‐temperature modification crystallizes monoclinic with a = 10.300(3) Å, b = 6.052(1) Å, c = 14.710(4) Å, β = 111.83(2)° (–120 °C) and the room‐temperature modification with a = 10.679(5) Å, b = 5.963(2) Å, c = 14.624(5) Å, β = 111.57(3)°, Z = 6, P21. According to single crystal structure determination, potassium trifluoromethanesulfonate consists of three different potassium‐oxygen‐coordination polyhedra, linked by sulfur atoms of the trifluoromethanesulfonate groups. This results in a channel structure with all lipophilic trifluoromethane groups pointing into these channels. By means of DSC, the transition temperature and enthalpy have been determined to be –33 °C and 0.93 ± 0.03 kJ/mol, respectively. The enthalpy of melting (237 °C) for potassium trifluoromethanesulfonate is 13.59 kJ/mol, the potassium ionic conductivity is 3.68 · 10–6 Scm–1 at 205 °C.  相似文献   

12.
2‐(Methylthio)aniline (H2L1) and 2‐(phenylthio)aniline (H2L2) were treated with n‐butyllithium to yield the corresponding anilides [LiHL1] and [LiHL2]. Recrystallization from diethyl ether and THF afforded the solvates [LiHL1(Et2O)] and [LiHL2(THF)2]. The X‐ray crystal structure determination revealed dimeric molecules which exhibit a centrosymmetric Li2N2 ring. In the case of [LiHL1(Et2O)] the SMe group is involved in Li coordination and in the case of [LiHL2(THF)2] the SPh group is part of an intramolecular N–H ··· S hydrogen bridge. The sodium anilides [NaHL1(DME)] and [NaHL2(DME)] were obtained from the reaction of H2L1 and H2L2 with sodium amide in DME as solvent. Like in the case of the lithium amides the sodium derivatives [NaHL1(DME)] and [NaHL2(DME)] display centrosymmetric Na2N2 cores. The coordination sphere of the sodium atoms is completed by DME molecules, which act as chelating ligands. In the case of [NaHL1(DME)] the DME molecules enable additionally a linkage of the dimeric subunits to give a chain structure. The potassium derivatives [KNHL1] and [KNHL2(DME)] were obtained from H2L1 and H2L2 and potassium hydride in DME as solvent. [KNHL1] displays a distinct structure based on [(KNHL1)2] dimers, which are linked by additional [KNHL1] units to give a 3D coordination polymer with {4.8.16(3)} topology. [KNHL2(DME)] forms dimers linked by bridging DME molecules to give a chain‐like coordination polymer.  相似文献   

13.
In the present study, the reaction of thiols with alkyl sulfonyl halides was carried out in a nitrogenous base to compare the reactivity of ‐SH with that of ‐OH, which, however, led to the formation of disulfides. The reaction achieved as a result offers the use of an inexpensive reagent, quantitative yields of the product, and simplicity for the formation of the S‐S bond.  相似文献   

14.
Thermal or photolytic reactions of bioactive S‐nitrosothiols and related thiols in the presence of radical generators in deaerated DMSO or aqueous solutions under argon or saturated with nitric oxide (NO) produced nitroxides and an oxyaminyl radical, which were well characterized by EPR spectra. Nitroxides containing a thiyl substituent were obtained. Possible mechanisms are proposed. Bioactive S‐nitrosothiols such as S‐nitrosoglutathione, S‐nitroso‐N‐acetylpenicillamine and related thiols such as glutathione and N‐acetylpenicillamine were used for the investigation. Radical generators utilized as transient radical sources were 2,2‐azobisisobutyronitrile, 2,2‐azobis(2‐methylpropionamidine) dihydrochloride, tert‐butyl peroxide and benzoyl peroxide. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

15.
《Analytical letters》2012,45(4):805-815
Abstract

This paper describes a kinetic spectrophotometric method for the determination of L‐ascorbic acid (AA) and thiols (RSH). Absorbance of Fe(II)‐phen complex formed during the reaction of AA or RSH with Fe(III)‐phen was continuously measured at 510 nm by double‐beam spectrophotometer with flow cell. For determination some thiols, the catalytic effect of Cu2+ ions was used. AA and RSH can be determined in concentration ranges from 4.0×10?6 to 4.0×10?5 M and from 8.0×10?6 to 8.0×10?5 M, respectively. The applicability of the proposed method was demonstrated by determination of chosen compounds in pharmaceutical dosage forms.  相似文献   

16.
The kinetics of Ir (III) chloride-catalyzed oxidation of D-glucose by iodate in aqueous alkaline medium was investigated at 45°C. The reaction follows first-order kinetics with respect to potassium iodate in its low concentration range but tends to zero order at its higher concentration. Zero-order kinetics with respect to [D-glucose] was observed. In the lower concentration range of Ir (III) chloride, the reaction follows first kinetics, while the order shifts from first to zero at its higher concentration range. The reaction follows first-order kinetics with respect to [OH?] at its low concentration but tends towards zero order at higher concentration. Variation in [Cl?] and ionic strength of the medium did not bring about any significant change in the rate of reaction. The first-order rate constant increased with a decrease in the dielectric constant of the medium. The values of rate constants observed at four different temperatures were utilized to calculate the activation parameters. Sodium salt of formic acid and arabinonic acid have been identified as the main oxidation products of the reaction. A plausible mechanism from the results of kinetic studies, reaction stoichiometry, and product analysis has been proposed.  相似文献   

17.
A procedure was developed for the synthesis of previously unknown β-chlorovinyl derivatives of PtIV chloride complexes by chloroplatination of terminal alkynes catalyzed by PtII chloride complexes. The reaction is highly stereoselective and gives only the products of trans-anti-addition of platinum and chlorine atoms. The regioselectivity of the catalytic reaction formally corresponds to Markovnikov’s rule, e.g., in alkynes containing electron-donating substituents, platinum attacks the terminal carbon atom. The σ-vinyl derivatives of PtIV chloride complexes were characterized by IR spectroscopy and 1H and 13C NMR spectroscopy. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2380–2384, October, 2005.  相似文献   

18.
Thiosilicates of the Rare‐Earth Elements: III. KLa[SiS4] and RbLa[SiS4] – A Structural Comparison Pale yellow, platelet shaped, air‐ and water resistant single crystals of KLa[SiS4] derived from the reaction of lanthanum (La) and sulfur (S) with silicon disulfide (SiS2) in a molar ratio of 2 : 3 : 1 with an excess of potassium chloride (KCl) as flux and source of potassium ions in evacuated silica ampoules at 850 °C within seven days. The analogous reaction utilizing a melt of rubidium chloride (RbCl) instead also leads to yellow comparable single crystals of RbLa[SiS4]. The potassium lanthanum thiosilicate crystallizes monoclinically with the space group P21/m (a = 653.34(6), b = 657.23(6), c = 867.02(8) pm, β = 107.496(9)°) and two formula units per unit cell, while the rubidium lanthanum thiosilicate has to be assigned orthorhombically with the space group Pnma (a = 1728.4(2), b = 667.23(6), c = 652.89(6) pm) and four formula units in its unit cell. In both compounds the La3+ cations are surrounded by 8+1 sulfide anions in the shape of tricapped trigonal prisms. The Rb+ cations in RbLa[SiS4] show a coordination number of 9+2 relative to the S2? anions, which form pentacapped trigonal prisms about Rb+. This coordination number, however, is apparently too high for the K+ cations in KLa[SiS4], so that they only exhibit a bicapped trigonal prismatic environment built up by eight S2? anions. The isolated thiosilicate tetrahedra [SiS4]4? of the rubidium compound are surrounded by La3+ both edge‐ and face‐capping, but terminal as well as edge‐ and face‐spanning by Rb+. In the potassium compound there is no change for the La3+ environment about the [SiS4]4? tetrahedra, but the K+ cations are only able to attach terminal and via edges. The whole structure is built up by anionic equation/tex2gif-stack-1.gif{La[SiS4]}? layers that are separated by the alkali metal cations. In direct comparison the two thiosilicate structures can be regarded as stacking variants.  相似文献   

19.
Three thiols with three aromatic rings and different structure – terphenyl-4-methanethiol (TPMT), terphenyl-4-thiol (TPT), and anthracene-2-thiol (AT) – have been used to form self-assembled monolayers (SAM) on vapour-deposited and flame-annealed Au films on glass substrates. All three SAMs effectively block the anodic formation of Au oxide, indicating densely packed layers which prevent the access of water and hydrated ions through the organic layer to the metal surface. The film improves its inhibiting properties with duration of exposure to the thiol solutions, reaching completion after 1 hour [1]. The charge-transfer reaction of the Fe(CN)6 3–/Fe(CN)6 4– system is blocked for TPMT films with an insulation of the π-electron system from the Au surface by the methylene group. TPT and especially AT films show the current density of the redox reactions. It is proposed that the charge transfer occurs via the aromatic molecules of the SAMs to the Au surface. Electronic Publication  相似文献   

20.
A Pt–NiCo nanomaterial has been synthesized for developing the sensitive electrochemical determination of biological thiols that include L ‐cysteine (CySH), homocysteine (HCySH), and gluthione (GSH) with high sensitivity and long‐term stability, in which the Pt nanoparticles are well supported on amorphous NiCo nanofilms. The electrochemical oxidation of thiols has been successfully facilitated on the optimized Pt–NiCo nanostructures, that is, two oxidation peaks of CySH have been clearly observed at potentials of +0.06 and +0.45 V. The experimental results demonstrate that the first peak for CySH oxidation may be attributed to a direct oxidation from CySH to L ‐cystine (CySSCy), whereas the second peak possibly results from a sequential oxidation from CySSCy to cysteic acid (CySO3H), together with a direct oxidation of CySH into CySO3H. The enhanced electrocatalytic activities at the Pt23–NiCo nanostructures have provided a methodology to determine thiols at a very low potential of 0.0 V with relatively high sensitivity (637 nA μM cm?2), a low detection limit (20 nM ), and a broad linear range. The striking analytical performance, together with the characteristic properties of the Pt–NiCo nanomaterial itself, including long‐term stability and strong antipoisoning ability, has established a reliable and durable approach for the detection of thiols in liver cancer cells, Hep G2.  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-23

京公网安备 11010802026262号