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 812 毫秒
1.
Raman spectra have been measured for aqueous Al2(SO4)3 solutions from 25 to hydrothermal conditions at 184°C under steam saturation. The Raman spectrum at 184°C contained four polarized bands in the S–O stretching wavenumber range, which suggest that a new sulfato complex, where sulfate acts as a bridging ligand (possibly bidentate or tridentate), is formed in solution, in addition to a 1:1 aluminium(III) sulfato complex, where sulfate is monodentate, which is the only ion pair identified at room temperature. Under hydrothermal conditions, it was possible to observe the hydrolysis of aluminium(III) aqua ion by measuring the relative intensity of bands due to SO2? 4 and HSO4 ?, according to the coupled equilibrium reaction [Al(OH2)6]3+ + SO4 2? ? [Al(OH2)5OH]2+ + HSO4 ?. The precipitate in equilibrium with the solution at 184°C could be characterized as a hydronium alunite, (H3O)Al3(SO4)2(OH)6, by chemical analysis, X-ray diffraction, and Raman and infrared spectroscopy.  相似文献   

2.
The enthalpy of formation at 298.15 K of the polymer Al13O4(OH)28(H2O)3+8 and an amorphous aluminium trihydroxide gel was studied using an original differential calorimetric method, already developed for adsorption experiments, and aluminium-27 NMR spectroscopy data. ΔHf “Al13” (298.15 K) = ? 602 ± 60.2 kJ mole?1 and ΔHf Al(OH)3 (298.15 K) = ? 51 ± 5 kJ mole?1. Using theoretical values of ΔGR “Al13” and ΔGR Al(OH)3, we calculated ΔGf “Al13” (298.15 K) = ? 13282 kJ mole?1; ΔSf “Al13” (298.15 K) = + 42.2 kJ mole?1; ΔGf Al(OH)3 (298.15 K) = ? 782.5 kJ mole?1; and ΔSf Al(OH)3 (298.15 K) = + 2.4 kJ mole?1.  相似文献   

3.
The formation of various hydrolytic and mixed hydrolytic complexes of the aluminium(III) ion in the presence of glycine and L-alanine, has been studied in 0.5 mol dm?3 (Na)NO3 medium at 25deg;C, by emf method. The concentration ratios of amine acids to aluminium(III) were varied from 1 : 1 to 10 : 1. The least-squares treatment of the data obtained, in the absence of the amino acids, indicates the formation of the dimer, [Al2(OH)2]4+, and monomer, [AlOH]2+, with the stability constants log β22 = ?7.03 ± 0.03 and log β11 = ?5.65 ± 0.09, respectively. At pH values higher then ~4.0 formation of the trimer [Al3(OH)4]5+ (log β34 = ?12.60 ± 0.08) becomes significant. In the presence of amino acids the evidence has been found for the formation of [Al2(OH)4]2+ (log β24 = ?15.65 ± 0.09). Besides the formation of the pure hydrolytic complexes, equilibria in the title systems can be explained by assuming the main reaction products to have the compositions [Al(OH)3Gly] (log β131 = ?7.53 ± 0.04), [Al2(OH)2(Gly)2] (log β222 = 6.56 ± 0.09) and [Al(OH)3Ala] (log β131 = ?7.70 ± 0.03), [Al2(OH)2Ala2] (log β222 = 7.23 ± 0.07).  相似文献   

4.
Al13 is one of the novel nanospecies in partially neutralized Al(III) solution and Al-OH sol or precipitate could be generated simultaneously in the neutralization. Unfortunately, the precipitate is believed to be harmful to the formation of Al13 due to the consumption of Al(OH)4 , which was regarded as the precursor of Al13. In this paper, the feasibility and potential of transformation of freshly formed Al-OH precipitate into Al13 species were studied by using ferron colorimetric method and 27Al NMR spectroscopy. The Al-OH precipitates were produced by two ways: injection of base solution into Al(III) solution gradually and mixing of Al(III) and base solutions instantaneously. The re-dissolving behaviors of the freshly formed precipitates were examined under different basicities (OH/Al molar ratio) and temperatures. It has been shown that Al13 could be formed through the re-dissolution of intermediate Al-OH precipitate generated in partially neutralized Al(III) solution. A possible formation mechanism of Al13 was suggested. Easily transformable precipitate was developed when the OH/Al molar ratio was less than 2.5. Rapid re-dissolution of freshly formed precipitate was favorable for Al13 formation, which could be enhanced by heating.  相似文献   

5.
[AlO4Al12(OH)24(H2O)12]7+ (Al13) formation in electrolysis process is studied. The results detected by27Al NMR spectroscopy show that high content of Al13 polymer is formed in the partially hydrolyzed aluminum solution prepared by controlled electrolysis process. In the produced electrolyte of total Al concentration ([AlT]) 2.0 mol · L−1 with a basicity (B = OH/Al molar ratios) of 2.0, the content of Al13 polymer is over 60% of total Al. Dynamic light scattering shows that the size distribution of the final electrolyte solutions ([AlT] = 2.0 mol · L−1) is trimodal with B = 2.0 and bimodal with B = 2.5. The aggregates of Al13 complexes increase the particle size of partially hydrolyzed aluminum solution.  相似文献   

6.
The ‘formal’ hydrolysis ratio (h = C(OH)added/C(Al)total) of hydrolysed aluminium-ions is an important parameter required for the exhaustive and quantitative speciation-fractionation of aluminium in aqueous solutions. This paper describes a potentiometric method for determination of the formal hydrolysis ratio based on an automated alkaline titration procedure. The method uses the point of precipitation of aluminium hydroxide as a reference (h = 3.0) in order to calculate the initial formal hydrolysis ratio of hydrolysed aluminium-ion solutions. Several solutions of pure hydrolytic species including aluminium monomers (AlCl3), Al13 polynuclear cluster ([Al13O4(OH)24(H2O)12]7+), Al30 polynuclear cluster ([Al30O8(OH)56(H2O)26]18+) and a suspension of nanoparticulate aluminium hydroxide have been used as ‘reference standards’ to validate the proposed potentiometric method. Other important variables in the potentiometric determination of the hydrolysis ratio have also been optimised including the concentration of aluminium and the type and strength of alkali (Trizma-base, NH3, NaHCO3, Na2CO3 and KOH). The results of the potentiometric analysis have been cross-verified by quantitative 27Al solution nuclear magnetic resonance (27Al NMR) measurements. The ‘formal’ hydrolysis ratio of a commercial basic aluminium chloride has been measured as an example of a practical application of the developed technique.  相似文献   

7.
聚合Al13晶体的制备及表征   总被引:7,自引:0,他引:7  
近几十年来 , 环境污染的日益严重使人们对健康问题和全球生态系统越来越关注 ?由于一个元素的生物可给性在很大程度上取决于它存在的物理化学形态和浓度 , 准确测定环境和生物体系中的痕量元素的不同形态是研究这些元素的生物毒性 ? 生物有效性和传输机理的关键 ?形态分析成了  相似文献   

8.
The forced hydrolysis reaction of aqueous aluminum ion (Al3+) is of critical importance in Al chemistry, but its microscopic mechanism has long been neglected. Herein, density functional calculations reveal an external OH‐induced barrierless proton dissociation mechanism for the forced hydrolysis of Al3+(aq). Dynamic reaction pathway modeling results show that the barrierless deprotonations induced by the second‐ or third‐shell external OH proceed via the concerted proton transfer through H‐bond wires connected to the coordinated waters, and the inducing ability of the external OH decreases with increasing hydration layers between Al(H2O)63+ and the external OH. The OH‐induced forced hydrolysis mechanism of Al3+(aq) is quite different from its self‐hydrolysis mechanism without OH. The inducing ability is a unique characteristic of OH, rather than other anions such as F or Cl.  相似文献   

9.
10.
The structural, electronic, energy, and vibrational characteristics of the Al13X? and Al13X 2 ? clusters, with an aluminum-centered (Alc) icosahedral cage Al13 and with one or two outer-sphere ligands X=H, F, Cl, Br, OH, NH2, CH3, C6H5, have been calculated within the B3LYP approximation of the density functional theory using the 6-31G* and 6-311+G* basis sets. In all Al13X? radicals, the unpaired electron is localized at the cage atom Al* located opposite the Al-X bond. This Al* atom is the most favorable site for attaching the second X ligand of any nature (trans-addition rule). According to the previously suggested molecular model of the valence state of the [Al 13 ? ] “superatom,” the calculated energies D 1(Al 13 ? -X) of addition of the first ligand to the Al 13 ? anion are about 1 eV lower than the corresponding energies of addition of the second ligand D 2(XAl 13 ? -X). The structure of the Al13 cage depends on the nature of the nature of the substituent X and can radically change in going from anions to their neutral congeners. In the lowest-lying Al13X isomer with electronegative substituents X (Hal, OH, NH2, CH3, etc.), the aluminum cage has a marquee structure (1, symmetry C s) with a hexagonal base and a pentagonal “roof.” For Al13X analogues with electropositive ligands X (Al, Li, Na), a tridentate isomer (T, C 3v ) with the X substituent coordinated to a face of the Al13 icosahedron is preferable. In the case of moderately electronegative X ligands (of the H type), the marquee (1) and icosahedral (T) isomers are close in energy. The stretching vibration frequencies of isomers 1 and T differ significantly in magnitude and intensity so that vibrational spectroscopy methods can be especially applicable to their experimental identification.  相似文献   

11.
The chemistry of mono or ortho silicic acid (Si(OH)4) is barely considered in most chemistry texts. Mention is usually only made of its autocondensation in forming hydrated amorphous silica and its reaction with ammonium molybdate in forming the molybdosilicic acid complex. Reference should now be made to its unique inorganic chemistry with aluminium (Al) and specifically aluminium hydroxide (Al(OH)3(s)) in forming hydroxyaluminosilicates (HAS(s)). The competitive condensation or substitution of Si(OH)4 into a framework of Al(OH)3(s) results in the formation of either HASA or HASB. Which type of HAS(s) predominates depends upon the ratio of Si:Al in preparative solutions with the formation of HASB requiring a two-fold excess of Si(OH)4 over Al. The Si:Al ratio of HASA is 0.5 and the existence of HASA is a prerequisite to the formation of HASB in which the ratio of Si:Al is 1.0. HASA is composed of only octahedrally co-ordinated Al, AlVI, whereas HASB is composed of equal quantities of AlVI and tetrahedrally coordinated Al, AlIV, and is formed by a Si(OH)4-fuelled dehydroxylation reaction. HAS(s) are significantly more ‘kinetically’ stable than Al(OH)3(amorphous) with HASB predicted to predominate at pH > 4.0 and [Si(OH)4] > 0.1 mmol/L. HAS(s) are critical secondary mineral phases in the biogeochemical cycle of Al and Si(OH)4 and the formation of HAS(s) have played a major role in precluding Al3+(aq) from biochemical evolution. In the future Si(OH)4 and the formation of HAS(s) are predicted to be of significant importance in providing protection for humans against a potentially burgeoning exposure to biologically available Al.  相似文献   

12.
The reaction of Be · aq2+ with OH? leeds not only to loss of protons by the metalaquo ion but also to structural changes in the solvation sphere. These can be studied by following the pH variations during the first decisecond after mixing the solutions of metal salt and alkali hydroxide. The equilibrium Be2+ ? BeOH+ is reached within 5 milliseconds if acid free Beryllium solutions are used. If the metal solution is strongly acidic, however, the establishment of the equilibrium needs more time because of the slowness of the process H+ + BeOH+ → Be2+ (k ~ 105 M?1, s?1). The extraction of two protons produces in the first instance an unstable Be(OH) species which transforms into the stable isomer Be(OH)2 (solvatation isomerism) in a first-order reaction of half-life of 7 ms. This isomerisation causes almost complete disappearance of BeOH+ from the equilibrium Be2+ ? BeOH+ ? Be(OH)2. (KAKIHANA & SILLEN state that the relaxed solutions contain only Be2+, Be(OH)2, Be3(OH) and some Be2OH3+.) The formation of the polynuclear species Be3(OH) needs about 30 seconds to go to completion.  相似文献   

13.
Silver(III) has a half-life at pH 11 of several hundred seconds in aqueous solutions in the presence of 0.1–1.0 M concentrations of certain basic oxoanions (Oxo) (phosphate, carbonate, borate, pyrophosphate, and arsenate). This compares with a lifetime of a few seconds at pH 11 in the absence of these oxoanions. UV-visible spectra and kinetic data for these solutions are interpreted as evidence for the following equilibria in the pH range 9–13.Ag(OH)4?1 + H2O ? Ag(OH)3H2O + OH? (1)Ag(OH)4?1 + Oxo ? Ag(OH)3Oxo + OH? (2)Ag(OH)3Oxo + H2O ? Ag(OH)2(Oxo)H2O + OH? (3) Values of K3 lie in the range 10?3 < K3 <104 M for the systems studied. K2 is estimated to be ~102 for phosphate and slightly smaller for the other systems. Ag(OH)4? undergoes an unusual reaction with pyrophosphate at pH ~ 8 to form a novel silver(II) complex, [Ag(P2O7)2]6?, for which EPR and electronic absorption spectral parameters are reported.  相似文献   

14.
The mechanisms of the redox reactions between a polymer containing Al(III) sulfonated phthalocyanine pendants, (AlIII(?NHS(O2)trspc)2?)2, and radicals have been investigated in this work. Pulse radiolysis and photochemical methods were used for these studies. Oxidizing radicals, OH?, HCO3?, (CH3)2COHCH2?, and N3?, as well as reducing radicals, eaq?, CO2??, and (CH3)2C?OH, respectively accept or donate one electron forming pendent phthalocyanine radicals, AlIII(?NHS(O2)trspc ?)? or 3?. The kinetics of the redox processes is consistent with a mechanism where the pendants react with radicals formed inside aggregates of five to six polymer strands. Electron donating radicals, that is, CO2?? and (CH3)2C?OH, produce one‐electron reduced phthalocyanine pendants that, even though they were stable under anaerobic conditions, donated charge to a Pt catalyst. While the polymer was regenerated in the Pt catalyzed processes, 2‐propanol and CO2 were respectively reduced to propane and CO. The reaction of SO3?? radicals with the polymer stood in contrast with the reactions of the radicals mentioned above. A first step of the mechanism, the coordination of the SO3?? radical to the Al(III), was subsequently followed by the formation of a SO3?? ‐ phthalocyanine ligand adduct. The decay of the SO3?? ‐ phthalocyanine ligand adduct in a ~102 ms time domain regenerates the polymer, and it was attributed to the dimerization/disproportionation of SO3?? radicals escaping from the aggregates of polymer. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

15.
The reactions of Al(III), Ga(III) and In(III) nitrates with 2-quinaldic acid (qaH) afforded [Al2(OH)2(qa)4]·2H2O (1), [Ga(qa)2(H2O)2]NO3 (2) and [In(qa)2(NO3)(H2O)] (3), respectively, in high yields. The crystal structures of 1, 2 and 3 have been determined by single-crystal X-ray crystallography. The structure of 1 features a di-hydroxo bridged [Al2(μ-OH)2]4+ dimer in which each Al(III) is further ligated by two bidentate chelate qa? ligands. Complexes 2 and 3 are mononuclear with the M(III) ions in octahedral environments surrounded by two bidentate chelate qa? and two H2O in 2 or one H2O and a terminal NO3? in 3. Characteristic IR as well as thermal analysis and solid-state fluorescence are discussed.  相似文献   

16.
Raman spectroscopic measurements were performed at ambient temperature onaqueous silica-bearing solutions (0.005 < m Si < 0.02; 0 < pH < 14). The spectraare consistent with the formation of monomeric Si(OH)o 4, SiO(OH) 3 andSiO2(OH)2– 2 species at acid to neutral, basic, and strongly basic pH, respectively.Raman spectra of aqueous Al-bearing solutions at basic pH confirm thepredominance of the Al(OH) 4 species in a wide concentration range (0.01 < m Al < 0.1).Raman spectra of basic solutions (12.4 < pH < 14.3), containing both Al andSi, exhibit a strong decrease in intensities of SiO(OH) 3, SiO2(OH)2– 2, andAl(OH) 4 bands in comparison with Al-free Si-bearing and Si-free Al-bearingsolutions of the same metal concentration and pH, suggesting the formation ofsoluble Al—Si complexes. The amounts of complexed Al and Si derived fromthe measurements of the Al and Si band intensities in strongly basic solutions(pH 14) are consistent with the formation, between Al(OH) 4 andSiO2(OH)2– 2, of the single Al—Si dimer SiAlO3(OH)3– 4 according to the reactionSiO2(OH)2– 2 + Al(OH) 4 SiAlO3(OH)3– 4 + H2OAt lower pH ( 12.5) the changes in band intensities are consistent with theformation of several, likely more polymerized, Al—Si complexes.  相似文献   

17.
The complex species formed between vanadium(III)?C2,2??-bipyridine (Bipy) and the small blood serum bioligands lactic (HLac), oxalic (H2Ox), citric (H3Cit) and phosphoric (H3PO4) acids were studied in aqueous solution by means of electromotive forces measurements emf(H) at 25?°C and 3.0?mol?dm?3 KCl as the ionic medium. The data were analyzed using the least-squares computational program LETAGROP, taking into account the hydrolytic products of vanadium(III) and the binary complexes formed. Formation of the complexes [V(Bipy)(Lac)]2+, [V(Bipy)(Lac)2]+, [V(OH)2(Bipy)(Lac)] and [V2O(Bipy)2(Lac)2]? were observed in the vanadium(III)?CBipy?CHLac system. Also, the species [V(Bipy)(HOx)]2+, [V(Bipy)(Ox)]+, [V(OH)(Bipy)(Ox)], [V(OH)2(Bipy)(Ox)]? and [V(OH)3(Bipy)(Ox)]2? were found in the vanadium(III)?CBipy?CH2Ox system, the complexes [V(Bipy)(HCit)]+, [V(Bipy)(Cit)], [V(OH)(Bipy)(Cit)]? and [V(OH)2(Bipy)(Cit)]2? were found in the vanadium(III)?CBipy?CH3Cit system, and the species [V(Bipy)(H2PO4)]2+ and [V(Bipy)(HPO4)]+ were detected in the vanadium(III)?CBipy?CH3PO4 system. The stability constants of these complexes were determined.  相似文献   

18.
SDS micelles flocculate in the presence of Al3+, creating an aggregate with pollutant-removing properties. The fraction of SDS micelles flocculating depends on the concentrations of SDS and Al3+. This paper describes how this fraction also changes with pH. There are two reasons for this dependence: a change in pH has a strong effect in the solution chemistry of Al3+, converting it into the compound [Al13O4(OH)24]7+, a strong flocculant, or precipitating it as Al(OH)3; and at low pH protons may compete with Al3+ as binding counter-ions for micelles. An increase in pH allows flocculation of SDS at high concentrations of Al3+, which under unmodified pH does not occur. Micelle flocculates are reported in this work to exist between pH 5 and pH 8, suggesting the potential use of Adsorptive Micellar Flocculation for the removal of anionic pollutants from waste waters not necessarily limited to acidic solutions. Received: 18 August 2000 Accepted: 5 October 2000  相似文献   

19.
Crystal and molecular structures of three Al(III) complexes of the tripod ligand 2,2′,2″-nitrilotriphenolate ( I ) are presented. They all show 5-coordinate Al in approximately trigonal bipyramidal geometry, with an external nucleophile X occupying the second axial position. X is OH? in[Al( I )(OH)]?[Hquin]+ (quin = quinuclidine), N in [Al( I )(py)] (py = pyridine), and one of the O-atoms of a second molecule in the dimeric [(Al( I ))2]. Correlated variations in the axial bond lengths of the trigonal bipyramid are observed: [(Al( I ))2]: Al–Nint. = 2.094 Å, Al–Oext. = 1.850 Å; [Al( I )(py)]: Al–Nint. = 2.153 Å, Al–Next., = 1.992 Å; [Al( I )(OH)]?: Al–Nint. = 2.278 Å, Al–Oext. = 1.765 Å. They are interpreted in terms of a dissociative reaction path at the Al(III) centre.  相似文献   

20.
The anode-cathode interplay is an important but rarely considered factor that initiates the degradation of aqueous zinc ion batteries (AZIBs). Herein, to address the limited cyclability issue of V-based AZIBs, Al2(SO4)3 is proposed as decent electrolyte additive to manipulate OH-mediated cross-communication between Zn anode and NaV3O8 ⋅ 1.5H2O (NVO) cathode. The hydrolysis of Al3+ creates a pH≈0.9 strong acidic environment, which unexpectedly prolongs the anode lifespan from 200 to 1000 h. Such impressive improvement is assigned to the alleviation of interfacial OH accumulation by Al3+ adsorption and solid electrolyte interphase formation. Accordingly, the strongly acidified electrolyte, associated with the sedated crossover of anodic OH toward NVO, remarkably mitigate its undesired dissolution and phase transition. The interrupted OH-mediated communication between the two electrodes endows Zn||NVO batteries with superb cycling stability, at both low and high scan rates.  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