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 609 毫秒
1.
The sorption of 109Cd and 65Zn microquantities with NaX and NaA zeolites in the presence of divalent lanthanides (Ln2+) (Ln=Tm, Dy, and Nd) from tetrahydrofuran (THF) solutions was studied. It was found that in contrast to 137Cs+, 109Cd and 65Zn, as well as 85Sr2+ microquantities were sorbed by zeolites very poorly: ∼99% of the species remained in solution. The distribution coefficients (K d ) of 109Cd and 65Zn were ∼0.3 and ∼0.4 ml/g, respectively. When Tm2+ was used as a reducer, it partially oxidized in solution to Tm3+ to form precipitates of the composition TmI3·3THF. A study of the co-crystallization of 109Cd and 65Zn, as well as 85Sr2+ microquantities with the TmI3·3THF solid phase in the presence of Tm2+ from THF solutions showed that in contrast to 85Sr2+, 109Cd and 65Zn microquantities co-crystallized with the TmI3·3THF solid phase. The co-crystallization coefficients (D) for 109Cd and 65Zn depended on the ratio Tm3+/Tm2+ in solution and increased with its increasing. The assumption was made that in the presence of Tm2+, 109Cd2+ and 65Zn2+ form M+ ions, which quickly react with M2+ ions to form dimers of the composition M2 3+ (M=Cd and Zn).  相似文献   

2.
E. coli is able to exchange 2 H+ from the external medium for one K+ of a cell with the concurrent synthesis of ATP. The 2 H+/K+ exchange and the coupled ATP synthesis are in need of both the essential Δ\?gmH+ (around 24 kJ) and the high K+ gradient (103) directed from a cell to the medium. The reverse pump cycle is actuated as well as the direct one [3] merely through an increase of osmolarity in the environmental medium. The reverse process can be blocked by the insignificant ncrease in external pH or in K+ concentration, as well as by DCCD or by protonophore. The stoichiometry of the entire pump cycle is ATP: 2 H+ :K+. These observations, as well as the previous findings [8,9] that 2 H+ from a cell and one external K+ are transported through F1·F0 and TrkA respectively, suggest that F1·F0 and TrkA form the structural association in bacterial membrane for the joint employment of phosphate bond energy, the supercomplex (F1 ·F0TrkA operating as an electrogenic proton-potassium pump with stable stoichiometry ATP:2 H+ :K+.  相似文献   

3.
Summary A radioactively contaminated marine sediment core stemming from Irish Sea has been characterized by radiometric and mass spectrometric techniques as for 237Np, 241Am, 239Pu, 240Pu, 241Pu, 137Cs and 154Eu. The data obtained with independent methods in the framework of a QA/QC program as compared with the source term discharges, as well as with those reported in literature, are in good agreement.  相似文献   

4.
The Equations of Motion method has been applied in the calculation of potential energy curves for the X2Σ+g, A2Πu and B2Σ+u states of N+2. Results are also reported for a new dissociative 2Σ+g state. The theoretical curves are directly compared with the experimental ones as well as in terms of spectroscopic constants. The applicability of the Equations of Motion method to this type of problem is critically examined and discussed with regard to the choice of basis set, numerical effort and agreement with experiment.  相似文献   

5.
The 195Pt chemical shifts of several organoplatinum compounds in solution have been determined. The δ(195Pt) values of the phosphine-PtII and -Pt0 compounds lie in separate ranges, and allow the metal-diene systems to be characterized either as metallacyclopentene or as η2-bonded diene. Although the two isomers of bis(η3-allyl)Pt (VIII) formally should be regarded as PtII compounds their 195Pt shifts clearly lie in the region for Pt0 compounds. The large separation between the 195Pt signals and the difference in 195Pt-T1 values for the two isomers of VIII are in accord with their having different geometries around the metal.  相似文献   

6.
Summary The 239+240Pu content of the marine sediments and seawater of the costal sea of Korea was measured. In marine sediments the 239+240Pu concentrations were in the range of 0.11-1.91 Bq/kg dry weight and in the coastal sea of Korea the ratio of 239+240Pu/137Cs was 0.27. The correlations between 239+240Pu and 137Cs concentrations and the content of organic matter (C, O, H, N, S) as well as the grain size of marine sediment were investigated by regression analysis. The distribution coefficient of 239+240Pu was 1.22. 105. The 239+240Pu concentration in seawater increased with seawater depth. However, the 137Cs concentration in seawater did not change considerably with depth.  相似文献   

7.
The Flory theory of solution thermodynamics is used to predict exess volumes for systems containing a series of n-alkanes mixed with liquids of higher P* parameter and internal pressure, i.e., cyclopentane, cyclohexane, carbon tetrachloride, benzene and dioxane as well as of lower P*, i.e., decamethyltetrasiloxane. Trends of V E , e.g. changes of sign with alkane carbon number are well predicted and indicate the importance of the P* contribution in V E . Mixtures of hexane isomers with liquids of much higher P* typically have large excess enthalpies through zero as an S-shaped curve against composition which is negative on the side of the high P* component. This behaviour is interpreted as arising from increasingly large negative P* contributions in V E .  相似文献   

8.
The equilibrium constants and the thermodynamic parameters for the interaction of CoLx (L1 = 5-OMe-salabza, L2 = salabza, L3 = 5-Br-salabza and L4 = 5-NO2-salabza) as acceptors, with phosphines (PBu3, PPh2Me) as donors in dichloromethane were studied. This was performed by using UV-Vis spectrophotometry titration for 1:1 adduct formation of the selected complexes at various temperatures (T = 283–298 K). The trend of the adduct formation of the Co(II) complexes with a given phosphine donor decreases as CoL1 > CoL2 > CoL3 > CoL4. The stability of the resulting adducts with different Co(II)-schiff base complexes found to decrease in the order PBu3 > PPh2Me.  相似文献   

9.
Negative ion chemical ionization (NICI) mass spectra with methane as reagent gas and the ion abundance ratios of the negative to the positive base peak for 51 polycyclic aromatic hydrocarbons and related compounds were measured and evaluated for highly sensitive detection and isomer differentiation. Either [M ? H]?, M?˙ or MH? was the base peak, except for one compound with [M ? H2]?˙ as its base peak. The numbers of compounds with [M ? H]?, M?˙ or MH? as their base peaks were 17, 26 and 7, respectively. Many of the compounds with [M ? H]? as the base peak had an aliphatic part in their structure. The average value of N/P (negative/positive ion abundance ratio at the base peaks) was < 1. Many of the compounds with M?˙ as the base peak had a relatively high electron affinity. A correlation between electron affinities and ion abundances was found. In most cases, the N/P ratios were > 1, and even reached 400 in benzo [a] pyrene. Many of the compounds with MH? as their base peaks had a phenyl group, in which cases the N/P ratios were < 1. In the case of compounds with 18 or fewer carbon atoms, in particular, it was easy to distinguish isomers by comparing their NICI mass spectra. The N/P values served as a guideline in sensitive detection. Nine compounds achieved an N/P of ≥50.  相似文献   

10.
Summary Leach characteristics of 137Cs and 60Co radionuclides from spent mix bead ion exchange resins and both ordinary Portland cement and cement mixed with two kind of natural sorbents (bentonite and clinoptilolite) have been studied using the International Atomic Energy's (IAEA) standard leach method. The waste immobilization performance of low-level wastes in natural sorbent mixtures was determined. The solidification matrix was a standard Portland cement mixed with 290-350 (kg/m3) spent mix bead exchange resins, with or without 1-10% of bentonite or/and clinoptilolite. The leaching rates from the cement-bentonite matrix were measured after 300 days as 60Co: (1.20-9.72) . 10-5 cm/d and 137Cs: (1.00-9.22) . 10-4 cm/d. From the leaching data, the apparent diffusivity of cobalt and cesium in cement-bentonite or/and clinoptilolite matrix with a waste load of 350 kg/m3 of spent mix bead exchange resin was calculated as 60Co: (1.0-5.9) . 10-6 cm2/d and 137Cs: (0.48-2.4) . 10-4 cm2/d. The compressive strength of these samples is determined according to the ASTM standards.  相似文献   

11.
Extraction by polyhedral complex compound of the formula H+ [(π-(3)-1,2-B9C2H11)2Co] further on referred to as dicarbolide-H+ and its chloro-derivate H+[B18C4H15Cl7Co] further on referred to as Cl-dicarbolide-H+ in nitrobenzene was used for the analysis of137Cs in urine and faeces after internal contamination. The dependence of distribution ratio on the acidity of analysed solutions was determined. The effect of urine dilution was assessed as well as the effect of various concentrations of the extraction agents on the distribution ratio of137Cs. The effect of phase ratio at the different concentrations of isotopic carrier was assessed, as well as the effect of potassium ions on the decrease of the distribution ratio at the extraction of137Cs by dicarbolide-H+ or its chloro-derivate. The possibility of isolation of137Cs by extraction and the isolation of137Cs by ion-exchange absorbents and by ammonium molybdophosphate was compared. The values of distribution coefficient were determined at various concentrations of nitric acid and the isotopic carrier. The dependence of coextraction of some activated radionuclides and fission products by dicarbolide-H+ on the nitric acid concentration in the solute was determined. The effect of mass of contaminated faeces on the value of the distribution ratio of137Cs by the extraction was evaluated. As a result, a suggestion was given for the rutine isolation procedure of137Cs extraction with dicarbolide-H+ from the excreta contaminated by a mixture of radionuclides.  相似文献   

12.
133Cs NMR spectroscopy was used to study the interaction of 133Cs+ with monovalent cations in aqueous solution. The interaction between 133Cs+ and the cations causes a repulsive polarization in the electron environment of 133Cs+, which is detected by a change in its resonance frequency. The resulting 133Cs chemical shift seemed to be dependent on the activities and the single-ion enthalpies and entropies of hydration of the cations. The hydration cospheres of an ion are the barriers through which 133Cs+ must pass in order to come into contact with the cation. These hydration cosphere barriers can be related to the single-ion enthalpy and entropy of hydration of the cation. The experiments revealed that 133Cs+ is not nearly as sensitive a probe as 129Xe, presumably due to its more tightly bound hydration cospheres. However, 133Cs NMR was moderately successful in predicting trends in the chemical shift behavior induced by the cations studied.  相似文献   

13.
The generation and distribution of radiation defects in single-crystal samples of indium selenide with an electron number density of 5 × 1014−1 × 1016 cm−3 during irradiation with γ-rays (doses of 107 and 108 R) and 25-MeV electrons (fluences of 1015 particle/cm2) was studied, as well as the behavior of the mobility.  相似文献   

14.
The kinetics of sorption on zirconium hydrophosphates with various water contents was studied. The self-diffusion coefficients of Ni2+ and H+ and the effective diffusion coefficients of Ni2+ ions corresponding to the exchange Ni2+ → H+ were determined. The self-diffusion coefficients of ions in samples with an 85% water content were as high as 4.17 × 10?11 m2/s (Ni2+) and 5.06 × 10?10 m2/s (H+). Water loss caused by drying the sorbents resulted in a decrease in the rate of ion exchange.  相似文献   

15.
An ion chromatography‐inductively coupled plasma mass spectrometric (IC‐ICP‐MS) method for the speciation of sulfur compounds, namely sulfite [SO32?], sulfate [SO42?] and thiosulfate [S2O32?], was described. Ionic sulfur compounds were well separated in about 3 min by ion chromatography with a Hamilton PRP‐X100 column as the stationary phase and 60 mmol L?1 NH4NO3 and 0.1% v/v formaldehyde (HCHO) solution (pH = 7) as the mobile phase. The analyses were carried out using dynamic reaction cell (DRC) ICP‐MS. The sulfur‐selective chromatogram was determined at m/z 48 as 32S16O+ by using its reaction with O2 in the reaction cell. The method avoided the effect of polyatomic isobaric interferences at m/z 32 caused by 16O16O+ and 14N18O+ on 32S+ by detecting 32S+ as the oxide ion 32S16O+ at m/z 48, which is less interfered. The detection limit of various species studied was in the range of 3.6–4.6 ng S mL?1. The accuracy of the method has been verified by comparing the sum of the concentrations of individual sulfur compounds obtained by the present procedure with the total concentration of sulfur in several natural water samples. The recovery was in the range of 97–102% for various compounds studied.  相似文献   

16.
The kinetics of the extraction of phenylsuccinic acid (PSA) enantiomers by hydroxypropyl-β-cyclodextrin (HP-β-CD) in a modified Lewis cell was studied, in which HP-β-CD dissolved in 0.1 mol L?1 NaH2PO4/H3PO4 buffer solution (pH = 2.5) was selected as the chiral extractant. PSA enantiomers were extracted from organic phase to aqueous phase in the extraction module. The theory of extraction accompanied by a chemical reaction has been used to obtain the intrinsic kinetics of this extraction module. The different parameters affecting the extraction rate such as agitation speed, interfacial area, initial concentration of PSA enantiomers in organic phase as well as HP-β-CD concentration in aqueous phase were separately studied. The experimental results demonstrate that the extraction reactions are fast. The reactions were found to be first order with respect to PSA and second order with respect to HP-β-CD with forward rate constants of 3.4 × 10?2 m6 mol?2 s?1 for R-PSA and 9.96 × 10?3 m6 mol?2 s?1 for S-PSA. These data will be useful in the design of extraction processes.  相似文献   

17.
The reaction between Au(I), generated by reaction of thallium(I) with Au(III), and peroxydisulphate was studied in 5 mol dm?3 hydrochloric acid. The reaction proceeds with the formation of an ion‐pair between peroxydisulphate and chloride ion as the Michealis–Menten plot was linear with intercept. The ion‐pair thus formed oxidizes AuCl2? in a slow two‐electron transfer step without any formation of free radicals. The ion‐pair formation constant and the rate constant for the slow step were determined as 113 ± 20 dm?3 mol?1 and 5.0 ± 1.0 × 10?2 dm3 mol?1 s?1, respectively. The reaction was retarded by hydrogen ion, and formation of unreactive protonated form of the reductant, HAuCl2, causes the rate inhibition. From the hydrogen ion dependence of the reaction rate, the protonation constant was calculated to be as 0.6 ± 0.1 dm3 mol?1. The activation parameters were determined and the values support the proposed mechanism. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 589–594, 2002  相似文献   

18.
The kinetic behavior of CrO3 in its reaction with wood has been elucidated. Various reactions take place between CrO3 and the lignin and cellulose in wood. CrO3 reacts with cellulose in a two-step reaction: the first step is an adsorption of CrVI onto the cellulose to form CrVI/cellulose activated complexes. The second step is a CrVI → CrIII reduction taking place on the cellulose surface. The CrIII formed is only physically adsorbed to the cellulose or very weakly bound as small amounts of CrIII can be released into the reaction medium. The CrVI adsorbed by cellulose appears mainly to be reduced to CrIII. The reaction of CrVI with lignin has been shown to be the composition of the three successive reaction of Cr2O72?, HCrO4?, and CrO42? with the guaiacyl units of lignin. Insoluble and stable CrVI/lignin complexes in which chromium maintains its hexavalent oxidation state are formed. Rate constants and energies of activation for all the reactions have been determined. The fixation of CrO3-derived compounds on wood has been explained as the combination of the various reactions investigated. The results indicate that 60% of Cr is fixed irreversibly to the lignin of wood as CrVI and 40% is weakly bound, probably just precipitated, on the cellulose surface as CrIII of which small amounts can be released in a water medium. The complex CrVI and CrIII species forming complexes with the guaiacyl units have been identified.  相似文献   

19.
Evidence of a one-electron transfer process in a carbene reaction has been observed for the first time. The example is the quenching of the photoexcited triplet state of diphenylcarbene (3*DPC) by electron donors. Measurement of the fluorescence lifetime as a function of donor concentration yielded the bimolecular rate constant, 3*k. An explanation is offered as to why 3* and 1DPC react efficiently with amines as well as alcohols, whereas the ground triplet, 3DPC, does not.  相似文献   

20.
The kinetics of the hydrogen–deuterium (H–D) exchange at both the methine (alpha) and methylene (gamma) positions of glutamic acid in deuterated hydrochloric acid solution has been studied in the temperature range of 383–433 K by 1H NMR detection. The reaction rates of H–D exchange at the two positions were described by applying multivariable linear regression (MLR) analysis and are determined as v = k[Glu]3.3[D3O+]1.5 mol L?1 h?1 with k = 3.52 × 1016 × exp (–1.37 × 105/RT) mol?3.8 L h?1 for the alpha position as well as v = k[Glu]1.0[D3O+]0.45 mol L?1 h?1 with k = 1.77 × 1012 × exp (–0.99 × 105/RT) mol?0.45 L h?1 for the gamma position. The Arrhenius activation energy (Ea) at the gamma position is less than that at the alpha position, which implies that the deuteration reaction at the gamma position proceeded more easily.  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