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.
Electrochemical behaviour of Ti(IV) species in the ionic liquid (IL) 1-butyl-2,3-dimethylimidazolium tetrafluoroborate (BMMImBF4) was studied by means of chronopotentiometry (CP) and cyclic voltammetry (CV) in melts with different concentrations of TiF4 (2-35 mol%) within a temperature range of 65-180 °C. The electrochemical reduction of Ti(IV) was suggested to proceed via the sequence of one-electron steps with the formation of poorly soluble low valence intermediates (LVI). LVIs undergo further solid-state electrochemical reduction to Ti metal. Thin Ti coatings on a Pt substrate were thus obtained and characterized by ESEM method. FTIR spectroscopy was used for identification of the electrochemical active species of Ti(IV). The reaction 2BF4 + TiF4 ⇔ TiF62− + 2BF3↑ takes place in the concentrated solutions of TiF4 at elevated temperatures.  相似文献   

2.
Reduction process of uranium(IV) and uranium(III) in molten fluorides   总被引:1,自引:0,他引:1  
This study focused on the electroreduction process of uranium cations in molten fluorides. It involved cyclic voltammetry, chronopotentiometry with and without current reversal, and square wave voltammetry.The results indicate a two-step reduction process for uranium(IV). The first step U(IV)/U(III) exchanging one electron corresponds to a soluble/soluble system and is limited by U(IV) diffusion with DU(IV) = 1.25 ± 0.35 × 10−5 cm2 s−1 in LiF-NaF at 720 °C.In order to perform a thorough study of the second step U(III)/U(0) in the reduction process, the melt was chemically reduced in U(III) with U metal as reducing agent. Alternatively to the use of LiF-NaF where U metal is unstable at 720 °C, the chemical reduction of U(IV) in U(III) was performed in a LiF-CaF2-UF4 solution containing U metal at 810 °C. It has been confirmed that the reduction of U(III) proceeds in one step exchanging three electrons and by a diffusion controlled process with DU(III) = 2.2 ± 0.7 × 10−5 cm2 s−1 in LiF-CaF2 at 810 °C.  相似文献   

3.
Co-Sn alloys were prepared by an electrochemical route in molten LiCl-KCl between 400 and 550 °C. The Sn(IV)/Sn(II), Sn(II)/Sn(0) and Co(II)/Co(0) redox couples were studied by cyclic voltammetry and/or chronopotentiometry over the temperature range. The diffusion coefficient values of Co(II) ions were measured. For example, it was found that the DCo(II) values deduced from chronopotentiometry range from DCo(II) = 1.65 × 10−5 cm2 s−1 at 400 °C to 4.95 × 10−5 cm2 s−1 at 550 °C. The standard potential of the Co(II)/Co(0) redox couple in molten LiCl-KCl was measured at 400 °C: vs Cl2/Cl. Finally, Co-Sn alloys were prepared in potentiostatic mode. The influence of the temperature of molten LiCl-KCl, the applied potential and the deposition time on the morphology and the composition of the Co-Sn alloys were also investigated. For T > 450 °C, the following tendency has been observed: the more negative the potential, the higher the Sn content in the deposited alloy. Thus, depending on the operating conditions, pure CoSn or CoSn2 can be prepared.  相似文献   

4.
Electrochemical synthesis of Ni-Sn alloys in molten LiCl-KCl   总被引:2,自引:0,他引:2  
The electrochemical formation of Ni-Sn was investigated in molten LiCl-KCl in the temperature range 380-580 °C. Before, an electrochemical study of the Ni2+/Ni0, Sn4+/Sn2+ and Sn2+/Sn0 redox couples was performed by cyclic voltametry and chronopotentiometry in a wide temperature range. It had been pointed out that in the case of the Sn4+/Sn2+ redox couple, an insoluble compound is probably formed for T < 460 °C. For higher temperature, this compound becomes soluble and then, the shape of the cyclic voltammogram is analogue to the one usually observed when a diffusion-controlled process is involved. The diffusion coefficient values of Ni2+ and Sn2+ ions were determined. For instance, DNi(II) and DSn(II) values deduced from chronopotentiometry were about 2.1 × 10−5 and 2.7 × 10−5 cm2 s−1 at 440 °C, respectively. Then, Ni-Sn alloys have been formed in potentiostatic mode. The electrochemical route proposed in this paper leads to the formation of crystallized alloys with a well-defined composition depending on the operating conditions.  相似文献   

5.
The normal potential of the Ce(IV)/Ce(III) redox couple was determined by square wave voltammetry (SWV) at different temperatures in solutions with a constant ratio [CO32−]/[HCO3] ≈10 for high ionic strengths (3.29 mol dm−3 at 4.39 mol dm−3): varies from 259.5 to 198.0 mV/S.H.E. in the 15-50 °C range. Linear variations were found for versus (RT/F)ln(mCO32−), leading to the stoichiometry, Ce(CO3)68− for the Ce(IV) limiting complex. But the slopes of these linear variations were actually found in the range 1.8-1.9, not exactly 2. This was interpreted as dissociation of the Ce(IV) limiting complex following the reaction: Ce(CO3)56− + CO32− → Ce(CO3)68− and as dissociation of the Ce(III) limiting complex following the reaction: Ce(CO3)33− + CO32− → Ce(CO3)45−; for which maximum possible values of log10 KIV,6 and log10 KIII,4 were estimated via fitting in the 15-50 °C temperature range (log10 KIV,6 = 0.42 (0.97) and log10 KIII,4 = 0.88 (7.00) at 15 °C (50 °C). The normal potential was found to decrease linearly with T, these variations correspond to , with T0 = 298.15 K and . The apparent diffusion coefficient of Ce(IV) was determined by direct current polarography (DCP), cyclic voltammetry (CV) and square wave voltammetry. It was found to depend on the ionic strength and to be proportional to T.  相似文献   

6.
This work demonstrates that the microstructure of TiO2 film can be designed and controlled by adjusting the temperature and cycle number of cathodic deposition in a solution containing TiCl3 and NaNO3. The redox interactions between TiCl3 and NO3 are investigated by in situ ultraviolet-visible (UV-vis) absorption spectroscopy. Linear sweep voltammetry (LSV) is employed to study the NO3 reduction and to clarify the deposition behavior of TiO2 in the designed plating solution. The decrease in TiO2 deposition rate with the TiO2 thickness may be due to the poor electron conductivity of TiO2 depressing the generation rate of OH from the NO3 reduction. The morphology and size of TiO2 aggregates are strongly influenced by varying the deposition temperature from 5 to 50 °C and a maximal rate of TiO2 deposition is obtained at 25-35 °C. TiO2 deposited at 25 °C is the roughest with a roughness factor (Ra) of ca. 67 nm. This study provides a useful method to control the morphology and deposition rate of TiO2 film for practical photoelectrochemical applications.  相似文献   

7.
Titanium was electrodeposited from a nominal 0.24 M TiCl4 in l-methyl-3-butyl-imidazolium bis (trifluoro methyl sulfone) imide ([BMIm]BTA) at room temperature on a Au(1 1 1) substrate. The process of electrodeposition was studied by cyclic voltammetry, chrono amperometry and in situ scanning tunneling microscopy (STM). In a first step TiCl4 is reacted to TiCl2, which is subsequently reduced to metallic Ti. Two dimensional (2D) clusters form preferentially on the terraces in the under potential deposition range. 2D clusters presumably of TiCl3 precipitates grow and coalesce to cover the whole substrate with a 2D film at a substrate potential below −1.1 V versus ferricenium/ferrocene ([Fc]+/[Fc]) redox couple. At a potential of −1.8 V a dense layer of three dimensional (3D) clusters of titanium of 1-2 nm thickness is formed. The features of the I-U tunneling spectra and the relative reduction of the effective tunneling barrier by 0.8 eV with respect to gold clearly indicate the metallic character of Ti deposits. Observation of circular holes on the Au(1 1 1) substrate after dissolution of the deposited Ti indicates the formation of Au-Ti surface alloying.  相似文献   

8.
The electrochemical behaviour of the Nd(III)/Nd(0) system has been investigated in several molten media and more particularly in LiF-CaF2. A preliminary study based both on thermodynamic and experimental data showed that it is not possible to observe the Nd(III)/Nd(0) system in LiF-KF and LiF-NaF melts; because the K+ and Na+ cation reduction waves hide the Nd3+ reduction wave. Then, the Nd(III)/Nd(0) system has been investigated at 810 °C using solutions of NdF3 in fluoride solvents without K+ and Na+ ions, such as LiF-CaF2, by cyclic voltammetry, chronopotentiometry and square wave voltammetry. Experimental results show that neodymium trifluoride is reduced in Nd(0) in a one-step process exchanging three electrons (Nd(III) + 3e → Nd(0)). The electrode process is shown to be diffusion controlled. Nd(III) diffusion coefficient is in the range of 1.1-1.3 × 10−5 cm2 s−1 at 810 °C.  相似文献   

9.
FTIR spectroscopy was used to identify the oxochloride species of tantalum(V) in ionic liquids and to confirm the correlations between their presence in electrolytes and the changes in the route of electrochemical reduction of tantalum(V). Electrochemical behaviour of the mixtures (x)1-butyl-1-methyl-pyrrolidinium chloride-(1 − x)TaCl5 at x = 0.80, 0.65, and 0.40 was investigated over the temperature range 90-160 °C with respect to the electrochemical deposition of tantalum and was discussed in terms of spectroscopic data. The mechanism of electrochemical reduction of tantalum(V) in the basic and acidic electrolytes depends strongly on the structure and composition of the electro active species of tantalum(V) defined by the molar composition of ionic liquids and on the competition between tantalum(V) chloride and oxochloride species. In the basic mixture at x = 0.80, with octahedral [TaCl6] ions as the electrochemically active species only the first reduction step Ta5+ → Ta4+ at −0.31 V was observed. The competitive reduction of tantalum(V) oxochloride species occurs at more anodic potential (−0.01 V) than the reduction of the chloride complexes and can restrict the further reduction of tantalum(IV). In the basic ionic liquid at x = 0.65, the cyclic voltammograms exhibit reduction peaks at −0.31 V and −0.51 V attributed to the diffusion controlled process as [TaCl6] + e → [TaCl6]2− and [TaCl6]2− + e → [TaCl6]3−. The further irreversible reduction of tantalum(III) to metallic state may occur at −2.1 V. In the acidic ionic liquids, at x = 0.40 the electrochemical reduction of two species occurs, TaCl6 and Ta2Cl11 and it is limited by two electron transfer for both of them at −0.3 V and −1.5 V, respectively.  相似文献   

10.
Lieyu Hu 《Polymer》2006,47(6):2099-2105
Branched poly(N-allylethylenimine) (BPAEI), a solid state polymer electrolyte host, was synthesized by allylation of branched poly(ethylenimine) (BPEI). Allylation was essentially complete with the 2 and 1° nitrogen atoms of BPEI being mono-allylated and di-allylated, respectively, and with little or no quaternization. BPAEI can be radically cross-linked with and without lithium trifluoromethanesulfonate (LiTf) present to form free-standing, homogeneous, minimally hygroscopic films. BPAEI has a glass transition temperature (Tg) of −65 °C, as measured by differential scanning calorimetry (DSC), which increases with the concentration of initiator upon cross-linking using V-50 (2,2-azobis(2-amidino-propane) dihydrochloride) to −15 °C at a 10:1 nitrogen to initiator molar ratio (N:initiator). BPAEI with 20:1 N:Li+ (molar ratio) LiTf has a Tg of −48 °C, which increases with the concentration of radical initiator upon cross-linking using V-50 to 3 °C at 10:1 N:initiator. At compositions near 60:1 N:initiator, an unusual decrease in the rate at which Tg changes with cross-linking was observed, both with and without LiTf present, indicating that some undefined morphological changes occur. The effect of this morphological change resulted in the highest Ac conductivities at 60:1 N:initiator for all LiTf concentrations studied. At 20:1 N:Li+ LiTf and 60:1 N:initiator, the room temperature Ac conductivity was 1×10−8 S/cm which increased to 1×10−5 S/cm at 80 °C, the highest conductivity observed in the concentration ranges studied. Infrared spectroscopy (IR) showed that the concentrations of the individual ionic species present were largely independent of either LiTf concentration or cross-linking density, suggesting that changes in ion mobility, likely resulting from morphological changes, substantially control the ionic conductivity.  相似文献   

11.
Perovskite solid solutions of (La0.6Sr0.4)(X1−yMgy)O3−δ (X = Ti, Zr, Al) were prepared by a coprecipitation method using corresponding aqueous solutions and ammonium carbonate solution. The freeze-dried powders were sintered in air at 1000-1500 °C for 1-36 h. Single phase solid solutions were produced in the compositions of (La0.6Sr0.4)(Zr0.6Mg0.4)O3−δ and (La0.6Sr0.4)(Al0.9Mg0.1)O3−δ where (3 − δ) < 3. For the compositions of X = Ti and Zr for y = 0.1 where (3 − δ) > 3, two phases including perovskite solid solution were produced at 1400-1500 °C. The stability of perovskite solid solution was closely related to the fraction of lattice oxygen atom (3 − δ). A relatively high conductivity was measured for (La0.6Sr0.4)(Al0.9Mg0.1)O3−δ (σ = 4.15 × 10−4 S/cm at 600 °C, activation energy 113.4 kJ/mol). The influence of fraction of oxide ion vacancy on the activation energy was small for δ = 0.1-0.3 of perovskite solid solution.  相似文献   

12.
An anhydrous proton conductor, Sn0.95Al0.05P2O7 (SAPO), composed of polystyrene-b-poly(ethylene/propylene)-b-polystyrene (SEPS), was developed and characterized using morphological, structural, and electrochemical analyses. In the composite membrane with 20 wt% SEPS, a homogeneous distribution of SAPO particles in the matrix was obtained in the thickness range of 65-90 μm, yielding a proton conductivity of 3.4 × 10−3 S cm−1 at 200 °C, tensile strength of 4.6 MPa and an elongation at break of 711.0% at room temperature. Fuel cell tests verified that the open-circuit voltage was maintained at a constant value of approximately 1 V between 100 and 250 °C. The peak power densities achieved with unhumidified H2 and air were 77.0 mW cm−2 at 100 °C, 121.0 mW cm−2 at 150 °C, and 163.1 mW cm−2 at 225 °C.  相似文献   

13.
Micro- and mesoporous carbide-derived carbon (CDC) was synthesised from molybdenum carbide (Mo2C) powder by gas phase chlorination in the temperature range from 400 to 1200 °C. Analysis of XRD results show that C(Mo2C), chlorinated at 1200 °C, consist mainly on graphitic crystallites of mean size, La = 9 nm and Lc = 7.5 nm. The first-order Raman spectra showed the graphite-like absorption peak at ∼1587 cm−1 and the disorder-induced (D) peak at ∼1348 cm−1. The low-temperature N2 adsorption experiments were performed and a specific surface area up to 1855 m2 g−1 and total pore volume up to 1.399 cm3 g−1 were obtained. Sorption measurements showed the presence of both micro- and mesopores after chlorination at 400-900 °C and only mesopores after chlorination at 1000°-1200 °C. Stepwise formation of micro- and mesopores was achieved and the peak pore size can be shifted from 0.8 nm up to 4 nm by increasing the chlorination temperature.  相似文献   

14.
Ti-SBA-15C catalysts were prepared by supporting Ti on SBA-15 with chemical vapor deposition (CVD) using TiCl4 as titanium source, and were characterized by XRD, N2 adsorption, FT-IR and ICP. The results show that SBA-15E prepared by ethanol solution extracting template has higher concentration of surface Si-OH groups than SBA-15 calcined at 550 °C, resulting in high Ti content on Ti-SBA-15EC prepared by CVD. The temperature and time of TiCl4 deposition affect the Ti content and catalytic activity of Ti-SBA-15EC for the epoxidation of propylene with cumene hydroperoxide (CHP). Ti-SBA-15EC prepared by CVD at 700 °C for 1.5 h exhibits more excellent performance than Ti-SBA-15C, Ti-SBA-15 prepared hydrothermally and Ti/SBA-15 (impregnation method), and the 87.3% conversion of CHP and 96.4% selectivity to propylene oxide can be obtained at 80 °C for 4 h. The performance of Ti-SBA-15EC is decreased hardly for the epoxidation of propylene after used repeatedly 6 times.  相似文献   

15.
An alumina precursor was prepared by the aluminium sulphate (0.20 M) and excess urea reaction in boiling aqueous solution. The precursor was calcined at 900 °C for 2 h and then δ-Al2O3 powder having volumetric agglomeration degree of 80% was obtained. Cylindrical compacts having diameter of 14 mm were prepared under 32 MPa by axial pressing using oleic acid as binder. Each compact was fired isothermally at various temperatures between 950 and 1400 °C. The firing time was changed from zero to 2 h. The fired compacts were examined by scanning electron microscopy (SEM) and nitrogen adsorption techniques. The specific surface areas (S/m2 g− 1) of the samples were calculated using the Brunauer, Emmett, and Teller (BET) procedure. The rate constant (k) and mechanism-characteristic parameter (n) were obtained for different temperatures between 950 °C and 1150 °C from the application of the neck-growth sintering rate (NGRM) model on the surface area reduction data. An Arrhenius equation and the parameter n for the sintering were found in the forms of k = (7.648 × 106 h− 1) exp (− 186,234 J mol− 1 / RT) and n = 4.0 × 10− 7 T3-1.7 × 10− 3 T2 + 2.3 T − 1030.8 respectively. The parameter n changes in the interval 0.61 <  n < 1.34 with rising temperature having maximum at about 1025 °C. Based on the SEM images and NGRM data, the intra-particle sintering was discussed.  相似文献   

16.
A series of novel fluorinated poly(aryl ether)s containing phthalazinone moieties (FPPEs) have been prepared by a modified synthetic procedure for optical waveguide applications. The obtained random copolymers exhibited excellent solubility in polar organic solvents, high glass transition temperatures (Tgs: 185-269 °C), good thermal stabilities (the temperatures of 1% weight loss: 487-510 °C) and good optical properties. By adjusting the feed ratio of the reactants, the refractive indices of TE and TM modes (at 1550 nm) could be well controlled in the range of 1.575-1.498 and 1.552-1.484, respectively. The optical losses of the FPPEs exhibited relatively low values (less than 0.27 dB/cm at 1310 nm). Additionally, the thermo-optic coefficient (dn/dT) values of the FPPEs at 1310 nm and 1550 nm (TE mode) ranged from −0.97 × 10−4 °C to −1.33 × 10−4 °C and from −0.96 × 10−4 °C to −1.29 × 10−4 °C, respectively.  相似文献   

17.
7Li and 19F NMR linewidths and impedance spectra are reported for low-dimensional CmOn (I):LiBF4 mixtures. Data for the ionophilic polymer C18O5 is compared with that for the ionophobic C18O1 and the block copolymer C16O1O5(21%) (21 mol% of C16O5). In C18O5:LiBF4 (1:1) narrow 7Li linewidths, which were observed in the liquid crystal phase above the side chain melting temperature (∼50 °C), persist in the crystal down to ca. 0 °C and broaden below −20 °C. However, in C18O1:LiBF4 (1:0.6) narrow 7Li linewidths were also observed down to −20 °C suggesting highly mobile neutral aggregates of salt since this system is non-conductive. In the copolymer C16O1O5(21%):LiBF4 (1:0.7) the linewidths were even narrower down to −70 °C with weak temperature dependence. In all systems 19F linewidths were significantly broader than 7Li linewidths. The complex plane plots obtained by impedance spectroscopy exhibit characteristic minima identified with ‘grain boundary’ resistance and, following heat treatment, minima with weak temperature dependence identified with ‘internal crystal’ resistance, Ri, and conductivities, σi ≥ 10−4 S cm−1. Four-component mixtures of copolymers CmO1O5 and CmO1O4 with LiBF4 and ‘salt-bridge’ poly(tetramethylene oxide)-dodecamethylene copolymers gave conductivities of ca. 4 × 10−4 S cm−1 at 20 °C with weak temperature dependence. A novel carrier-hopping mechanism of lithium transport decoupled from side chain melting in the crystalline state is postulated.   相似文献   

18.
Lead zirconate titanate (PZT) precursor sols were prepared using a triol based sol–gel route. Inorganics salts metal alkoxides lead acetate trihydrate [Pb(OOCCH3)2·3H2O], titanium (IV) isopropoxide [Ti(OCH(CH3)2)4], and zirconium n-propoxide [ZrOC3H7)4] were used as starting materials. Thin films were deposited by spin coating onto Pt/Ti/SiO2/Si substrates. The samples were pre-heated (pyrolysis) on a calibrated hotplate over the temperature range of 200–400 °C for 10 min then firing at a temperature of 600 °C for 30 min. Randomly-oriented PZT thin films pre-heated at 400 °C for 10 min and annealed at 600 °C for 30 min showed well-defined ferroelectric hysteresis loops with a remanent polarization of 27 μC/cm2 and a coercive field of 115 kV/cm. The dielectric constant and dielectric loss of the PZT films were 621 and 0.040, respectively. The microstructures of the thin films are dense, crack-free and homogeneous with fine grains about 15–20 nm in size.  相似文献   

19.
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   

20.
《Electrochimica acta》2002,48(4):331-340
The behavior of a small-scale flow alkaline fuel cell (AFC) built-up for on-site production of HO2 using commercial gas-diffusion electrodes has been studied. It produces a spontaneous current due to the oxidation of H2 to H2O at the H2-diffusion anode and the reduction of O2 to HO2 at the O2-diffusion cathode, while a fresh 1.0-6.0 mol dm−3 KOH electrolyte at 15.0-45.0 °C is injected through it. Under circulation of HO2+KOH solutions in open circuit, the flow AFC behaves as a two-electron reversible system. When it is shorted with an external load (Rext), steady cell voltage-current density curves are found. The use of O2/N2 mixtures to fed the cathode causes a loss of its performance, being required to supply pure O2 to yield a maximum HO2 electrogeneration. The current density and HO2 productivity increase with raising OH concentration, temperature and pressure of O2 fed. At Rext=0.10 Ω, a current efficiency close to 100% is obtained, and current densities >100 mA cm−2 are achieved for 1.0 mol dm−3 KOH at 45.0 °C and for higher KOH concentrations at 25.0 °C. The flow AFC can work under optimum conditions up to 6.0 mol dm−3 KOH and 45.0 °C for possible industrial applications.  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