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.
The electrochemical behaviour of the Li+/Li couple is examined briefly in molten dimethylsulfone (DMSO2) at 150°C. It has been found that the Li+/Li system is reversible. However, electrodeposited lithium exhibits some instability which severely limits the use of the lithium anode in molten DMSO2 based secondary cells. Therefore we have undertaken the study of the electroformation of the LiAl alloy in 1 mol kg−1 LiClO4/DMSO2 at 150°C by investigating the electrochemical incorporation of lithium into an aluminium electrode by potentiostatic and galvanostatic techniques. The results obtained from electrochemical measurements and X-ray diffraction experiments have proved that incorporation of Li in Al in molten DMSO2 leads to the formation of the β LiAl alloy. Analysis of the chronocoulometric curves has allowed the different processes limiting the rate of incorporation of Li and Al to be specified. Moreover, the diffusion coefficient of lithium into aluminium has been determined from chronoamperometric measurements: 0.7 × 10−10DLi(α) ⪕ 1.4 × 10−10 cm2 s−1. Finally, the galvanostatic study has shown that the β LiAl alloy can be considered a promising anodic material for molten dimethylsulfone-based secondary batteries.  相似文献   

2.
Improved durability, enhanced interfacial stability, and room temperature applicability are desirable properties for all-solid-state lithium metal batteries (ASSLMBs), yet these desired properties are rarely achieved simultaneously. Here, in this work, it is noticed that the huge resistance at Li metal/electrolyte interface dominantly impeded the normal cycling of ASSLMBs especially at around room temperature (<30 °C). Accordingly, a supramolecular polymer ion conductor (SPC) with “weak solvation” of Li+ was prepared. Benefiting from the halogen-bonding interaction between the electron-deficient iodine atom (on 1,4-diiodotetrafluorobenzene) and electron-rich oxygen atoms (on ethylene oxide), the O-Li+ coordination was significantly weakened. Therefore, the SPC achieves rapid Li+ transport with high Li+ transference number, and importantly, derives a unique Li2O-rich SEI with low interfacial resistance on lithium metal surface, therefore enabling stable cycling of ASSLMBs even down to 10 °C. This work is a new exploration of halogen-bonding chemistry in solid polymer electrolyte and highlights the importance of “weak solvation” of Li+ in the solid-state electrolyte for room temperature ASSLMBs.  相似文献   

3.
The thermal instabilities of deposited lithium with electrolytes in lithium-ion batteries are simulated by the reactions between metallic lithium with organic esters and ethers. Exothermic onset temperatures and enthalpy changes are measured and analyzed by differential scanning calorimetry. In this study, heat of reactions in lithium with eight different formations of esters and ethers are determined which are consistent to the data of lithiated graphite (LiC6) reacted with electrolytes in literature. Furthermore, violently exothermic reactions with enthalpy larger than 1,000 J g?1 and onset temperature lower than 120 °C are further conducted by the confinement test to verify the worst scenarios and consequences of lithium-ion batteries encountered any kind of abuses. Thermal instability of metallic lithium with organic esters in descending order determined to be Li + EB (70 °C)>Li + MB (73.1 °C)>Li + EA (90.8 °C). Finally, thermal hazard data such as onset temperature, maximum self-heat rate, maximum temperature, and maximum pressure of lithium reacted with esters and ethers are compared, evaluated, and some conclusion and suggestions are made.  相似文献   

4.
《Analytical letters》2012,45(4):847-866
Abstract

The fluorescence emission from the lithium/1,4-dihydroxyanthraquinone system shows a great enhancement in the presence of certain water-miscible solvents. This is justified from the donicity (nucleophilic properties) of the solvent that facilitates the solvation of the lithium cation in solution and the stabilization of an nondissociated ion-pair between the solvated lithium cation and the 1,4-dihydroxyanthraquinonate anion. A very sensitive analytical method was proposed for the spectrofluorimetric determination of lithium based on its reaction with 1,4-dihydroxyanthraquinone (quinizarin) in a dimethylsulfoxide medium (90%) and in presence of sodium hydroxide. The fluorescence is measured at an excitation wavelength of 602 nm and an emission wavelength of 670 nm and it is stable at 25°C at least 6 h. The calibration curve is linear over the concentration ranges of 2–40 μg/l of lithium in an aqueous matrix and 3–50 μg/l in a serum matrix; the RSD's in the determination of 20 μg/l of Li+ were 2.6% and 3.2%, respectively. The proposed procedure was satisfactorily applied to the determination of lithium in drugs, dietetic products and human serum.  相似文献   

5.
Solvation of the lithium ion by acetone was studied in acetone-nitromethane solutions by Raman, infrared, and7Li and35Cl NMR spectroscopic techniques. It was confirmed that the 390-cm?1 IR acetone band is split by the lithium ion and that a 369-cm?1 IR band, attributed by other authors to Li+-nitromethane vibration, is due to the vibration of acetone in the lithium inner solvation shell. The frequency of the Li+-nitromethane vibrational band is strongly anion dependent due to contact-ion-pair formation. Several different techniques indicate that Li+ is solvated by four acetone molecules, and approximate equilibrium-constant values for the stepwise solvation reaction were calculated. The influence of weak complexing agents on Li+ ClO 4 ? ion-pair formation was investigated.  相似文献   

6.
Lithium metal is a promising anode material for next-generation high-energy-density batteries but suffers from low stripping/plating Coulombic efficiency and dendritic growth particularly at sub-zero temperatures. Herein, a poorly-flammable, locally concentrated ionic liquid electrolyte with a wide liquidus range extending well below 0 °C is proposed for low-temperature lithium metal batteries. Its all-anion Li+ solvation and phase-nano-segregation solution structure are sustained at low temperatures, which, together with a solid electrolyte interphase rich in inorganic compounds, enable dendrite-free operation of lithium metal anodes at −20 °C and 0.5 mA cm−2, with a Coulombic efficiency of 98.9 %. As a result, lithium metal batteries coupling thin lithium metal anodes (4 mAh cm−2) and high-loading LiNi0.8Co0.15Al0.05O2 cathodes (10 mg cm−2) retain 70 % of the initial capacity after 100 cycles at −20 °C. These results, as a proof of concept, demonstrate the applicability of locally concentrated ionic liquid electrolytes for low-temperature lithium metal batteries.  相似文献   

7.
A cubic Li5La3Nb2O12 phase with a garnet framework was synthesized by the sol–gel process, in which lithium hydroxide, niobium oxide and acetic lanthanum were used as starting materials, while water was used as solvent. Pure garnet-like Li5La3Nb2O12 powders were obtained after heating the gel precursor at 700 °C for 6 h with 10 % excess lithium salt. The calcination temperature is nearly 250 °C lower than that by the solid state reaction. The phase transforms from cubic to tetragonal symmetry with loss of lithium at 717 °C, but the garnet framework remains stable to above 900 °C. A pellet annealed at 900 °C for 6 h had a room-temperature Li+-ion conductivity σLi (22 °C) = 1.0 × 10?5 S cm?1, a little higher than that attained by solid-state synthesis. The Li5La3Nb2O12 compound was chemically stable against two commonly used cathode materials, LiMn2O4 and LiCoO2, up to 900 °C and against metallic lithium.  相似文献   

8.
The anionic polymerization of 1.3-cyclohexadiene (1.3-CHD) was investigated in temperatures that ranged from 25 to ?77°C. Initiation by lithium naphthalene (N?·,Li+) in tetrahydrofuran at ?20°C yields polymers with fairly narrow molecular weight distribution. The M?w of these polymers so prepared is ca. 20,000. Polymerization of 1.3-CHD conducted at room temperature is accompanied by the dehydrogenation and disproportionation of the monomer, especially when N?·,K+ acts as initiator. Oligomers are formed when hexamethylphosphoramide is used as a solvent. The mechanism of the initiation of the polymerization of 1.3-CHD by N?·,Li+ was elucidated and the rate constants at ?20°C in tetrahydrofuran of the elementary reactions were determined. It was established that the dianions formed by disproportionation of N?·,Li+ act as effective initiators for 1.3-CHD. The adducts formed constitute the cyclohexanyl and naphthyl carbanionic groups. The former carbanions (λmax ~ 275 nm) propagate the polymerization. The initially formed dimeric adducts are stabilized by the separation of the carbanionic end groups by the additional monomer units. Chain transfer to the monomer limits the growth of the polymers. The isomerization of the cyclohexadienyl anions, formed as result of chain transfer, may be followed by the elimination of lithium hydride. The latter reaction represents a termination step. Addition of 1.4-CHD to the reaction mixture enhances the chain transfer and the termination.  相似文献   

9.
The behaviour of pH glass sensors is determined strongly by the kind and composition of glasses. For the heat stability, the kind and concentration of the alkaline oxide as modifier play an important role. It could be shown by potentiometric and impedimetric measurements that lithium oxide-containing glass membranes are much more stable than those with sodium oxide concerning the emf of the electrochemical cell. They tolerate several 5-h heat treatments at temperatures of 100°C. The impedance plots show differences in electrochemical kinetics between both glass membranes, which can be caused by different leaching due to differences in ionic radii of Li+ and Na+, respectively.  相似文献   

10.
Second‐order rate constants for the reactions of acceptor‐substituted phenacyl (PhCO?CH??Acc) and benzyl anions (Ph?CH??Acc) with diarylcarbenium ions and quinone methides (reference electrophiles) have been determined in dimethylsulfoxide (DMSO) solution at 20 °C. By studying the kinetics in the presence of variable concentrations of potassium, sodium and lithium salts (up to 10?2 mol L?1), the influence of ion‐pairing on the reaction rates was examined. As the concentration of K+ did not have any influence on the rate constants at carbanion concentrations in the range of 10?4–10?3 mol L?1, the acquired rate constants could be assigned to the reactivities of the free carbanions. The counter ion effects increase, however, in the series K+<Na+<Li+, and the sensitivity of the carbanion reactivities toward variation of the counter ion strongly depends on the structure of the carbanions. The reactivity parameters N and sN of the free carbanions were derived from the linear plots of log k2 against the electrophilicity parameters E of the reference electrophiles, according to the linear‐free energy relationship log k2(20 °C)=sN(N+E). These reactivity parameters can be used to predict absolute rate constants for the reactions of these carbanions with other electrophiles of known E parameters.  相似文献   

11.
The paper reported a green and efficient extraction strategy to lithium isotope separation. A 4-methyl-10-hydroxybenzoquinoline (ROH), hydrophobic ionic liquid—1,3-di(isooctyl)imidazolium hexafluorophosphate ([D(i-C8)IM][PF6]), and hydrophilic ionic liquid—1-butyl-3-methylimidazolium chloride (ILCl) were used as the chelating agent, extraction medium and ionic associated agent. Lithium ion (Li+) first reacted with ROH in strong alkali solution to produce a lithium complex anion. It then associated with IL+ to form the Li(RO)2IL complex, which was rapidly extracted into the organic phase. Factors for effect on the lithium isotope separation were examined. To obtain high extraction efficiency, a saturated ROH in the [D(i-C8)IM][PF6] (0.3 mol l?1), mixed aqueous solution containing 0.3 mol l?1 lithium chloride, 1.6 mol l?1 sodium hydroxide and 0.8 mol l?1 ILCl and 3:1 were selected as the organic phase, aqueous phase and phase ratio (o/a). Under optimized conditions, the single-stage extraction efficiency was found to be 52 %. The saturated lithium concentration in the organic phase was up to 0.15 mol l?1. The free energy change (ΔG), enthalpy change (ΔH) and entropy change (ΔS) of the extraction process were ?0.097 J mol?1, ?14.70 J mol K?1 and ?48.17 J mol?1 K?1, indicating a exothermic process. The partition coefficients of lithium will enhance with decrease of the temperature. Thus, a 25 °C of operating temperature was employed for total lithium isotope separation process. Lithium in Li(RO)2IL was stripped by the sodium chloride of 5 mol l?1 with a phase ratio (o/a) of 4. The lithium isotope exchange reaction in the interface between organic phase and aqueous phase reached the equilibrium within 1 min. The single-stage isotope separation factor of 7Li–6Li was up to 1.023 ± 0.002, indicating that 7Li was concentrated in organic phase and 6Li was concentrated in aqueous phase. All chemical reagents used can be well recycled. The extraction strategy offers green nature, low product cost, high efficiency and good application prospect to lithium isotope separation.  相似文献   

12.
High-alkaline protease (HAP) has been entrapped in Manugel DMB (an alginate gel) and assayed with two sizes and types of substrates: neutral protein casein and synthetic chromogenic tripeptide substrate, Z-Gly-Pro-Cit-PNA. Increasing the concentration of calcium chloride used for capsule formation decreased the measured enzyme activity with both substrates. Capsules were found to be stable in water for long periods of time, but they dissolved in both phosphate and carbonate-bicarbonate buffers. The pH vs activity profiles of encapsulated enzyme showed pH optima between 10 and 11 with both substrates. The calcium alginate matrix surrounding the enzyme was quite effective in stabilizing the enzyme at 20–25 °C and even more so at 4°C. Enzyme stability at 50 °C was quite impressive, some enzyme activity being evident even after remaining for 1 wk at this temperature in water. Increasing concentrations of sodium dodecyl sulfate (SDS) were also found to inhibit the protease progressively, whereas a polyhexamethylene biguanidium chloride (PHMBH+Cl-) and SDS:PHMBH+Cl- combination showed the opposite effect. Optical microscopy, especially polarized light microscopy, provided a sensitive physical means of ascertaining some of the structural properties (sphericity, disorganization or organization, distinct layer enveloping the capsules, intensity of the maltese cross) of the capsules with and without enzyme before and after different chemical treatments and the presence or absense of the substrate.  相似文献   

13.
The electrochemical behavior of the Li+/Li couple was studied at polycrystalline tungsten, platinum, copper and aluminum electrodes in tri‐1‐butylmethylammonium bis((trifluoromethyl)sulfonyl)imide ionic liquid mixed with a little propylene carbonate at 30 °C. Lithium cations were introduced into the ionic liquid by dissolution of lithium bis((trifluoromethyl)sulfonyl)imide which is highly soluble in ionic liquid. Propylene carbonate was used to reduce the viscosity of this ionic liquid in order to enhance the mass transfer and to additionally improve the stability of lithium deposits. At the tungsten and copper electrodes, the cyclic voltammetric behavior of a Li+/Li couple is a quasi‐reversible reaction. At the platinum electrode, the behavior becomes very complicated because of the alloy formation. Coulombic efficiency was used to evaluate the stability of lithium deposits at each electrode. The aluminum electrode showed the best efficiency due to the formation of Li‐Al alloy. However, lowest efficiency was obtained at the platinum electrode because of the low redox reversibility of the lithium in the Li‐Pt alloy. The diffusion coefficient of lithium cation in this solution was 1.0 ± 0.1 × 10?;7 cm2 s?;1 as determined by chronopotentiometry. The best coulombic efficiency obtained at the Al electrode is 97% but dropped to about 90% after 12 hours. The self‐discharge current of the lithium deposits at the Al electrode was 0.4 μA/cm2 during the experimental period.  相似文献   

14.
The heat capacity and density of solutions of lithium and sodium nitrates in N-methylpyrrolidone (MP) at 298.15 K are studied by calorimetry and densimetry. The standard partial molar heat capacities and volumes (C? p,2° and V? 2°) of LiNO3 and NaNO3 in MP are calculated. The standard heat capacities C? p,i ° and volumes V? i ° of Li+ and Na+ ions in MP at 298.15 K are determined on the basis of a proposed scale of ionic contributions of C? p,2° and V? 2° values. The obtained data are discussed in relation to certain features of solvation in solutions of the investigated salts.  相似文献   

15.
The lithium, potassium, and ammonium salts of bis (2‐ethylhexyl) sulphosuccinic acid have been prepared from the sodium salt (AOT) by applying ion‐exchange technique. The critical micellization concentrations (cmc) of the surfactants with four different counterions have been determined at a temperature range of 10°C to 40°C using surface tension as well as electrical conductivity measurements. Observed data have been utilized to evaluate the ionization degree (counter ion association constant),α, and various thermodynamic parameters of micellization viz, free energy, enthalpy, entropy changes of micelle formation, and also the surface parameters (Γmax, Amin) in aqueous media. The value of cmc decreases with hydrated ionic size of the counter ions (except K+) and follows the order NH4 +>Na+>Li+>K+. While large negative free energy change (ΔG0 m) and the positive entropy change (ΔS0 m) favor the micellization process thermodynamically, nature of their variation with counterion supports the involvement of counterion size factor in micellization process via a change in the hydrophilicity of surfactant head group.  相似文献   

16.
Extraction of lithium ions from salt‐lake brines is very important to produce lithium compounds. Herein, we report a new approach to construct polystyrene sulfonate (PSS) threaded HKUST‐1 metal–organic framework (MOF) membranes through an in situ confinement conversion process. The resulting membrane PSS@HKUST‐1‐6.7, with unique anchored three‐dimensional sulfonate networks, shows a very high Li+ conductivity of 5.53×10?4 S cm?1 at 25 °C, 1.89×10?3 S cm?1 at 70 °C, and Li+ flux of 6.75 mol m?2 h?1, which are five orders higher than that of the pristine HKUST‐1 membrane. Attributed to the different size sieving effects and the affinity differences of the Li+, Na+, K+, and Mg2+ ions to the sulfonate groups, the PSS@HKUST‐1‐6.7 membrane exhibits ideal selectivities of 78, 99, and 10296 for Li+/Na+, Li+/K+, Li+/Mg2+ and real binary ion selectivities of 35, 67, and 1815, respectively, the highest ever reported among ionic conductors and Li+ extraction membranes.  相似文献   

17.
Polysiloxanes with covalently attached oligo ethylene oxide and di-t-butylphenol ( I ), naphthol ( II ), and hexafluoropropanol ( III ) were synthesized. The crosslinked polymers with a hexamethylene spacer were also prepared. The ion conductivities of the Li, Na, and K salts were measured as a function of temperature. The highest conductivities for K and Na of I at 30°C were 5.5 × 10?5 and 5.0 × 10?5 S/cm, respectively, when the ratio of the ion to ethylene oxide unit was 0.014. On the other hand, Li conductivity was 8.0 × 10?6 S/cm when the ratio between Li and ethylene oxide unit was 0.019. The maximum conductivities of Li ions of II and III were in the order of 10?6 and 10?7 S/cm at 30°C, respectively. When the polymers were crosslinked by a hexamethylene residue, the ion conductivities decreased while the degree of crosslinking increased. The temperature dependence of the cation conductivities of these systems could be described by the Williams-Landel-Ferry (WLF) and the Vogel-Tammann-Fulcher (VTF) equation. The results demonstrate that ion movement in these polymers is correlated with the polymer segmental motion. The order of ionic conductivity was K+ > Na+ ? Li+. This suggests that steric hindrance and π-electron delocalization of the anions attached to polymer backbone have a large effect on ion-pair separation and their ionic conductivities. Thermogravimetric analysis of the polymers indicated that the degradation temperature for I and II were about 100°C higher than for poly(siloxane-g-ethylene oxide). This is due to the antioxidant properties of sterically hindered phenols and naphthols. © 1993 John Wiley & Sons, Inc.  相似文献   

18.
To obtain thermally stable and mechanically strong sodium and lithium conducting polymers, we prepared Na+ and Li+ poly(phenylene terephthalamide sulfonate salts) (MW ~ 5500). We also synthesized oligo(ethylene oxide) (3, 5, or 7 units of ethylene oxide) substituted ethylene carbonate and poly[oxymethylene-oligo(oxyethylene)]. These are high boiling point liquids with high dielectric constants as well as metal chelating properties. Polyelectrolyte systems were prepared by mixing Na+ or Li+ poly(phenylene terephthalamide sulfonate) salts with various amounts of modified ethylene carbonate and/or poly[oxymethylene-oligo(oxyethylene)]. Films (0.1–0.5 mm thick) obtained from the blends were found to have considerable mechanical strength; forming free standing films. The ionic conductivities of the Na+ and Li+ polyelectrolyte systems were 10?6?10?5 S/cm at 25°C. Thermal properties of these blend systems were investigated in detail. © 1994 John Wiley & Sons, Inc.  相似文献   

19.
The cationic polymerization of N-vinyl carbazole, initiated by Ph3C+ AsF6? and Ph3C+ PF6? in methylene dichloride at 20 and 0°, has been studied in some detail. Reactions were very fast and rates of monomer consumption were measured using an adiabatic calorimetric technique. Initiation was relatively “slow” but complete, and termination was deduced to be insignificant during kinetic lifetimes. Values for kp were found to vary with the initial initiator concentration; this dependence is discussed in terms of current theories regarding equilibria between ion pairs and free ions in non-aqueous solvents. kp+ values estimated from two methods of extrapolation are 9.5 · 105 M?1 sec?1 at 20° and 4.8 · 105 M?1 sec?1 at 0°. Finally, it has been found that ion pairs are much less reactive than free ions in this system.  相似文献   

20.
Anionic polymerization of methyl methacrylate (MMA) was carried out in tetrahydrofuran (THF) or THF/toluene mixture at ?78°C initiated by triphenylmethyl sodium or lithium as initiators. Highly syndiotactic PMMA of low polydispersity (M w/m n = 1.11–1.17) could be prepared with triphenylmethyl lithium in THF or THF/toluene mixture at ? 78°C. Moreover, PMMA macromonomer having one vinylbenzyl group per polymer chain was prepared by the couplings of living PMMA initiated by triphenylmethyl lithium with p-chloromethyl styrene (CMS) at ?78°C. The coupling reaction of living PMMA initiated by triphenylmethyl sodium with CMS was scarcely occurred.  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