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 959 毫秒
1.
Hydrated layered crystalline barium phenylarsonate, Ba(HO3AsC6H5)2·2H2O was used as host for intercalation of n-alkylmonoamine molecules CH3(CH2)n-NH2 (n = 1-4) in aqueous solution. The amount intercalated (nf) was followed batchwise at 298 ± 1 K and the variation of the original interlayer distance (d) for hydrated barium phenylarsonate (1245 ppm) was followed by X-ray powder diffraction. Linear correlations were obtained for both d and nf as a function of the number of carbon atoms in the aliphatic chain (nc): d = (2225 ± 32) + (111 ± 11)nc and nf = (2.28 ± 0.15) − (11.50 ± 0.03)nc. The exothermic enthalpies of intercalation increased with nc, which was derived from the monomolecular amine layer arrangements with the longitudinal axis inclined by 60° to the inorganic sheets. The intercalation was followed by titration with amine at the solid/liquid interface and gave the enthalpy/number of carbons correlation: ΔH = −(7.25 ± 0.40) − (1.67 ± 0.10)nc. The negative Gibbs free energies and positive entropic values reflect the favorable host/guest intercalation processes for this system.  相似文献   

2.
Lattice energies and thermochemical radii of the anions OR (R = 2-Me; 2,6-Me2; 2,4,6-Me3; 2,6-t-Bu2) in alkali metal phenoxides, MOR (M = Li, Na, K, Rb and Cs) were investigated based on the corresponding standard molar enthalpies of formation determined by reaction-solution calorimetry. The results obtained at 298.15 K were as follows: (MOR, cr)/kJ mol−1 = −398.4 ± 1.1 (LiO-2-MePh), −423.4 ± 1.6 (LiO-2,6-Me2Ph), −457.3 ± 7.1 (LiO-2,4,6-Me3Ph), −346.6 ± 1.4 (NaO-2-MePh), −399.1 ± 1.5 (NaO-2,6-Me2Ph), −422.4 ± 7.1 (NaO-2,4,6-Me3Ph), −496.6 ± 7.1(NaO-2,6-t-Bu2Ph), −367.8 ± 1.2 (KO-2-MePh), −399.1 ± 1.4 (KO-2,6-Me2Ph), −368.8 ± 1.2 (RbO-2-MePh), −403.6 ± 1.3 (RbO-2,6-Me2Ph), −387.0 ± 1.6 (CsO-2-MePh) and −413.6 ± 1.3 (CsO-2,6-Me2Ph). Estimates of thermochemical raddi, lattice energies and standard enthalpies of formation of not experimentally measured alkali metal phenoxides was successfully done with a model based on the Kapustinskii equation and the set of values experimentally determined.  相似文献   

3.
A large, covalent macrocycle that can be served as an artificial allosteric model was prepared in a reasonable yield (36%) through the template-directed synthesis. The macrocycle contains two topologically discrete subcavities, each of which consists of four amide NHs of pyridine-2,6-dicarboxamide units. The macrocycle strongly binds two molecules of N,N,N′,N′-tetramethylterephthalamide in positive cooperative manner by hydrogen-bonding interactions. The association constants were calculated to be K1 = 1480 ± 90 and K2 = 5580 ± 150 M−1 with the Hill coefficient (h) of 1.6 at 25 °C in CDCl3.  相似文献   

4.
In this work, the variations of the relaxation times are investigated above and below the glass transition temperature of a model amorphous polymer, the polycarbonate. Three different techniques (calorimetric, dielectric and thermostimulated currents) are used to achieve this goal. The relaxation time at the glass transition temperature was determined at the temperature dependence convergence of the relaxation times calculated with dynamic dielectric spectroscopy (DDS) for the liquid state and thermostimulated depolarisation currents (TSDC) for the vitreous state. We find a value of τ(Tg) = 110 s for PC samples. The knowledge of the temperature dependence, τ(T), and the value τ(Tg) enables to determine the glass-forming liquid fragility index, m. We find m = 178 ± 5.  相似文献   

5.
In the search for gallium bioactive compounds five Ga(III) complexes, [GaIII(L-H)2](NO3), with tridentate salicylaldehyde semicarbazone derivatives as ligands (L) have been synthesized and characterized in the solid state and in solution by different techniques. The crystal structure of [GaIII(L4-H)2](NO3)·2H2O, where L4 is 3-ethoxysalicylaldehyde semicarbazone, was solved by X-ray diffraction methods. The gallium(III) ion is in a distorted octahedral environment, coordinated to two nearly planar and mutually perpendicular 3-ethoxysalicylaldehyde semicarbazonato anions acting as tridentate ligands through their phenol and carbonyl oxygen atoms and their azomethine nitrogen atom. Their biological potential has been explored by evaluating their activity on Mycobacterium tuberculosis, causative agent of tuberculosis, and their cytotoxicity on tumor cell lines. Three different human tumor cell lines were selected that show different degrees of resistance to metallodrugs: ovarian A2780 (low resistance), breast MCF7 (medium resistance) and prostate PC3 (high resistance) cells. Although the complexes have not shown activity on M. tuberculosis, complexation with gallium has led to the enhancement of the cytotoxic potencies of the organic compounds. Those complexes that contain a bromide substituent at the phenolate ring have shown the highest cytotoxicities. In particular, [GaIII(L2-H)2](NO3), where L2 is 5-bromosalicylaldehyde semicarbazone,·has shown a remarkable cytotoxicity on A2780 tumor cell line with an IC50 value of the same order than cisplatin (IC50 Ga-L2 = 2.4 ± 0.3 μM; IC50 cisplatin = 2.0 ± 0.1 μM, 72 h incubation at 37 °C). Interestingly, this complex has also shown moderate cytotoxicity against MCF7 and PC3 cells (IC50 MCF7 = 30 ± 6; IC50 PC3 = 18 ± 3 μM). Therefore, this gallium compound could be considered a promising wide spectrum potential anti-tumor agent.  相似文献   

6.
Olga P. Kryatova 《Tetrahedron》2004,60(21):4579-4588
Three complexes of benzo-15-crown-5 (B15C5) with protonated primary amines [PhCH2NH3(B15C5)](ClO4), [p-C6H4(CH2NH3)2(B15C5)2](ClO4)2, and [(CH2)4(NH3)2(B15C5)2](SCN)2 were isolated and studied in acetonitrile solutions by NMR, and in the solid state by X-ray crystallography. In all complexes, one B15C5 molecule was bound with each R-NH3+ moiety with characteristic small separation of 1.84-1.86 Å between the nitrogen of the R-NH3+ group and the O5 mean plane of the crown residue. No sandwich-type complexes with a 1:2 R-NH3+/B15C5 stoichiometry were observed. Binding affinities of B15C5 in acetonitrile were similar for all ammonium cations studied: K1=550±10 M−1 for [PhCH2NH3]+; K1=1100±100 and K2=400±30 M−1 for [p-C6H4(CH2NH3)2]2+; and K1=1100±100 and K2=300±30 M−1 for [H3N(CH2)4NH3]2+. The complexation is primarily enthalpy-driven (ΔH°=−4.9±0.5 kcal/mol, ΔS°=−3.8±1.0 eu for PhCH2NH3+-B15C5), as determined by variable temperature 1H NMR titrations.  相似文献   

7.
The hydrogen peroxide-oxidation of o-phenylenediamine (OPD) catalyzed by horseradish peroxidase (HRP) at 37 °C in 50 mM phosphate buffer (pH 7.0) was studied by calorimetry. The apparent molar reaction enthalpy with respect to OPD and hydrogen peroxide were −447 ± 8 kJ mol−1 and −298 ± 9 kJ mol−1, respectively. Oxidation of OPD by H2O2 catalyzed by HRP (1.25 nM) at pH 7.0 and 37 °C follows a ping-pong mechanism. The maximum rate Vmax (0.91 ± 0.05 μM s−1), Michaelis constant for OPD Km,S (51 ± 3 μM), Michaelis constant for hydrogen peroxide Km,H2O2 (136 ± 8 μM), the catalytic constant kcat (364 ± 18 s−1) and the second-order rate constants k+1 = (2.7 ± 0.3) × 106 M−1 s−1 and k+5 = (7.1 ± 0.8) × 106 M−1 s−1 were obtained by the initial rate method.  相似文献   

8.
Phase relations in the ternary system Ce-Pt-Si have been established for the isothermal section at 800 °C based on X-ray powder diffraction, metallography, scanning electron microscopy (SEM) and electron probe microanalysis (EPMA) techniques on about 120 alloys, which were prepared by various methods employing arc-melting under argon or powder reaction sintering. Nineteen ternary compounds were observed. Atom order in the crystal structures of τ18-Ce5(Pt,Si)4 (Pnma; a=0.77223(3) nm, b=1.53279(8) nm c=0.80054(5) nm), τ3-Ce2Pt7Si4 (Pnma; a=1.96335(8) nm, b=0.40361(4) nm, c=1.12240(6) nm) and τ10-CePtSi2 (Cmcm; a=0.42943(2) nm, b=1.67357(5) nm, c=0.42372(2) nm) was determined by direct methods from X-ray single-crystal CCD data and found to be isotypic with the Sm5Ge4-type, the Ce2Pt7Ge4-type and the CeNiSi2-type, respectively. Rietveld refinements established the atom arrangement in the structures of Pt3Si (Pt3Ge-type, C2/m, a=0.7724(2) nm, b=0.7767(2) nm, c=0.5390(2) nm, β=133.86(2)°), τ16-Ce3Pt5Si (Ce3Pd5Si-type, Imma, a=0.74025(8) nm, b=1.2951(2) nm, c=0.7508(1) nm) and τ17-Ce3PtSi3 (Ba3Al2Ge2-type, Immm, a=0.41065(5) nm, b=0.43221(5) nm, c=1.8375(3) nm). Phase equilibria in Ce-Pt-Si are characterised by the absence of cerium solubility in platinum silicides. Cerium silicides and cerium platinides, however, dissolve significant amounts of the third component, whereby random substitution of the almost equally sized atom species platinum and silicon is reflected in extended homogeneous regions at constant Ce content such as for τ13-Ce(PtxSi1−x)2, τ6-Ce2Pt3+xSi5−x or τ7-CePt2−xSi2+x.  相似文献   

9.
Long-range electron transfer (ET) matrix elements (VPS), rate constants (kET) and reorganization energies for ET from phthalimide radical (pha) moiety to methyl aminoacetate radical (aa) moiety in pa–(gly)n = 0–6–aa (pa = C6H4(CO)2N–(CH2CO), gly = glycine, aa = HNCH2COOCH3) ionic molecules have been investigated using two-state variational method (TSVM) and classical rate model. Calculations on VPS reveal that the overlap between the frontier orbitals of two diabatic states is quite small, which leads to a small value of VPS. kET has a minimum at the range n = 1–3 for β-strand conformation, but linearly increases as the peptide chain length (n) increases for pro II-helix conformation. These results are in good agreement with the experimental predictions. Relevant ET mechanisms are elucidated. The transition energies for charge transfer in such systems are also calculated to test the influences of local dipoles on the potentials of the donor and acceptor. For comparison electron couplings in [pa–(gly)n = 1,3–aa]+ cations are calculated and the effects of electron correlation on inner reorganization energies in pha + pha−/+ self-exchange reactions are examined at different levels of theory respectively. Calculated results are discussed also.  相似文献   

10.
The crystal structures of two new Sc(III) porphyrins, [Sc(TPP)Cl]·2.5(1-chloronaphthalene), (5,10,15,20-tetraphenylporphyrin)-chloro-scandium(III)·2.5(1-chloronaphthalene) solvate, (Mo Kα, 0.71073 Å, triclinic system  = 9.9530(2) Å, b = 15.4040(3) Å, c = 17.7770(3) Å, α = 86.5190(10)°, β = 89.7680(10)°, γ = 86.9720(10)°, 13101 independent reflections, R1 = 0.0712) and the dimeric [μ2-(OH)2(Sc(TPP))2], bis-(μ-hydroxo)-(5,10,15,20-tetraphenylporphyrin) scandium(III) (Mo Kα, 0.71073 Å, monoclinic system C2, a = 24.2555(16) Å, b = 11.1598(7) Å, c = 25.6468(17) Å, β = 91.980(2)°, 13084 independent reflections, R1 = 0.0485) are reported. In [Sc(TPP)Cl] the metal is five-coordinate and the porphyrin is domed with the metal displaced by 0.63 Å from the mean porphyrin towards the axial Cl ligand. The average Sc-N bond length is 2.143(3) Å, which is shorter than the average bond length of previously reported structures. Two of the phenyl rings are nearly orthogonal to the porphyrin core and the other two are significantly tilted because of contacts with 1-chloronaphthalene solvent molecules, and the phenyl rings of neighbouring porphyrins. In [μ2-(OH)2(Sc(TPP))2] both porphyrins are domed, with the metal displaced from the mean porphyrin plane towards the bridging hydroxo ligands. The average Sc-N bond length is 2.197(12) Å, which is in the upper range of Sc-N bond lengths in known Sc(III) porphyrins but not dissimilar to the average Sc-N bond lengths in another other bis-μ2-hydroxo Sc(III) porphyrin, [μ2-(OH)2(Sc(OEP))2]. One porphyrin is rotated relative to the upper porphyrin by 25° due to steric contacts between the phenyl substituents. We have used these new structures to re-evaluated our previously reported molecular mechanics force field parameters for modelling Sc(III) porphyrins using the MM2 force field; the training set was augmented from two to seven structures by using all available Sc(III) porphyrin structures and the two new structures. The modelling reproduces the porphyrin core very accurately; bond lengths are reproduced to within 0.01 Å, bond angles to within 0.5° and torsional angles to within 2°. The optimum parameters for modelling the Sc(III)-N bond lengths, determined by finding the minimum difference between the crystallographic and modelling mean bond lengths with the aid of artificial neural network architectures, were found to be 0.90 ± 0.03 mdyn Å−1 for the bond force constant and2.005 ± 0.005 Å for the strain-free bond length. Modelling the seven Sc(III) porphyrins with the new parameters gives an average Sc-N bond length of 2.182 ± 0.018 Å, indistinguishable from the crystallographic mean of 2.181 ± 0.024 Å.  相似文献   

11.
Copolymerization of an excess of methyl methacrylate (MMA) relative to 2-hydroxyethyl methacrylate (HEMA) was carried out in toluene at 80 °C according to both conventional and controlled Ni-mediated radical polymerizations. Reactivity ratios were derived from the copolymerization kinetics using the Jaacks method for MMA and integrated conversion equation for HEMA (rMMA = 0.62 ± 0.04; rHEMA = 2.03 ± 0.74). Poly(ethylene glycol) α-methyl ether, ω-methacrylate (PEGMA, Mn = 475 g mol−1) was substituted for HEMA in the copolymerization experiments and reactivity ratios were also determined (rMMA = 0.75 ± 0.07; rPEGMA ∼ 1.33). Both the functionalized comonomers were consumed more rapidly than MMA indicating the preferred formation of heterogeneous bottle-brush copolymer structures with bristles constituted by the hydrophilic (macro)monomers. Reactivity ratios for nickel-mediated living radical polymerization were comparable with those obtained by conventional free radical copolymerization. Interactions between functional monomers and the catalyst (NiBr2(PPh3)2) were observed by 1H NMR spectroscopy.  相似文献   

12.
From literature data presently available, the decomposition temperature and the nature of the decomposition reaction of the ternary compound α-AlFeSi (also designated as αH or τ5) are not clearly identified. Moreover, some uncertainties remain concerning its crystal structure. The crystallographic structure and thermochemical behaviour of the ternary compound α-AlFeSi were meticulously studied. The crystal structure of α-AlFeSi was examined at room temperature from X-ray single crystal intensity data. It presents hexagonal symmetry, space group P63/mmc with unit cell parameters (293 K) a=12.345(2) Å and c=26.210(3) Å (V=3459 Å3). The average chemical formula obtained from refinement is Al7.1Fe2Si. From isothermal reaction-diffusion experiments and Differential Thermal Analysis, the title compound decomposes peritectically upon heating into θ-Fe4Al13(Si), γ-Al3FeSi and a ternary Al-rich liquid. Under atmospheric pressure, the temperature of this reversible transformation has been determined to be 772±12 °C.  相似文献   

13.
Organic-inorganic hybrid compounds Ni(II)5(OH)6(C6H8O4)2(1), Ni(II)5(OH)6(C8H12O4)2(2) and Co(II)5(OH)6(C8H12O4)2(3) have a similar layered structure as determined ab initio from synchrotron powder diffraction analysis. The metal sites are octahedrally coordinated by O atoms. The slabs are built from edge-sharing octahedra in such a way that channels with an average size of 4 Å are formed. Bis-bidentate and bridging dicarboxylate anions lead to a 3D framework. The compounds (1) and (2) order antiferromagnetically below TN=26.5 and 19.3 K, respectively, while (3) is ferrimagnetic with TC=16.2 K. Crystal data for compounds are as follows: (1)a=11.6504(1) Å, b=6.8021(3) Å, c=6.3603(1) Å, α=73.52(1)°, β=99.69(1)°, γ=96.16(1)°, RB=0.070, 668 reflections; (2)a=13.9325(1) Å, b=6.7893(1) Å, c=6.3534(4) Å, α=73.63(1)°, β=95.14(1)°, γ=91.80(1)°, RB=0.052, 804 reflections; (3)a=13.9806(1) Å, b=6.9588(1) Å, c=6.3967(1) Å, α=73.05(1)°, β=94.51(1)°, γ=92.19(1)°, RB=0.048, 410 reflections. The space group is P−1 for the three compounds.  相似文献   

14.
A series of titanocene(III) alkoxides L2Ti(III)OR where L = Cp, R = Et(1b), tBu(1a), 2,6-Me2C6H3(1c), 2,6-tBu2-4-Me-C6H2(1d), or L = Cp*, R = Me(2e), tBu(2a), Ph(2f) was synthesized and subjected to reaction with [CpM(CO)3]2 [M = Mo, W], [CpRu(CO)2]2, and Co2(CO)8. The Ti(III) precursors 1a, 1c, 2a, 2e, and 2f reacted with [CpM(CO)3]2 [M = Mo, W] to form heterobimetallic complexes L2Ti(OR)(μ-OC)(CO)2MCp [M = Mo, W], of which Ti and M are linked by an isocarbonyl bridge. Reactions of these Ti(III) complexes with Co2(CO)8 resulted in formation of Ti-Co1 heterobimetallic complexes, from 2a, 2e, or 2f, or Ti-Co3 tetrametallic complexes, Cp2Ti(OtBu)(μ-OC)Co3(CO)9 from 1a, 1b, or 1c. The products were characterized by NMR, IR, and X-ray crystallography. Reaction mechanisms were proposed from these results, in particular, from steric/electronic effects of titanium alkoxides.  相似文献   

15.
Subsolidus phase relationships in the In2O3-WO3 system at 800-1400°C were investigated using X-ray diffraction. Two binary-oxide phases—In6WO12 and In2(WO4)3—were found to be stable over the range 800-1200°C. Heating the binary-oxide phases above 1200°C resulted in the preferential volatilization of WO3. Rietveld refinement was performed on three structures using X-ray diffraction data from nominally phase-pure In6WO12 at room temperature and from nominally phase-pure In2(WO4)3 at 225°C and 310°C. The indium-rich phase, In6WO12, is rhombohedral, space group (rhombohedral), with Z=1, a=6.22390(4) Å, α=99.0338(2)° [hexagonal axes: aH=9.48298(6) Å, c=8.94276(6) Å, aH/c=0.9430(9)]. In6WO12 can be viewed as an anion-deficient fluorite structure in which 1/7 of the fluorite anion sites are vacant. Indium tungstate, In2(WO4)3, undergoes a monoclinic-orthorhombic transition around 250°C. The high-temperature polymorph is orthorhombic, space group Pnca, with a=9.7126(5) Å, b=13.3824(7) Å, c=9.6141(5) Å, and Z=4. The low-temperature polymorph is monoclinic, space group P21/a, with a=16.406(2) Å, b=9.9663(1) Å, c=19.099(2) Å, β=125.411(2)°, and Z=8. The structures of the two In2(WO4)3 polymorphs are similar, consisting of a network of corner sharing InO6 octahedra and WO4 tetrahedra.  相似文献   

16.
Tren amine cations [(C2H4NH3)3N]3+ and zirconate or tantalate anions adopt a ternary symmetry in two hydrates, [H3tren]2·(ZrF7)2·9H2O and [H3tren]6·(ZrF7)2·(TaOF6)4·3H2O, which crystallise in R32 space group with aH = 8.871 (2) Å, cH = 38.16 (1) Å and aH = 8.758 (2) Å, cH = 30.112 (9) Å, respectively. Similar [H3tren]2·(MX7)2·H2O (M = Zr, Ta; X = F, O) sheets are found in both structures; they are separated by a water layer (Ow(2)-Ow(3)) in [H3tren]2·(ZrF7)2·9H2O. Dehydration of [H3tren]2·(ZrF7)2·9H2O starts at room temperature and ends at 90 °C to give [H3tren]2·(ZrF7)2·H2O. [H3tren]2·(ZrF7)2·H2O layers remain probably unchanged during this dehydration and the existence of one intermediate [H3tren]2·(ZrF7)2·3H2O hydrate is assumed. Ow(1) molecules are tightly hydrogen bonded with -NH3+ groups and decomposition of [H3tren]2·(ZrF7)2·H2O occurs from 210 °C to 500 °C to give successively [H3tren]2·(ZrF6)·(Zr2F12) (285 °C), an intermediate unknown phase (320 °C) and ZrF4.  相似文献   

17.
Thermodynamics of chromium acetylacetonate sublimation   总被引:1,自引:0,他引:1  
The equilibrium sublimation pressure Cr(acac)3(s) = Cr(acac)3(g) was measured in the range 320 ≤ T (K) ≤ 476 by two procedures. One of them is Knudsen's effusion procedure with mass spectrometric analysis of the composition of the gas phase, which proved to be good in measuring low pressure. The second is mass spectrometric procedure “calibrated volume method” (CVM), which helped us to expand the possibilities of the effusion method toward high pressure range. Experimental data are in good agreement with each other.For this process were obtained ln(P (Pa)) = 39.197 − 15 308.5/T, enthalpy ΔsubH° (T) = 127.28 ± 0.22 kJ mol−1 and entropy ΔsubS° (T) = 230.1 ± 0.5 J mol−1 K−1.  相似文献   

18.
Three non-isostructural metal(II) coordination polymers (metal=copper, cobalt, cadmium) were synthesized under the same mild hydrothermal conditions (T=408 K) by mixture of the corresponding metal acetate with 2-carboxyethylphosphonic acid and 1,10-phenanthroline (1:1:1 M ratio) and their structures were determined by single-crystal X-ray diffraction. Cu2(HO3PCH2CH2COO)2(C12H8N2)2(H2O)2 and Cd2(HO3PCH2CH2COO)2(C12H8N2)2 are triclinic (space group P-1) with a=7.908(5) Å, b=10.373(5) Å, c=11.515(5) Å, α=111.683(5)°, β=95.801(5)°, γ=110.212(5)° (T=120 K), and a=8.162(5) Å, b=9.500(5) Å, c=11.148(5) Å, α=102.623(5)°, β=98.607(5)°, γ=113.004(5)° (T=293 K), respectively. In contrast, [Co2(HO3PCH2CH2COO)2(C12H8N2)2(μ-OH2)](H2O) is orthorhombic (space group Pbcn) with a=21.1057(2) Å, b=9.8231(1) Å, c=15.4251(1) Å (T=120 K). For these three compounds, structural features, including H-bond network and the π-π stacking interactions, and thermal stability are reported and discussed. None of the materials present a long-range magnetic order in the range of temperatures investigated from 300 K down to 1.8 K.  相似文献   

19.
The kinetic and thermodynamic parameters for regioisomerisation of 2-methyl- and 2,6-dimethyl-derivatives of tricarbonyl[η4-tropone]iron complexes have been studied by 1H NMR spectrometry over a range of 40 °C. Regioisomerisation of these complexes proceeds by an intramolecular first-order process and results in the almost complete conversion of the less stable complexes (48) to more stable regioisomers (15). The activation energies and half lifes for the conversion (4 → 1) and (8 → 5) were found to be ΔG#=92 kJ mol−1; τ1/2=12.8 h, and ΔG#=107 kJ mol−1; τ1/2=26.8 h, respectively, at 23 °C. Complex 1 reacts with (1R,2S,5R)-menthol in sulphuric acid solution, followed by neutralisation with sodium carbonate to give a separable mixture of diastereomeric tricarbonyl[(2,3,4,5-η)-(1R,2S,5R)-6-menthyloxy-2-methyltropone]iron complexes, 9 and 10. The corresponding dimethylated complex 5 fails to react under these conditions.  相似文献   

20.
Francesco Crea 《Talanta》2007,71(2):948-963
In this paper we investigated the interactions between dioxouranium(VI) and oxalate using (H+-glass electrode) potentiometry and titration calorimetry. Potentiometric measurements were carried out in NaCl aqueous solutions and at T = 25 °C in a wide range of experimental conditions (concentrations, ligand/metal molar ratio, pH, titrants) at low ionic strength values (I ≤ 0.090 mol L−1, without supporting electrolyte) and at I = 1.0 mol L−1; different procedures were employed for the acquisition of experimental data and careful analysis of these data performed. In all cases the speciation model that best fits experimental data takes into account the formation of the binary mononuclear species UO2(ox)0, UO2(ox)22−, UO2(ox)34− widely reported in literature, the ternary hydroxyl mononuclear species UO2(ox)OH, UO2(ox)(OH)22−, UO2(ox)2OH3−, UO2(ox)3OH5−, the protonated ternary mononuclear species UO2(ox)3H3− and the binuclear species (UO2)2(ox)56−.Calorimetric measurements were carried out following similar procedures and in the same experimental conditions as employed for the potentiometric measurements at I = 1.0 mol L−1 in NaCl. The stability of UO22+-oxalate2− complexes is fairly high and their main contribution to stability is entropic in nature. Some linear empirical relationships were found which make it possible to calculate (i) the contribution of a single bond: and ; (ii) chelate stabilisation per ring: and and (iii) the mean stability of negatively charged Na+-ion pair complexes: logTK = (0.46 ± 0.02)·|z| (z = charge of complex species), ΔG° = −(2.60 ± 0.1)·|z| kJ mol−1 and TΔS° = 2.5 ± 0.5 kJ mol−1. Both potentiometric and calorimetric results provide evidence of the penta-coordination of the species UO2(ox)34−. SIT parameters were calculated from the data at I = 0 and I = 1.02 mol kg−1. Comparisons are made with literature data. An insoluble dioxouranium(VI) ternary complex was synthesised (at I = 1.0 mol L−1 in NaCl) and characterised by thermoanalysis and elemental analysis.  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