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 671 毫秒
1.
The Se(VI)-analogues of ettringite and monosulfate, selenate-AFt (3CaO·Al2O3·3CaSeO4·37.5H2O), and selenate-AFm (3CaO·Al2O3·CaSeO4·xH2O) were synthesised and characterised by bulk chemical analysis and X-ray diffraction. Their solubility products were determined from a series of batch and resuspension experiments conducted at 25 °C. For selenate-AFt suspensions, the pH varied between 11.37 and 11.61, and a solubility product, log Kso=61.29±0.60 (I=0 M), was determined for the reaction 3CaO·Al2O3·3CaSeO4·37.5H2O+12 H+⇔6Ca2++2Al3++3SeO42−+43.5H2O. Selenate-AFm synthesis resulted in the uptake of Na, which was leached during equilibration and resuspension. For the pH range of 11.75 to 11.90, a solubility product, log Kso=73.40±0.22 (I=0 M), was determined for the reaction 3CaO·Al2O3·CaSeO4·xH2O+12 H+⇔4Ca2++2Al3++SeO42−+(x+6)H2O. Thermodynamic modelling suggested that both selenate-AFt and selenate-AFm are stable in the cementitious matrix; and that in a cement limited in sulfate, selenate concentration may be limited by selenate-AFm to below the millimolar range above pH 12.  相似文献   

2.
This paper presents polarographic and voltammetric studies of the reduction of the herbicide imazamethabenz methyl (2/3-methyl-(4-isopropyl-4-methyl-5-oxo-2-imidazolin-2-yl)-p-toluate), on mercury and carbon electrodes. The electrochemical studies were performed in strongly acidic media (0.1-2.7 M H2SO4) as well as in the pH range of 1-12. The overall reduction process involves the uptake of two electrons. The results obtained in polarography show that there is the reduction of two species, related via an acid-base equilibrium, and having very close reduction potentials. The voltammetric results obtained with a glassy carbon electrode were very similar to those observed on mercury electrodes. The reducible group in the molecule is the imidazolinone ring. In strongly acidic media (pH < pKa), the reaction mechanism proposed is the reduction of the protonated herbicide by an electrochemical-chemical-electrochemical (ECE) process, being the r.d.s. the second electron transfer. At pH > pKa the neutral form of the herbicide is reduced and the second electron transfer becomes reversible or quasi-reversible. In basic media, the species reduced is the deprotonated imazamethabenz methyl and the r.d.s. is the second electron transfer.  相似文献   

3.
The Li+ ion-exchange reaction of K+-type α-K0.14MnO1.93·nH2O containing different amounts of water molecules (n = 0-0.15) with a large (2 × 2) tunnel structure has been investigated in a LiNO3-LiCl molten salt at 300 °C. The Li+ ion-exchanged products were examined by chemical analysis, X-ray diffraction, and transmission electron microscopy measurements. The K+ ions and the hydrogens of the water molecules in the (2 × 2) tunnels of α-MnO2 were exchanged by Li+ ions in the molten salt, resulting in the Li+-type α-MnO2 containing different amounts of Li+ ions and lithium oxide (Li2O) in the (2 × 2) tunnels with maintaining the original hollandite structure.The electrochemical properties and structural variation with initial discharge and charge-discharge cycling of the Li+ ion-exchanged α-MnO2 samples have been investigated as insertion compounds in the search for new cathode materials for rechargeable lithium batteries. The Li+ ion-exchanged α-MnO2 samples provided higher capacities and higher Li+ ion diffusivity than the parent K+-type materials on initial discharge and charge-discharge cyclings, probably due to the structural stabilization with the existence of Li2O in the (2 × 2) tunnels.  相似文献   

4.
The electrochemical behaviour of fluorinated bis(cyclopentadienyl) mono(β-diketonato) titanium(IV) complexes, of general formula [Cp2Ti(R′COCHCOR)]+ClO4 with Cp = cyclopentadienyl and R′, R = CF3, C4H3S; CF3, C4H3O; CF3, Ph (C6H5); CF3, CH3; CH3, CH3; Ph, Ph and Ph, CH3 is described. Both metal and ligand based redox processes are observed. The chemically and electrochemically reversible TiIV/TiIII couple is followed by an irreversible ligand reduction at a considerably more negative (cathodic) potential. A comparison of the ligand reduction in its free and chelated state indicates that the β-diketonato ligand (R′COCHCOR) in [Cp2Ti(R′COCHCOR)]+ClO4 is electroactive at more negative potentials. A theoretical density functional theory (DFT) study shows that a highly localized metal centred frontier orbital dominates the TiIV/TiIII redox chemistry resulting in a non-linear relationship between the formal redox potential (E°′) and the sum of the group electronegativities of the R and R′ groups, χR + χR′, of the ligand. Linear relationships, however, are obtained between the DFT calculated electron affinity (EA) of the complexes and χR + χR′, the pKa of the free β-diketones R′COCH2COR and the carbonyl stretching frequency, vCO, of the complexes. The DFT calculated electronic structure of the second reduced species [Cp2Ti(β-diketonato)] shows that it is best described as Ti(III) coupled to a β-diketonato radical.  相似文献   

5.
A model was proposed to calculate several thermodynamic parameters for the initial-stage sintering of an alumina powder obtained after calcinations at 900 °C for 2 h of a precursor. The precursor was synthesized by an alumina sulphate-excess urea reaction in boiling aqueous solution. The cylindrical compacts of the powder with a diameter of 14 mm were prepared under 32 MPa by uniaxial pressing using oleic acid (12% by mass) as binder. The compacts were fired at various temperatures between 900 and 1400 °C for 2 h. The diameter (D) of the compacts before and after firing was measured by a micrometer. The D value after firing was taken as a sintering equilibrium parameter. An arbitrary sintering equilibrium constant (Ka) was calculated for each firing temperature by assuming Ka = (Di − D) / (D − Df), where Di is the largest value before sintering and Df is the smallest value after firing at 1400 °C. Also, an arbitrary change in Gibbs energy (ΔG a°) was calculated for each temperature using the Ka value. The graphs of ln Kavs. 1 / T and ΔG a° vs. T were plotted, and the real change in enthalpy (ΔH°) and the real change in entropy (ΔS°) were calculated from the slopes of the obtained straight lines, respectively. Inversely, real ΔG° and K values were calculated using the real ΔH° and ΔS° values in the ΔG° = − RT ln K = ΔH° − TΔS° relation. The best fitting ΔH° and ΔS° values satisfying this relation were found to be 157,301 J mol− 1 and 107.6 J K 1 mol 1, respectively.  相似文献   

6.
A new chiral open-framework indium phosphite In3(HPO3)6(H3DETA)·2H2O 1 has been synthesized under hydrothermal conditions using achiral diethylenetriamine (DETA) as template and characterized by single crystal X-ray diffraction, powder X-ray diffraction, IR spectroscopy, TGA, ICP-AES and elemental analyses. Single crystal X-ray diffraction analysis reveals that the 3D inorganic framework is built up by alternation of InO6 octahedra and HPO3 pseudo-pyramids. Interestingly, the right-handed inorganic helical ribbons are presented in the 3D open-framework while fascinating hydrogen-bonded helices are self-assembled under hydrothermal conditions between organic template molecules and H2O molecules. It is worth noting that the guest hydrogen-bonded helical chain and the host helical ribbon of inorganic framework show the same chirality and intertwine with each other along the b axis. Further understanding of the hydrogen bonds between guest molecules and inorganic framework demonstrates that the flexible DETA molecule may conduce to the formation of the chiral framework. Crystal data: In3(HPO3)6·(H3DETA)·2H2O, Monoclinic, P21 (No. 4), a = 9.2713(19) Å, b = 14.453(3) Å, c = 9.996(2) Å, β = 116.16(3)°, V = 1202.2(4) Å3, and Z = 2, Flack parameter: 0.0(4).  相似文献   

7.
Well-known reversibility of platinum voltammograms in supporting aqueous electrolytes formally contradicts the equilibrium defined by the ionic product of water: Kw = [H+]·[OH] ≅ 1 × 10−14, since at any pH value, concentration one of these ions in solution should be too low (≤10−7 M) to ensure reversibility of voltammograms in the whole potential region of investigation at usual potential scan rates. This contradiction could be overcome by allowing dissociative water adsorption. There are no direct experimental proofs of this process on platinum, however all available relevant data obtained by in situ physical methods appear to provide indirect evidences for this hypothesis.  相似文献   

8.
Thin films of carbonate or sulphate green rusts were synthesised from potentiostatic oxidation of solutions containing ferrous species and bicarbonate or sulphate ions at slightly alkaline pHs and ambient temperature. The thin films were characterised by means of electrochemical quartz crystal microbalance, scanning electron microscopy, X-ray diffraction and infrared reflection-absorption spectroscopy. The composition of carbonate or sulphate green rusts was studied through chemical titration, inductively coupled plasma-optical emission spectroscopy (ICP-OES) and gravimetry and is as follows:
[FeII(2R)FeIII2(OH)(4R−2R′+6)(H2O)(2R′−2)]2R′+·[R′CO3,(2R-{3 or 4}R′ + 2)·H2O]2R′− and [FeII(2R)FeIII2(OH)(4R−2R″+6)(H2O)(2R″−2)]2R″+·[R″SO4,(4R − 4R′ + 4)·H2O]2R″−  相似文献   

9.
Ni modified K2CO3/MoS2 catalyst was prepared and the performance of higher alcohol synthesis catalyst was investigated under the conditions: T = 280–340 °C, H2/CO (molar radio) = 2.0, GHSV = 3000 h 1, and P = 10.0 MPa. Compared with conventional K2CO3/MoS2 catalyst, Ni/K2CO3/MoS2 catalyst showed higher activity and higher selectivity to C2+OH. The optimum temperature range was 320–340 °C and the maximum space-time yield (STY) of alcohol 0.30 g/ml h was obtained at 320 °C. The selectivity to hydrocarbons over Ni/K2CO3/MoS2 was higher, however, it was close to that of K2CO3/MoS2 catalyst as the temperature increased. The results indicated that nickel was an efficient promoter to improve the activity and selectivity of K2CO3/MoS2 catalyst.  相似文献   

10.
The combination of sonolysis and photolysis in the presence of hydrogen peroxide (H2O2) in a 7-L external-loop airlift sonophotoreactor was used to treat the aqueous solution of p-aminophenol. The central composite design (CCD) and response surface methodology (RSM) were employed to evaluate the interaction effects of the initial H2O2 concentration (x1 = 100–900 mg/L), the ultrasonic power (x2 = 25–65 W), the air flow rate (x3 = 1–5 L/min), and the initial concentration of p-aminophenol (x4 = 10–50 mg/L) on the p-aminophenol degradation and total organic carbon (TOC) reduction efficiencies as well as to optimize operating conditions. The coefficients of determination (R2) and adjusted-R2 obtained from the analysis of variance (ANOVA) were 0.9900 and 0.9812 for the p-aminophenol degradation; and 0.9742 and 0.9516 for the TOC removal, respectively, ensuring a satisfactory adjustment of the quadratic regression model with experimental results. The linear, square, and interaction effects of x1, x2, x3, and x4 were also calculated. Genetic algorithm optimization was employed to maximize the mineralization efficiency. 79% TOC reduction efficiency after 90 min and 86.5% p-aminophenol removal efficiency after 30 min were achieved under recirculating batch mode at operating conditions of x1 = 740 mg/L, x2 = 65 W, x3 = 5 L/min, and x4 = 24 mg/L.  相似文献   

11.
The effects of sintering temperature and the addition of CuO on the microstructure and piezoelectric properties of 0.95(K0.5Na0.5)NbO3-0.05Li(Nb0.5Sb0.5)O3 were investigated. The KNN-5LNS ceramics doped with CuO were well sintered even at 940 °C. A small amount of Cu2+ was incorporated into the KNN-5LNS matrix ceramics and XRD patterns suggested that the Cu2+ ion could enter the A or B site of the perovskite unit cell and replace the Nb5+ or Li+ simultaneously. The study also showed that the introduction of CuO effectively reduced the sintering temperature and improved the electrical properties of KNN-5LNS. The high piezoelectric properties of d33 = 263 pC/N, kp = 0.42, Qm = 143 and tan δ = 0.024 were obtained from the 0.4 mol% CuO doped KNN-5LNS ceramics sintered at 980 °C for 2 h.  相似文献   

12.
The standard rate constant of a simple electrode reaction Ox + ne ↔ Red, in which both Ox and Red are solution soluble, can be determined by the variation of frequency in the square-wave voltammetry with inverted scan direction: log ks = log f01/2 + log D1/2 − 0.60 ± 0.01. In this equation log f0 is the abscissa of the intersection of straight lines Ep,2 = a log f + b and Ep,2 = E0, where Ep,2 is the potential of the second peak of the net response, E0 is the standard potential, a = 2.3RT/2(1 − α)nF, b = E0 − 2a log ks + a log D − 0.0353/(1 − α)n and D is a common diffusion coefficient.  相似文献   

13.
The autoprotolysis constant KHS of formic acid/water mixtures as solvent has been calculated from acid-base potentiometric titration curves. A correlation of the acidity scale pKHS of each medium versus pure water has been implemented owing to the Strehlow R0(H+) electrochemical redox function. The results show that formic acid/water mixtures are much more dissociated than pure water; such media are sufficiently dissociated to allow electrochemical measures without addition of an electrolyte. It has also been shown that for a same H+ concentration the activity of protons increases with formic acid concentration. For more than 80 wt.% of formic acid the acidity is sufficiently increased to locate the whole acidity scale pKHS in the super acid medium of the generalized acidity scale pHH2O.  相似文献   

14.
In this study, single crystal V3O7·H2O nanobelts were successfully synthesized using a simple hydrothermal route, in which templates or catalysts were absent. The synthesized V3O7·H2O nanobelts are highly crystalline and have lengths up to several tens of micrometers. The width and thickness of the nanobelts are found to be about 30-50 and 30 nm, respectively. A lithium battery using V3O7·H2O nanobelts as the positive electrode exhibits a high initial discharge capacity of 409 mAh g−1, corresponding to the formation of LixV3O7·H2O (x = 4.32). Such a high degree of electrochemical performance is attributed to the intrinsic properties of the single-crystalline V3O7·H2O nanobelts.  相似文献   

15.
Po-Yu Chen 《Electrochimica acta》2005,50(12):2533-2540
The selective extraction of Cs+ and Sr2+ from aqueous solutions by using the ionophores calix[4]arene-bis(tert-octylbenzo-crown-6) (BOBCalixC6) and dicyclohexano-18-crown-6 (DCH18C6), respectively, was demonstrated in the hydrophobic, room-temperature ionic liquid (RTIL), tri-1-butylmethylammonium bis((trifluoromethyl)sulfonyl)imide (Tf2N). The electrochemistry of Cs+ coordinated by BOBCalixC6 and Sr2+ coordinated by DCH18C6 was examined at a mercury film electrode (MFE) in this ionic liquid by using cyclic staircase voltammetry, sampled current voltammetry at a rotating electrode, and chronoamperometry. Both BOBCalixC6·2Cs+ and DCH18C6·Sr2+ exhibit well-defined reduction waves at approximately −2.4 and −2.9 V versus the ferrocene/ferrocenium (Fc/Fc+) couple, respectively, in which the coordinated ions are reduced to their respective amalgams, permitting the recycling of the ionophores. The diffusion coefficients of BOBCalixC6·2Cs+ and DCH18C6·Sr2+ are (2.7 ± 0.1) × 10−9 and (2.1 ± 0.1) × 10−9 cm2 s−1, respectively, at 30 °C. The coulometric efficiency for the reduction and stripping of Cs at mercury pool electrodes was about 90% and was independent of the deposition time, whereas the efficiency for Sr was slightly less than 90% at short times and decreased with the deposition time, probably due to the formation of a passive layer of Sr(Tf2N)2.  相似文献   

16.
The kinetic parameters of Hg22+, 1·0 M HClO4/Hg and Fe(Ox)33−, Fe(Ox)34−, 0·5 M K2 (Ox), 0·7 N H2SO4/Hg reactions have been obtained by faradaic rectification at radio frequencies. The value of Ia°, the exchange current density for discharge of 1·0 mM and 0·5 mM of mercurous ions on mercury in 1·0 M perchloric acid is 3·9 A/cm2 and 1·1 A/cm2 respectively. The rate constant ka° for the reaction is of the order of 0·9 × 10−2 cm/s at 35°C and the transfer coefficient is 0·56. The values of kinetic parameters obtained for the discharge of mercurous ions have been compared with data obtained by others using different methods. The value of α for Fe(Ox)33−/Fe(Ox)24−, 0·5 M K2(Ox)/Hg reaction is 1·0.  相似文献   

17.
Biodiesel defined as mono-alkyl esters of vegetable oils and animal fats, has had a considerable development and great acceptance as an alternative fuel for diesel engines. Density and viscosity are two important physical properties to affect the utilization of biodiesel as fuel. In this work, mixtures of biodiesel and ultra low sulfur diesel (ULSD) were used to study the variation of density (ρ) and kinematic viscosity (η) as a function of percent volume (V) and temperature (T), experimental measurements were carried out for six biodiesel blends at nine temperatures in the range of 293.15-373.15 K. Both, density and viscosity increases because of the increase in the concentration of biodiesel in the blend, and both of them decrease as temperature increases. One empirical correlation was proposed to estimate the density: ρ = α·V + β·T + δ; and three empirical correlations were developed to predict the kinematic viscosity: η = exp[ln(γ) + ?·V + ω/T + λ·V/T2], η = exp[ln(γ) + ω/T + λ·V/T2] and η = exp[ln(γ) + ω/T + λ·V/T]. The corresponding parameters were optimized by the Levenberg-Marquardt method. The estimated values of density and viscosity are in good agreement with the experimental data because absolute average prediction errors of 0.02% and 2.10% were obtained in the Biodiesel(1) + ULSD(2) system studied in this work.  相似文献   

18.
The electrogenerated chemiluminescence (ECL) of Ru(bpy)32+ (bpy = 2,2′-bipyridyl) with tertiary aliphatic amines as co-reactants, was theoretically and experimentally studied as a function of the pre-equilibria involved in the ammonium proton lost and in relation to the nature of the rate determining step. Transient potential steps were used with a 3-mm glassy carbon disk electrode or carbon fiber ultramicroelectrodes array to investigate emission behavior in a variety of aqueous solution types, containing phosphate, tartrate and phthalate acid-base systems at differing pH values. The emission of Ru(bpy)32+ resulting from the reaction with n-tripropylamine (TPrA), tri-isobutylamine (TisoBuA), n-tributylamine (TBuA), methyl-di-n-propylamine (MeDPrA) and triethylamine (TEtA) in varying acid-base media was interpreted on the basis of the quoted pre-equilibria, ammonium pKa being known. The nature of the rate determining steps changes depending on pH. Above pH ≈ 5 the amine neutral radical formation is the rate determining step and, is independent of pH with rate constant close to 103 s−1; below pH ≈ 5 the rate determining step becomes the deprotonation of the ammonium ion, operated by different bases present in solution. Different amines in the same acid-base system showed analogous ECL behavior, conditioned by the chosen acid base system. A single amine in different acid-base systems showed different kinetic behaviors, due to the dissociation constants of the chosen buffers. The concentration of the acid-base system also played an important role and influenced emission intensity and shape. ECL emission were simulated by finite difference methods, implementing a previously proposed mechanism by including the relevant pre-equilibria. Simulation may also give estimates of the pKa values of the ammonium ions. An ion pair formation between R3N+ and the mostly charged species present in solution is hypothesized to explain the contradictory experimental results concerning the reaction mechanism of the proton lost of the radical cation.  相似文献   

19.
Members of the solid-solution series Ce1−xSrxPO4−δ (x = 0, 0.01, 0.02) with mixed protonic and electronic transport have been synthesized by a nitrate-decomposition method followed by sintering at 1450 °C. Impedance spectroscopy is employed to estimate the bulk electrical conductivity in wet (∼0.03 atm) and dry atmospheres of O2 and 10%H2:90%N2. Conductivity increases with dopant concentration (x), oxygen partial pressure (pO2) and water vapour partial pressure (pH2O) reaching ∼3.5 × 10−3 S cm−1 at 600 °C for x = 0.02 in wet O2. Activation energies (Ea) for the bulk conductivity of Ce0.98Sr0.02PO4−δ below 650 °C are 0.44 and 0.78 eV for wet oxidising and wet reducing conditions, respectively. A moderate but positive pO2+n power-law dependence (n < 1/10) of conductivity is exhibited in the pO2 range 10−2.5 to 10−1 atm, consistent with mixed ionic and p-type electronic transport. Thermogravimetric analysis indicates that the Sr-doped materials are stable in a CO2 atmosphere in the temperature range 25–1200 °C.  相似文献   

20.
Dieter Heymann 《Carbon》2005,43(11):2235-2242
The mean lifetimes of polyyne C8H2 in hexane were determined at 50, 60, 80, and 100 °C and in methanol at 60 °C. The reactions are second order at all temperatures: ln k2 = 20.5 ± 1.5-10303 ± 520T−1 and the corresponding activation energy is 85.7 ± 6.3 kJ mol−1 (7164 cm−1). Extrapolation suggests that solutions at 1 mM concentration are significantly unstable at room temperature. Quantum chemical calculations show that polyynes CmH2 + CnH2 (m + n = 16) could be products, but these were not detected. Alternatively, C16H2 isomers could form. IR spectra of the solid residues from hexane and methanol solutions were obtained.  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