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.
Precise conductance measurements of solutions of lithium chloride, lithium bromide, lithium iodide, lithium perchlorate, lithium tetrafluoroborate, lithium hexafluoroarsenate, tetrabutylammonium bromide, and tetrabutylammonium tetraphenylborate in N,N-dimethylacetamide are reported at 25°C in the concentration range 0.005–0.015 mol-dm–3. The conductance data have been analyzed by the 1978 Fuoss conductance equation in terms of the limiting molar conductance (0), the association constant (K a), and the association diameter (R). The limiting ionic conductances have been estimated from an appropriate division of the limiting molar conductivity of the reference electrolyte Bu4NBPh4. Slight ionic association was found for all these salts in this solvent medium. The results further indicate significant solvation of Li+ion, while the other ions are found to be unsolvated in N,N-dimethylacetamide.  相似文献   

2.
The crystal structures of vanadates Li1-2xCo1+xVO4 with x = 0 and 0.25 have been studied by a full pattern analysis. It has been shown that in cubic spinel LiCoVO4 (space group Fd3m), the 8a tetrahedral sites contain a majority of vanadium and a small amount of lithium; all cobalt, lithium, and a small amount of vanadium occupy the 16d octahedral sites. Li0.5Co1.25VO4 crystals belong to the rhombic system (Imma space group) with unit cell parameters a = 5.939(1), b = 5.810(1), and c = 8.303(1). On substitution of lithium by cobalt according to the scheme 2Li+ Co2+ + , half of the lithium and 70% of the vacancies formed are in the 4a octahedral sites, and onethird of lithium and most of cobalt occupy the 4d octahedral sites. The 4e tetrahedral sites are completely occupied by vanadium and lithium in a ratio of 0.92/0.08. The interatomic distances in LiCoVO4 and Li0.5Co1.25VO4 are calculated, and the sizes of lithium ion transport channels are evaluated.  相似文献   

3.
Large quantities of single phase, polycrystalline LiIrSn4 have been synthesised from the elements by melting in sealed tantalum tubes and subsequent annealing. LiIrSn4 crystallises with an ordered version of the PdGa5 structure: I4/mcm, a=655.62(8), . The lithium atoms were clearly localised from a neutron powder diffraction study: RP=0.147 and RF=0.058. Time-dependent electrochemical polarisation techniques, i.e. coulometric titration, chronopotentiometry, chronoamperometry and cyclic voltammetry were used to study the kinetics of lithium ion diffusion in this stannide. The range of homogeneity (Li1+ΔδIrSn4, −0.091?δ?+0.012) without any structural change in the host structure and the chemical diffusion coefficient (∼10−7-10−9 cm2/s) point out that LiIrSn4 is a first example of a large class of intermetallic compounds with lithium and electron mobility. Optimised materials from these ternary lithium alloys may be potential electrode material for rechargeable lithium batteries.  相似文献   

4.
Conjugate addition of achiral lithium dimethylamide to the chiral iron cinnamoyl complexes (S,E)- and (S,Z)-[(η5-C5H5)Fe(CO)(PPh3)(COCHCHPh)] proceeds with high diastereoselectivity, with this protocol being used to establish unambiguously the absolute configuration of Winterstein’s acid (3-N,N-dimethylamino-3-phenylpropanoic acid) as (R). The highly diastereoselective conjugate addition of lithium N-benzyl-N-trimethylsilylamide to a range of α,β-unsaturated iron acyl complexes, followed by in-situ elaboration of the derived enolate by either alkylation or aldol reactions is also demonstrated, facilitating the stereoselective synthesis of both cis- and trans-β-lactams. This methodology has been used to effect the formal asymmetric syntheses of (±)-olivanic acid and (±)-thienamycin. Addition of chiral lithium amides derived from primary and secondary amines to the iron crotonyl complex [(η5-C5H5)Fe(CO)(PPh3)(COCHCHMe)] indicates that lithium N-α-methylbenzylamide shows low levels of enantiorecognition, while lithium N-3,4-dimethoxybenzyl-N-α-methylbenzylamide and lithium N-benzyl-N-α-methylbenzylamide show high levels of enantiodiscrimination. The high level of observed enantiorecognition was used to facilitate a kinetic resolution of (RS)-[(η5-C5H5)Fe(CO)(PPh3)(COCHCHMe)] with homochiral lithium (R)-N-3,4-dimethoxybenzyl-N-α-methylbenzylamide. Further mechanistic studies show that conjugate additions of (RS)-lithium N-benzyl-N-α-methylbenzylamide to either the (RS)- or homochiral iron crotonyl complex show 2:1 stoicheiometry, while homochiral lithium N-benzyl-N-α-methylbenzylamide shows 1:1 stoicheiometry.  相似文献   

5.
Drastic effects of Lewis acids E(C6F5)3 (E = Al, B) on polymerization of functionalized alkenes such as methyl methacrylate (MMA) and N,N-dimethyl acrylamide (DMAA) mediated by metallocene and lithium ester enolates, Cp2Zr[OC(OiPr)CMe2]2 (1) and Me2CC(OiPr)OLi, are documented as well as elucidated. In the case of metallocene bis(ester enolate) 1, when combined with 2 equiv. of Al(C6F5)3, it effects highly active ion-pairing polymerization of MMA and DMAA; the living nature of this polymerization system allows for the synthesis of well-defined diblock and triblock copolymers of MMA with longer-chain alkyl methacrylates. In sharp contrast, the 1/2B(C6F5)3 combination exhibits low to negligible polymerization activity due to the formation of ineffective adduct Cp2Zr[OC(OiPr)CMe2]+[OC(OiPr)CMe2B(C6F5)3] (2). Such a profound Al vs. B Lewis acid effect has also been observed for the lithium ester enolate; while the Me2CC(OiPr)OLi/2Al(C6F5)3 system is highly active for MMA polymerization, the seemingly analogous Me2CC(OiPr)OLi/2B(C6F5)3 system is inactive. Structure analyses of the resulting lithium enolaluminate and enolborate adducts, Li+[Me2CC(OiPr)OAl(C6F5)3] (3) and Li+[Me2CC(OiPr)OB(C6F5)3] (4), coupled with polymerization studies, show that the remarkable differences observed for Al vs. B are due to the inability of the lithium enolborate/borane pair to effect the bimolecular, activated-monomer anionic polymerization as does the lithium enolaluminate/alane pair.  相似文献   

6.
The influence of co-ions in the eluent on the separation factor () of lithium isotope separation has been studied by ion exchange chromatography. A strongly acid cation exchange resin (Dowex 50W-X8) was used for the separation of lithium isotopes. The co-ions used in eluent were H+, K+, Ba2+, Cu2+, Al3+ and Cr3+ as their chlorides. From the experiments, it was found that6Li was enriched in the resin phase and7Li in solution phase. At the same distribution coefficient (Kd=30), the separation factor increased linearly with the charge of co-ion (=1.0022 to 1.0039).  相似文献   

7.
The method of quasi-equilibrium galvanostatic curves was applied to study the thermodynamics of lithium deintercalation from the system Li x C6 (solid phase)/Li+ (solution) in the interval 293-323 K and the thermodynamic characteristics (G, S, H) of lithium intercalation compounds in a carbonized fabric in relation to the degree of intercalation x.  相似文献   

8.
A comparative analysis of 6,7Li NMR spectra is performed for the samples of monoclinic lithium titanate obtained at different synthesis temperatures. In the 7Li NMR spectra three lines are found, which differ in quadrupole splitting frequencies v Q and according to ab initio EFG calculations are assigned to three crystallographic sites of lithium: Li1 (v Q ~ 27 kHz); Li2 (v Q ~ 59 kHz); Li3 (v Q ~ 6 kHz). The dynamics of lithium ions is studied in a wide temperature range from 300 K to 900 K. It is found that the narrowing of 7Li NMR spectra as a result of thermally activated diffusion of lithium ions in the low-temperature Li2TiO3 sample is observed at a higher temperature in comparison with a sample of high-temperature lithium titanate. Based on the analysis of 6Li NMR spectra it is assumed that there is mixed occupancy of lithium and titanium sites in the corresponding layers of the crystal structure of low-temperature lithium titanate, which hinders lithium ion transfer over regular crystallographic sites.  相似文献   

9.
Crystal Structure and Electric Conductivity of Spinel-Type Li2–2xMn1+xCl4 Solid Solutions The electric conductivity of the fast lithium ion conductors Li2–2xMn1+xCl4 was measured by impedance spectroscopic methods. The conductivities obtained, e.g. ~ 4 × 10?1 Ω?2 cm?1 at 570 K, depend only little on the lithium content. The crystal structure of Li1.6Mn1.2Cl4 was determined by neutron powder and X-ray single crystal diffraction (space group Fd3 m, Z = 8, a = 1 049.39(6) pm, Rw = 1.4% on the basis of 170 reflections). The lithium deficient chloride crystallizes in an inverse spinel structure like the stoichiometric compound Li2MnCl4 according to the formula (Li0,8)[Li0,4Mn0,6]2Cl4 with vacancies ( ) at the tetrahedral sites. The decrease of the Moct? Cl distances with the increase of x reveals that the ionic radius of Mn2+ in chlorides is equal or even smaller than that of Li+ opposite to fluorides and oxides. The ? Cl distances of spinel type chlorides are 237 ( tet) and 274 pm ( oct), respectively. The mechanism of the ionic conductivity is discussed.  相似文献   

10.
The lithium complex with the acenaphthylene dianion [Li(Et2O)2]22:3[Li(3:3-C12H8)]2 (1) was synthesized by the reduction of acenaphthylene with lithium in diethyl ether. According to the X-ray diffraction data, compound 1 has a reverse-sandwich structure with the bridging dianion 2:3[Li(3:3-C12H8)]2. Two lithium atoms in complex 1 are located between two coplanar acenaphthylene ligands of the 2:3[Li(3:3-C12H8)]2 2– dianion and are 3-coordinated with the five- and six-membered rings. The lanthanum complex with the acenaphthylene dianion [LaI2(THF)3]2(2-C12H8) (2) was synthesized by the reduction of acenaphthylene in THF with the lanthanum(iii) complex [LaI2(THF)3]2(2-C10H8) containing the naphthalene dianion. The 1H NMR spectrum of complex 2 in THF-d8 exhibits four signals of the acenaphthylene dianion, whose strong upfield shifts compared to those of free acenaphthylene indicate the dianionic character of the ligand. The highest upfield chemical shift belongs to the proton bound to the C atom on which, according to calculation, the maximum negative charge is concentrated.  相似文献   

11.
The garnets Li3Nd3W2O12 and Li5La3Sb2O12 have been prepared by heating the component oxides and hydroxides in air at temperatures up to 950 °C. Neutron powder diffraction has been used to examine the lithium distribution in these phases. Both compounds crystallise in the space group with lattice parameters a=12.46869(9) Å (Li3Nd3W2O12) and a=12.8518(3) Å (Li5La3Sb2O12). Li3Nd3W2O12 contains lithium on a filled, tetrahedrally coordinated 24d site that is occupied in the conventional garnet structure. Li5La3Sb2O12 contains partial occupation of lithium over two crystallographic sites. The conventional tetrahedrally coordinated 24d site is 79.3(8)% occupied. The remaining lithium is found in oxide octahedra which are linked via a shared face to the tetrahedron. This lithium shows positional disorder and is split over two positions within the octahedron and occupies 43.6(4)% of the octahedra. Comparison of these compounds with related d0 and d10 phases shows that replacement of a d0 cation with d10 cation of the same charge leads to an increase in the lattice parameter due to polarisation effects.  相似文献   

12.
The isotope composition of lithium charge carriers is experimentally found to severely affect transport in solid electrolytes -Li3BO3, Li3N, Li3AlN2, Li5SiN3, Li6MoN4, Li6WN4, and LiCl. The lithium cation conduction of these decreases with increasing content of 6Li or 7Li and reaches a minimum at [6Li] = [7Li]. The activation energy for conduction increases, reaches a maximum in the same compositions, and then diminishes. Rates of spin–lattice relaxation of 7Li nuclei in electrolytes are studied by an NMR method at 15–35 MHz. The calculated activation energy for short-range motion (to one interatom distance) of lithium charge carriers in crystal lattices of electrolytes is lower than that for ionic conduction by 2–3 times, which is attributed to two types of correlation (electrostatic, isotopic) of charge carriers.  相似文献   

13.
The lithium intercalation into nanostructured films of mixed tin and titanium oxides is studied. X-ray diffraction and Moessbauer spectroscopy analyses reveal that films consist of a rutile solid solution (Sn, Ti)O2 and an amorphous tin oxide enriched with Sn2+ ions. The films specific capacity during the first cathodic polarization in a 1 M lithium imide solution in dioxolane is 200–700 mA h/g, of which nearly one half is the irreversible capacity. During the second cycle, the latter is 15% of that in the first cycle. As the films are thin (<1 m), their capacity does not depend on the current density at 1–80 mA/g. During the electrode cycling, the capacity decreases by 2 mA h/g each cycle. The effective lithium diffusion coefficient, determined by a pulsed galvanostatic method, is 10–11 cm2/s; it slightly increases with the film lithiation. During the first cycle, the amorphous phase of oxides is reduced to tin metal, the solid solution (Sn, Ti)O2 decomposes, SnO2 disperses to become an x-ray amorphous phase, and TiO2 precipitates as a rutile phase. Lithium reversibly incorporates into the tin metal, yielding Li y Sn, and into a disperse SnO2 phase, yielding Li x SnO2.  相似文献   

14.
The title compounds Li6PO5Br (Fand Li6PO5Cl (F represent the first oxidic argyrodites in general and the first lithiumoxoargyrodites in particular. The overall crystal structure corresponds to the cubic high temperature (HT) modification of all known cubic argyrodites, however, with a seemingly small but important difference concerning the lithium positions. In all other HT argyrodites with similar lithium content the 24 lithium atoms per unit cell are disordered over a 48 fold position in close vicinity to a 24 fold one causing a high mobility of the Li+. In the title compounds, however, they occupy the 24 fold one in a strictly ordered manner thus establishing a planar triangular first sphere coordination environment. This detail is of great importance for the amount of the specific lithium ionic conductivity and for the possible phase transition to an LT (low temperature) modification accompanied by an ordering of the disordered lithium atoms. Apparently the latter transition is suppressed in the title compounds because the Li+ are already frozen out in the cubic (HT = LT) form. The initially open question how this structural peculiarity influences the ionic conductivity (strengthening or weakening in comparison to oxygen free argyrodites?) is answered by a series of impedance measurements. The specific lithium ionic conductivity of the title compounds in the range 313 K < T < 518 K is significantly lower than in oxygen free argyrodites.  相似文献   

15.
Tin (IV) chloride reacts with sulfolane (S) to form a cis-octahedral adduct SnCl4·S2. Solutions of lithium chloride and tin (IV) chloride in sulfolane contain the complex ions SnCl 5 and SnCl 6 2– at 11 and 21 mole ratios of constituents, respectively. The complexes are characterized by conductimetry and by Mössbauer, IR, and Raman spectroscopy.  相似文献   

16.
Hollow titanium dioxide (TiO2) microspheres were synthesized in one step by employing tetrabutyl orthotitanate (TBOT) as a precursor through a facile solvothermal method in the presence of NH4HCO3. XRD analysis indicated that anatase TiO2 can be obtained directly without further annealing. TiO2 hollow microspheres with diameters in the range of 1.0–4.0 μm were confirmed through SEM and TEM measurements. The specific surface area was measured to be 180 m2 g?1 according to the nitrogen adsorption–desorption isotherms. Superior photocatalytic performance and good lithium storage properties were achieved for resultant TiO2 samples. The H2 evolution rate of the optimal sample is about 0.66 mmol h?1 after loaded with 1 wt.% Pt (20 mg samples). The reversible capacity remained 143 mAh g?1 at a specific current of 300 mA g?1 after 100 charge–discharge cycles. This work provides a facile strategy for the preparation of hollow titanium dioxide microspheres and demonstrates their promising photocatalytic H2 evolution and the lithium storage properties.
Graphical abstract Hollow titanium dioxide spheres are directly synthesized via a facile template-free solvothermal method with the presence of NH4HCO3 based on inside-out Ostwald ripening (see picture), and demonstrated both as a photocatalyst for water splitting and a promising anode material for lithium-ion batteries. Superior photocatalytic performance and excellent lithium storage properties are achieved for resultant TiO2 hollow microspheres.
  相似文献   

17.
Solutions of silver and lithium tetrafluoroborate in acrylonitrile, over a range ofconcentrations between 0.5 and 4 mol-kg–1, have been studied byFourier-transform Raman spectroscopy. The spectral regions studied include the solvent(C=dN) fundamental and the anion B-F symmetric stretching band. In AgBF4solutions the absence of ionic pairing was demonstrated and the anion 1(A 1)remains as a single narrow band located at 764.7±0.1 cm–1. Consequently, thesilver ion solvation number does not change in the range of concentrations studied,having a constant value of 3.54±0.10. However, a high level of ionic pairingwas observed in the corresponding solutions of LiBF4. Three components weredetected in the tetrafluoroborate 1(A 1) band located at 766.0±0.4, 773.4±1.1,and 782.7±0.9 cm–1, and assigned to spectroscopically free anions, ion pairs,and dimers, respectively. The solvation number of the lithium ion, which shouldbe three in the limit of infinite dilution, decreases as the salt concentrationincreases as a result of the ionic pairing. However, the ionic pairing of LiBF4 inacrylonitrile is less than that previously observed in lithiumtrifluoromethanesulfonate (triflate) or lithium perchlorate.  相似文献   

18.
Hydrogen storage : In lithium amidoboranes an initial molecule of H2 is released by the formation of LiH, followed by a redox reaction of the dihydrogen bond formed between LiHδ? and NHδ+. In this dehydrogenation process, an intermolecular N? B bond forms through the catalytic effect of a Li cation. After releasing the first molecule of H2, a Li cation binds to a nitrogen atom, lowering the energy barrier for the second H2 loss per lithium amidoborane dimer (see figure).

  相似文献   


19.
The structure of 5-nitraminotetrazole lithium salt monohydrate was determined by X-ray diffraction analysis. Crystals are monoclinic, space group P21/c; a = 8.3789(5), b = 10.1872(6), c = 6.6709(5) ; = 106.63(1)°; V = 545.60(98) 3; Z = 4; calc = 1.875 g/cm3. The anion has a planar nitrimine structure with a delocalized negative charge. Each lithium cation (c.n. 5) is bound to three anions and two hydration water molecules. Both oxygen atoms of the nitro groups and the N(3) atom of the tetrazole ring are involved in cation coordination. The geometrical characteristics of the anion are similar to those found for other monosalts of 5-nitraminotetrazole.  相似文献   

20.
The reduction of 1-chloro-1,2,3,4,5-pentaphenylsilole, (C4Ph4SiPhCl, 1) with 2 equiv lithium gave the pentaphenylsilole anion [C4Ph4SiPh] (2), silole dianion [C4Ph4Si]2− (3), and hexaphenylsilole C4Ph4SiPh2 (4). 2, 3, and 4 from the reaction mixture were characterized by 29Si NMR spectroscopy. The 29Si chemical shift of 3.7 ppm for 2 is shifted upfield as compared to that of previously reported t -butyltetraphenylsilole anion Li[C4Ph4SitBu], but shifted downfield compared to that of the other silole monoanion such as Li[C4Me4SiSiMe3], indicating the delocalization of silole anion through the 5-membered ring. Derivatization of the reaction mixture with iodomathane gave C4Ph4SiPh2 (4), C4Ph4SiMePh (5), and C4Ph4SiMe2 (6), which were characterized by 1H, 13C, and 29Si NMR spectroscopy. The silole dianion 3 could be either from the continuous reduction of 1 with lithium or from the disproportionation of 2. The reduction of 1 with excess lithium in THF gave the silole dianion [C4Ph4Si]2− in about 70% yield.  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