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.
[PW11O39]7– heteropolyanion (HPA) stabilizes Ti(IV) in aqueous solution at Ti:PW11 ratios from 1 to 12 and pH 1–3. Ti(IV) is completely precipitated under these conditions in the absence of HPA. Differential dissolution phase analysis, optical, IR,31P and17O NMR spectra show that one Ti(IV) ion is incorporated into the Keggin lattice. The other ions, most probably, are located on the HPA surface in the form of oligomeric hydroxo fragments: [PW11TiIVO40·Tin–1 IVOxHy]k–. Both types of Ti(IV) ions bind peroxo groups on interaction of the complex with H2O2.  相似文献   

2.
Unsaturated heteropolyanions (HPA) [PW11O39]7− stabilize TiIV hydroxo complexes in aqueous solutions (Ti: PW11 [PW11O39]7−⪯12, pH 1–3). Spectral studies (optical,17O and31P NMR, and IR spectra) and studies by the differential dissolution method demonstrated that TiIV hydroxo complexes are stabilized through interactions of polynuclear TiIV hydroxo cations with heteropolyanions [PW11TiO40 5− formed. Depending on the reaction conditions, hydroxo cations Ti n−1O x H y m+ either add to oxygen atoms of the W−O−Ti bridges of the heteropolyanions to form the complex [PW11TiO40·Ti n−1O x H y ] k− (at [HPA]=0.01 mol L−1) or interact with TiIV of the heteropolyanions through the terminal o atom to give the polynuclear complexes [PW11O39Ti−O−Ti n−1O x H y ]q− (at [HPA]=0.2 mol L−1). When the complexes of the first type were treated with H2O2, TiIV ions added peroxo groups. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 914–920, May, 1997.  相似文献   

3.
Reverse osmosis was used for the separation of various types of heteropolyanions (HPA): [PW11O39M(H2O)] k (M = CoII, FeIII, CrIII), [(PW11O39Fe)2O]10– , and [PW11O39 · Fe n O x H y ] p from contaminant ions NO3 and Na+ that are usually introduced into the solution in the synthesis of HPA.Translated fromIzvestiya Akodemii Nauk. Seriya Khimicheskaya, No. 4, pp. 1009–1011, April, 1996.  相似文献   

4.
The 31P NMR method shows that four forms of titanium(IV)-monosubstituted Keggin-type heteropolytungstate (Ti–HPA) exist in MeCN: the dimer (Bu4N)7[{PTiW11O39}2OH] (in the abbreviated form, (PW11Ti)2OH or H1), its conjugate base (PW11Ti)2O (1), and two monomers, PW11TiO (2) and PW11TiOH (H2). The ratio between the forms depends on the concentrations of H+and H2O. Dimer H1is produced from 2in MeCN when H+(1.5 mol) is added, and monomer H2is the key intermediate in this process. The catalytic activity of Ti–HPA in the oxidation of thioethers by H2O2correlates with their activity in peroxo complex formation and decreases in the order H2> H1> 2. The reaction of 2with H2O2in MeCN occurs slowly to form the inactive peroxo complex PW11TiO2(A). The addition of H2O2to H1and H2most likely results in the formation of the active hydroperoxo complex PW11TiOOH (B). Complexes Aand Btransform into each other when H+or OH(1 mol) is added per 1 mol of Aor B, respectively. The activity of Btoward thioethers in the stoichiometric reaction is proven by 31PNMR and optical spectroscopy.  相似文献   

5.
Gold nanoparticles loaded onto Keggin‐type insoluble polyoxometalates (CsxH3?xPW12O40) showed superior catalytic performances for the direct conversion of cellobiose into gluconic acid in water in the presence of O2. The selectivity of Au/CsxH3?xPW12O40 for gluconic acid was significantly higher than those of Au catalysts loaded onto typical metal oxides (e.g., SiO2, Al2O3, and TiO2), carbon nanotubes, and zeolites (H‐ZSM‐5 and HY). The acidity of polyoxometalates and the mean‐size of the Au nanoparticles were the key factors in the catalytic conversion of cellobiose into gluconic acid. The stronger acidity of polyoxometalates not only favored the conversion of cellobiose but also resulted in higher selectivity of gluconic acid by facilitating desorption and inhibiting its further degradation. On the other hand, the smaller Au nanoparticles accelerated the oxidation of glucose (an intermediate) into gluconic acid, thereby leading to increases both in the conversion of cellobiose and in the selectivity of gluconic acid. The Au/CsxH3?xPW12O40 system also catalyzed the conversion of cellulose into gluconic acid with good efficiency, but it could not be used repeatedly owing to the leaching of a H+‐rich hydrophilic moiety over long‐term hydrothermal reactions. We have demonstrated that the combination of H3PW12O40 and Au/Cs3.0PW12O40 afforded excellent yields of gluconic acid (about 85 %, 418 K, 11 h), and the deactivation of the recovered H3PW12O40–Au/Cs3.0PW12O40 catalyst was not serious during repeated use.  相似文献   

6.
1H NMR was applied to study the interaction of chloral hydrate in deuterionitrobenzene solution with tetrabutylammonium salts of the heteropoly acids (HPA) belonging to five structural types: Keggin (H3PW12O40, H3PMo12O40, H4SiW12O40), Dawson (-H6P2W18O62, -H6P2Mo18O62, -H4S2Mo18O62), H6P2W21O71(H2O)3, H6As2W21O69(H2O), and H21B3W39O132. The surface of the HPA anions is nonuniform in acid-base properties. A general rule for all HPA was found, namely, that the HPA acidity increases with a decrease in the specific anion charge (per W or Mo atom).  相似文献   

7.
The electrochemical transfer behaviour of vanadium-containing heteropolytungstate anions [PW12−xVxO40](3+x)− (x = 1−4) across the water | nitrobenzene interface has been investigated by cyclic voltammetry and chronopotentiometry with cyclic linear current scanning. The transfer of PW11V1O4−40, HPW10V2O4−40, H2PW10V2O3−40, H3PW9V3O3−40 and H4PW8V4O3−40 across the water | nitrobenzene interface can be observed within the potential window. The effects were observed of pH in the water phase on the transfer behaviour and the formation of vanadium-containing heteropolytungstate anions in solution. Heteropolytungstate anions become more stable due to their involving the vanadium atom. The degree of protonation and the dissociation constant of the trivalent vanadium-containing heteropolytungstate anion of protonation increase with increasing vanadium content. The transfer processes are diffusion-controlled. The standard transfer potential, the standard Gibbs energy and the dissociation constant for vanadium-containing heteropolytungstate anions have been obtained and the transfer mechanisms are discussed.  相似文献   

8.
Combining the ion cyclotron resonance method and a Knudsen effusion source, we obtained a series of MoxOy + (x = 1 – 5, y = 1 – 15) molybdenum oxide cluster ions. We studied the dependence of the concentrations of these ions on the trapping time and their reactions with carbon monoxide. It is shown that MoxOy + ions with x > 3 contain a cyclic Mo3O9 fragment in their structure. The oxygen bond energies in MoxOy + ionic clusters are estimated.  相似文献   

9.
The formation of Pd(II)-containing and mixed Pd(II),Cu(II), Pd(II),Fe(III), and Pd(II),V(V) complexes with heteropolyanion PW9O9– 34was studied using 31P, 183W, 51V NMR, visible UV and IR spectroscopy, and the differentiating dissolution methods. In an aqueous solution and at optimal pH (3.7), the monometallic complexes [Pd3(PW9O34)2]12–and [Pd3(PW9O34)2Pd n O x H y ] q(n av= 3), the bimetallic complexes [Pd2Cu(PW9O34)2]12–, [Pd2Fe(PW9O34)2]11–, and [PdFe2(PW9O34)2]10–, and a mixture of the [Pd3(PW9O34)2Pd n O x H y ] q(n av 10) + [(VO)3(PW9O34)2]9–complexes are formed. The title complexes were isolated from solution as Cs+solid salts belonging to the same [M3(PW9O34)2] structural type.  相似文献   

10.
Adsorption of H3PW12O40 from water and organic oxygen-containing solvents (AcOH, Me2CO, MeOH) by carbon mesoporous materials, viz., Sibunit and catalytic filamentous carbons (CFC), was studied. The amount of irreversibly sorbed heteropolyacid is 50—100 mg g–1 of support and decreases in the series of solvents: H2O > Me2CO > AcOH > MeOH. The adsorption capacity of CFC depends on the specific surface, total pore volume, and microstructure of the CFC fiber.  相似文献   

11.
Vanadium-containing H6+xP2Mo18−xVxO62 (x = 0, 1, 2 and 3) Wells-Dawson heteropolyacid (HPA) and H3+xPMo12−xVxO40 (x = 0, 1, 2 and 3) Keggin HPA catalysts were applied to the vapor-phase dehydrogenation of cyclohexanol. The catalytic oxidation activity showed a volcano-shaped curve with respect to vanadium substitution for both families of HPA catalysts. The Wells-Dawson HPA showed a better catalytic oxidation performance than the Keggin HPA at the same level of vanadium substitution.  相似文献   

12.
Heteropoly acid (HPA) H8(PW11TiO39)2xH2O (I) is synthesized by three different ways and studied by chemical analysis, potentiometric titration, mass-spectrometry, IR, 31P, 183W, and 17O NMR spectroscopy, thermogravimetry, and transmission electron microscopy. Anion I consists of two subparticles of the Keggin structure bridged by Ti-O-Ti. The dimeric anion exists in HPA aqueous solutions at [I] > 0.02 M. At pH > 0.6 it splits to a [PW11TiO40]5− monomer stable up to pH ∼ 6. When heated (150–400)°C, I splits into H3PW12O40 and, apparently, H3PW10Ti2O38 without phase separation. Thermolysis products are soluble and when dissolved in water turn again into I. Complete decomposition of I to oxides occurs at ∼450°C.  相似文献   

13.
The surface acidity of H3PW12O40 and NaxH3-xPW12O40(x = 1-3) was studied using the adsorption of 2,2,6,6-tetramethyl-4-oxopiperidine-1-oxyl (TEMPON) and 2,2,6,6-tetramethyl-4-hydroxypiperidine-1-oxyl (TEMPOL) radicals. It was found that the amount of surface proton sites determined from the adsorption of TEMPON decreased with the degree of substitution of Na+ cations for protons. A correlation between amount of strong surface proton sites and catalytic activity of NaxH3-xPW12O40(x = 0-3) in the dealkylation reaction of 2,6-di-tert-butyl-4-methylphenol was found.Translated from Kinetika i Kataliz, Vol. 46, No. 1, 2005, pp. 131–136.Original Russian Text Copyright © 2005 by Timofeeva, Ayupov, Volodin, Pak, Volkova, Echevskii.  相似文献   

14.
A series of insoluble cesium partly substituted Wells–Dawson type heteropolyacids, CsxH6−xP2W18O62 (x = 1.5–6.0), were synthesized and characterized using the techniques including UV–vis/DRS, FT-IR, XRD, XPS, and N2 porosimetry. As the unique and reusable solid acid catalysts, CsxH6−xP2W18O62 salts were applied to produce diphenolic acid by the condensation reaction of phenol with bio-platform molecule, levulinic acid. For comparison, cesium partly substituted Keggin type heteropolyacids (CsxH3−xPW12O40, x = 1.0–3.0), HCl, HZSM-5, and MCM-49 were also tested. Influences on the catalytic activity and selectivity were considered for factors including solvent, molar ratio of phenol to levulinic acid, amount of catalyst, reaction temperature, stirring speed, and reaction time. The experimental results demonstrated that both Cs1.5H4.5P2W18O62and Cs2.5H0.5PW12O40 exhibited excellent catalytic performance under solvent-free conditions. Furthermore, both selectivity and activity of Cs1.5H4.5P2W18O62 were higher than those of Cs2.5H0.5PW12O40. Reasons for the different catalytic behaviors between two types of cesium partly substituted heteropolyacids were investigated.  相似文献   

15.
Acid anilinium dodecatungstenphosphate of composition (C6H5NH3)2H[PW12O40] 2H2O is synthesized and characterized by X-ray diffraction analysis, IR spectroscopy, and thermogravimetry. The crystals are monoclinic: space group P21/m, a = 9.940(2) , b = 15.412(3) , c = 16.201(3) , = 100.42(3)°, Z = 2, calcd = 4.221 g/cm3. The structure contains heteropolyanions [PW12O40]3–, cations [C6H5NH3]+ and H+, and molecules of water of crystallization. The heteropolyanions and anilinium cations in crystal are linked by the electrostatic interaction and hydrogen bonds.__________Translated from Koordinatsionnaya Khimiya, Vol. 31, No. 4, 2005, pp. 273–279.Original Russian Text Copyright © 2005 by Kaziev, Dutov, Quinones, Koroteev, Belskii, Stash, Kuznetsova.  相似文献   

16.
A set of oxygen-containing molybdenum oxide clusters Mo x O y (x = 1–3; y = 1–9) was obtained with the use of a combination of a Knudsen cell and an ion trap cell. The reactions of positively charged clusters with C1–C4 alcohols were studied using ion cyclotron resonance. The formation of a number of organometallic ions, the products of initial insertion of molybdenum oxide ions into the C–O and C–H bonds of alcohols, and polycondensation products of methanol and ethanol were found. The reactions of neutral molybdenum oxide clusters Mo x O y (x = 1–3; y = 1–9) with protonated C1–C4 alcohols and an ammonium ion were studied. The following limits of proton affinity (PA) were found for neutral oxygen-containing molybdenum clusters: (MoO) < 180, (Mo2O4, Mo2O5, and Mo3O8) = 188 ± 8, PA(MoO2) = 202 ± 5, PA(MoO3, Mo2O6, and Mo3O9) > 207 kcal/mol.  相似文献   

17.
Excess molar volumes V E and excess molar heat capacities C P E at constant pressure have been measured, at 25°C, as a function of composition for the four binary liquid mixtures propylene carbonate (4-methyl-1,3-dioxolan-2-one, C4H6O3; PC) + benzene (C6H6;B), + toluene (C6H5CH3;T), + ethylbenzene (C6H5C2H5;EB), and + p-xylene (p-C6H4(CH3)2;p-X) using a vibrating-tube densimeter and a Picker flow microcalorimeter, respectively. All the excess volumes are negative and noticeably skewed towards the hydrocarbon side: V E (cm3-mol–1) at the minimum ranges from about –0.31 at x1=0.43 for {x1C4H6O3+x2p-C6H4(CH3)2}, to –0.45 at x1=0.40 for {x1C4H6O3+x2C6H5CH3}. For the systems (PC+T), (PC+EB) and (PC+p-X) the C P E s are all positive and even more skewed. For instance, for (PC+T) the maximum is at x 1,max =0.31 with C P,max E =1.91 J-K–1-mol–1. Most interestingly, C P E of {x1C4H6O3+x2C6H6} exhibits two maxima near the ends of the composition range and a minimum at x 1,min =0.71 with C P,min E =–0.23 J-K–1-mol–1. For this type of mixture, it is the first reported case of an M-shaped composition dependence of the excess molar heat capacity at constant pressure.Communicated at the Festsymposium celebrating Dr. Henry V. Kehiaian's 60th birthday, Clermont-Ferrand, France, 17–18 May 1990.  相似文献   

18.
We have measured the synchrotron‐induced photofragmentation of isolated 2‐deoxy‐D ‐ribose molecules (C5H10O4) at four photon energies, namely, 23.0, 15.7, 14.6, and 13.8 eV. At all photon energies above the molecule′s ionization threshold we observe the formation of a large variety of molecular cation fragments, including CH3+, OH+, H3O+, C2H3+, C2H4+, CHxO+ (x=1,2,3), C2HxO+ (x=1–5), C3HxO+ (x=3–5), C2H4O2+, C3HxO2+ (x=1,2,4–6), C4H5O2+, C4HxO3+ (x=6,7), C5H7O3+, and C5H8O3+. The formation of these fragments shows a strong propensity of the DNA sugar to dissociate upon absorption of vacuum ultraviolet photons. The yields of particular fragments at various excitation photon energies in the range between 10 and 28 eV are also measured and their appearance thresholds determined. At all photon energies, the most intense relative yield is recorded for the m/q=57 fragment (C3H5O+), whereas a general intensity decrease is observed for all other fragments— relative to the m/q=57 fragment—with decreasing excitation energy. Thus, bond cleavage depends on the photon energy deposited in the molecule. All fragments up to m/q=75 are observed at all photon energies above their respective threshold values. Most notably, several fragmentation products, for example, CH3+, H3O+, C2H4+, CH3O+, and C2H5O+, involve significant bond rearrangements and nuclear motion during the dissociation time. Multibond fragmentation of the sugar moiety in the sugar–phosphate backbone of DNA results in complex strand lesions and, most likely, in subsequent reactions of the neutral or charged fragments with the surrounding DNA molecules.  相似文献   

19.
Through a combination of Raman spectroscopy, multi-element NMR spectroscopy and chemical analysis, the differences between the action of carbonate and carbamate as agents for dissolving Cs3PMo12O40xH2O(s) (CPM) and ZrMO2O7(OH)2(H2O)2(s) (ZM) have been elucidated. Alkaline H2NCO2/HCO3/CO32− solutions, derived from the dissolution of ammonium carbamate (NH4H2NCO2; AC), dissolve CPM by base hydrolysis of the PMo12O403− Keggin anion, ultimately forming [MoO4]2− and PO43− when excess base is present. If the initial concentration of H2NCO2/HCO3/ CO32− is lowered, base hydrolysis is incomplete and the dissolved species include [Mo7O24]6− and [P2Mo5O23]6−, and undissolved solid Cs3PMo12O40, CsxNH7−xPMo11O39, and CsxNH6−xMo7O24 remain. Na2CO3 solutions dissolve Cs3PMo12O40 through a similar mechanism, but the dissolution rate is much lower. We attribute this difference to the different buffering effects of H2NCO2/HCO3/CO32− and CO32−/HCO3 solutions, and the instability of carbamic acid, the protonated form of H2NCO2 (which rapidly decomposes into NH3 and CO2). The ability of NH3 to produce NH4+ and OH, together with the evolution of CO2 gas, drive the reaction forward. Low temperature measurements under conditions where pure H2NCO2 is kinetically stable, allowed the rates of dissolution of CPM by H2NCO2 and CO32− to be compared directly, confirming the faster dissolution by H2NCO2. Compared to CPM, the dissolution of ZM by H2NCO2/HCO3/CO32− is a much slower process and is driven by the formation of soluble ZrIV-carbonate complexes and MoO42−. The driving force for the dissolution of ZM is the superior complexing ability of carbonate over carbamate; consequently solutions containing a higher carbonate concentration dissolve ZM faster.  相似文献   

20.
The acid properties of heteropoly acids of the following three structure types were studied by conductometry in acetic acid: Keggin (H3PW12O40, H3PMo12O40, H4SiW12O40, H3PW11ThO39; and H5PW11XO40, where X(IV) = Ti or Zr), Dawson (-H6P2W18O62and -H6P2Mo18O62), and H6P2W21O71(H2O)3. These compounds are electrolytes that dissociate in only the first step of this solvent. The thermodynamic dissociation constants of the heteropoly acids were calculated by the Fuoss–Kraus method. The Hammett acidity functions H 0of the solutions of H5PW11XO40, H3PW12O40, H4SiW12O40, and H6P2W21O71(H2O)3in 85% acetic acid at 25°C were determined by the indicator method. All of the test heteropoly acids were found to be strong acids.  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