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 15 毫秒
1.
Nanocrystalline ZnMn2O4 is prepared by a polymer-pyrolysis route and used as a novel anode for lithium ion batteries. XRD and HRTEM studies reveal that the products are highly phase-pure and 30–60 nm in size. Galvanostatic cycling of ZnMn2O4 electrode at 100 mA g−1 (about 0.52 mA cm−2) between 0.01 and 3.0 V up to 50 cycles exhibits almost stable cycling performance between 10 and 50 cycles with only an average capacity fade of 0.20% per cycle and the electrode still maintains a capacity of 569 mAh g−1 after 50 cycles.  相似文献   

2.
EPR studies are carried out on Cr3+ ions doped in d-gluconic acid monohydrate (C6H12O7·H2O) single crystals at 77 K. From the observed EPR spectra, the spin Hamiltonian parameters g, |D| and |E| are measured to be 1.9919, 349 (×10−4) cm−1 and 113 (×10−4) cm−1, respectively. The optical absorption of the crystal is also studied at room temperature. From the observed band positions, the cubic crystal field splitting parameter Dq (2052 cm−1) and the Racah interelectronic repulsion parameter B (653 cm−1) are evaluated. From the correlation of EPR and optical data the nature of bonding of Cr3+ ion with its ligands is discussed.  相似文献   

3.
Herein, we report a facile method for synthesizing MoCo-layered double hydroxide (LDH) nanosheets employing Prussian blue analog (PBA) as the precursor. The introduction of Mo in Co-LDH modulates the electronic structure, increases the number of active sites and electrochemical surface area to improve the hydrogen evolution, oxygen evolution, and overall water splitting activity. As a result, PBA-derived Mo0.25Co0.75-LDH nanosheets demonstrated 10 mA cm?2 current density at only 220 mV and 115 mV overpotentials for OER and HER, respectively. The overall water splitting was attained at 1.52 V cell voltage for 10 mA cm?2 current density.  相似文献   

4.
The kinetics of phenylalanine (phe) oxidation by permanganate has been investigated in absence and presence of cetlytrimethylammonium bromide (CTAB) using conventional spectrophotometric technique. The rate shows first- and fractional-order dependence on [MnO4] and [phe] in presence of CTAB. At lower values of [CTAB] (≤10.0 × 10−4 mol dm−3), the catalytic ability of CTAB aggregates are strong. In contrast, at higher values of [CTAB] (≥10.0 × 10−4 mol dm−3), the inhibitory effect was observed in absence of H2SO4. We find that anions (Br, Cl and NO3) in the form of sodium salts are strong inhibitors for the CTAB catalyzed oxidation. Kinetic and spectrophotometric evidences for the formation of an intermediate complex and an ion-pair complex between phe and MnO4, CTAB and MnO4, respectively, are presented. A mechanism consistent with kinetic results has been discussed. Complex formation constant (Kc) and micellar binding constant (Ks) were calculated at 30 °C and found to be Kc = 319 mol−1 dm−3 and Ks = 1127 mol−1 dm−3, respectively.  相似文献   

5.
The basic copper arsenate mineral strashimirite Cu8(AsO4)4(OH)4·5H2O from two different localities has been studied by Raman spectroscopy and complemented by infrared spectroscopy. Two strashimirite mineral samples were obtained from the Czech (sample A) and Slovak (sample B) Republics. Two Raman bands for sample A are identified at 839 and 856 cm−1 and for sample B at 843 and 891 cm−1 are assigned to the ν1 (AsO43−) symmetric and the ν3 (AsO43−) antisymmetric stretching modes, respectively. The broad band for sample A centred upon 500 cm−1, resolved into component bands at 467, 497, 526 and 554 cm−1 and for sample B at 507 and 560 cm−1 include bands which are attributable to the ν4 (AsO43−) bending mode. In the Raman spectra, two bands (sample A) at 337 and 393 cm−1 and at 343 and 374 cm−1 for sample B are attributed to the ν2 (AsO43−) bending mode. The Raman spectrum of strashimirite sample A shows three resolved bands at 3450, 3488 and 3585 cm−1. The first two bands are attributed to water stretching vibrations whereas the band at 3585 cm−1 to OH stretching vibrations of the hydroxyl units. Two bands (3497 and 3444 cm−1) are observed in the Raman spectrum of B. A comparison is made of the Raman spectrum of strashimirite with the Raman spectra of other selected basic copper arsenates including olivenite, cornwallite, cornubite and clinoclase.  相似文献   

6.
Recombination rate coefficients of protonated and deuterated ions KrH+, KrD+, XeH+ and XeD+ were measured using Flowing Afterglow with Langmuir Probe (FALP). Helium at 1600 Pa and at temperature 250 K was used as a buffer gas in the experiments. Kr, Xe, H2 and D2 were introduced to a flow tube to form the desired ions. Because of small differences in proton affinities of Kr, D2 and H2 mixtures of ions, KrD+/D3+ and KrH+/H3+ are formed in the afterglow plasma, influencing the plasma decay. To obtain a recombination rate coefficient for a particular ion, the dependencies on partial pressures of gases used in the ion formation were measured. The obtained rate coefficients, αKrD+(250 K) = (0.9 ± 0.3) × 10−8 cm3 s−1 and αXeD+(250 K) = (8 ± 2) × 10−8 cm3 s−1 are compared with αKrH+(250 K) = (2.0 ± 0.6) × 10−8 cm3 s−1 and αXeH+(250 K) = (8 ± 2) × 10−8 cm3 s−1.  相似文献   

7.
Degradation of polyoxyethylene chain of non-ionic surfactant (TritonX-100) by chromium(VI) has been studied spectrophotometrically under different experimental conditions. The reaction rate bears a first-order dependence on the [Cr(VI)] under pseudo-first-order conditions, [TritonX-100]  [Cr(VI)] in presence of 1.16 mol dm−3 perchloric acid. The observed rate constant (kobs) was 3.3 × 10−4 to 3.5 × 10−4 s−1 and the half-life (t1/2) was 33–35 min for chromium(VI). The effects of total [TritonX-100] and [H+] on the reaction rate were determined. Reducing nature of non-ionic TritonX-100 surfactant is found to be due to the presence of –OH group in the polyoxyethylene chain. It was observed that monomeric and non-ionic micelles of TritonX-100 were oxidized by chromium(VI). When [TritonX-100] was less than its critical micelle concentration (cmc) the kobs values increased from 0.76 × 10−4 to 1.5 × 10−4 s−1. As the [TritonX-100] was greater than the cmc, the kobs values increases from 2.1 × 10−4 to 8.2 × 10−4 s−1 in presence of constant [HClO4] (1.16 mol dm−3) at 40 °C. A comparison was made of the oxidative degradation rates of TritonX-100 with different metal ion oxidants. The order of the effectiveness of different oxidants was as follows: permanganate > diperiodatoargentate(III) > chromium(VI) > cerium(IV).  相似文献   

8.
Raman spectroscopy complimented with infrared spectroscopy has been used to characterise the antimonate mineral bindheimite Pb2Sb2O6(O,OH). The mineral is characterised by an intense Raman band at 656 cm−1 assigned to SbO stretching vibrations. Other lower intensity bands at 664, 749 and 814 cm−1 are also assigned to stretching vibrations. This observation suggests the non-equivalence of SbO units in the structure. Low intensity Raman bands at 293, 312 and 328 cm−1 are assigned to the OSbO bending vibrations. Infrared bands at 979, 1008, 1037 and 1058 cm−1 may be assigned to δOH deformation modes of SbOH units. Infrared bands at 1603 and 1640 cm−1 are assigned to water bending vibrations, suggesting that water is involved in the bindheimite structure. Broad infrared bands centred upon 3250 cm−1 supports this concept. Thus the true formula of bindheimite is questioned and probably should be written as Pb2Sb2O6(O,OH,H2O).  相似文献   

9.
The spectra and kinetic behavior of solvated electrons (esol) in alkyl ammonium ionic liquids (ILs), i.e. N,N-diethyl-N-methyl-N-(2-methoxyethyl)ammonium bis(trifluoromethanesulfonyl)imide (DEMMA-TFSI), N,N-diethyl-N-methyl-N-(2-methoxyethyl)ammonium tetrafluoroborate (DEMMA-BF4), N,N,N-trimethyl-N-propylammonium bis(trifluoromethanesulfonyl)imide (TMPA-TFSI), N-methyl-N-propylpiperidinium bis(trifluoromethanesulfonyl)imide (PP13-TFSI), N-methyl-N-propylpyrrolidinium bis(trifluoromethanesulfonyl)imide (P13-TFSI), and N-methyl-N-butylpyrrolidinium bis(trifluoromethanesulfonyl)imide (P14-TFSI) were investigated by the pulse radiolysis method. The esol in each of the ammonium ILs has an absorption peak at 1100 nm, with molar absorption coefficients of 1.5–2.3×104 dm3 mol−1 cm−1. The esol decayed by first order with a rate constant of 1.4–6.4×106 s−1. The reaction rate constant of the solvated electron with pyrene (Py) was 1.5–3.5×108 dm3 mol−1 s−1 in the various ILs. These values were about one order of magnitude higher than the diffusion-controlled limits calculated from measured viscosities. The radiolytic yields (G-value) of the esol were 0.8–1.7×10−7 mol J−1. The formation rate constant of esol in DEMMA-TFSI was 3.9×1010 s−1. The dry electron (edry) in DEMMA-TFSI reacts with Py with a rate constant of 7.9×1011 dm3 mol−1 s−1, three orders of magnitude higher than that of the esol reactions. The G-value of the esol in the picosecond time region is 1.2×10−7 mol J−1. The capture of edry by scavengers was found to be very fast in ILs.  相似文献   

10.
TiO2 photocatalytic mineralization of β-naphthol: influence of some inorganic ions, ethanol, and hydrogen peroxide. In this work, the photocatalytic oxidation of β-naphthol in aqueous suspensions of TiO2 was investigated at room temperature, by following the formation of CO2. The disappearance of β-naphthol fits a Langmuir-Hinshelwood kinetic model. The activation energy for the degradation reaction of β-naphthol is estimated at 10.2 kJ/mol. The effects of some additives such as ethanol, H2O2, and inorganic ions (Cl, SO42−, HCO3, NO3, Fe3+, Cu2+, and Cr3+) on the photomineralization of β-naphthol were examined. The inhibition of the anions for this reaction was in the order : NO3 < HCO3 < SO42− < Cl. This can be due to a partial blockage of catalyst active sites by these ions or their reaction with an oxidizing radical such as OH. The most photoactive systems for β-naphthol degradation were found in the presence of ferric ions, while the addition of Cr3+ strongly inhibited the photocatalytic decomposition of β-naphthol.  相似文献   

11.
Combining a temperature variable 22-pole ion trap with a cold effusive beam of neutrals, rate coefficients k(T) have been measured for reactions of CO2+ ions with H, H2 and deuterated analogues. The neutral beam which is cooled in an accommodator to TACC, penetrates the trapped ion cloud with a well-characterized velocity distribution. The temperature of the ions, T22PT, has been set to values between 15 and 300 K. Thermalization is accelerated by using helium buffer gas. For reference, some experiments have been performed with thermal target gas. For this purpose hydrogen is leaked directly into the box surrounding the trap. While collisions of CO2+ with H2 lead exclusively to the protonated product HCO2+, collisions with H atoms form mainly HCO+. The electron transfer channel H+ + CO2 could not be detected (<20%). Equivalent studies have been performed for deuterium. The rate coefficients for reactions with atoms are rather small. Within our relative errors of less than 15%, they do not depend on the temperature of the CO2+ ions nor on the velocity of the atoms (k(T) lays between 4.5 and 4.7 × 10−10 cm3 s−1 with H as target, and 2.2 × 10−10 cm3 s−1 with D). For collisions with molecules, the reactivity increases significantly with falling temperature, reaching the Langevin values at 15 K. These results are reported as k = α (T/300 K)β with α = 9.5 × 10−10 cm3 s−1 and β = −0.15 for H2 and α = 4.9 × 10−10 cm3 s−1 and β = −0.30 for D2.  相似文献   

12.
This work proposes a new procedure for on-line electro-oxidative leaching and spectrophotometric determination of uranium in ore samples. By associating a conventional flow injection system, used for uranium determination with Arsenazo III, with an on-line system for electro-oxidative leaching, a fully integrated system was assembled. The systems were integrated after achieving optimum conditions for uranium determination and leaching. According to the results obtained in the present work, a current density of 280 mA cm−2 generated enough hypochlorite ions in the electrolyte solution (3.6 mol L−1 HCl + 2% (w/v) NaCl) to promote quantitative oxidation of U(IV) to U(VI) thus improving the extraction efficiency. The slurry density did not significantly affect the performance of the system and the increasing temperature resulted in a decrease in extraction efficiency. This methodology was applied in the determination of U3O8 in four ore samples and the results obtained agreed with those obtained by ICP-MS after conventional wet acid digestion of the samples.  相似文献   

13.
Room temperature rate coefficients and product distributions are reported for the reactions initiated in D2O with dications of the alkaline-earth metals Mg, Ca, Sr and Ba. The measurements were performed with a selected-ion flow tube (SIFT) tandem mass spectrometer and electrospray ionization (ESI). Mg2+ reacts with water by a fast electron transfer leading to charge separation with a rate coefficient of 1.4 × 10−9 cm3 molecule−1 s−1. Ca2+ reacts with D2O in a first step to form the adduct Ca2+(D2O), with an effective bimolecular rate coefficient of 2.3 × 10−11 cm3 molecule−1 s−1, which then undergoes rapid charge separation by deuteron transfer to form CaOD+ and D3O+ in a second step with k = 7.9 × 10−10 cm3 molecule−1 s−1. The CaOD+ ion reacts further by clustering up to five more D2O molecules. Sr2+ clusters up to eight D2O molecules and Ba2+ up to seven D2O molecules, with the first addition of D2O being rate determining in each case and the last addition being distinctly slower, as might be expected from a transition in the occupation of the added water molecules from an inner to an outer hydration shell.  相似文献   

14.
In this study, H2Ti3O7 nanowires were successfully synthesized via a hydrothermal process and post-treatments. The diameter of the nanowires is found to be about 30 nm and the length up to several micrometers. A lithium battery using H2Ti3O7 nanowires as the active material of the positive electrode exhibits a discharge capacity of 100 mA hg−1 and still keeps stable after 200 cycles at a current density as high as 40 Ag−1, demonstrating excellent high rate performance.  相似文献   

15.
In the present work, blends of poly(ethylene oxide) (PEO), poly(acrylonitrile-co-methyl acrylate) (PANMA) and poly(4-vinylphenol-co-2-hydroxyethyl methacrylate) (PVPh-HEM) were studied by DSC, FTIR and electrochemical impedance spectroscopy (EIS). PEO/PANMA blends were found to be immiscible, while PEO/PVPh-HEM blends are miscible and PVPh-HEM/PANMA exhibits partial miscibility behaviour. The ternary PEO/PANMA/PVPh-HEM blends exhibited miscible compositions for PVPh-HEM and PEO-rich systems. The miscibility observed is a direct consequence of the hydrogen bond interactions among the polymer chains, in which the phenol groups in PVPh-HEM interact with both PEO and PANMA chains. The proton conductivity of a selected membrane based on the ternary blend containing 60% PEO and doped with H3PO4 aqueous solution reached 8 × 10−3 Ω−1 cm−1 at room temperature and 3 × 10−2 Ω−1 cm−1 at 80 °C.  相似文献   

16.
The coordination of nitric oxide (NO) to cobalt(II) phthalocyanine (CoPc) in dimethyl sulphoxide (DMSO) has been studied. CoPc coordinates with NO in a 1:1 ratio, forming a CoPc(NO) species. The IR band observed at 1680 cm−1 is assigned to the coordinated NO. In the presence of excess NO, pseudo first order kinetics were followed. The observed rate constant, kf, was determined to be 15.0±0.3 dm−3 mol−1 s−1 and the equilibrium constant was K=5.4±0.4×104dm3 mol−1. Solution or adsorbed CoPc catalyses the reduction of NO. The products of reduction include NH3 and NH2OH.  相似文献   

17.
A new type of polyphenylene, ionic liquid (IL) 1,3-methylimidazolium hexafluorophosphate substituted, has been prepared by electrodeposition on Au electrode surface via pulse galvanostatic method in 1-butyl-3-methylimidazolium hexafluorophosphate solution. The obtained polymer film had a spherulitic morphology with smallest grains of around 500 nm. Infrared spectrometry revealed that polyphenylene was deposited to a certain extent. The capacitive behavior of the IL substituted polyphenylene was investigated by cyclic voltammetry (CV) and galvanostatic charge–discharge method in 0.2 mol L−1 H2SO4 aqueous solutions or pure IL [bmim]PF6. The specific capacitance of the polymer at the charge–discharge current density of 1 mA cm−2 equaled 206 F g−1 in acidic aqueous solution or 164 F g−1 in [bmim]PF6. Additionally, excellent charge–discharge cycle stability (over 85% value of specific capacitance remained after 600 charge–discharge cycles) and power characteristics of the polymer electrode were observed in both electrolytes.  相似文献   

18.
Recombination of HCO+ and DCO+ ions with electrons was studied in afterglow plasma. The flowing afterglow with Langmuir probe (FALP) apparatus was used to measure the recombination rate coefficients and their temperature dependencies in the range 150–270 K. To obtain a recombination rate coefficient for a particular ion, the dependencies on partial pressures of gases used in the ion formation were measured. The variations of αHCO+(T) and αDCO+(T) seem to obey the power law: αHCO+(T) = (2.0 ± 0.6) × 10−7 (T/300)−1.3 cm3 s−1 and αDCO+(T) = (1.7 ± 0.5) × 10−7 (T/300)−1.1 cm3 s−1 over the studied temperature range.  相似文献   

19.
A series of concentrated aqueous solutions of ferric chloride with different chloride:iron(III) ratios has been studied by means of EXAFS to determine the structure around the iron(III) ion of the dominating species in such solutions. The dominating species in dilute acidic aqueous solution of ferric chloride, at less than 1 mmol·dm?3, are the hydrated iron(III) and chloride ions, while in concentrated aqueous solution and in solutions with an excess of chloride ions, up to 1.0 mol·dm?3, it is the trans-[FeCl2(H2O)4]+ complex. Possible higher chloroferrate(III) or dimeric [Fe2Cl6] complexes at room temperature, as proposed in the literature, were not observed in any of the studied solutions in spite of an excess of chloride ions of 1 mol·dm?3.  相似文献   

20.
The flow injection (FI) method is used for the first time in a partially-closed environment (PCE) to study the electrodissolution of iron in 0.5 mol dm−3 H2SO4 solution. With the precise control of the amount of the solution injected, the physicochemical microenvironment at the electrode/electrolyte interface is manipulated, thus the mechanism of anodic dissolution is examined in detail. At the beginning stage of the passive region, when 0.5 mol dm−3 H2SO4 solution or distilled water is injected quantitatively into the PCE periodically, current oscillations are induced. Unlike the spontaneous oscillations, the periods of the artificial oscillations can be changed as designed. The local perturbation caused by the FI offers valuable information about the onset of the current oscillations. This method will have some potential applications in the study of the anodic dissolution mechanism of the metallic materials.  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