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.
Solvated electron-Na+ pairs, e?,Na+, and 1,1-diphenylethylene reversibly recombine in THF, the capture constant = 3 × 107 M?1 s?1 and the detachment constant = 46 s?1; the e?,Na+, formed by flash-photolysis of Na+,C?(Ph)2CH2CH2C?(Ph2, Na+ survive for 0.1 s in this solvent at ambient temperature without any detectable decay.  相似文献   

2.
The optical pumping method of alkali molecules by atom—molecule exchange collisions is applied to obtain the magnetic shielding difference σ(Na) — σ(Na2) = (29 ± 16) × 10?6 between Na atoms and Na2 molecules and the scalar nuclear spin—spin coupling constants ds = (306 ± 30)s?1 of 23Na39K.  相似文献   

3.
Picosecond absorption spectroscopy has been used to examine the primary photoproduct of Cr(CO)6 in both neat and mixtures of THF and cyclohexane. The primary intermediate observed is shown to be the solvated pentacarbonyl in which one solvent molecule occupies the coordination site created by the photoelimination of CO. The rate of exchange of cyclohexane from (cyclohexane)Cr(CO)5 by THF to form (THF)Cr(CO)5 was found to be bimolecular, k = (4 ± 1) × 107 mole?1 s?1, with an enthalpy and entropy of activation of 1 ± 1 kcal/mole and ?20 ± 4 eu, respectively.  相似文献   

4.
The spinel LiMn2O4 cathode material has been considered as one of the most potential cathode active materials for rechargeable lithium ion batteries. The sodium-doped LiMn2O4 is synthesized by solid-state reaction. The X-ray diffraction analysis reveals that the Li1?x Na x Mn2O4 (0?≤?x?≤?0.01) exhibits a single phase with cubic spinel structure. The particles of the doped samples exhibit better crystallinity and uniform distribution. The diffusion coefficient of the Li0.99Na0.01Mn2O4 sample is 2.45?×?10?10 cm?2 s?1 and 3.74?×?10?10 cm?2 s?1, which is much higher than that of the undoped spinel LiMn2O4 sample, indicating the Na+-ion doping is favorable to lithium ion migration in the spinel structure. The galvanostatic charge–discharge results show that the Na+-ion doping could improve cycling performance and rate capability, which is mainly due to the higher ion diffusion coefficient and more stable spinel structure.  相似文献   

5.
Rate constants for a series of alcohols, ethers, and esters toward the sulfate radical (SO4?) have been directly determined using a laser photolysis set‐up in which the radical was produced by the photodissociation of peroxodisulfate anions. The sulfate radical concentration was monitored by following its optical absorption by means of time resolved spectroscopy techniques. At room temperature the following rate constants were derived: methanol ((1.6 ± 0.2) × 107 M?1 s?1); ethanol ((7.8 ± 1.2) × 107 M?1 s?1); tert‐butanol ((8.9 ± 0.3) × 105 M?1 s?1); diethyl ether ((1.8 ± 0.1) × 108 M?1 s?1); MTBE ((3.13 ± 0.02) × 107 M?1 s?1); tetrahydrofuran (THF) ((2.3 ± 0.2) × 108 M?1 s?1); hydrated formaldehyde ((1.4 ± 0.2) × 107 M?1 s?1); hydrated glyoxal ((2.4 ± 0.2) × 107 M?1 s?1); dimethyl malonate (CH3OC(O)CH2C(O)OCH3) ((1.28 ± 0.02) × 106 M?1 s?1); and dimethyl succinate (CH3OC(O)CH2CH2C(O)OCH3) ((1.37 ± 0.08) × 106 M?1 s?1) where the errors represent 2σ. For the two latter species, we also measured the temperature dependence of the corresponding rate constants. A correlation of these kinetics with the bond dissociation energy is also presented and discussed. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 539–547, 2001  相似文献   

6.
In the present study, iron oxide (α-Fe2O3) thin films with good adhesion on stainless steel substrates are deposited by liquid phase deposition (LPD) technique, which is additive and binder-free. Iron oxyhydroxide (FeOOH) thin films are formed by means of a ligand-exchange equilibrium reaction of metal-fluoro complex ions and an F?ions consuming reaction by using boric acid (H3BO3) as a scavenging agent. These films are annealed at 500 °C to get α-Fe2O3 thin films. The transformation from hydrophobic to hydrophilic nature of the films is observed due to annealing. The films are characterized by different techniques. The α-Fe2O3 film is checked for electrochemical supercapacitive performance in Na2SO3 solutions of various concentrations. Specific capacitance is calculated from cyclic voltammetry at numerous scan rates (5–200) mV s?1. The highest obtained value of specific capacitance is 582 F g?1 at 5 mV s?1 for 0.5 M Na2SO3 electrolyte. The maximum values of specific power and specific energy are 6.9 and 53.4 Wh kg?1 from the charge-discharge curve at the current density 2 mA cm?2 in 0.5 M Na2SO3 electrolyte.  相似文献   

7.
Evidence is given that the yields of photo-oxidation for free pyrene anions (Π?) and for anions associated with sodium cations (Π?Na+) in liquid THF and THP are connected to the effective threshold values for the photo- ionization of free and associated pyrene anions. Our experimental results indicate that the phenomena observed by Fisher and Rämme after flash-photolysis of sodium pyrenide in THF which have led to the postulation of transient; sodium atoms Na0 may be interpreted as being caused by benzene or toluene negative ions.  相似文献   

8.
Nickel(O) Complexes with the Anionic Ligands E (C6H5)?3 (E = Si, Ge, Sn) Complexes of the type MeIXNi(EPH3)X(THF)Y are formed from Ni(COD)2 by substitution with MeIEPh3 (E = Si, Ge, Sn) in THF (COD = Cyclooctadiene-1,5). In the case of the ligands GePh?3 and SnPh?3 nickel(O) is fourfold coordinated, but in the case of SiPh?3 it is only two-fold or threefold coordinated. Products of the reaction between Ni(COD)2 and LiPbPh3 are Li2Ni(COD)Ph2(THF)5 and Ph3PbPbPh3. The 1H-n.m.r., 29Si-n.m.r., and 119Sn-Mössbauer spectra of the complexes MeIXNi(EPh3)X(THF)Y are compared with the spectra of the corresponding alkali compounds MeIEPh3. The magnetic anisotropy effects of the atomes Ge, Sn, Pb and Ni are of high importance for 1H- and 29Si-chemical shifts. The donor action of SnPh?3 is shown by the Mössbauer spectrum of Na4Ni(SnPh3)4(THF)4. But there is no direct evidence of π-back donation in the compound.  相似文献   

9.
The decomposition of polybromostyryl carbanions (PBS?), obtained by anionic polymerization of 4-bromostyrene in tetrahydrofuran (THF), was investigated in the dark in a temperature range of ?6–?21°C. It was accompanied by the evolution of bromine anions and by the formation of polymeric allylic carbanions (λmax = 575 nm; εmax = 6800 eq?1·L·cm?1). The reaction mechanism was elucidated. The rate constant of the unimolecular rate-determining step of the process was 1.3 × 10?5 s?1 and 9.7 × 10?5 s?1 at ?21 and ?6°C, respectively. Its apparent energy of activation Eapp = 18.38 Kcal/mol. The polybromostyrenes with allylic carbanions at their ends may decompose further. Their “dark” decomposition yielded 1,3-butadiene-1,3-diphenyl-macromers. The mechanisms of decomposition of the PBS? carbanions and the dark decomposition of the polybromostyryl allylic carbanions are analogous. The rate constant of the latter process was 2.5 × 10?6 s?1 at ?6°C. The anionic polymerization of prepared macromers can be initiated in THF at ?78°C by α-methylstyryl carbanions, which do not react, however, with PBS? carbanions. “Comblike” polymacromers were prepared in which each branch had a molecular weight of about 50,000. The overall molecular weight of the polymacromer was estimated to be about 1 × 106. It has been assumed that the 2–1 mode of addition to the diene group of the macromer is predominant during its polymerization. The 3–4 mode of addition followed by proton shift represents the termination step. The 4–3 mode of addition was ruled out on the basis of spectroscopic evidence.  相似文献   

10.
P2‐type layered oxides suffer from an ordered Na+/vacancy arrangement and P2→O2/OP4 phase transitions, leading them to exhibit multiple voltage plateaus upon Na+ extraction/insertion. The deficient sodium in the P2‐type cathode easily induces the bad structural stability at deep desodiation states and limited reversible capacity during Na+ de/insertion. These drawbacks cause poor rate capability and fast capacity decay in most P2‐type layered oxides. To address these challenges, a novel high sodium content (0.85) and plateau‐free P2‐type cathode‐Na0.85Li0.12Ni0.22Mn0.66O2 (P2‐NLNMO) was developed. The complete solid‐solution reaction over a wide voltage range ensures both fast Na+ mobility (10?11 to 10?10 cm2 s?1) and small volume variation (1.7 %). The high sodium content P2‐NLNMO exhibits a higher reversible capacity of 123.4 mA h g?1, superior rate capability of 79.3 mA h g?1 at 20 C, and 85.4 % capacity retention after 500 cycles at 5 C. The sufficient Na and complete solid‐solution reaction are critical to realizing high‐performance P2‐type cathodes for sodium‐ion batteries.  相似文献   

11.
Carbon-coated Na3V2(PO4)3 (NVP) was prepared by a standard sol–gel procedure. The apparent diffusion coefficients of sodium ions in the rhombohedral NVP have been determined by different techniques such as galvanostatic intermittent titration technique (GITT) and cyclic voltammetry (CV). It was found that the apparent diffusion coefficients range from 6?×?10?13 cm2 s?1 to 2?×?10?15 cm2 s-1. These sodium ion apparent diffusion coefficients follow a similar trend as observed for lithium ions in the closely related monoclinic modification of Li3V2(PO4)3, demonstrating a minimum at the potential where the ion extraction/insertion occurs.  相似文献   

12.
Rate coefficients of reactions of Cl atoms with cyclic ethers, tetrahydropyran (THP), tetrahydrofuran (THF), and dihydrofurans (2,5‐DHF and 2,3‐DHF) have been measured at 298 K using a relative rate method. The relative rate ratios for THP and THF are 0.80 ± 0.05 and 0.80 ± 0.08, respectively, with n‐hexane as the reference molecule. The relative rate ratios for THF and 2,5‐DHF with n‐pentane as the reference molecule are 0.95 ± 0.07 and 1.73 ± 0.06, respectively, and for 2,5‐DHF with 1‐butene as reference is 1.38 ± 0.05. The average values of the rate coefficients are (2.52 ± 0.36), (2.50 ± 0.39), and (4.48 ± 0.59) × 10?10 cm3 molecule?1 s?1 for THP, THF, and 2,5‐DHF, respectively. The errors quoted here for relative rate ratios are 2σ of the statistical variation in different sets of experiments. These errors, combined with the reported errors of the reference rate coefficients using the statistical error propagation equation, are the quoted errors for the rate coefficients. In the case of 2,3‐DHF, after correcting for the dark reaction with CH3COCl and assuming no interference from other radical reactions, a relative rate ratio of 0.85 ± 0.16 is obtained with respect to cycloheptene, corresponding to a rate coefficient of (4.52 ± 0.99) × 10?10 cm3 molecule?1 s?1. Unlike cyclic hydrocarbons, there is no increase with increasing number of CH2 groups in these cyclic ethers whereas there is an increase in the rate coefficient with unsaturation in the ring. An attempt is also made to correlate the rate coefficients of cyclic hydrocarbons and ethers with the molecular size as well as HOMO energy.  相似文献   

13.
The values of pseudo-first-order rate constants (k obs) for the acetolysis of phthalic anhydride (PAn) increase from 6.60?×?10?7 to 31.5?×?10?7?s?1 with the increase in temperature from 30 to 50?°C. These values of k obs give activation parameters ?H* and ?S* as 14.4?±?0.4?kcal?mol?1 and ?39.1?±?1.3?cal?K?1?mol?1, respectively. The values of k obs remain essentially unchanged with the increase in the content of CS (CS?=?CH3CN or THF) from 0 to 40?% v/v in mixed AcOH?CCS solvents. These observations have been explained qualitatively.  相似文献   

14.
Contributions to the Chemistry of Phosphorus. 143. Li4P26 and Na4P26, the First Salts with Hexacosaphosphid(4?) Ions The hexacosaphosphides Li4P26 ( 1 ) and Na4P26 ( 2 ) are formed besides other polyphosphides in the reaction of white phosphorus with lithium dihydrogenphosphide or sodium. 1 also results from the decomposition of Li2HP7 in tetrahydrofuran at room temperature and can be obtained pure as a crystalline solvent adduct Li4P26 · 16 THF. According to 2D?31P-NMR spectroscopic investigations the P264? ion is a conjucto-phosphane of two P7(5)?-and two P9(3)?-unit groups with structures analogous to norbornane and deltacyclane, respectively.  相似文献   

15.
Electron capture processes in the H+?Na(3s) and H+?Na(3p) collisions are experimentally investigated in the 0.3–3 keV energy range using a crossed beam experiment. The excited Na(3p) target is produced with a well-defined alignment using laser pumping. The time of flight technique enables the identification of all the H(n)+Na+ channels populated in the collision. Total cross section ratios σ3p (n=2)/σ3s (n=2),σ3p (n=3)/σ3s (n=2) and σ3s (n=3)/σ3s (n=2) for the production of H(n=2) and H(n=3) are measured in the H+?Na (3s) and H+?Na (3p) collisions. They reveal a strong dominance of the production of H(n=2) in the H+?Na(3p) collision, especially for energies below 1 keV.  相似文献   

16.
Dissociative ionisation of Na2 via the 3s 3d 1Σ g and1Π g states has been studied in the near threshold energy regime up to 120 meV above the three particle (Na+ + Na(3s) +e ?) break up limit. A pulsed, cold molecular beam, pulsed laser 2 colour 3 photon resonantly enhanced multiphoton ionisation, and kinetic energy analysis of the fragments by a time of flight method (KETOF) is used. As series of vibrational levels in the two intermediate 3s 3d Rydberg states are excited, slow Na+ fragments are observed with a maximum kinetic energy given by the excess energy of the 2 + 1 photon process above threshold, thus confirming a direct dissociative ionisation process. The intensity distribution of the Na+ fragments shows a very pronounced maximum at zero kinetic energy, its shape differing somewhat for the1Σ g and1Π g intermediate states. Also observed is a strong signal of fast fragments arising from a typical 4 photon process which leads to dissociation of Na 2 + molecules in their electronic ground state.  相似文献   

17.
Mono‐ and dianions of 2‐tert‐butyl‐3a2‐azapentabenzo[bc,ef,hi,kl,no]corannulene ( 1 a ) were synthesized by chemical reduction with sodium and cesium metals, and crystallized as the corresponding salts in the presence of 18‐crown‐6 ether. X‐ray diffraction analysis of the sodium salt, [{Na+(18‐crown‐6)(THF)2}3{Na+(18‐crown‐6)(THF)}( 1 a 2?)2], revealed the presence of a naked dianion. In contrast, controlled reaction of 1 a with Cs allowed the isolation of singly and doubly reduced forms of 1 a , both forming π‐complexes with cesium ions in the solid state. In [{Cs+(18‐crown‐6)}( 1 a ?)]?THF, asymmetric binding of the Cs+ ion to the concave surface of 1 a ? is observed, whereas in [{Cs+(18‐crown‐6)}2( 1 a 2?)], two Cs+ ions bind to both the concave and convex surfaces of the dianion. The present study provides the first successful isolation and characterization of the reduced products of heteroatom‐containing buckybowl molecules.  相似文献   

18.
Reactions of SmII, TbIII, TmII, YbII, and LuIII iodides with 2,2′-bipyridyllithium in THF afford [Li(THF)4][Ln(bipy) n ] complexes (n=3 or 4) containing trivalent lanthanides. X-ray structural analysis demonstrated that in the crystalline state, the Yb derivative has the ionic structure, [Li(THF)4]+[Yb(bipy)3]?. In THF solutions, the reversible ligand exchange between metal atoms occurs to yield neutral compounds [Ln(bipy) n?1(THF) x ] and [Li(bipy)(THF) y ]. A decrease in the temperature shifts the equilibrium to ionic pairs.  相似文献   

19.
The kinetics of the reduction of octacyanomolybdate(V) and octacyanotungstate(V) by sulphite ions has been studied over a wide pH range. The reaction is catalysed by alkali metal ions. The rate law is found to be of the form:
The third order rate constants at [OH?] = 0.05 mol dm?3 for the reduction of Mo(CN)83? and W(CN)3?8 were determined as 6.2 x 103dm6mol?2 s?1 and 22.3 dm6mol?2s?1 respectively at 298 K for A+ = Na+ while Ka for the hydrogen sulphite ion was determined as 2.4 x 10?8 mol dm?3. It was established that the reaction proceeds via an outer-sphere mechanism. An explanation for the alkali metal ion catalysis is proposed.  相似文献   

20.
Mixed Ligand Complexes of Nickel(0) and Cobalt(I) with the Anionic Ligands E(C6H5)3? (E ? Ge, Sn, Pb) Complexes of the general formula MINi(PPh3)3(EPh3)(THF)x (E ? Ge[Ia], Sn[Ib], Pb[Ic]) and MI3Ni(PPh3)(EPh3)3(THF)x (E ? Ge[IIa], Sn[IIb]) are formed from (Ph3P)2Ni(C2H4) by substitution with MIEPh3. The analogous complexes of the ligand SiPh3? could not be prepared, because of the formation of SiPh4 from LiSiPh3 and coordinated PPh3. Attempts to synthesize a nickel(II) complex of the ligand SnPh3? had no success, only possible decomposition products of these compounds, like (nBu2PPh)2NiII(Ph)Cl and NaxNi°(PPh3)4?x(SnP4)x(THF)Y, were isolated. NaCoI(PPh3)2(SnPh3)2(THF)7 (IV) was prepared by the reaction of Co(PPh3)3Cl and NaSnPh3. 1H-NMR and 119Sn Mössbauer spectra show a higher donor action of SnPh3? in IIb than in Ib. This causes a stronger π-back donation Ni → P in the case of IIb. IV is a paramagnetic compound, the vis-spectrum is discussed using simple crystal field theory.  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