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.
Determination of trace levels of palladium(II) is described. The method relies upon the extraction of palladium(II)-biacetylmonoxime 2-pyridylhydrazone (BMPH) from aqueous acidic solution into chloroform to form a purple-reddish complex. The molar absorptivity of the Pd-BMPH complex is about 7500 liters mol?1 cm?1 at 560 nm in the chloroform extract. The highly colored chloroform extract is suitable for spectrophotometric determination. The method devised has been applied to the determination of palladium in PdCaCO3 catalyst with good results.  相似文献   

2.
The reactions (I) Hg2Cl2(s) + Br2(g) and (II) HgCl2(s) + HgBr2(s) have been investigated by an X-ray method. Both the reactions yield two forms of the mixed halide HgClBr, designated as α-HgClBr and β-HgClBr. The cell parameters of the two are as follows:α-HgClBr: a = 6.196 A?, b = 13.12 A?, c = 4.37 A?, z = 4, ? = 5.91 g/cm3. The powder pattern and cell parameters are similar to that of HgCl2. Therefore it is probable that the chlorine atoms, in the linear halogenHghalogen molecules of HgCl2 structure have been replaced by bromines, and since the radius of the bromine atom is larger than that of chlorine, the lattice is larger in this case.β-HgClBr: a = 6.78 A?, b = 13.175 A?, c = 4.17 A?, z = 4, ? = 5.40. These parameters are the same as those reported in the literature for β-Hg(ClBr)2, and its X-ray powder pattern is similar to HgCl2. Therefore this phase also has linear halogenHghalogen molecules but the distribution of Cl and Br atoms is perhaps random.Heating the products (I) and (II) up to the melting point increases the amount of α phase and decreases the β phase, whereas crystallization increases the β phase. DTA study has supported the X-ray findings.  相似文献   

3.
A technique for stripping determination of mercury traces in air employing a glassy carbon electrode is described. The sample is passed at 2 liters min?1 for 2 hr through an absorber containing 0.2 M potassium permanganate and 10% wv sulfuric acid (1:1). After reduction with hydroxylamine hydrochloride, the determination is carried out in 0.12 M potassium thiocyanate at pH 2.0 ± 0.2 in the presence of 0.2 ppm of cupric ions. Calibration curves were found to be linear in the range 20 ppb-1 ppm Hg(II) in the cell. The accuracy of the method was tested over simulated samples and it was found to be better than 95%; the relative standard deviation was 5% or less. The limit of detection of mercury in air was approximately 10 μg m?3.  相似文献   

4.
The product of the reduced inertia Jr of a dipolar molecule and the square of the far infrared absorption maximum wave-number νmax that it displays in dilute solution is shown, using a three-parameter Mori/itinerant oscillator model, to be related to the mean-square action on the molecule and hence by a simple argument to the solute volume of rotation V. Estimates of V have been made for seventeen rigid molecules (mainly substituted phenyls) and their νmax values have been measured in decalin solution at 293 and 110 K. Provided allowance is made for translation—rotation coupling in four cases, the derived relation Irν2max = (8aa2c2kT)?1V2P(0) is found to be approximately obeyed at both temperatures with P(0) (the solute-independent mean-square torque acting on a molecule of unit V) having a value of 4.0 × 1017 (N m?2)2 at both temperatures. It should now be possible to predict μmax for other solutes in decalin if their structures are known.  相似文献   

5.
The Hg(II)/xylenol orange (XO) complex has been studied in the presence of the watersoluble alcohols methanol, ethanol, propanol, iso-propanol, ethylene glycol, and glycerol. Only one binary complex Hg(II)/XO = 1:2 is formed in a HCit-Na2HPO4 pH 6.6–6.8 buffer in a 25% EtOH medium. Hg(II) can be determined (0.16–3.0 ppm) with a molar absorptivity of 3.2 × 104 liter mol−1 cm−1.A reaction mechanism of chelation is suggested according to the experimental results. The proposed model is based on the assumption of a 1:11 intramolecular sigmatropics process allowed by the orbital symmetry, followed by an ion association with the cationic species of the solution. The constants of the reaction have been calculated and the study of the interferences and the role of strong electrolytes carried out.  相似文献   

6.
The reaction between Cu(II) and pyridine-2-aldehyde guanylhydrazone nitrate (PAG) was investigated by using the techniques of spectrophotometry and polarography.A spectrophotometric determination of Cu(II) with PAG can be performed in a concentration range varying from 0.39–5.50 ppm of copper; the molar absorptivity is 1.2 × 104 liters mole?1 cm?1 at 380 nm and the relative error is ±0.6%.The complex gave rise to a single well defined cathodic wave (E12 = ?0.34 V for pH = 8.1). The reduction process is diffusion controlled, involves two electrons, and is irreversible.The stoichiometry 1:1 of the complex was established by different methods and by both techniques. The apparent stability constants were computed as log K = 4.4 ± 0.1 and 4.7 from spectrophotometric and polarographic techniques, respectively.  相似文献   

7.
A new thermodynamic treatment of continuous association is presented, where the various equilibria between i-mers are replaced by a single equilibrium between an OH groups in the bonded and the non-bonded states, linked in both cases to an indefinite ensemble of molecules. The treatment leads to an association constant K which differs from those considered in the theories of Kretschmer and Wiebe and of Wiehe and Bagley. For pure alcohols the association constant can be estimated from the vapor pressure of the alcohol and that of the homomorphous hydrocarbon. The fraction γ of free OH groups determined in this way is markedly smaller than those calculated from the other theories. For the normal alcohols the product KVA is approximately constant at a given temperature, VA being the molar volume. This can be expected from the increasing of the standard entropy of the non-bonded molecules when the molecular volume increases. For secondary and tertiary alcohols the product KVA is significantly lower due to steric hindrances. However for all the alcohols considered here the enthalpy of the hydrogen bond remains nearly constant — δH being equal to 24.8 ± 2 kJ mol?1.  相似文献   

8.
9.
Soluble complex-formation of mercury(II) thiocyanate has been studied oscillometrically. The titration of mercury(II) nitrate with thiocyanate gives one inflection corresponding to the formation of Hg(SCN)2, while in the reverse titration the formation of Hg(SCN)+ is also indecated. The method is useful for a rapid determination of very small quantities of mercury or thiocyanate in highly dilute solutions. The titrations can be effected in presence of nitric acid provided its total acidity in the system does not exceed about 1300N. Further Work on the mercury(II) -halide and mercury(II)-cyanide reactions is in progress.  相似文献   

10.
Mass spectrometric studies of the ions present in H2/O2/N2 flames with potassium and chlorine added have demonstrated that ionization can occur in the forward steps of K + Cl ? K+ + Cl? (II), KCl + M ? K+ + Cl? + M (IV), where M is any third body. Variations of [K+] with time in these systems have been measured and establish that the rate coefficients (in ml molecule?1 s?1) of the ion-producing steps are k2 = 5 × 10?10T?12 exp(?10 500/T) and k4 = 2.2 × 107T?3.5 × exp(?60 800/T). Coefficients for ion-ion recombination have been obtained from k2 and k4 using the equilibrium constants of (II) and (IV) and are k?2 = 1.7 × 10?9T?12 and k?4 = 1.1 × 10?17T?3, with each one in the ml molecule?1 s?1 system of units. Replacement of the N2 in one of these flames with sufficient Ar to maintain the temperature constant leaves the measured k2 and k?2 unchanged, but lowers the observed k4 and k?4. This confirms that ion-recombination in the backward step in (II) is a two-body process, whereas in (IV) it is termolecular.  相似文献   

11.
The complexes [Zn(en)3]X2·n H2O, where en = ethylenediamine, X = Cl?, Br? or 12SO2?4, n = 1 or 0.5, and [Zn(tn)2]X2·n H2O, where tn=1,3-diaminopropane, X=Cl?, Br? or 12SO2?4, n = 0 or 0.25, have been synthesized and their thermal investigations carried out. The complexes were characterized by elemental analysis and IR spectral data. These complexes have been observed to decompose through several isolable as well as non-isolable complex species as intermediates during heating. [Zn(tn)2]SO4 undergoes solid-state phase transition in the temperature range 126–145°C. ZnenSO4 and ZntnX2 (X = Cl?, Br? or 12SO2?4) have been synthesized pyrolytically in the solid state from their corresponding mother diamine complexes. ZnenSO4 and ZntnX2 (X = Cl?, Br? or 12SO2?4) complexes decompose through non-isolable hemidiamine species. ZnX2 (X = Cl? or Br?) complexes of tn undergo melting after formation of the monodiamine species. In contrast, the corresponding en complexes undergo melting at non-stoichiometric composition. Diamine (en or tn) is found to be bridging in all monodiamine (en or tn) complexes; whilst their mother complexes possess chelated en or tn. The thermal stability sequence of en and tn complexes of Zn(II) is ZnCl2 < ZnBr2 < ZnSO4. ΔH values are reported for some steps of decomposition. Possible mechanistic paths have been reported for each step of decomposition.  相似文献   

12.
The magnetic and electric properties of V2O3+x were investigated by measurements of magnetic susceptibility, electrical resistivity, magnetotorque, Mössbauer of doped 57Fe, and NMR of 51V, and the results were compared with those of the (V1?xTix)2O3 system or highly pressured V2O3. The results obtained are as follows: (1) The metallic state shows an antiferromagnetic ordering at TN (x). The value of TN for metallic V2O3, obtained by interpolation to x = 0, shows the coincidence between V2O3+x and the (V1?xTix)2O3 system. (2) Magnetic susceptibility of V2O3+x is expressed as χM(V2O3+x) = (1?x)χM(V3+) + M(V4+). χM(V4+) obeys the Curie-Weiss law M(V4+) = 0.77T + 17). (3) In the insulating phase, the electrical resistivity ? is expressed as a common equation: ? = 10?1.8exp(EkT). This implies that the substitution of Ti or nonstoichiometry (V+4 + metal vacancies) has little influence on the carrier mobility (or bandwidth). (4) There is a critical length in the c-axis (? 14.01 Å) where the metal-insulator transition takes place. This suggests that the length of the c-axis plays an important role in the metal-insulator transition of V2O3-related compounds.  相似文献   

13.
The phase equilibria in the V2O3Ti2O3TiO2 system have been determined at 1473°K by the quench method, using both sealed tubes and controlled gaseous buffers. For the latter, CO2H2 mixtures were used to vary the oxygen fugacity between 10?10.50 and 10?16.73 atm. Under these conditions the equilibrium phases are: a sesquioxide solid solution between V2O3 and Ti2O3 with complete solid solubility and an upper stoichiometry limit of (V, Ti)2O3.02; an M3O5 series which has the V3O5 type structure between V2TiO5 and V0.69Ti2.31O5 and the monoclinic pseudobrookite structure between V0.42Ti2.58O5 and Ti3O5; series of Magneli phases, V2Tin?2O2n?1TinO2n?1, n = 4–8; and reduced rutile phases (V, Ti)O2?x, where the lower limit for x is a function of the V(V + Ti) ratio. The extent of the different solid solution areas and the location of the oxygen isobars have been determined.  相似文献   

14.
Vanadium(II) ions form with the pyridine-2-carboxylate ligand a deep blue, tris-substituted complex absorbing at 660 nm (ε = 7.2 × 103 M?1) cm?1) with a shoulder at 450 nm. Reversible spectroelectrochemistry and cyclic voltammetry were observed for this complex, with E12 = ?0.448 V vs NHE, and ΔSrcθ = ?6 cal · mol?1 · deg?1. Electron transfer kinetics with [CO(en)3]3+ led to k12 = 3100 M?1 s?, ΔH = 12.4 kcal · mol?1 and ΔS = ?0.9 cal · mol?1 · deg?1 (I = 0.10 M). For the related [Co(NH3)6]3+ complex, k13 = 1.9 × 104 M?1 s?1. The self-exchange rate constant and activation parameters were analysed in terms of relative Marcus theory.  相似文献   

15.
The reaction of aluminum EDTA with calmagite was found to be first-order in calmagite, half-order in aluminum EDTA, and zero-order in excess EDTA. The three-halves-order apparent rate constant at pH 9.48 and 20°C with ionic strength 0.30 was 0.232 ± 0.014112 · mol12 · s?1. The activation energy was 3.2 ± 0.3 kcal/mol. The pH dependence was complex between 8 and 10. The mechanism proposed is a rapid dissociation of a dimer of aluminum EDTA to a monomer followed by a rate-determining displacement of the EDTA by calmagite. The reaction of Co(II) EDTA with calmagite was confirmed to be first-order in Co(II) EDTA and first-order in calmagite with an activation energy of 7.5 ± 0.6 kcal/mol.  相似文献   

16.
Negative chemical ionisation mass spectrometry is used as a probe to identify reactions between hydrocarbon radicals and cornplexed cobalt(II) centres in the gas phase. Methane NCI mass spectra of a series of cobalt(II) complexes containing O4, O2N2 and N4 donor atom sets are characterised by adduct ions of the form [M + CnH2n+1]? at m/z values above the molecular ion, [M]?. Formation of such ionic species has been rationalised in terms of a one-electron oxidative-addition mechanism involving attack by hydrocarbon plasma-derived alkyl radicals at the metal centre prior to electron capture: CoIILn + R? → RCoIIILne? [CoILn]?. The competing resonance electron attachment reaction: CoIILne? also occurs within the ion source.  相似文献   

17.
The mechanism of ascorbic acid (DH2) oxidation with molecular oxygen catalysed by the polynuclear complex of Cu2+ with poly-4-vinylpyridine (PVP), partially quaternized by dimethylsulphate, has been studied. The half-conversion time of the reaction of DH2 with Cu(II) PVP under anaerobic conditions is independent of [Cu2+]. At pH 3.5, t0.5 (sec) = 0.8 + 5 × 10?4 [DH2]. The formation of an intermediate cupric-ascorbate complex is suggested (Kc ≈ 104 M?1). Free radicals of ascorbic acid are detected by the ESR-method combined with a flow technique. The small steady-state concentration of radicals indicates that their decay occurs inside the macromolecular complex. The rate constant of the PVP Cu(II) DH? ternary complex dissociation is ≈0.4 sec?1 (pH 3.5). The reaction of Cu(I) PVP with O2 is not accompanied by formation of O2? outside the macromolecule bulk. The rate constant of this reaction is 1.3 ± 0.15) × 102 M?1 sec?1 (pH 3.5). The cyclic mechanism of the catalytic reaction is suggested to include interchange of the redox state of copper-ions. About 23 of the total copper ion exists in the form Cu(I) PVP during the reaction at pH 3.5. The rate of DH2 oxidation under these conditions is limited by the rate of Cu(I) PVP reaction with O2. At pH 4.5 the overall reaction rate is limited by the rate of interaction of Cu(II) PVP with DH?.  相似文献   

18.
Single crystals of PdPSe were shown to be n-type semiconductors. Weak Pauli paramagnetic behavior was observed, which is consistent with the presence of delocalized electrons. Electrical measurements showed a room-temperature resistivity ? = 70 ohm-cm, activation energy of resistivity Ea = 0.32 eV, and Hall mobility μ = 34 cm2 V?1 sec?1. Photoelectronic measurements in aqueous solutions of I?I?3 indicate that PdPSe has high quantum efficiencies below 800 nm. The indirect optical band gap is 1.28(2) eV.  相似文献   

19.
Copper(I), copper(II), and thallium(III) hexafluoromolybdates(V), prepared by the oxidation of the metals in acetonitrile with molybdenum hexafluoride (A. Prescott, D.W.A.Sharp, and J.M. Winfield, J. Chem. Soc., Dalton Trans., 1975, 963) have been investigated by cyclic voltametry. Half wave potentials, E12 V vs. Agp+/Ag were obtained using a evacuable cell equipped with anexternal Agp+/Ag electrode, enabling strict anerobic conditions to be maintained. A number of reversible or quasi-reversible electron transfer processes have been observed, enabling comparison with synthetic work to be made. Results for CuI and CuII hexafluoromolybdates(V) are in accord with preparative experience. MoF6. MoVI/MoVE12 +1.600V, oxidises Cu metal to CuII in MeCN, and CuII is reduced by CuO to CuI , CulI/CuIE12 = +0.750 or +0.710V for CuI and CuII solutes respectively, CuI/CuOE12 = ?0.720V not reversible. A wave at E12 = ?0.350V is assigned to MoV/MoIV by analogy with AgI hexafluoromolybdate (D.W.A. Sharp, unpublished work). E12 data for I2 in MeCN, I2/I3- = 0.280, I3?/I? = -0.116V, suggest that reduction of MoF6? by I is not likely, in contrast to the situation in SO2 (A.J. Edwards and R.D. Peacock, Chem. Ind., 1960, 1441). Reduction of MoF6? by Cuo in MeCN should be feasible, but appears to be very slow. Pure TλIII hexafluoromolybdate(V) is obtained from Tλo and MoF6 only when the mole ratio MoF6:Tλ>5:1. Smaller ratios produce yellow solids in which Mo:Tλ is ca. 2:1. TλIII is a stronger oxidising agent than CuII in MeCN, as oxidation of CuI by TλIII is rapid and quantitative. However a reversible electron transfer wave assignable to TλIII/TλI is not observed in the expected fange +1.600 to +0.710V possibly because of solute-electrode interactions.  相似文献   

20.
Single crystals of a new compound, FeV2O6H0.5, have been obtained by hydrothermal synthesis at 650°C and 2 kbar. An electron microprobe analysis indicated that the chemical formula is FeV2O6. This compound has an orthorhombic symmetry, space group P212121 with Z = 4. The unit cell dimensions are
a = 4.891 A?, b = 9.553 A?, c = 8.786 A?
These parameters are related to a′, b′, c′ of the diaspore-type VO2 by the relations: a ? a′, b ? b′, c ? 3c′. The structure, based on single crystal data, has been determined from Patterson and Fourier syntheses. A structural refinement gave a final R-factor of 2.4%. The new structure can be deduced from that of the diaspore-type VO2, by displacement of one third of the cations from an octahedral site to an adjacent unoccupied tetrahedral site, which is located in a channel parallel to the c axis. The calculation of the cation and oxygen valences indicated that some O2? were in fact OH?. This conjecture was supported by thermogravimetric analysis. The chemical formula was indeed Fe3+2V3+V4+V5+2O11(OH). Some intensities of the powder diffraction lines changed strongly when the preparation temperature was reduced from 650°C to 300°C. This was explained by an increase of the hydrogen ratio in the formula FeV2O6Hx, which implies a structural change. The different phases FeV2O6Hx can be considered as solid solutions between two extreme phases with different structure: one with the present one and the other with a diaspore-type structure.  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