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 343 毫秒
1.
Interpretation of the results of determinations of free fluoride (Ff) and total fluoride (Ft) obtained with fluoride ISE while conducting elemental chemical analysis of bulk material of newly synthesized inorganic fluoride compounds is of crucial importance for the purpose of determination of purity and stoichiometry of these compounds. Knowledge of the properties and behavior of these compounds in aqueous media is therefore essential. Observations are presented on the determinations of the amounts of Ft and Ff in fluorinated compounds, in the particular hexafluoropnictate salts (PnF6, Pn = P, As, Sb, Bi) as found in aqueous media and in some compounds with XeF2, AsF3 ligands. A critical look at the determined amounts of Ff, Ftand calculated amounts of bound fluoride (Fb) is provided.  相似文献   

2.
The basic copper arsenate mineral strashimirite Cu8(AsO4)4(OH)4·5H2O from two different localities has been studied by Raman spectroscopy and complemented by infrared spectroscopy. Two strashimirite mineral samples were obtained from the Czech (sample A) and Slovak (sample B) Republics. Two Raman bands for sample A are identified at 839 and 856 cm−1 and for sample B at 843 and 891 cm−1 are assigned to the ν1 (AsO43−) symmetric and the ν3 (AsO43−) antisymmetric stretching modes, respectively. The broad band for sample A centred upon 500 cm−1, resolved into component bands at 467, 497, 526 and 554 cm−1 and for sample B at 507 and 560 cm−1 include bands which are attributable to the ν4 (AsO43−) bending mode. In the Raman spectra, two bands (sample A) at 337 and 393 cm−1 and at 343 and 374 cm−1 for sample B are attributed to the ν2 (AsO43−) bending mode. The Raman spectrum of strashimirite sample A shows three resolved bands at 3450, 3488 and 3585 cm−1. The first two bands are attributed to water stretching vibrations whereas the band at 3585 cm−1 to OH stretching vibrations of the hydroxyl units. Two bands (3497 and 3444 cm−1) are observed in the Raman spectrum of B. A comparison is made of the Raman spectrum of strashimirite with the Raman spectra of other selected basic copper arsenates including olivenite, cornwallite, cornubite and clinoclase.  相似文献   

3.
Raman spectroscopy complimented with infrared spectroscopy has been used to characterise the antimonate mineral bindheimite Pb2Sb2O6(O,OH). The mineral is characterised by an intense Raman band at 656 cm−1 assigned to SbO stretching vibrations. Other lower intensity bands at 664, 749 and 814 cm−1 are also assigned to stretching vibrations. This observation suggests the non-equivalence of SbO units in the structure. Low intensity Raman bands at 293, 312 and 328 cm−1 are assigned to the OSbO bending vibrations. Infrared bands at 979, 1008, 1037 and 1058 cm−1 may be assigned to δOH deformation modes of SbOH units. Infrared bands at 1603 and 1640 cm−1 are assigned to water bending vibrations, suggesting that water is involved in the bindheimite structure. Broad infrared bands centred upon 3250 cm−1 supports this concept. Thus the true formula of bindheimite is questioned and probably should be written as Pb2Sb2O6(O,OH,H2O).  相似文献   

4.
Tetra-n-butylammonium halides and, to a lesser extent alkali fluorides, enhance the addition/reduction ratio in the reaction of trialkylaluminium with benzaldehyde in ether, and lower the reactivity of the organoaluminium compounds. These results are consistent with the existence of complexes between salts and organoaluminium compounds [MX•R3 Al (1/1) and MX•2R3 Al (1/2)]. The more stable the complex, the more important are the effects I Br Cl F and NaF KF Bu4 NF. The 1/1 complexes are more stable and less reactive than their corresponding 1/2 complexes.  相似文献   

5.
The surface state of optically pure polydisperse TiO2 (anatase and rutile) was determined by infra-red (IR) spectroscopy analysis in the temperature range of 100–453 K. Anatase A300 spectrum, contrary to rutile R300 one, has a broad three-component absorption band with peaks at 1048, 1137 and 1222 cm−1 in the spectral range of δ(Ti–O–H) deformation vibrations. For rutile R300 we observed a very weak band at 1047 cm−1, and for the thermal treated rutile R900 these bands were not appeared at all. The analysis of temperature dependencies for the mentioned absorption bands revealed the spectral shift of 1222 cm−1 band towards the high frequencies, when the temperature increased, but the spectral parameters of 1137 and 1048 cm−1 bands remained the same. The temperature of 1222 cm−1 band maximum shift was 373–393 K and correlated with DSC data. Obtained results allowed to assign 1222 cm−1 band to the deformation vibrations of OH-groups, bounded to the surface adsorbed water molecules by weak hydrogen bonds (5 kcal/mol). During the temperature growth these molecules desorbed, which also resulted in the intensity decreasing of stretching OH-groups vibration IR-bands at 3420 cm−1. The destruction and desorption of surface water complexes led to Ti–O–H bond strengthening. IR bands at 1137 and 1048 cm−1 were attributed to the stronger bounded adsorbed water molecules, which are also characterized with stretching OH-groups vibration bands at 3200 cm−1. These surface structure were additionally stabilized by hydrogen bonds with the neighbouring TiO2 lattice anions and other OH-groups, and desorbed at higher temperatures.  相似文献   

6.
In an excitation range of 620–760 nm, resonance Raman spectra of aluminum dimers (Al2) in an argon matrix have been obtained for the first time. Temperature annealing experiments were performed to remove Raman lines attributed site effects caused by the Al2/Ar matrix. We observe a single fundamental at 293.3 (5) cm−1 along with a progression up to 1149 (1) cm−1. Taking successive differences of band centers we obtain spectroscopic constants for the ground state fundamental, ωe=297.5 (5) cm−1, the anharmonicity, ωexe=1.68 (8) cm−1. Our results are in close agreement with previous experimental results for Al2 which designate the ground state as a 3Πu state, and may be considered as confirmation of this assignment.  相似文献   

7.
The synthesis of the protected TTF tetrathiolate 2,3,6,7-tetrakis(2-cyanoethylthio)tetrathiafulvalene (TCE-TTF), as well as those of five new radical cation salts [TCE-TTF](X), X = PF6, CF3SO3, BF4, obtained by electrocrystallization technique is presented. Five crystal structures of these materials based on their fully oxidized donor molecules are described. The flexibility of the cyanoethylene arm yields two conformations cis and trans to the molecule. Then compounds with PF6 and BF4 anions crystallized as two different phases. All these materials are insulators, and the magnetic studies of one phase of [TCE-TTF](PF6) revealed an antiferromagnetic behavior.  相似文献   

8.
The kinetics of the reaction of the CH3CHBr, CHBr2 or CDBr2 radicals, R, with HBr have been investigated in a temperature-controlled tubular reactor coupled to a photoionization mass spectrometer. The CH3CHBr (or CHBr2 or CDBr2) radical was produced homogeneously in the reactor by a pulsed 248 nm exciplex laser photolysis of CH3CHBr2 (or CHBr3 or CDBr3). The decay of R was monitored as a function of HBr concentration under pseudo-first-order conditions to determine the rate constants as a function of temperature. The reactions were studied separately from 253 to 344 K (CH3CHBr + HBr) and from 288 to 477 K (CHBr2 + HBr) and in these temperature ranges the rate constants determined were fitted to an Arrhenius expression (error limits stated are 1σ + Student’s t values, units in cm3 molecule−1 s−1, no error limits for the third reaction): k(CH3CHBr + HBr) = (1.7 ± 1.2) × 10−13 exp[+ (5.1 ± 1.9) kJ mol−1/RT], k(CHBr2 + HBr) = (2.5 ± 1.2) × 10−13 exp[−(4.04 ± 1.14) kJ mol−1/RT] and k(CDBr2 + HBr) = 1.6 × 10−13 exp(−2.1 kJ mol−1/RT). The energy barriers of the reverse reactions were taken from the literature. The enthalpy of formation values of the CH3CHBr and CHBr2 radicals and an experimental entropy value at 298 K for the CH3CHBr radical were obtained using a second-law method. The result for the entropy value for the CH3CHBr radical is 305 ± 9 J K−1 mol−1. The results for the enthalpy of formation values at 298 K are (in kJ mol−1): 133.4 ± 3.4 (CH3CHBr) and 199.1 ± 2.7 (CHBr2), and for α-C–H bond dissociation energies of analogous compounds are (in kJ mol−1): 415.0 ± 2.7 (CH3CH2Br) and 412.6 ± 2.7 (CH2Br2), respectively.  相似文献   

9.
The differential pulse polarographic behavior of two series of organotin(IV) compounds having the general formula RxSnCl4 − x (R = Me, Ph; x = 1, 2, 3, 4) has been investigated in DMSO. The peaks obtained are recommended for the trace determination of these compounds. Linear calibration curves are obtained over the concentration range 10−4 –10−6 M.  相似文献   

10.
The oxygen ions of the β-VOPO4 catalyst were exchanged with an tracer by a reduction–oxidation method and by a catalytic oxidation of but-1-ene using 2. The bands at 992 and 900 cm−1 were more shifted to lower frequencies than those at 1076 and 1002 cm−1. Applying the correlation between the Raman bands and stretching vibrations in the literature, the exchanged oxygen species were estimated. The results suggest that the P–O–V vacancies corresponding to 992 and 900 cm−1 were responsible for reoxidation and the V=O oxygen corresponding to the 1002 cm−1 band of β-VOPO4 was not. The (VO)2P2O7 was oxidized to β-VOPO4 by O2 above 823 K. The insertion position of oxygen was determined at the bands at 992 and 900 cm−1 of β-VOPO4 using 2, which is the same as the exchanged position.  相似文献   

11.
Fourier transform infrared spectroscopy is used to study the molecular interaction between gramicidin D and bilayer membranes, dioctadecadimethylammonium bromide (2C18N+2C1Br), and 1,2-di-palmytoyl- -α-phosphatidylcholine (DPPC). Frequencies and bandwidths of the symmetric CH2 stretching band measured as a function of temperature are used to study variation of packing of alkyl chains. The bilayer membrane prepared from 2C18N+2C1Br is found to have a gel to liquid crystal phase transition at 42°C. The presence of gramicidin in the membrane causes an increase in the mobility of the alkyl chain and also a decrease in the abruptness of the transition.The frequencies of amide I and II bands of gramicidin reflecting secondary structures of polypeptides are used to identify its conformation in membranes and to study the interaction between gramicidin and the matrices. Gramicidin is found to have hydrophobic interaction with 2C18N+2C1Br, whereas it has both hydrophobic and hydrophilic interactions with DPPC.  相似文献   

12.
The naphthalene absorption in the spectral region of 3500–4100 Å consists of two perpendicularly polarized band systems. This can be shown by polarization measurements. The arguments given do not allow a decision as to whether or not the short axis polarized band is the 3Ag3Biu+ transition or an induced vibronic band of the 3B3g3Biu+ transition generated by a b3g vibration mode. The T-T polarization of phenanthrene as a function of the excitation wavenumber is quantitatively compared with the equivalent S1 ← S0 fluorescence polarization. There is a good agreement between the calculated and the measured curve.  相似文献   

13.
Layered crystalline zirconium phenylphosphonate, Zr(O3PC6H5)2, changed its interlamellar distance of 1481 pm after intercalation of n-alkylmonoamines, CH3---(CH2)n---NH2 (n=0–6). The infrared spectra of the precursor host and the corresponding intercalated compounds presented vibrations associated with PO3 groups in the 1163–1039 cm−1 range and additional bands related to C---H stretching bands in the 2950–2850 cm−1 interval were observed after amine insertion. The thermogravimetric curves showed a mass loss assigned to the phenyl group; however, the amine intercalated fraction was not quantitatively determined. A peak in the 31P NMR spectrum centered at −6 ppm for the host was observed. The surface area was 42.0±0.2 m2 g−1 and the scanning electron micrograph gave images consistent with lamellar structural features. The layered compound was calorimetrically titrated with amine in ethanol, requiring three independent operations: (i) titration of matrix with amine, (ii) matrix salvation, and (iii) dilution of the amine solution. From those thermal effects the variation in enthalpy was calculated as: −41±1.00,−33.28±0.50,−34.40±0.80,−10.40±0.40,−12.40±0.42,−16.10±0.08 and −7.0±0.04 kJ mol−1, for n=0–6, respectively. The exothermic enthalpic values reflected a favorable energetic process of amine–host intercalation in ethanol. The negative Gibbs free energy results supported the spontaneity of all these intercalation reactions. The positive favorable entropic values, as carbon chain size increased, are in agreement with the free solvent molecules in the medium, as the amines are progressively bonded to the crystalline lamellar inorganic matrix at the solid/liquid interface.  相似文献   

14.
13,14-Benzo-5,8-dihydro-2,4,9,11-tetramethyl-1,5,8,12-tetraazacyclotetradeca-1,3,9,11-tetraene and 13,14-benzo-3,10-di(p-[X]benzoyl)-5,8-dihydro-2,4,9,11-tetramethyl-1,5,8,12-tetraazacyclotetradeca-1,3,9,11-tetraene, wherein X = CH3, H, Cl, NO2 and OCH3, were synthesized and characterized. IR spectra of the benzoylated compounds showed an intense band assigned to the stretching modes of C==O in the region 1632 ~ 1638 cm−1. Hammett plot of the energies of π → π* for the compounds was linear with a positive slope (+0.923). 1H NMR signals exhibited a deshielding effect due to benzoyl groups, while methyl protons only exhibited the shielding effect by the magnetic anisotropy of the group. The substituent effect on 13C NMR was similar to that on proton NMR. The cyclic voltammograms of the compounds mostly showed two one-electron irreversible oxidation peaks and one or two reduction waves between −0.280 and −2.500 V depending on the substituents. Hammett plots of 1st and 2nd oxidation potentials (Eop(1) and Eop(2)) were linear with slopes of +0.036 and +0.045, respectively. The structure of the compound 4 (monoclinic, C2/c, a = 26.0059 (19), b = 6.9058 (5), c = 33.082 (2), β = 109.5930°, Z = 8) was determined using X-ray diffraction method.  相似文献   

15.
Some acyl-thiourea derivatives containing isatin group were synthesized and their interactions with anions were investigated using UV–vis spectroscopy and 1H NMR titrations in DMSO and DMSO-d6, respectively. These compounds have a same molecular framework, functionalising with different groups lead to different anion binding strength of these receptors. Receptor 1 showed a higher binding affinity for AcO than for F, due to the cooperative multiple hydrogen bond interactions of AcO with the acyl-thiourea group and N–H group in the indole unit of receptor 1. Displacing the N–H proton in the indole unit with –CH3 group, receptor 2 showed no obviously discriminative responses for F, AcO and H2PO4 due to lack of such additional binding. In the case of receptor 3, which was functionalised with strong electron-withdrawing group, it showed selectively chromogenic response for F based on double deprotonation of the receptor in DMSO, whereas AcO and H2PO4 induced single deprotonation only.  相似文献   

16.
This paper reports the results of a variety of experiments carried out for understanding the solvation behavior of potassium thiocyanate in methanol–water mixtures. Electrical conductivity, speed of sound, viscosity, and FT-Raman spectra of potassium thiocyanate solutions in 5 and 10% methanol–water (w/w) mixtures were measured as functions of concentration and temperature. The conductivity and structural relaxation time suggest the ion–solvent and solvent-separated ion–ion associations increase as the salt concentration increases in the mixtures. The Raman band shifts due to the C–O stretching mode of methanol for the solvent mixtures reveal the formation of methanol–water complexes. The significant changes in the Raman bands for the C–N, C–S and O–H stretching modes indicate the presence of SCN−solvent interactions through the N-end, “free” SCN and the solvent-shared ion pairs as potassium thiocyanate is added to the methanol–water mixtures. The relative changes corresponding to H–O–H bending and C–O stretching frequencies indicate that K+ is preferentially solvated by water in these solvent mixtures. The appearance and increase of the intensity of a broad band at ≈940 cm−1 upon salt addition was attributed to the SCN–H2O–K+ solvent-shared ion pairs. No Raman spectral evidence for K+(H2O)n species was observed. The preferential solvation of K+ and SCN in the methanol−water mixtures was verified by the application of the Kirkwood−Buff theory of solutions. This theory confirms that K+ is strongly preferentially solvated by water, whereas SCN is preferentially solvated by the methanol component.  相似文献   

17.
The rate constants and product ion branching ratios were measured for the reactions of various small negative ions with O2(X 3Σg) and O2(a 1Δg) in a selected ion flow tube (SIFT). Only NH2 and CH3O were found to react with O2(X) and both reactions were slow. CH3O reacted by hydride transfer, both with and without electron detachment. NH2 formed both OH, as observed previously, and O2, the latter via endothermic charge transfer. A temperature study revealed a negative temperature dependence for the former channel and Arrhenius behavior for the endothermic channel, resulting in an overall rate constant with a minimum at 500 K. SF6, SF4, SO3 and CO3 were found to react with O2(a 1Δg) with rate constants less than 10−11 cm3 s−1. NH2 reacted rapidly with O2(a 1Δg) by charge transfer. The reactions of HO2 and SO2 proceeded moderately with competition between Penning detachment and charge transfer. SO2 produced a SO4 cluster product in 2% of reactions and HO2 produced O3 in 13% of the reactions. CH3O proceeded essentially at the collision rate by hydride transfer, again both with and without electron detachment. These results show that charge transfer to O2(a 1Δg) occurs readily if the there are no restrictions on the ion beyond the reaction thermodynamics. The SO2 and HO2 reactions with O2(a) are the only known reactions involving Penning detachment besides the reaction with O2 studied previously [R.S. Berry, Phys. Chem. Chem. Phys., 7 (2005) 289–290].  相似文献   

18.
EPR studies are carried out on Cr3+ ions doped in d-gluconic acid monohydrate (C6H12O7·H2O) single crystals at 77 K. From the observed EPR spectra, the spin Hamiltonian parameters g, |D| and |E| are measured to be 1.9919, 349 (×10−4) cm−1 and 113 (×10−4) cm−1, respectively. The optical absorption of the crystal is also studied at room temperature. From the observed band positions, the cubic crystal field splitting parameter Dq (2052 cm−1) and the Racah interelectronic repulsion parameter B (653 cm−1) are evaluated. From the correlation of EPR and optical data the nature of bonding of Cr3+ ion with its ligands is discussed.  相似文献   

19.
A detailed study of the electrochemical reduction of diacetylbenzene A in aqueous medium between Ho = −5 and pH 14 is presented. The reactants are strongly adsorbed, so that the reactions are of a surface nature. From Ho = −5 to pH 6, a global 2e reduction yielding an enediol-type intermediate occurs. Analysis using the theory of the square schemes with protonations at equilibrium shows that, up to pH 4, the reaction is controlled by the first electron uptake, the paths being successively H+e and eH+. The elementary electrochemical surface rate constants are 9.6 × 107 s and 1.2 × 106 s for AH+ and A respectively. From pH 6 to 14, a le adsorption wave, corresponding to the formation of (a) monoradical(s), appears and is followed by a le wave due to the reduction of the radical(s). A dimerization occurs, due to the coupling A + AH, as in the case of the monocarbonyl compounds. The rate of this surface process, kd = 5 × 1013 cm2 mol−1 s−1, is markedly smaller than the rate of the homogeneous reaction obtained in alkaline ethanol by Savéant et al. for the coupling of the radicals of benzaldehyde, benzophenone and acetophenone.  相似文献   

20.
The quaternary aluminium hydrides SrAlGeH and BaAlGeH were synthesized from either hydrogenating the intermetallic AlB2-type precursors SrAlGe and BaAlGe or reacting SrH2 with a mixture of Al and Ge in the presence of pressurized hydrogen. Their structures were characterized by X-ray and neutron powder diffraction of the corresponding deuterides. The compounds crystallize with the trigonal SrAlSiH structure type (space group P3m1, Z = 1, a = 4.2435(2) and 4.3450(2) Å, c = 4.9710(3) and 5.2130(4) Å for SrAlGeH and BaAlGeH, respectively) and feature a two-dimensional polyanion [AlGeH]2− which represents a corrugated hexagon layer built from three-bonded Al and Ge atoms. H is terminally attached to Al. Polyanions [AlGeH]2− are electron precise and, according to electronic structure calculations, the quaternary hydrides display band gaps with sizes between 0.7 and 0.8 eV. Infrared and inelastic neutron scattering spectroscopy show Al–H stretching and bending mode frequencies at around 1250 and 870 cm−1, respectively. SrAlGeH and BaAlGeH are thermally stable up to at least 500 °C. When exposed to air the hydrides decompose rapidly to amorphous, orange colored materials.  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