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 31 毫秒
1.
Tarazi L  George A  Patonay G  Strekowski L 《Talanta》1998,46(6):1413-1424
The spectral features of the near-infrared (NIR) dye TG-170 in different solutions and its complexation with several metal ions were investigated. The absorbance maxima of the dye are at λ=819, 805, and 791 nm in dimethyl sulfoxide (DMSO), methanol, and a buffer of pH 5.9, respectively. These values match the output of a commercially available laser diode (780 nm), thus making use of such a source practical for excitation. The emission wavelengths of the dye are at λem =822, 812, and 803 nm in DMSO, methanol, and the buffer, respectively. The molar absorptivity and fluorescence quantum yield increase accordingly. The addition of either an Al(III) ion or Be(II) ion resulted in fluorescence quenching of the dye. The Stern–Volmer quenching constant, KSV, was calculated from the Stern–Volmer plot to be KSV=3.11×105 M−1 for the Al(III) ion and KSV=1.17×106 M−1 for the Be(II) ion. The molar ratio of the metal to the dye was established to be 1:1 for both metal ions. The stability constant, KS, of the metal–dye complex was calculated to be 4.37×104 M−1 for the Al–dye complex and 1.94×106 M−1 for the Be–dye complex.  相似文献   

2.
Reartes GB  Liberman SJ  Blesa MA 《Talanta》1987,34(12):1039-1042
The acidity constants of benzidine (Bz) in aqueous solutions determined potentiometrically at 25° were Ka1 = (1.11 ± 0.08) × 10−5, Ka2 = (1.45 ± 0.12) × 10−4. The apparent mixed constants in 0.1M sodium nitrate are Ka1 = (5.37 ± 0.28) × 10−6 and Ka2 = (1.14 ± 0.09) × 10−4. The ultraviolet spectra were recorded as a function of pH and analysed with these constants to obtain the absorption spectra of H2Bz2+, HBz+ and Bz; the corresponding wavelengths of maximal absorption are 247, 273 and 278 nm, and molar absorptivities 1.63 × 104, 1.76 × 104 and 2.26 × 104 1.mole−1.cm−1.  相似文献   

3.
Kudo Y  Usami J  Katsuta S  Takeda Y 《Talanta》2003,59(6):1213-1218
Ion-pair formation constants (KMLA mol−1 dm3) of Na+– and K+–18-crown-6 ether (18C6) complexes with MnO4 in water (w) were determined potentiometrically at 25 °C. Simultaneously, extraction constants (Kex mol−2 dm6) of the permanganates with 18C6 from w into 1,2-dichloroethane at 25 °C were obtained from the spectrophotometric determination of distribution ratios of the permanganates. These Kex values were divided into KMLA and other three component equilibrium constants and thereby extraction-selectivity and -ability were discussed in comparison with corresponding metal picrate–18C6 extraction systems reported before.  相似文献   

4.
Gu Z  Wang X  Gu X  Cheng J  Wang L  Dai L  Cao M 《Talanta》2001,53(6):194-1170
Fulvic acids (FAs) were extracted by alkali extraction from different soil samples in China, then purified using resins and characterized by Fourier transform infrared spectroscopy. The complexing ability of FAs was investigated by measuring the stability constants of rare earth elements (REEs) (La3+, Ce3+, Sm3+, Gd3+, Y3+) with FAs by the ion exchange technique. The results indicated that maximum binding ability forY3+ (4.414.44) was higher than other REEs (La3+, Ce3+, Sm3+, Gd3+) (0.721.03). There were two types of binding sites in the functional groups of fulvic acids. The complexing reaction followed two steps. The stability constants (K1 and K2) of REEs with FAs were calculated from experimental data by division of Scatchard plots into two straight-line segments. Y3+ (log K1=5.72±0.05, log K2=4.83±0.01) also has higher stability constants than the other four REEs (log K1=4.37±0.16, log K2=3.62±0.28).  相似文献   

5.
J. Femi Iyun  Ade Adegite 《Polyhedron》1989,8(24):2883-2888
At 25°C, I = 1.0 M (CF3SO3Li++CF3SO3H), [H+] = 0.034–0.274 M and λ = 453 nm, the rate equation for the oxidation of Ti(H2O), 63+ by bromine was found to be: −d/[Br2]T/dt=kK/[Br2][TiIII]/[H+]+K+kK/[Br3][TiIII]/[H++K, where k = 9.2 × 10−3 M −1 s −1 and K = 4.5 × 10−3 M. At [H+] = 1.0 M, [Br] = 0.05–0.4 M, the apparent second-order rate constant decreases as [Br] increases.

The pH-dependence of the oxidation of TiIII-edta by bromine is interpreted in terms of the change in identity of the TiIII-edta species as the pH of the reaction medium changes. The second-order rate constants were fitted using a non-linear least-square computer program with (1/k0edta)2 weighting into an equation of the form: k0edta =k1+k2K1[H+]−1+k3K1K2[H+]−2/1+K1[H+[H+−1+K1K2[H+]−2, with K1 and K2 fixed as earlier determined at 9.55 × 10−3 and 2.29 × 10−9 M, respectively, for the oxidation of bromine. k1=k2=(3.1±0.32)×103M−1s−1 k3=(2.3±0.45)×106N−1s−1.

It is proposed that these electron transfer reactions proceed by univalent changes with the production of Br2.− as a transient intermediate. An outer-sphere mechanism is proposed for these reactions. The homonuclear exchange rate for TiIII-edta+TiIV-edta is estimated at 32 M−1 s−1.  相似文献   


6.
Blanco SE  Ferretti FH 《Talanta》1998,45(6):1103-1109
A UV spectrometric method was developed to determine the molar absorptivity (C) and formation constant (Kc) of the association complex of unsubstituted chalcone in cyclohexane, in the concentration range from 4.00·10−4 to 2.00·10−2 mol dm−3. The thermodynamic and spectroscopic magnitudes such as Kc and C contribute to the understanding of the physicochemical behavior of several ,β-unsaturated carbonylic compounds, of low solubility in water, as it is the case of numerous flavonoids of chemical and biological importance. The studied association complex, formed by two chalcone molecules, is characterized by the constants C (300.8 nm)=4.98·104 dm3 mol−1 cm−1 and Kc=5.58·103. The method proposed is convenient for the study of solute–solute molecular associations particularly those due to dipole–dipole interactions.  相似文献   

7.
Formation constants for recrystallized thymol blue were determined in water, using the SQUAD and SUPERQUAD programs. The best model correlating spectrophotometric, potentiometric and conductimetric data was fitted with the dissociation of HL=L2−+H+−log K=8.918±0.070 and H3L2=2L2−+3H+−log K=29.806±0.133 with the SUPERQUAD program at variable low ionic strength (1.5×10−4–3.0×10−4 M); and HL=L2−+H+−log K=8.9±0.000, H3L2 =2L2−+3H+−log K=30.730±0.032, H4L2=2L2−+4H+−log K=32.106±0.033 with SQUAD at 1.1 M ionic strength.  相似文献   

8.
Saran L  Cavalheiro E  Neves EA 《Talanta》1995,42(12):2027-2032
The highly neutralized ethylenediaminetetraacetate (EDTA) titrant (95–99% as Y4− anion) precipitates with Ag+ cations to form the Ag4Y species, in aqueous medium, which is well characterized from conductometric titration, thermal analysis and potentiometric titration of the silver content of the solid. The precipitate dissolves in excess Y4− to form a complex, AgY3−. Equilibrium studies at 25°C and ionic strength 0.50 M (NaNO3) have shown from solubility and potentiometric measurements that the formation constant (95% confidence level) β1 = (1.93 ± 0.07) × 105 M−1 and the solubility products are KS0 = [Ag +]4[Y4−] = (9.0 ± 0.4) × 10−18 M5 and KS1 = [Ag +]3[AgY3−] = (1.74 ± 0.08) × 10−12 M4. The presence of Na+, rather than ionic strength, markedly affects the equilibrium; the data at ionic strength 0.10 M are: β1 = (1.19 ± 0.03) × 106 M−1, KS0 = (1.6 ± 0.4) × 10−19 M5 and KS1 = (1.9 ± 0.5) × 10−13 M4; at ionic strength tending to zero; β1 = (1.82 ± 0.05) × 107 M−1, KS0 = (2.6 ± 0.8) × 10−22 M5 and KS1 = (5 ± 1) × 10−15 M4. The intrinsic solubility is 2.03 mM silver (I) in 0.50 M NaNO3. Well-defined potentiometric titration curves can be taken in the range 1–2 mM with the Ag indicator electrode. Thermal analysis revealed from differential scanning calorimetry a sharp exothermic peak at 142°C; thermal gravimetry/differential thermal gravimetry has shown mass loss due to silver formation and a brown residue, a water-soluble polymeric acid (decomposition range 135–157°C), tending to pure silver at 600°C, consistent with the original Ag4Y salt.  相似文献   

9.
In this work, the lowest excited singlet states of acridine (Acr), acridinium (AcrH+) and 10-methylacridinium (AcrMe+) are quenched by sulfur-containing amino acids and carboxylic acids in aqueous solution. Both steady-state and time-resolved fluorescence techniques were used to monitor the quenching of fluorescence. Stern–Volmer plots of the fluorescence intensity showed a static component (KS) to the quenching. The experimental KS values were compared to theoretical KS values for outer-sphere complexes based on Debye–Hückel theory and the Fuoss equation. The general agreement between experimental and theoretical KS values indicate that the static quenching can be attributed to non-fluorescing ion pairs associated as simple outer-sphere complexes. The computed values of the interionic distances of the ion pairs are consistent with the ion pairs of the ZAZQ=−1 and −2 cases being solvent-separated ion pairs while those of the ZAZQ=−3 case are contact ion pairs. The effect of the reactants’ charges on the quenching rate constants (dynamic component) was observed for the reactions of AcrMe+ with the anionic forms of the quenchers (having charges ZQ=−1, −2 and −3). The rate constants (extrapolated to ionic strength, μ=0) for the quenching processes were determined to be 0.3–5.3×1010 M−1 s−1 depending on the ionic charge (ZQ) of the quencher used. These trends in the quenching rate constants are rationalized with a quenching scheme for electron transfer. Analogous quenching rate constants for alanine and glycine were found to be at least an order of magnitude lower. Photoinduced electron transfer from the sulfur atom of the quencher molecule to the acridine excited singlet state is suggested to be the most likely mechanism of the process under discussion.  相似文献   

10.
The equilibrium constant K for the ion-pair formation fac-[Co(pic)3]3+ + C2O22− fac-[Co(pic)3]3+/C2O42−1 where pic = 2-aminomethylpyridine, has been determined spectrophotometrically at 0.35 M (KCl) ionic strength and 25.0°C, using four different calculation approaches. The best results were obtained when the concentration of the minor component (the cobalt complex ion) was not neglected in comparison with the oxalate ion concentration. The value of K (5.3 M−1) increases when the supporting electrolyte is LiCl (K = 8.2 M−1). The effect of the ionic strength variation from 0.35 to 2.0 M (LiCl) was also investigated.  相似文献   

11.
In this work it is reported that the kinetic modelling of the separation of cadmium from phosphoric acid by non-dispersive solvent extraction. Using Cyanex 302 as selective extractant, the extraction step was carried out in a hollow fibre module containing polypropylene fibres, whereas the concentration step required a ceramic module with tubular channels due to the high acidity of the backextraction agent. Application of the methodology previously reported by the authors led to the development of a kinetic model with three design parameters, i.e., equilibrium constant of the extraction reaction (K'e = 6 × 103 mol−2/l−2), membrane mass transport coefficient of the extraction module Kme=8.33×10−8 m/s) and of the backextraction module (Kms=3.33×10−8 m/s), that described satisfactorily the behaviour of the separation-concentration system. Thus, in this work a new application of the non-dispersive solvent extraction technology is presented, characterising at the same time the behaviour and parameters of a new type of contactor, i.e., a tubular ceramic module.  相似文献   

12.
As a part of the European EUROCORE and GRIP (Greenland Ice Core Project) operations aimed at recovering deep ice cores at Summit (Central Greenland), we have for the first time successfully performed ion chromatography measurements in the field and investigated in detail the soluble impurities, including Na+, NH+4, K+, Mg2+, Ca2+, F, CH3COO, CH2 OHCOO, HCOO, CH3SO3, Cl, NO2, SO42− and C2O42−, trapped in ice deposited over some 200 000 years in Greenland.  相似文献   

13.
From extraction experiments and γ-activity measurements, the extraction constant corresponding to the equilibrium H3O+(aq) + 1· Na+(nb)  1·H3O+(nb) + Na+(aq) taking place in the two-phase water–nitrobenzene system (1 = hexaethyl p-tert-butylcalix[6]arene hexaacetate; aq = aqueous phase, nb = nitrobenzene phase) was evaluated as log Kex (H3O+, 1·Na+) = −0.6 ± 0.1. Further, the stability constant of the 1·H3O+ complex in water saturated nitrobenzene was calculated for a temperature of 25 °C as log βnb (1·H3O+) = 6.8 ± 0. 2. By using quantum mechanical DFT calculations, the most probable structure of the 1·H3O+ complex species was predicted. In this complex, the hydroxonium ion H3O+ is bound partly to three carbonyl oxygen atoms by strong hydrogen bonds and partly to three alternate phenoxy oxygens by somewhat weaker hydrogen bonds.  相似文献   

14.
The substitution reactions of 2,3-, 2,4-, 3,4-, or 3,5-dichlorobenzoyl chloride (Cl2C6H3COCl) and 2,3-, 2,4-, 3,4-, or 3,5-dichlorobenzoate ion (Cl2C6H5COO) or benzoate ion (C6H5COO) in a two-phase H2O/CH2Cl2 medium using pyridine 1-oxide (PNO) as an inverse phase transfer catalyst were investigated. The reaction of Cl2C6H3COCl and PNO in CH2Cl2 to produce the ionic intermediate, 1-(dichlorobenzoyloxy)-pyridinium chloride (Cl2C6H3COONP+Cl) is the rate-determining step. In the PNO-catalyzed two-phase reaction of Cl2C6H3COCl and C6H5COONa, the order of reactivities of Cl2C6H3COCl toward reaction with PNO is (2,3-, 2,4-)>3,5->3,4-2,6-Cl2C6H3COCl, whereas it is 3,5->(2,3-, 3,4-)>2,4-Cl2C6H3COCl in the PNO-catalyzed two-phase reaction of Cl2C6H3COCl and the corresponding Cl2C6H3COONa. The order of reactivities of Cl2C6H3COO ions towards the reaction with 1-(benzoyloxy)-pyridinium (C6H5COONP+) ion is (3,4-, 3,5-)>(2,3-, 2,4-Cl2C6H3COO).  相似文献   

15.
Reaction of cis-[Ptph2(SMe2)2] with Me2PCH2PMe2 (dmpm) gave cis-[PtPh2(dmpm-P)2] (1) or cis,cis-[Pt2Ph4(μ-dmpm)2] (2) and reaction of 1 with [Pt2Me4(μ-SMe2)2] gave cis,cis-[Ph2Pt(μ-dmpm)2PtMe2] (3). Reaction of 1 with trans-[PtClR(SMe2)2] gave cis,trans-[Ph2Pt(μ-dmpm)2PtClR], R = Me (5) or Ph (6), and in polar solvents, these isomerized to give [Ph2Pt(μ-dmpm)2PtR]+Cl. When R = Me, further isomerization via the phenyl group transfer gave [PhMePt(μ-dmpm)2PtPh]+Cl. Oxidative addition of methyl iodide occurred reversibly at the cis-[PtMe2P2 unit of 3 to give cis,fac-[Ph2Pt(μ-dmpm)2PtIMe3] but complex 2 failed to react with MeI. A comparison with similar known complexes of Ph2PCH2PPh2 (dppm) is made and differences are attributed primarily to the lower steric hindrance of dmpm.  相似文献   

16.
Ramirez AA  Linares CJ  Barrero FA  Ceba MR 《Talanta》1986,33(12):1021-1025
Mixtures of La(III) and Mg(II) form with purpurin (P; 1,2,4-trihydroxyanthraquinone) the mixed-metal complex LaMg2P5, which is extracted with methyl isobutyl ketone at pH 7.5. The molar absorptivity of the complex is 6.1 × 104 l.mole−1.cm−1 at 570 nm and its conditional extraction constant 5 × 1013 l4.mole−4. Similar complexes have been found to be formed with yttrium, cerium, praseodymium, neodymium and samarium. The use of these complexes for the spectrophotometric determination of yttrium and lanthanides up to 15 μg has been investigated.  相似文献   

17.
Marczenko Z  Lobiński R 《Talanta》1988,35(12):1001-1004
The formation and extraction of the ion-associates of the vanadium(V)-3,5-dinitrocatechol (DNC) anionic chelate complex with various basic dyes have been studied and a new sensitive extraction—spectrophotometric method for the determination of vanadium based on the system V(V)-DNC-Brilliant Green has been developed. Beer's law is obeyed up to a vanadium concentration of 0.3 μg/ml and the molar absorptivity is 1.7 × 1O5 l.mole−1. cm−1 at 630 nm. The molar ratios of the components and the form of the vanadium(V) cation in the extracted compound have been determined, and the formula [VO(OH)(DNC)2−2][BG+]2 is proposed. Titanium, molybdenum, tungsten, EDTA and thiocyanate interfere seriously. The method becomes specific after a preliminary separation of vanadium by its extraction as the BPHA complex from H2SO4-HF medium, and is 40 times more sensitive than the spectrophotometric BPHA method. The proposed method has been applied to determination of traces of vanadium (about 10−5%) in alums.  相似文献   

18.
Asbury GR  Klasmeier J  Hill HH 《Talanta》2000,50(6):738-1298
The analysis of explosives with ion mobility spectrometry (IMS) directly from aqueous solutions was shown for the first time using an electrospray ionization technique. The IMS was operated in the negative mode at 250°C and coupled with a quadrupole mass spectrometer to identify the observed IMS peaks. The IMS response characteristics of trinitrotoluene (TNT), 2,4-dinitrotoluene (2,4-DNT), 2-amino-4,6-dinitrotoluene (2-ADNT), 4-nitrotoluene (4-NT), trinitrobenzene (TNB), cyclo-1,3,5-trimethylene-2,4,6-trinitramine (RDX), cyclo-tetramethylene-tetranitramine (HMX), dinitro-ethyleneglycol (EGDN) and nitroglycerine (NG) were investigated. Several breakdown products, predominantly NO2 and NO3, were observed in the low-mass region. Nevertheless, all compounds with the exception of NG produced at least one ion related to the intact molecule and could therefore be selectively detected. For RDX and HMX the [M+Cl] cluster ion was the main peak and the signal intensities could be greatly enhanced by the addition of small amounts of sodium chloride to the sprayed solutions. The reduced mobility constants (K0) were in good agreement with literature data obtained from experiments where the explosives were introduced into the IMS from the vapor phase. The detection limits were in the range of 15–190 μg l−1 and all calibration curves showed good linearity. A mixture of TNT, RDX and HMX was used to demonstrate the high separation potential of the IMS system. Baseline separation of the three compounds was attained within a total analysis time of 6.4 s.  相似文献   

19.
Excess molar enthalpies HE and excess molar volumes VE have been measured, as a function of mole fraction x1, at 298.15 K and atmospheric pressure for the five liquid mixtures (x11,4-C6H4F2 + x2n-ClH2l+2), l = 7, 8, 10, 12 and 16. In addition, HE and excess molar heat capacities CPE at constant pressure have been determined for the two liquid mixtures (x1C6F6 + x2n-ClH2l+2), l = 7 and 14, at the same temperature and pressure. The instruments used were flow microcalorimeters of the Picker design (the HE version was equipped with separators) and a vibrating-tube densimeter, respectively.

The excess enthalpies of the five difluorobenzene mixtures are all positive and quite large; they increase with increasing chain length l of the n-alkane from HE(x1 = 0.5)/(J mol−1) = 1050 for l = 7 to 1359 for l = 16. The corresponding excess volumes VE are all positive and also increase with increasing l: VE(x1 = 0.5)/(cm3 mol−1) = 0.650 for l = 7 and 1.080 for l = 16. Interestingly, the excess enthalphies of the corresponding mixtures with hexafluorobenzene are only about 5% larger, whereas the excess volumes of (x1C6F6 + x2n-ClH2l+2) are roughly twice as large as those of their counterparts in the series containing 1,4-C6H4F2. Specifically, at 298.15 K HE(x1 = 0.5)/(J mol−1) = 1119 for (x1C6F6 + x2n-C7H16) and 1324 for (x1C6F6 + x2n-C14H30), and for the same mixtures VE(x1 = 0.5)/(cm3 mol−1) = 1.882 and 2.093, respectively. The excess heat capacities for both systems are negative and of about the same magnitude as the excess heat capacities of mixtures of fluorobenzene with the same n-alkanes (Roux et al., 1984): CPE(x1 = 0.5)/(J K−1 mol−1) = −1.18 for (x1C6F6 + x2n-C7H16), and −2.25 for (x1C6F6 + x2n-C14H30). The curve CPE vs. (x1 for x1C6F6 + x2n-C14H30) shows a sort of “hump” for x1 0.5, which is presumed to indicate emerging W-shape composition dependence at lower temperatures.  相似文献   


20.
Using in-situ microbalance and infrared spectroscopy techniques the double ammonia proton complexation was traced. The results confirm the formation of N2H7+ dimer in solid Dawson acid H6P2W18O62, previously reported only for N2H7I and for (N2H7)4SiW12O40. The formation of such dimers was evidenced by the microbalance results, the molar ratio of ammonia to proton was measured as 2:1 at 10.7 kPa and 298 K. The formation of NH4+ monomer (band at 1410 cm−1) and N2H7+ dimer (1460 cm−1) was revealed by IR spectroscopy. Enthalpy of ammonia sorption on Dawson structure was calculated to be −127.9 kJ mol−1, indicating the lower acid strength of Dawson-type compared to that of the Keggin-type heteropolyacids, like H4SiW12O40.  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