共查询到20条相似文献,搜索用时 15 毫秒
1.
The electrochemical behaviour of the spinel-like LiMn2O4 was studied in non-aqueous and aqueous saturated alkali nitrate electrolytes in comparison with the layered manganese dioxide
δ-MnO2. The results obtained by galvanostatic and cyclic voltammetry techniques showed that the insertion of Li+/e– or H+/e– depends on both the host lattice and the type of electrolyte. The spinel-like LiMn2O4 preferably allowed the insertion of Li+/e– in non-aqueous and aqueous saturated LiNO3 electrolytes, as observed from the similarity of the electrochemical behaviour in these electrolytes and the stability of
the structure. This was explained by the presence of a three-dimensional network of vacant tetrahedral and half-filled octahedral
sites in LiMn2O4, which guarantee high mobility of Li+ ions. The layered manganese dioxide could be inserted by Li+/e– only in non-aqueous electrolytes.
The work described herein was carried out at the Institut für Anorganische und Analytische Chemie, Technische Universit?t
Berlin, Germany 相似文献
2.
T. L. Kulova A. M. Skundin V. M. Kozhevin D. A. Yavsin S. A. Gurevich 《Russian Journal of Electrochemistry》2010,46(8):877-881
The process of lithium insertion into thin gold films is studied. It is found that lithium is reversibly inserted with the formation of compounds, which are close to Li1.5Au in their average composition. The long-term cycling leads to the breakdown of gold layer (and corresponding decrease in the electrode capacity), which is associated with considerable volume changes due to lithium insertion. 相似文献
3.
4.
Reaction of trans-1,2-dideuterioethene oxide () with carbon dioxide, using copper and nickel catalysts, and subsequent analysis of the product ethene carbonate-d2 () by microwave spectroscopy, shows that the copper-catalyzed reaction is stereo-specific (retention) whereas the nickel-catalyzed reaction is non-stereospecific. 相似文献
5.
《Solid State Sciences》2012,14(9):1251-1255
Chemical manganese dioxide (CMD) is synthesized by the SEDEMA process and adopted as a precursor for lithium manganese oxide with a spinel structure (LMO). LMO is also prepared from electrolytic manganese dioxide (EMD) as a reference for comparison. X-ray diffraction (XRD) shows that CMD is composed of γ-MnO2, and scanning electron microscopy (SEM) with transmission electron microscopy (TEM) shows that the nanorods cover a spherical core with a diameter < 1 μm. The LMO prepared from CMD shows a much better rate capability and cycle life performance than that from EMD at high temperatures and high current densities. The excellent electrochemical performance is attributed to the structural stability during charge and discharge and the morphology of the LMO, a loose aggregation of the octahedral particles with a uniform size (<1 μm) and shape, which originated from that of CMD. 相似文献
6.
7.
8.
Lithium insertion into titanium dioxide (anatase) electrodes: microstructure and electrolyte effects
Dina Fattakhova Ladislav Kavan Petr Krtil 《Journal of Solid State Electrochemistry》2001,5(3):196-204
Insertion characteristics of anatase electrodes were studied on single-crystal and polycrystalline electrodes of different
microstructures. The lithium incorporation from propylene carbonate solution containing LiClO4 and Li(CF3SO2)2N was studied by means of cyclic voltammetry (CV), the quartz crystal microbalance (QCM) and the galvanostatic intermittent
titration technique (GITT). The electrode microstructure affects both the accessible coefficient x and the reversibility of the process. The highest insertion activity was observed for electrodes composed of crystals with
characteristic dimensions of ∼10–8 m. The insertion properties deteriorate for higher as well as for smaller crystal sizes. Enhanced insertion was observed in
Li(CF3SO2)2N-containing solutions. Lithium insertion is satisfactorily reversible for mesoscopic electrodes; the reversibility in the
case of compact polycrystalline and single-crystal electrodes is poor. The reversibility of the insertion improves with increasing
electrolyte concentration. The lithium diffusion coefficient decreases with increasing x and ranges between 10–15 and 10–18 cm2 s–1.
Electronic Publication 相似文献
9.
Reduction of colloidal manganese dioxide by manganese(II) 总被引:1,自引:0,他引:1
Perez-Benito JF 《Journal of colloid and interface science》2002,248(1):130-135
The reduction of colloidal MnO(2) by Mn(2+) in aqueous HClO(4) has been studied by a spectrophotometric method. The reaction product is Mn(III). The reaction is of first order in both colloidal MnO(2) and H(+), whereas it presents a fractional order (0.58+/-0.02) in Mn(2+). The reaction is retarded by addition of NaClO(4), but is not affected by addition of tert-butanol. The corresponding activation energy is 29.5+/-1.3 kJ mol(-1). The reaction is catalyzed by Na(4)P(2)O(7), and the pyrophosphate-catalyzed reaction is of first order in both colloidal MnO(2) and pyrophosphate and of fractional order (0.64+/-0.01) in Mn(2+), whereas its rate presents a complex dependence on the concentration of H(+). The pyrophosphate-catalyzed reaction is accelerated by addition of both NaClO(4) and tert-butanol. The corresponding activation energy is 49.7+/-3.0 kJ mol(-1). Mechanisms in agreement with the experimental data are proposed for both the parent and the pyrophosphate-catalyzed reactions. 相似文献
10.
Norihito Kijima Yasuhiko Takahashi Junji Akimoto Junji Awaka 《Journal of solid state chemistry》2005,178(9):2741-2750
Lithium ion insertion and extraction reactions with a hollandite-type α-MnO2 specimen free from any stabilizing cations in its tunnel cavity were investigated, and the crystal structure of a Li+-inserted α-MnO2 specimen was analyzed by Rietveld refinement and whole-pattern fitting based on the maximum-entropy method (MEM). The pH titration curve of the α-MnO2 specimen displayed a monobasic acid behavior toward Li+, and an ion-exchange capacity of 3.25 meq/g was achieved at pH>11. The Li/Mn molar ratio of the Li+-inserted α-MnO2 specimen showed that about two Li+ ions can be chemically inserted into one unit cell of the hollandite-type structure. As the amount of Li content was increased, the lattice parameter a increased while c hardly changed. On the other hand, the mean oxidation number of Mn decreased slightly regardless of Li content whenever ions were exchanged. The Li+-inserted α-MnO2 specimen reduced topotactically in one phase when it was used as an active cathode material in a liquid organic electrolyte (1:1 EC:DMC, 1 mol/dm3 LiPF6) lithium cell. An initial discharge with a capacity of approximately 230 mAh/g was achieved, and the reaction was reversible, whereas the capacity fell steadily upon cycling. About six Li+ ions could be electrochemically inserted into one unit cell of the hollandite-type structure. By contrast, the parent α-MnO2 specimen showed a poor discharge property although no cationic residues or residual H2O molecules remained in the tunnel space. Rietveld refinement from X-ray powder diffraction data for a Li+-inserted specimen of (Li2O)0.12MnO2 showed it to have the hollandite-type structure (tetragonal; space group I4/m; a=9.993(11) and ; Z=8; Rwp=6.12%, Rp=4.51%, RB=1.41%, and RF=0.79%; S=1.69). The electron-density distribution images in (Li2O)0.12MnO2 showed that Li2O molecules almost fill the tunnel space. These findings suggest that the presence of stabilizing atoms or molecules within the tunnel of a hollandite-type structure is necessary to facilitate the diffusion of Li+ ions during cycling. 相似文献
11.
Shinichi Komaba Naoaki Yabuuchi Sachi Ikemoto 《Journal of solid state chemistry》2010,183(1):234-4419
To study crystallization process of spinel-type Li1+xMn2−xO4, in-situ high-temperature X-ray diffraction technique (HT-XRD) was utilized for the mixture consisting of Li2CO3 and Mn2O3 as starting material in the temperature range of 25-700 °C. In-situ HT-XRD analysis directly revealed that crystallization process of Li1+xMn2−xO4 was significantly affected by the difference in the Li/Mn molar ratio in the precursor. Single phase of stoichiometric LiMn2O4 formed at 700 °C. The formation of single phase of spinel was achieved at the lower temperature than the stoichiometric sample as Li/Mn molar ratio in the precursor increased. Lattice parameter of the stoichiometric LiMn2O4 at 25 °C was 8.24 Å and expanded to 8.31 Å at 700 °C, which corresponds to the approximately 3% expansion in the unit cell volume. From the slope of the lattice parameter change as a function of temperatures, linear thermal expansion coefficient of the stoichiometric LiMn2O4 was calculated to be 1.2×10−5 °C−1 in this temperature range. When the Li/Mn molar ratio in Li1+xMn2−xO4 increased (x > 0.1), the spinel phase segregated into the Li1+yMn2−yO4 (x > y) and Li2MnO3 during heating, which involved the oxygen loss from the materials. During the cooling process from 700 °C, and the segregated phase merged into Li1+xMn2−xO4 with oxygen incorporation. Such trend directly observed by in-situ HT-XRD was supported by thermal gravimetric analysis as reversible weight (oxygen) loss/gain at higher temperature (500-700 °C). 相似文献
12.
Despite the general consensus concerning the role played by sodium channels in the molecular mechanism of local anesthetics, the potency of anaesthetic drugs also seems to be related with their solubility in lipid bilayers. In this respect, this work represents a thermodynamic study of benzocaine insertion into lipid bilayers of different compositions by means of molecular dynamics simulation. Thus, the free energy profiles associated with benzocaine insertion into symmetric lipid bilayers composed of different proportions of dipalmitoylphosphatidylcholine and dipalmitoylphosphatidylserine were studied. From the simulation results, a maximum in the free energy (ΔG) profile was measured in the region of the lipid/solution interface. This free energy barrier appears to be very much dependent on the lipid composition of the membrane. On the other hand, the minimum free energy (ΔG) within the bilayer remained almost independent of the lipid composition of the bilayer. By repeating the study at different temperatures, it was seen how the spontaneity of benzocaine insertion into the lipid bilayer is due to an increase in the entropy associated with the process. 相似文献
13.
Hyun Kyoo Park Hana Rah Dong Jun Kang Uong Chun Sun-Geon Kim 《Journal of Sol-Gel Science and Technology》2013,67(3):464-472
Nanostructures of single crystallites of spinel LiMn2O4 (LMO) were prepared by the simple pyrolysis of aqueous solution of LiNO3 and Mn(NO3)2 in a confined space such as either droplets or mesopores. When the mixed nitrate solution was spray pyrolyzed at temperatures below 700 °C, 1-μm LMO spheres were obtained consisting of ~20-nm single crystallites randomly packed. Such LMO phase, once obtained, would sustain for further heat-treatment. Next, new spraying solution was prepared by adding the precursor for mesoporous silica (MPS) to the nitrates solution. By spray pyrolyzing such solution, LMO was impregnated inside pores of the MPS being structured. The silica could be removed by subsequent NaOH treatment to leave spherical LMO mesophase. The nitrates was also able to soak into the existing MPS having cylindrical pores and form short isolated LMO chains in the mesopores by the subsequent heating. After the same NaOH treatment, the LMO phase turned into bundles of very ‘long’, and often straight, chains, consisting of 8-nm LMO nanoparticles. This will be elucidated through further study. 相似文献
14.
Stress generation and fracture in lithium insertion materials 总被引:1,自引:0,他引:1
A mathematical model that calculates volume expansion and contraction and concentration and stress profiles during lithium
insertion into and extraction from a spherical particle of electrode material has been developed. The maximum stress in the
particle has been determined as a function of dimensionless current, which includes the charge rate, particle size, and diffusion
coefficient. The effects of pressure-driven diffusion and nonideal interactions between the lithium and host material have
also been described. The model predicts that carbonaceous particles will fracture in high-power applications such as hybrid-electric
vehicle batteries.
相似文献
John ChristensenEmail: |
15.
Kinetic data for the colloidal MnO2–thiourea redox system are reported for the first time. The reduction of water-soluble colloidal MnO2 by thiourea (sulfur containing reductant) in aqueous perchloric acid medium has shown that it proceeds in two stages, i.e., a fast stage followed by a relatively slow second stage. The log (absorbance) versus time plot deviates from linearity. The kinetics of both the stages was investigated spectrophotometrically. The first-order kinetics with respect to [thiourea] at low concentration shifts to zero-order at higher concentration. The reaction rate increases with [HClO4] and the kinetics reveals complex order dependence in [HClO4]. Addition of P2O
7
4−
and F− in the form of Na4P2O7 and NaF, respectively, has inhibitory effect on the reaction rate. The reaction proceeds through the fast adsorption of thiourea on the surface of the colloidal MnO2. A mechanism involving the protonated thiourea as the reactive reductant species is proposed. The observed results are discussed in terms of Michaelis–Menten/Langmuir–Hinshelwood model. From the observed kinetic data, colloidal MnO2–thiourea adsorption constant (K
ad1) and rate constant (k
1) were calculated to be 1.25×1010 mol−1 dm3 and 3.1×10−4 s−1, respectively. The variation of temperature does not have any effect on the reaction rate. 相似文献
16.
V. G. Koshechko V. E. Titov V. A. Lopushanskaya V. D. Pokhodenko 《Theoretical and Experimental Chemistry》1998,34(1):1-4
The effect of the electronic structure of the reagents, the nature of the supporting electrolyte, the electrode material,
the electrolysis conditions, and other factors on the electrochemical carboxylation of acyl chlorides (CH3COCl, p-X-C6H4COCl, where X=CH3, H, NO2) and on the yields of the respective α-oxocarboxylic acids were studied. It was shown that factors that accelerate the cathodic
reduction of the acid chlorides and stabilize the intermediates and products of the electrosynthesis lead to an increase in
the yield of the oxo acids.
L. V. Pisarzhevskii Institute of Physical Chemistry, National Academy of Sciences of Ukraine, 31 Nauki Prosp., Kiev 252039.
Translated from Teoreticheskaya i éksperimental’naya Khimiya, Vol. 34, No. 1, pp. 1–5, January–February, 1998. 相似文献
17.
Yu. D. Pankratiev V. M. Turkov M. Forissier J. L. Portefaix 《Reaction Kinetics and Catalysis Letters》1977,7(2):133-137
The heat of oxygen removal from the MnO2 surface (qs) during its reduction by carbon monoxide and the heat of reoxidation of the reduced surface by gaseous oxygen (qs) have been determined by a calorimetric method at 100 and 200°C. At 200°C both qs and qs are close to the enthalpy of MnO2 transformation to Mn2O3. At 100°C the reduction of MnO2 takes place without any distinct formation of a phase of the lower oxide, however, it leads to significant changes in both the surface and the subsurface layer of the oxide.
100° 200°C MnO2 (qs) (qs). 200°C qs, qs ( ) MnO2 Mn2O3. 100°C MnO2 , .相似文献
18.
G. Koulouris 《Journal of Radioanalytical and Nuclear Chemistry》1996,204(1):183-191
Selective sorption characteristics of electrolytic manganese dioxide (EMD) were investigated by dynamic tests using different types of water under varying conditions of235U activity, chemical species and flow rate in the presence of226Ra. EMD showed selectivity for ions having an effective ionic radius of about 1.4 Å indicating a selective sieve mechanism supporting a tunnel type structure. Sorption of235U under all experimental conditions is small (<1%) due to the larger sizes of urano-complexes (>3.54 Å). The Ra2+ ion (size 2.86 Å) was selectively extracted by EMD so that226Ra can be measured directly by gamma spectrometry using the 186 keV gamma line. 相似文献
19.
In this work a titration technique was used to determine the amphoteric surface properties of a series of heat-treated electrolytic manganese dioxide (EMD) samples (up to 500 degrees C). The surface of each sample was found to consist of independent acidic and basic hydroxyl sites, which could be characterized by their respective equilibrium constants and site concentrations. It was found that the acidic sites could not be characterized by a single equilibrium constant, but rather by a distribution indicating the subtle differences between individual sites, while a single equilibrium constant adequately represented the basic sites. For EMD, K(a) varied between 0.1 and 6.3x10(-5), with a corresponding [MnOH((a)T)] value varying between 9.1 and 6.4x10(-6) mol m(-2) over the pH range considered. K(b) and [MnOH((b)T)] were found to be 1.81x10(-9) and 1.93x10(-5) mol m(-2), respectively. With heat treatment, K(a) increased, suggesting a strengthening of the MnO bond via the removal of defects such as Mn(3+) ions and cation vacancies. The fact that K(b) also increased was initially counterintuitive because it suggested that the MnO bond had been weakened by heat treatment. However, assuming that the acidic and basic hydroxyl groups are independent, the trends in K(b) could be rationalized in terms of oxygen ion coordination in the progressively heat-treated samples. The number of surface sites (N(s)) was determined crystallographically and from the sum [MnOH((a)T)] + [MnOH((b)T)]. The data from both methods were of the same order of magnitude but exhibited different trends due to certain inadequacies in both methods. However, the data trends did indicate that the crystal planes at the particle surface could be changing with heat treatment due to a decrease in the value of N(s) determined from the surface titrations. Electrochemical analysis of the samples in 9 M KOH indicated that their performance degraded considerably with heat treatment. In comparison with the surface titration data, it was concluded that proton insertion into the structure occurred only through basic surface sites, the decreasing number of which could limit performance. 相似文献
20.
V. A. Job S. B. Kartha G. Krishnamurty Sheila Gopal 《Analytical and bioanalytical chemistry》1974,269(1):26-28
The fluoride content of aqueous and organic Purex process solutions can be determined with good reproducibility and accuracy. The method includes a distillation separation of fluoride and its subsequent electrometric determination using a fluoride-selective electrode. The analytical range covers 5 to 500 μg of fluoride per aliquot and the concentration limit is 5×10?5 M (1 ppm). Complexing cations like Al(III), Zr(IV), La(III), and U(VI) do not interfere. 相似文献