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 483 毫秒
1.
Structures of bromo-metal complexes in concentrated aqueous solutions of FeBr2 and of CoBr2 were investigated by X-ray diffraction analysis. The complexes possess an octahedral geometry coordinating Br along with H2O ligands. The frequency factors of metal-Br contacts per one atom of metal were 0.325 for the 2.7M (mol-dm–3) and 0.747 for the 4.5M FeBr2 solutions, and 0.280 for the 2.8M and 0.595 for the 4.3M CoBr2 solutions. The frequency factors suggested that the tendency of metal ions to forming monobromo complexes is in the order, Fe>Co>Ni相似文献   

2.
A laboratory method is presented for the preparation of ferrous bromide by heating ferric bromide or its crystal hydrate in toluene or xylene at reflux for 3–4 h. The yield is quantitative. The FeBr2 or57FeBr2 sample thereby obtained may be used in the synthesis of iron porphyrin complexes.Institute for Physicotechnical Research, Russian Academy of Sciences, 151034 Moscow. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 8, pp. 1927–1928, August, 1992.  相似文献   

3.
The complex formation of anionic copolymer poly(4-styrenesufonic acid-co-maleic acid) sodium salt (PSSA-co-MA) with surfactant tetradecyltrimethylammonium bromide (TTAB) and surface active ionic liquid (SAIL) 1-tetradecyl-3-methylimidazolium bromide ([C14mim][Br]) has been studied in aqueous media by utilizing various techniques such as tensiometry, conductometry and fluorimetry. Conductometric and tensiometric curve of all the investigated systems demonstrate four break points corresponding to four transition states. All the thermodynamic and surface properties of surfactant-polyelectrolyte systems have been determined by conductance and surface tension measurements respectively. The value of cmc decreases with increasing the concentration of polyelectrolyte for studied systems. But it has been observed that, the lowering in cmc values is more in [C14mim][Br]-PSSA-co-MA system than TTAB-PSSA-co-MA system, although the differences in cmc are not much significant. The lowering in cmc of [C14mim][Br]-PSSA-co-MA system shows that ionic attractions between cationic head group [C14mim]+ and anionic parts (SO3? and COOH), are stronger than those in TTA+ and anionic parts (SO3? and COOH). The results indicated that the [C14mim][Br]-PSSA-co-MA complexes are comparatively more surface active than TTAB-PSSA-co-MA complexes. The fluorescence probe behaviour also confirms cmc value and provides aggregation number (Nagg). Finally all the findings of [C14mim][Br] and TTAB have been compared.  相似文献   

4.
Density measurements were carried out for aqueous solutions of two cationic surfactants: dodecylethyldimethylammonium bromide (C12(EDMAB)) and benzyldimethyldodecylammonium bromide (BDDAB). On the basis of the obtained results of the measurements the CMC and partial molar volumes of the surfactants studied were also determined. The obtained CMC values were also analyzed with those accounted on the basis of the surface tension data from the previous paper [J. Harkot, B. Jańczuk, J. Colloid Interface Sci. (2008), submitted for publication]. The values of CMC determined from the surface tension and density measurements for C12(EDMAB) are equal to 9.9×10−3 and 1.5×10−2 M and for BDDAB to 5.25×10−3 and 5.3×10−3 M, respectively. These obtained values are very similar. However, in the literature it is difficult to find the CMC values for C12(EDMAB) and BDDAB determined by these two methods used by us—especially from the density measurements for BDDAB and surface tension measurements for C12(EDMAB). In the case of the apparent molar volumes of C12(EDMAB) there is a good agreement between the values obtained by us and those found in the literature. The CMC values for C12(EDMAB) and BDDAB were also determined on the basis of their surface tension and free energy of electrostatic interactions between the polar heads of these surfactants and compared with those obtained from the surface tension and density measurements. It was found that the theoretically obtained CMC values were close to those determined from the density and surface tension data for the C12(EDMAB) and that the ratios of the CMC values of the surfactants to their concentration at which the water surface tension decreased by about 20 mN/m proved that the presence of the aryl group in the BDDAB head instead of the methyl group caused that its micellization process was more inhibited in relation to its adsorption at air–water interface than that of C12(EDMAB).  相似文献   

5.
Density functional theory (DFT) at the B3LYP/6‐31+G* and B3LYP/AUG‐cc‐pVDZ levels was employed to study O6‐methylation of guanine due to its reactions with methyl chloride and methyl bromide and to obtain explanation as to why the methyl halides cause genotoxicity and possess mutagenic and carcinogenic properties. Geometries of the various isolated species involved in the reactions, reactant complexes (RCs), and product complexes (PCs) were optimized in gas phase. Transition states connecting the reactant complexes with the product complexes were also optimized in gas phase at the same levels of theory. The reactant complexes, product complexes, and transition states were solvated in aqueous media using the polarizable continuum model (PCM) of the self‐consistent reaction field theory. Zero‐point energy (ZPE) correction to total energy and the corresponding thermal energy correction to enthalpy were made in each case. The reactant complexes of the keto form of guanine with methyl chloride and methyl bromide in water are appreciably more stable than the corresponding complexes involving the enol form of guanine. The nature of binding in the product complexes was found to be of the charge transfer type (O6mG+ · X?, X?Cl, Br). Binding of HCl, HBr, and H2O molecules to the PCs obtained with the keto form of guanine did not alter the positions of the halide anions in the PCs, and the charge transfer character of the PCs was also not modified due to this binding. Further, the complexes obtained due to the binding of HCl, HBr, and H2O molecules to the PCs had greater stability than the isolated PCs. The reaction barriers involved in the formation of PCs were found to be quite high (~50 kcal/mol). Mechanisms of genotoxicity, mutagenesis and carcinogenesis caused by the methyl halides appear to involve charge transfer‐type complex formation. Thus the mechanisms of these processes involving the methyl halides appear to be quite different from those that involve the other strongly carcinogenic methylating agents. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

6.
The 1?:?2 M-ratio reaction between cuprous bromide and pyrazole derivatives in toluene results in mononuclear Cu(I) complexes [CuBr(pyrazole)2]. The complexes have been characterized by 1H NMR spectroscopy and elemental analysis. The molecular structure, established by single-crystal X-ray diffraction, features a trigonal planar geometry around copper, with monodentate pyrazole derivatives. All the Cu(I) complexes are luminescent in the solid state at ambient temperature. Intense blue or blue-green emission in the solid state is observed for these complexes, with the maxima ranging from 431 to 493 nm. The observed photoluminescence could be ascribed to the metal-to-ligand charge-transfer excited states, probably mixed with some halide-to-ligand character. The microsecond lifetime scale of the complexes implies that these transitions arise from the triplet excited states.  相似文献   

7.
The UV-Vis light absorption spectroscopy of bromophenol blue (BPB) in a series of cetyltrimethylammonium bromide (CTAB) aqueous solutions and microemulsions has been determined. There exist association interactions between the BPB and cationic surfactant CTAB. By establishing the appropriate association models and measuring the absorbance in different concentrations of BPB aqueous solutions and microemulsions that have different R values, the association constants K were obtained and the values of thermodynamics functions of association ΔrGm were calculated. It shows that the formation of microemulsion has inhibitory effects on the association reaction.  相似文献   

8.
Abstract

The protonation constants of tetraethylenepentaamineheptaacetic acid, TPHA, were determined by potentiometric titration in aqueous solution at an ionic strength of 0.10 M KNO3 and at 25°C. The formation constants of various metal-TPHA complexes were also obtained by titrating mixtures of metal to ligand in molar ratios of 1 :1 and 2:1. Calculations were performed with the computer program BEST. Individual stability constants are reported for Co(II). Ni(II), Cu(II), Zn(II), Cd(II), Hg(II) and Pb(II) with TPHA as well as their related pro-tonated species. The stabilities of the 1:1 complexes parallel to those of similar complexes with DTPA and TTHA. However the 2: 1 complexes have significantly larger log K ML's than their TTHA counterparts. The extra stability of the 2:1 metal-TPHA complexes is explained in terms of ligand denticity and steric effects. Mercury(II)-TPHA complexes exhibited the highest formation constants and the copper-TPHA complexes had slightly higher log K ML's than those for Co(II), Ni(II), Zn(II), Cd(II) and Pb(II).  相似文献   

9.
The electrical conductances of the solutions of tetrabutylammonium bromide (Bu4NBr), and tetrapentylammonium bromide (Pen4NBr) in 2-ethoxyethanol (1) + water (2) mixed solvent media containing 0.25, 50 and 0.75 mass fractions of 2-ethoxyethanol (w 1) have been reported at 308.15, 313.15, 318.15 and 323.15 K. The conductance data have been analyzed by the 1978 Fuoss conductance–concentration equation in terms of the limiting molar conductance (Λ0), the association constant (K A) and the association diameter (R). These two electrolytes are found to exist essentially as free ions in the solvent mixtures with w 1 = 0.25 and 0.50 over the entire temperature range; however, slight ionic association was observed in the mixed solvent medium richest in 2-ethoxyethanol. The electrostatic ion–solvent interaction is found to be very weak for the tetraalkylammonium ions in the aqueous 2-ethoxyethanol mixtures investigated.  相似文献   

10.
The potential of the cell Zn–Hg(2 phase)|ZnBr2(m)|AgBr|Ag was measured from 0 to 35.9°C and for molalities from 0.125 to 4.0. From these results and values of the standard cell potential, activity coefficients of aqueous zinc bromide were calculated. Good agreement was found with the emf and isopiestic results of Stokes and Stokes, where the ranges of temperature and molality overlap. The activity coefficients of solutions more concentrated that 1.0m are much lower than those of strong 1–2 electrolytes, and this anomaly becomes more pronounced with increasing temperature.  相似文献   

11.
Available data on the solubility of silver chloride in aqueous solutions of HCl, NaCl, KCl, LiCl, and NH4Cl, along with potentiometric measurements of the activity of Ag+ in aqueous NaCl–NaClO4 mixtures, have been analyzed to obtain the thermodynamic properties of the (AgCl)0, AgCl 2 , AgCl 3 2– , and AgCl 4 3– complexes. Results obtained include the stability constants of the complexes and 25°C, the virial parameters needed to calculate the activity coefficients of the complexes and the heats of the reactions by which they are formed. These results are sued to calculate tables of recommended values for the solubility of silver chloride in the host solutions as a function of concentration from 0 to 150°C and to make a critical evaluation of the reliability of previously published data.  相似文献   

12.
A Raman spectral study of 14 solutions of varying bromide to zinc ratios was conducted up to 300°C and 9 MPa. The tetra-, tri-, di- as well as the mono-bromozinc complexes were identified. The signal from the ZnBr+ complex increased in intensity as temperature increased, for solutions of low bromide- to-zinc ratios. The ZnBr 4 2– species was favored at higher Br/Zn ratios, and higher temperatures favored the formation of the species ZnBr2 and ZnBr+ at the expense of ZnBr 4 2– and ZnBr 3 . Although solvated water is probably present in these zinc-bromo complexes, we found no evidence of O–Zn vibrations other than for Zn(H2O) 6 2+ . However, spectra of successive dilutions of solutions with high bromide to zinc ratios show a relative change in species populations thereby suggesting that water activity plays a decisive role in complex formation. For the first time trifluoromethanesulfonic acid (HTFMS) has been used as an internal standard in Raman spectroscopy. This permitted quantitative measurement of stepwise stability constants.  相似文献   

13.
 The self-association of n-hexyltrimethylammonium bromide (C6TAB) in aqueous solution has been studied by static and dynamic light scattering and NMR spectroscopy at 25 °C in the presence of added electrolyte, and critical aggregation concentrations, aggregation numbers and the degree of ionization have been calculated. Aggregation numbers determined from light scattering and from the application of mass-action theory to the concentration dependence of 1H NMR chemical shifts of four protons along the alkyl chain of C6TAB, were between three and four over the range of electrolyte concentration studied (0.2–0.7  molkg−1 NaBr). A structure for the small aggregates has been proposed from the NMR chemical shift data. Received: 4 June 2001 Accepted: 17 September 2001  相似文献   

14.
Complexation of maltol (MH) with Fe3+ ions in aqueous solutions was studied. The compositions of [FeMa]2+, [FeMa2]+, and [FeMa3] complexes were determined by the method of isomolar series, and their instability constants were calculated. The values of the latter were confirmed by the method of apparent deviation from the Bouger—Lambert—Beer law. An increase in the Ma : Fe3+ ratio from 1 to 3 decreases the instability constants of the complexes. The [FeMa3] complex can be considered as a basis for the antianemic drug with a prolonged effect.  相似文献   

15.
Specific conductivity of aqueous solutions of dodecyldimethylethylammonium bromide has been determined in the temperature range of 15-40°C. The critical micelle concentration (cmc) and ionization degree of the micelles, b, were determined from the data. Thermodynamic functions, such as standard Gibbs free energy, ΔG m°, enthalpy, ΔG m°, and entropy, ΔG m°, of micellization, were estimated by assuming that the system conforms to the mass action model. The change in heat capacity upon micellization, ΔG m°, was estimated from the temperature dependence of ΔG m°. An enthalpy-entropy compensation phenomenom for the studied system has been found. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

16.
The equilibrium and rate of solvent extraction of FeBr2 complexes from HBr solutions into benzene solutions of tributyl phosphate (TBP) have been investigated. It is found that two reactions control the iron(III)-TBP extraction from hydrobromic acid solutions. From HBr activity of 1.2–1.55 (molarity based) the reaction is inverse third order with respect to the aqueous phase acidity. From HBr activity of 1.7–6 (molarity based) the reaction is first order in HBr concentration in the aqueous phase. Both of these reactions are first order for both Fe(III) and TBP. The rate constants for these reactions were calculated and the rate-controlling steps are discussed.  相似文献   

17.
Stability constants for aqueous Ag+/Br?, Ag+/SCN?, and mixed Ag+/Br?/SCN? complexes are determined at 25° C by using data generated potentiometrically in solutions having ionic strengths of 0.4, 1.0, and 2.0 m. Monte Carlo numerical methods which yield apparent stability constants for these complexes as well as confidence limits are described in detail. Explicit consideration of speciation shows that under useful precipitation conditions (high bromide and low thiocyanate), a significant fraction of soluble silver is present as AgBrn (SCN)1?n?mm complexes. The most prevalent mixed complexes under these conditions are AgBr (SCN)? (log β11=8.0 ± 0.5) and AgBr2(SCN)2? (log β21=9.2 ± 0.3). The free energies of formation of the other tri- and tetra-coordinate mixed complexes are nearly indistinguishable (log β12=9.3 ± 0.5; log β31=9.0 ± 0.6; log β22=9.6 ± 0.9; log β13=10.3 ±0.5).  相似文献   

18.
The effect of magnesium bromide hexahydrate [MgBr2·6H2O] as a nondurable finish on the flammability of 100% cotton fabric, (woven construction, massing 150 g m−2) has been investigated. The laundered bone-dried, massed fabrics were impregnated with various concentrations of the aqueous above-mentioned salt solutions by means of squeeze rolls and drying in an oven at 110°C for 30 min. The specimens were then cooled in a desiccator, re-massed with an analytical balance and kept under standard conditions before the fulfillment of the vertical flame spread test. After several experiments the optimum add-on values to impart flame-retardancy expressed in g anhydrous magnesium bromide hexahydrate per 100 g fabric were determined to be about 5.6%. The ashes of the treated specimens were subjected to X-ray diffraction analysis (XRD), and the result was compared with data for pure MgO powder and/or MgBr2 specimens. Consequently the existence of MgO was detected in the ashes.  相似文献   

19.
For lithium halides, LiX (X = Cl, Br and I), hydrates with a water content of 1, 2, 3 and 5 moles of water per formula unit are known as phases in aqueous solid–liquid equilibria. The crystal structures of the monohydrates of LiCl and LiBr are known, but no crystal structures have been reported so far for the higher hydrates, apart from LiI·3H2O. In this study, the crystal structures of the di‐ and trihydrates of lithium chloride, lithium bromide and lithium iodide, and the pentahydrates of lithium chloride and lithium bromide have been determined. In each hydrate, the lithium cation is coordinated octahedrally. The dihydrates crystallize in the NaCl·2H2O or NaI·2H2O type structure. Surprisingly, in the tri‐ and pentahydrates of LiCl and LiBr, one water molecule per Li+ ion remains uncoordinated. For LiI·3H2O, the LiClO4·3H2O structure type was confirmed and the H‐atom positions have been fixed. The hydrogen‐bond networks in the various structures are discussed in detail. Contrary to the monohydrates, the structures of the higher hydrates show no disorder.  相似文献   

20.
Apparent molar heat capacities and volumes of amylamine (PentNH2) 0.02m, capronitrile (PentCN) 0.02m and nitropentane (PentNO2) 0.009m in decyltrimethylammonium bromide (DeTAB) micellar solutions, in water and in octane were measured at 25°C. By assuming that their concentration approaches the standard infinite dilution state, heat capacities and volumes were rationalized by means of previously reported equations following which the distribution constant between the aqueous and the micellar phase and heat capacity and volume of the additives in both phases are simultaneously derived. The present results are compared to those we have previously obtained for pentanol (PentOH). The thermodynamic properties of PentNH2 in water and in micellar phase are substantially identical to those of PentOH but different from those of PentCN and PentNO2 whereas the opposite behavior was observed in their pure liquid state and in octane. The nature of the solvent medium seems to affect the thermodynamic behavior of PentNH2. Also, the study of the apparent molar heat capacities of the amyl compounds investigated here in micellar solutions as a function of surfactant concentration shows evidence of a maximum at about 0.4m DeTAB, which can be attributed to a micellar structural transition. Accordingly, the solubilities of PentCN and PentNO2 as a function of the DeTAB concentration drop in the neighborhood of the concentration where heat capacities display the maximum.  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