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 703 毫秒
1.
2.
The anodization of mercury microelectrodes was investigated in synthetic samples containing several strong and weak electrolytes at different concentrations. In particular, the effects on mercury anodization due to the presence of NaOH, HClO4, NaCl, NaI, NaF, Na2SO4, NaHCO3, Na2CO3, tartaric and citric acids, were studied in solutions containing either each species or mixtures of them, and without addition of supporting electrolyte. Some of the electrode processes studied led to linear calibration plots e.g. 1 × 10−5 − 1 × 10−4M Cl, 1 × 10−6 − 1 × 10−5M I, 5 × 10−4 − 3 × 10−3M SO42−, 5 × 10−4 − 2 × 10−2M HCO3, with typical correlation coefficients of 0.998–0.999. The anodization of mercury microelectrodes was also investigated directly in wine, rain, tap and mineral water, without pretreatment and without addition of supporting electrolyte. In the real samples only the ions Cl and HCO3 could be quantified, and the values found were in agreement, within 3–5%, with the reference values obtained by using Italian standard methods for food.  相似文献   

3.
A predominantly localized electron pair scheme is outlined for describing the electron distribution and bonding in closo borane anions BnHn2− and related electron deficient deltahedral clusters, in which a skeletal electron pair is assigned to each vertex, one pair being regarded as delocalized just inside the roughly spherical surface on which the skeletal atoms lie. The scheme gives a clearer picture of the electron distribution than is conveyed by resonating 2- and 3-centre bonds in the polyhedron edges and faces, and allows the bond orders of the polyhedron edge links to be calculated readily. The consequence of formal removal of BH2+ units from closo species BnHn2− to generate nido species Bn−1Hn−14− and arachno species Bn−2Hn−26− is explored, and seen to allow rationalization of two features of such deltahedral-fragment clusters: (i) why a high-connectivity vertex is left vacant and (ii) why the frontier orbitals of such species concentrate electronic charge around their open faces. Moreover, in the case of D4‘h B4H46− (cf. C4H42−) and D5h B5H56− (cf. C5H5), the approach leads directly to the familiar picture for aromatic ring systems in which the highest filled, doubly degenerate π-bonding molecular orbital concentrates electronic charge in rings above and below the polygon on which the skeletal nuclei lie. It also leads to the expectation that arachno clusters with non-adjacent vacant vertices will be more stable than those with adjacent vacant vertices.  相似文献   

4.
《Chemical physics》1995,200(3):309-318
Dynamics of electronic polarization in the vicinity of charge carriers in molecular crystals is for the first time investigated here in connection with the carrier transport and intramolecular vibronic polarization. According to standard picture it has been assumed that the electronic polarization relaxation time is extremely short, as estimated from the relation τc = τd1h/Eexc, where Eexc is the energy of the first single exciton state. In the case of anthracene (Ac) crystals, the value of τe is about 2 × 10−16 s, i.e. by several orders of magnitude shorter than a typical hopping (residence) time of charge carriers τh = 10−14 -10−13 s. It is argued that typical time of full reconstruction of the electronic polarization after individual carrier hops equals, in the slow carrier regime, approximately to td2hEexc is the width of the lowest singlet-exciton band. In Ac, this means td2 ≈ 0.73 × 10−14 s. Physical implications of this relatively high value of td2 in connection with carrier transport and molecular (vibronic) polarization are discussed.  相似文献   

5.
Organic-rich natural waters from peat bogs in continental (Switzerland) and maritime (Shetland Islands, Scotland) areas were analysed for Cl, NO2, Br, NO3, HPO42−, SO42− and oxalate using ion chromatography. These anions can be determined simultaneously in the surface and pore water samples from the continental bogs using a 250-μl injection loop. Using this loop, the detection limits were ca. 5 ng/g for the monovalent anions and SO42− and 10 ng/g for HPO42− and oxalate. An organics-removal cartridge (Dionex OnGuard P) was used to remove humic materials. These cartridges did not significantly affect the measured concentrations of anions in blind standards. Analyses of deionized water treated with these cartridges are not significantly different from those for untreated deionized water. For the maritime bogs, the relatively high concentrations of Cl (more than 100μ/g in many samples) and SO42− (up to 50 μg/g) require two separate determinations for complete analyses. A 10-μl injection loop was used to determine Cl, Br and SO42−. A 250-μl injection loop was used to measure NO2, NO3, HPO 42− and oxalate. In each instance a Dionex OnGuard P cartridge was used to remove humic materials. In addition, a chloride-removal cartridge (Dionex OnGuard AG) was used to remove Cl when the larger injection loop was used. This cartridge has no significant effect on the measurement of HPO4-2− at concentrations of 20 ng/g. In each of the bog water chromatograms there were usually a number of unknown peaks. These are probably due mainly to organic anions.  相似文献   

6.
Monte Carlo simulation studies of statistical perturbation theory (SPT) have been carried out to investigate the solvent effects on the relative free energies of solvation and the difference in partition coefficients (log P) for K+ to Na+ ion mutation in the several solvents. We compared the relative free energies for interconversion of K+ to Na+, in H2O (TIP4P) in this study with those published works, that in H2O (TIP4P) is −16.55 kcal/mol in this study, those of the published works are −17.6, −17.3 and −17.31 kcal/mol and that of the experiment is −17.6 kcal/mol, respectively. Comparing the relative free energies for interconversion of K+ to Na+, in CH3OH in this study with those published works, that in CH3OH is −18.08±0.28 kcal/mol in this study, that of molecular dynamic simulation is −19.6±0.4 kcal/mol and that of the experimental work is −17.3 kcal/mol, respectively. There is good agreement among the several studies if we consider both methods of obtaining the solvation (or hydration) free energies and the standard deviations. For the present K+ and Na+ ions, the relative free energies of solvation vs Born's function of solvents are decreased with increasing Born's function of solvent except for CH3OH, THF and MEOME. There is also good agreement between the calculated structural properties in this study and the computer simulation, ab initio and experimental works.  相似文献   

7.
The self-diffusion coefficient of water in Nafion-117 membrane in different cationic forms was measured by the transient radiotracer method, which is based on an analytical solution of Fick's second law. The self-diffusion coefficient of water in the membrane was obtained from the analysis of time-dependent isotopic-exchange rates of tritium tagged water between sample of Nafion-117 membrane and equilibrating water. This transient method does not require the knowledge of partition coefficient of water, which is an essential parameter in the radiotracer permeation method. In present work, self-diffusion coefficients of water in the Nafion-117 membrane with H+, Li+, Ag+, Na+, K+, Rb+ and Cs+ monovalent counterions were obtained. The values of logarithm of self-diffusion coefficients were found to vary linearly as a function of polymer volume fraction except for membrane sample with H+ counterions. The self-diffusion coefficient of water in Nafion-117 membrane with H+ counterions was significantly different from the trend observed in the variation of self-diffusion coefficient of water as a function of polymer volume fraction in the membrane with other monovalent counterions. This observation seems to suggest that the physical structure of Nafion-117 membrane in H+ form may be quite different from the Nafion-117 membrane with other monovalent counterions. The high self-diffusion coefficient of water (1.67 × 10−6 cm2/s) in Nafion-117 membrane with Cs+ counterions indicates that the ionic clusters in Nafion-117 membrane are well connected even at low water content (8.2 wt.%) in the membrane.  相似文献   

8.
Natural salt minerals often contain inclusions of saturated salt solutions with diameters from 1 to> 100 μm. With the quantification of the composition of the fluid inclusions, the origin and metamorphism of the salt rocks can be interpreted. Hence, these data are important concerning the long-term safety of underground repositories in salt rocks [1]. For the extraction of the solutions in fluid inclusions with diameters 300 μm, an optical precision instrument was developed. For the simultaneous determination of Cl, Br, SO42−, Li+, Na+, K+, Mg2+ and Ca2+ two ion chromatographic systems with conductivity detection for cations and anions and additional photometric detection for Br were used. To prevent column overload, the Cl concentration must be less than 50 μg/ml in the measuring solution. The extracted samples (volumes> 0.1 μl) are diluted with demineralized water by a factor of 1 · 104 (20-μl sample loops). The practical limit of determination for the measured elements is 0.01–0.3 μg/ml in the measuring solutions. By calculation of the anion and cation charge balance (molar equivalence), a relative error of <5% for the analysis of fluid inclusions was found.  相似文献   

9.
The objective of this study was to investigate the retention of phosphate anions, H2PO4 and HPO42−, by nanofiltration. The first part of this study deals with the characterisation of the NF200 membrane used in permeation experiments with aqueous solutions of neutral organic and charged inorganic solutes. In the second part the effects of feed pressure, ionic strength, concentration and pH on the retention of phosphate anions were investigated. Results show that the membrane is negatively charged, its pore radius is around 0.5 nm and the retention order for the salts tested was R(Na2SO4) > R(NaCl) > R(CaCl2). The retentions of phosphate anions are in the order of 85% for H2PO4 and 96% for HPO42−. They are relatively high when compared to retentions of other anions with the same charge. The retentions of phosphate anions, particularly the monovalent species, depend on the chemical parameters (feed concentration, ionic strength, and pH) and applied pressure. The experimental data were analysed using the Speigler–Kedem model and the transport parameters, i.e., the reflection coefficient (σ) and solute permeability (Ps) have been determined.  相似文献   

10.
Amberlite XAD-2 has been functionalized by coupling it, through the ---N=N--- group, with Pyrocatechol Violet (PV), and the resulting resin has been characterized by elemental analysis, thermogravimetric analysis (TGA) and IR spectra. The resin has been used for preconcentrating Zn(II), Cd(II), Pb(II) and Ni(II) ions prior to their determination by flame atomic absorption spectrometry. The optimum pH values for quantitative sorption are 5, 5–7, 4, and 3 for Zn, Cd, Pb and Ni, respectively. The four metals can be desorbed (recovery ˜98%) with 4 M HNO3; also, 4 M HCl is equally suitable except for Zn. The sorption capacity of the resin is 1410, 1270, 620 and 1360 μg g−1 resin for Zn, Cd, Ni and Pb, respectively. The effect of F, Cl, NO3, SO42− and PO43− on the sorption of these four metal ions has been investigated. They are tolerable in the range 0.01–0.20 M, for Pb. In the sorption of Zn(II) and Ni(II), the tolerance limits of all these ions are upto 0.01 M, whereas for Cd(II), F, NO3, and PO43− have been found to be tolerable upto 0.50, 0.10 and 0.10 M, respectively. The preconcentration factors are 60, 50, 23 and 18 for Zn, Cd, Pb and Ni, respectively. Simultaneous collection and determination of the four metals are possible. Cations commonly present in drinking water do not affect the sorption of either metal ion if present at a concentration level similar to that of water. The method has been applied to determine Zn, Ni and Pb content of well-water samples (RSD ≤9%).  相似文献   

11.
We have previously determined an analytical ab initio six-dimensional potential energy surface for the HCl dimer, and in the present paper we use this potential, with the HCl bond lengths held fixed, in a full (four-dimensional) close-coupling calculation to determine the energies of the lowest 24 vibrational states. These vibrational states involve the intermolecular stretch ν4, the trans-bend tunneling vibration ν5, and the torsion ν6. The highest of the 24 levels is the (ν4ν5ν6)=(111) state, for which we calculate an energy of 200 cm−1 above the (000) state. As well as determining tunneling energies up to 5ν5=183 cm−1, we determine ν4=49 cm−1, 2ν4=93 cm−1, 3ν4=134 cm−1, 4ν4=172 cm−1, ν6=137 cm−1 and ν46=178 cm−1, together with tunneling energies in all these states. Making allowance for the HCl stretching zero-point energy we determine the dissociation energy D0 as 390 cm−1 on this analytical surface. We determine that below 300 cm−1 there are 72 vibrational (J=K=0) states, and below dissociation there are 162 vibrational (J=K=0) states, for this potential surface.  相似文献   

12.
Agnihotri NK  Singh VK  Singh HB 《Talanta》1993,40(12):1851-1859
Derivative photometric methods for trace analysis of Th(IV) and UO2(II), and their simultaneous determination in mixtures using 5,8-dihydroxy-1,4-naphthoquinone in a micellar medium are reported. Molar absorptivity and Sandell's sensitivity of 1:2 Th(IV) and 1:1 UO2(II) complexes at their λmax, 614.5 nm and 637.0 nm are, 1.19 × 104 1/mol/cm and 1.12 × 104 1/mol/cm and 1.95 × 10−2 μg/cm2 and 2.13 × 10−2 μg/cm2 μg/cm2, respectively. Calibration graph is linear over the range 9.28 × 10−2−18.56 μg/ml of Th(IV) and 9.52 × 10−2−19.04 μg/ml of UO2(II). Though presence of Th(IV) and UO2(II) causes interference in each others determination, 9.28 × 10−1−9.28 μg/ml Th(IV) and 9.52 × 10−1−9.52 μg/ml UO2(II) when present together, can be simultaneously determined using derivative spectra.  相似文献   

13.
The MNDO molecular orbital method is employed to calculate the proton affinities of fluorinated formaldehydes and acetones. Agreement with experimentally reported proton affinities is good. In the acetone series a decrease in proton affinity is calculated for each successive fluorine substituent. The calculated strength of the intramolecular hydrogen bond in the protonated fluoro-formaldehydes and acetones is 0.6—2.7 kcal mol−1, in good agreement with the experimental value of 2—3 kcal mol−1 in the protonated fluoroacetones. Examination of the calculated charge distribution shows that the trends in proton affinity can be understood qualitatively both in terms of initial-state and final-state effects caused by the fluorine substituents. Protonation at the fluorine atom is less stable by about 25 kcal mol−1 than protonation at the oxygen atom for monofluoroacetone.  相似文献   

14.
The spectrum of CD2HF was measured by high-resolution interferometric Fourier-transform IR (FTIR) spectroscopy (apodised instrumental band with:0.004 cm−1 fwhm) between 800 and 1200 cm−1 covering the four lowest fundamentals. A complete rotational analysis using a semi-automatic assignment procedure yields accurate band centres (ν9: 912.2028 cm−1, ν6:964.4994 cm−1, ν5: 1050.5104 cm−1, ν4: 1093.8632 cm−1) and a complete set of first-order Coriolis coupling constants. The most important couplings occur between ν9 and ν6a= 1.069 cm−1, ξc= −0.3535 cm−1) and between ν5 and ν4b= −0.80606 cm−1). The analysis was guided by and compared with results from our ab initio calculations for Coriolis constants and transition moments using CADPAC at TZP/MP2 level.  相似文献   

15.
Matos RC  Coelho EO  Souza CF  Guedes FA  Matos MA 《Talanta》2006,69(5):1208-1214
The importance of atmospheric hydrogen peroxide (H2O2) in the oxidation of SO2 and other compounds has been well established. A spectrophotometric method for the determination of hydrogen peroxide in rainwater is proposed. This method is based on selective oxidation of hydrogen peroxide using an on-line tubular reactor containing peroxidase immobilized on Amberlite IRA-743 resin. The hydrogen peroxide in the presence of phenol, 4-aminoantipyrine and peroxidase, produces a red compound (λ = 505 nm). Beer's law is obeyed in a concentration range of 1–100 μmol l−1 hydrogen peroxide with an excellent correlation coefficient (r = 0.9991), at pH 7.0, with a relative standard deviation (R.S.D.) <2%. The detection limit of the method is 0.7 μmol l−1 (4.8 ng of H2O2 in a 200 μl sample). Measurements of hydrogen peroxide in rain samples were carried out over the period from November 2003 to January 2005, in the central area of the Juiz de Fora city, Brazil. The concentration of H2O2 varied from values lower than the detection limit to 92.5 μmol l−1. The effects of the presence of nonseasalt (NSS) SO42−, NO3 and H+ in the concentration of hydrogen peroxide in the rainwater had been evaluated. The average concentrations of H2O2, NO3, NSS SO42− and SO42− are 23.4, 18.9, 7.9 and 10.3 μmol l−1, respectively. The pH values for 82% of the collected samples are greater than 5.0. The spectrophotometeric method developed in this work that uses enzyme immobilized on the resin ion-exchange compared with the amperometric method did not present any significant difference in the results.  相似文献   

16.
Optical electron transfer in the mixed-valence cation of biferrocenylacetylene (BF+) has been examined in CD2Cl2 solvent. The intervalence absorption line shape is relatively narrow at both low and high chromophore concentrations, but broader at intermediate concentrations. The transition energy for metal-to-metal charge transfer increases from ≈4440 cm−1 at infinite dilution to 5995 cm−1 for 3.8 mM BF+. Related effects exist due to added electrolyte. Neither the electrolyte nor chromophore concentration effects are expected from a simple reading of electron transfer theories. Nevertheless, both phenomena can be understood and within the context of theory upon careful consideration of the effects of ion-pairing (and tripling) equilibria upon electron-transfer energetics.  相似文献   

17.
A novel sequential injection (SI) method was developed for the determination of penicillamine (PA) and ephedrine (EP) based on the reaction of these drugs with tris(bipyridyl)ruthenium(II) (Ru(bpy)32+) and peroxydisulfate (S2O82−) in the presence of light. Derivatization of PA and EP with aldehydes has resulted in a significant enhancement of the chemiluminescence emission signal by at least 25 times for PA and 12 times for EP, leading to better sensitivities and lower detection limits for both drugs. The instrumental setup utilized a syringe pump and a multiposition valve to aspirate the reagents, (Ru(bpy)32+ and S2O82−), and a peristaltic pump to propel the sample. The experimental conditions affecting the derivatization reaction and the chemiluminescence reaction were systematically optimized using the univariate approach. Under the optimum conditions linear calibration curves between 0.2–24 μg mL−1 for PA and 0.2–20 μg mL−1 for EP were obtained. The detection limits were 0.1 μg mL−1 for PA and 0.03 μg mL−1 for EP. The procedure was applied to the analysis of PA and EP in pharmaceutical products and was found to be free from interferences from concomitants usually present in these preparations.  相似文献   

18.
The photochemical reaction of azide derivatives induced by ultraviolet (UV) laser in matrix-assisted laser desorption/ionization mass spectrometry (MALDI) is reported. A novel synthesized class of azide aromatic derivatives, spin-labeled photoaffinity non-nucleoside adenosine triphosphate (ATP) analogs which are useful probes in study of muscle contraction mechanism, is used in this investigation. In the negative ion MALDI spectra of these ATP analogs, “fingerprint” peaks corresponding to [M − 10 − 1], [M − 12 − 1], [M − 16 − 1], [M − 26 − 1], [M − 28 − 1], [M − 41 − 1], and [M − 42 − 1] were observed with relative intensities depending on the MALDI matrix. Only the [M − 16 − 1] is present in the similar mass spectra of the analog in which the azido group is replaced by a hydrogen. A model is suggested for the photochemical reactions of azide derivatives under UV laser irradiation. The photoreaction fingerprint information is diagnostically useful in characterization of azido compounds, especially for spin-labeled photoaffinity non-nucleoside ATP analogs.  相似文献   

19.
A method for low-molecular-mass anion screening is described using a buffer composed of 5-sulfosalicylate (SS) as a visualizing ion, hexadimethrine bromide as an electroosmotic flow modifier and Tris as a pH buffer component, at pH 8.6. All ions with effective mobility higher than 2610−9 m2 s−1 V−1 can be separated within 7.5 min under −30 kV. By using the moderately mobile SS (5410−9 m2 s−1 V−1), not only the sensitivity of the detection is improved due to its high UV absorptivity, but also a smaller overall overloading effect is achieved. Meanwhile, the resolution of the high mobility ions, which is normally critical, remains almost the same as compared to a chromate buffer. With an electrokinetic injection, the limit of detection (LOD) of the common ions is 2–13 nM and the detection range is linear up to 0.5–3 μM. With a hydrostatic injection the LOD is 0.15–1 μM and the detection range is linear up to 25–200 μM. The identification of ions is performed by comparing the mobility of the ions with that of standards, taking the apparent and effective mobility of HCO3, which is normally present in the sample solution, as a reference.  相似文献   

20.
We present a molecular dynamics study of the solvation properties of large spherical ions S+ and S of same size, in water, chloroform and acetonitrile solutions, and at a water–chloroform interface. According to the “extrathermodynamic” TATB hypothesis, such ions have identical free energies of transfer from water to any solvent. We find that this is not the case, because S interacts better than S+ with water (by about 20 kcal mol−1), while S+ is better solvated by acetonitrile (by about 2 kcal mol−1) and chloroform (about 8 kcal mol−1) solvents. The importance of “long-range” electrostatic interactions on the charge discrimination by solvent is demonstrated by the comparison of standard vs corrected methods to calculate: (i) the electrostatic potential at the centre of the solute; (ii) the interaction energies between the ions and the solvents; and (iii) the free energies of charging the neutral sphere S0 to S+ and S, respectively. These conclusions are obtained with several solvent models and simulation conditions. The question of ion pairing for the S+S, S+Cl and SNa+ pairs is also examined in the three solvents. Finally, simulations at a liquid–liquid water–chloroform interface represented explicitly, show that S+ and S are highly surface active, although they do not possess, like classical surfactants, an amphiphilic topology. Adsorption at the interface is found with different methodologies and at different ion concentrations. These results are important in the context of the “TATB hypothesis”, and for our understanding of solvation of large hydrophobic ions in pure liquids or in heterogeneous liquid environments.  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