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 515 毫秒
1.
The metastable type-II clathrate Na24-δGe136 was obtained from Na12Ge17 by applying a two-step procedure. At first, Na12Ge17 was reacted at 70 °C with a solution of benzophenone in the ionic liquid (IL) 1,3-dibutyl-2-methylimidazolium-bis(trifluoromethylsulfonyl) azanide. The IL was inert towards Na12Ge17, but capable of dissolving the sodium salts formed in the redox reaction. By annealing at 340 °C under an argon atmosphere, the X-ray amorphous intermediate product was transformed to crystalline Na24-δGe136 (δ≈2) and α-Ge in an about 1 : 1 mass ratio. The product was characterized by X-ray powder diffraction, chemical analysis, and 23Na solid-state NMR spectroscopy. Metallic properties of Na24-δGe136 were revealed by a significant Knight shift of the 23Na NMR signals and by a Pauli-paramagnetic contribution to the magnetic susceptibility. At room temperature, Na24-δGe136 slowly ages, with a tendency to volume decrease and sodium loss.  相似文献   

2.
Na12Ge17 is prepared from the elements at 1025 K in sealed niobium ampoules. The crystal structure reinvestigation reveals a doubling of the unit cell (space group:P21/c; a = 22.117(3)Å, b = 12.803(3)Å, c = 41.557(6)Å, β = 91.31(2)°, Z = 16; Pearson code: mP464), furthermore, weak superstructure reflections indicate an even larger C‐centred monoclinic cell. The characteristic structural units are the isolated cluster anions [Ge9]4— and [Ge4]4— in ratio 1:2, respectively. The crystal structure represents a hierarchical cluster replacement structure of the hexagonal Laves phase MgZn2 in which the Mg and Zn atoms are replaced by the Ge9 and Ge4 units, respectively. The Raman spectrum of Na12Ge17 exhibits the characteristic breathing modes of the constituent cluster anions at ν = 274 cm—1 ([Ge9]4—) and ν = 222 cm—1 ([Ge4]4—) which may be used for identification of these clusters in solid phases and in solutions. Raman spectra further prove that Na12Ge17 is partial soluble both in ethylenediamine and liquid ammonia. The solution and the solid extract contain solely [Ge9]4—. The remaining insoluble residue is Na4Ge4. By heating the solvate Na4Ge9(NH3)n releases NH3 and decomposes irreversibly at 742 K, yielding Na12Ge17 and Ge.  相似文献   

3.
Systematic studies in the quaternary system Na/Ge/Sb/Te yielded the new compound Na9Sb[Ge2Te6]2. Its crystal structure is isotypic to Na9Sb[Ge2Se6]2 (space group C2/c with a = 9.541(2), b = 26.253(7), c = 7.5820(18) Å and β = 122.233(15)°, Z = 2). The structure is characterized by Ge–Ge dumbbells that are octahedrally coordinated by Te, forming ethane‐like [Ge2Te6]6– anions. Cation sites are occupied by Na+ as well as shared by Na+ and Sb3+. Na9Sb[Ge2Te6]2 is formally obtained from the reaction of one equivalent Na8[Ge4Te10] and one equivalent NaSbTe2. In contrast to members of the metastable solid solution series (NaSbTe2)1–x(GeTe)x, Na9Sb[Ge2Te6]2 is a thermodynamically stable compound. It is a semiconductor with a bandgap of 1.51 eV.  相似文献   

4.
The GeIV chlorometallate complexes, [EMIM]2[GeCl6], [EDMIM]2[GeCl6] and [PYRR]2[GeCl6] (EMIM=1‐ethyl‐3‐methylimidazolium; EDMIM=2,3‐dimethyl‐1‐ethylimidazolium; PYRR=N‐butyl‐N‐methylpyrrolidinium) have been synthesised and fully characterised; the first two also by single‐crystal X‐ray diffraction. The imidazolium chlorometallates exhibited significant C?H???Cl hydrogen bonds, resulting in extended supramolecular assemblies in the solid state. Solution 1H NMR data also showed cation–anion association. The synthesis and characterisation of GeII halometallate salts [EMIM][GeX3] (X=Cl, Br, I) and [PYRR][GeCl3], including single‐crystal X‐ray analyses for the homologous series of imidazolium salts, are reported. In these complexes, the intermolecular interactions are much weaker in the solid state and they appear not to be significantly associated in solution. Cyclic‐voltammetry experiments on the GeIV species in CH2Cl2 solution showed two distinct, irreversible reduction waves attributed to GeIV–GeII and GeII–Ge0, whereas the GeII species exhibited one irreversible reduction wave. The potential for the GeII–Ge0 reduction was unaffected by changing the cation, although altering the oxidation state of the precursor from GeIV to GeII does have an effect; for a given cation, reduction from the [GeCl3]? salts occurred at a less cathodic potential. The nature of the halide co‐ligand also has a marked influence on the reduction potential for the GeII–Ge0 couple, such that the reduction potentials for the [GeX3]? salts become significantly less cathodic when the halide (X) is changed Cl→Br→I.  相似文献   

5.
The mixed silicide‐germanides Li12Si7–xGex, Na7LiSi8–zGez, and Li3NaSi6–vGev which could serve as potential precursors for Si1–xGex materials were synthesized and characterized by X‐ray diffraction methods. The full solid solution series Li12Si7–xGex (0 ≤ x ≤ 7) is easily accessible from the elements and features preferential occupation of the more negatively charged crystallographic tetrel positions by Ge, which is the more electronegative element. In case of Na7LiSi8–zGez a broad solid solution range of 1.3 ≤ x ≤ 8 is available but the ternary silicide Na7LiSi8 could not be obtained by the tested methods of synthesis. The solubility of Ge in Li3NaSi6–vGev is highly limited to a maximum of v ≈ 0.5, and again the formally more negatively charged tetrel positions are preferred by Ge. Additionally, the two crystallographic Li positions in Li12Si7 with unusually large displacement parameters can be partially substituted by Na in Li12–yNaySi7 with 0 ≤ y ≤ 0.6. The statistical mixing of Li and Na in this solid solution contrasts the typical ordering of Li and Na in most ternary tetrelides.  相似文献   

6.
The 1H nuclear magnetic resonance (1H-NMR) spectrum is a useful tool for characterizing the hydrogen bonding (H-bonding) interactions in ionic liquids (ILs). As the main hydrogen bond (H-bond) donor of imidazolium-based ILs, the chemical shift (δH2) of the proton in the 2-position of the imidazolium ring (H2) exhibits significant and complex solvents, concentrations and anions dependence. In the present work, based on the dielectric constants (ϵ) and Kamlet-Taft (KT) parameters of solvents, we identified that the δH2 are dominated by the solvents polarity and the competitive H-bonding interactions between cations and anions or solvents. Besides, the solvents effects on δH2 are understood by the structure of ILs in solvents: 1) In diluted solutions of inoizable solvents, ILs exist as free ions and the cations will form H-bond with solvents, resulting in δH2 being independent with anions but positively correlated with βS. 2) In diluted solutions of non-ionzable solvents, ILs exist as contact ion-pairs (CIPs) and H2 will form H-bond with anions. Since non-ionizable solvents hardly influence the H-bonding interactions between H2 and anions, the δH2 are not related to βS but positively correlated with βIL.  相似文献   

7.
8.
The behavior of acids (citric acid, nitric acid, oxalic acid, tartaric acid) as a mobile phase and imidazolium ionic liquids (the bromides, tetrafluoroborates and hexafluorophosphates of 1‐ethyl, 1‐butyl, and 1‐hexyl‐3‐methylimidazolium) as additives in ion exchange chromatography for cations (Na+, K+, Mg2+, Ca2+) separation were studied. The results showed that nitric acid and 1‐hexyl‐3‐methyl‐imidazolium hexafluorophosphate offered the most interesting features in the separation of cations, such as lower retention time and better resolution. The selected optimal conditions were achieved by adding 0.10 mM 1‐hexyl‐3‐methyl‐imidazolium hexafluorophosphate in 4.0 mM HNO3 mobile phase for the separation of four cations with the flow rate of 0.9 mL/min at room temperature (25°C). The linear regression equations of Na+, K+, Mg2+, Ca2+ were = 4.4763c  + 0.0209, = 3.8903c  – 0.0087, = 6.3974c  – 0.0173, and = 7.601c  – 0.0339 and the limits of detection of Na+, K+, Mg2+, Ca2+ were 0.296, 4.98, 0.0970, and 1.22 μg/L, respectively. In this work, four cations in samples were successfully detected.  相似文献   

9.
A series of nine different known ionic liquids or low melting salts was synthesised and purified. They are composed of the [NTf2] (bis(trifluoromethane)sulfonimide), [OTf] (trifluoro-methane-sulfonate), or [B(CN)4] (tetracyanidoborate) anion and [Ph4P]+ (tetraphenylphosphonium), [Ph3BzP]+ (triphenylbenzyl phosphonium), [nBu4P]+ (tetra-nbutylphosphonium), [nBuPh3P]+ (tri-phenyl-nbutylphosphonium), [nBu4N]+ (tetra-nbutylammonium), or the [PPN]+ (bis(triphenylphosphine)iminium) cation. Precise vapour pressure data and enthalpies of vaporisation were measured using the Quartz Crystal Microbalance (QCM) method and evaluated. Structure-property relations are established using the obtained data as well as literature known data of ILs with alkyl-substituted imidazolium cations. It turns out that ILs with the tetracyanidoborate anion have even higher values of the enthalpy of vaporisation than those with the common [NTf2] or [OTf] anion and therefore are even less volatile.  相似文献   

10.
Isotope clusters in library electron ionization mass spectra of germanes often appear a few u lower than theoretically expected from elemental composition; for example, the dominant peak of the Ge4H10+ pattern is shifted 8 u down. This phenomenon is due to combinations of three essential components: the molecular ion Ge n H2n+2+ and two products of hydrogen elimination, Ge n H+ and Ge n +. Using these components, isotope clusters can be accurately projected for germanium hydrides from Ge2H6 up to Ge5H12.  相似文献   

11.
In a combined experimental and theoretical approach, the interactions of valinomycin (Val), macrocyclic depsipeptide antibiotic ionophore, with sodium cation Na+ have been investigated. The strength of the Val–Na+ complex was evaluated experimentally by means of capillary affinity electrophoresis. From the dependence of valinomycin effective electrophoretic mobility on the sodium ion concentration in the BGE (methanolic solution of 20 mM chloroacetic acid, 10 mM Tris, 0–40 mM NaCl), the apparent binding (stability) constant (Kb) of the Val–Na+ complex in methanol was evaluated as log Kb = 1.71 ± 0.16. Besides, using quantum mechanical density functional theory (DFT) calculations, the most probable structures of the nonhydrated Val–Na+ as well as hydrated Val–Na+·H2O complex species were proposed. Compared to Val–Na+, the optimized structure of Val–Na+·H2O complex appears to be more realistic as follows from the substantially higher binding energy (118.4 kcal/mol) of the hydrated complex than that of the nonhydrated complex (102.8 kcal/mol). In the hydrated complex, the central Na+ cation is bound by strong bonds to one oxygen atom of the respective water molecule and to four oxygens of the corresponding C=O groups of the parent valinomycin ligand.  相似文献   

12.
In recent years, great progress has been made in the dissolution of cellulose with ionic liquids (ILs). However, the mechanism of cellulose dissolution, especially the role the IL cation played in the dissolution process, has not been clearly understood. Herein, the mixtures of cellulose with a series of imidazolium‐based chloride ionic liquids and 1‐butyl‐3‐methyl pyridinium chloride ([C4mpy]Cl) were simulated to study the effect that varying the heterocyclic structure and alkyl chain length of the IL cation has on the dissolution of cellulose. It was shown that the dissolution of cellulose in [C4mpy]Cl is better than that in [C4mim]Cl. For imidazolium‐based ILs, the shorter the alkyl chain is, the higher the solubility will be. In addition, an all‐atom force field for 1‐allyl‐3‐methyl imidazolium cation ([Amim]+) was developed, for the first time, to investigate the effect the electron‐withdrawing group within the alkyl chain of the IL cation has on the dissolution of cellulose. It was found that the interaction energy between [Amim]+ and cellulose was greater than that between [C3mim]+ and cellulose, indicating that the presence of electron‐withdrawing group in alkyl chain of the cation enhanced the interaction between the cation and cellulose due to the increase of electronegativity of the cations. These findings are used to assess the cationic effect on the dissolution of cellulose in ILs. They are also expected to be important for rational design of novel ILs for efficient dissolution of cellulose.  相似文献   

13.
The synthesis, structure, and physical properties of ionic liquids (IL) bearing the novel [Al(O–C6H4–CN)4] ion as counterion to the commonly used [NR4]+, [PR4]+ and imidazolium ions are reported. Both the influence of the alkyl chain length as well as the functionalization with cyano groups is studied. These ILs are easily obtained by reaction of Ag[Al(O–C6H4–CN)4] with the corresponding ammonium, phosphonium, and imidazolium halides. The stability towards electrophilic cations was investigated. All prepared salts have a window for the liquid phase of ca. 200 °C and are thermally stable up to 450 °C. The solid‐state structures reveal only weak cation ··· anion and anion ··· anion interactions in accord with the observed low melting points (glass transition points).  相似文献   

14.
The first example of a heteropolyoxomolybdate containing palladium(IV) was isolated and characterized by X‐ray crystallography. The palladium(IV) hexamolybdate, K0.75Na3.75[PdMo6O24H3.5]·17H2O, was isolated from an aqueous solution at pH 4.5 in the space group P\bar{1} , a 10.790(2), b 12.244(3), c 14.086(3) Å, α 113.77(1), β 90.41(1),γ 107.86(1)°, and the structure was determined using X‐ray diffraction methods, refining to a residual of 0.0301 for 5334 reflections. A formal “[PdMo6O24H3]5–” subunit exhibits the basic Anderson structure, with two [PdMo6O24H3]5– cluster anions in the structure bridged by a hydrogen atom (formally an H+) situated on a center of symmetry to give a “[Pd2Mo12O48H7]9–” dimeric anion. The palladium(IV) atom occupies a slightly distorted octahedral environment, with Pd–O distances ranging from 1.968 to 2.009 Å.  相似文献   

15.
Abstract . With the attempt to synthesize Nb-enclosed germanium Zintl cluster ion in ethylenediamine (en) / toluene solution, the reaction of K4Ge9/Na4Ge9 with Nb(mes)2 (mes = mesitylene) gives unexpected 18-crown-6 cleavage product of [K14Na2(NbO2)2(C8H16O5)8](Ge5) (en) · solv in the form of very air-sensitive, light-brown, plate-like crystals. This unique compound crystallizes in monoclinic Pn (No. 7) space group, the cleavage of crown linked the 14 potassium, 2 sodium, and 2 niobium atoms into the bulky 2+ charged cation, which balances the [Ge5]2– anion. This compound represents the rare example of heteronuclear metal alkoxide / Zintl ion hybrid.  相似文献   

16.
Both cation and anion in the title compound, C2H12BN2+·I3, lie on a crystallographic mirror plane and are bound in the lattice by N—H⋯I hydrogen bonds, forming layers. Methyl‐H–borane‐H dihydride [–C—H(δ+)⋯(δ)H—B–] inter­actions between mol­ecules crosslink adjacent layers, giving `sandwich' stacking along the a axis.  相似文献   

17.
Application of ionic liquids as low-volatility plasticizers for PMMA   总被引:1,自引:0,他引:1  
Room temperature ionic liquids (ILs) based on imidazolium salts, were found to be excellent plasticizers for poly(methyl methacrylate), with improved thermal stability, and the ability to reduce glass transition temperatures to near 0 °C. Because ILs have environmentally benign properties, they can be used in place of traditional chemicals in numerous products and processes. In this work, PMMA was formulated using dioctyl phthalate, DOP, as a traditional plasticizer, and properties were compared to PMMA plasticized with two ILs: butyl methylimidazolium/hexafluorophosphate, [bmim+][PF6], and hexyl methylimidazolium/hexafluorophosphate, [hmim+][PF6]. Formulations incorporated up to 30 vol.% DOP and 50 vol.% ILs. Bulk and plasticized polymers were characterized for glass transition temperature, elastic modulus, and the thermal stability of the plasticizers.  相似文献   

18.
NaK alloy in contact with 15-crown-5 hexane solution became potassium sodide K+(15-crown-5)2Na. After the evaporation of hexane the crystalline solid product was analyzed by X-ray diffraction and the lattice parameters were calculated. The potassium sodide thus obtained could be easily dissolved in tetrahydrofuran. A deep blue solution containing sodium anions and complexed potassium cations was formed with a very low concentration of solvated electrons, i.e. of the order of 10−7 M. Potassium anions were not detected in this case. A new crystalline potassium sodide K+(DCH-24-crown-8)2Na was obtained using NaK alloy and dicyclohexano-24-crown-8 hexane solution.  相似文献   

19.
During the reaction of an aqueous solution of (H3O)2[B12H12] with Tl2CO3 anhydrous thallium(I) dodecahydro‐closo‐dodecaborate Tl2[B12H12] is obtained as colorless, spherical single crystals. It crystallizes in the cubic system with the centrosymmetric space group Fm$\bar{3}$ (a = 1074.23(8) pm, Z = 4) in an anti‐CaF2 type structure. Four quasi‐icosahedral [B12H12]2– anions (d(B–B) = 180–181 pm, d(B–H) = 111 pm) exhibit coordinative influence on each Tl+ cation and provide a twelvefold coordination in the shape of a cuboctahedron (d(Tl–H) = 296 pm). There is no observable stereochemical activity of the non‐bonding electron pairs (6s2 lone pairs) at the Tl+ cations. By neutralization of an aqueous solution of the acid (H3O)2[B12H12] with PbCO3 and after isothermic evaporation colorless, plate‐like single crystals of lead(II) dodecahydro‐closo‐dodecaborate hexahydrate Pb(H2O)3[B12H12] · 3H2O can be isolated. This compound crystallizes orthorhombically with the non‐centrosymmetric space group Pna21 (a = 1839.08(9), b = 1166.52(6), c = 717.27(4) pm, Z = 4). The crystal structure of Pb(H2O)3[B12H12] · 3H2O is characterized as a layer‐like arrangement. The Pb2+ cations are coordinated in first sphere by only three oxygen atoms from water molecules (d(Pb–O) = 247–248 pm). But a coordinative influence of the [B12H12]2– anions (d(B–B) = 173–181 pm, d(B–H) = 93–122 pm) on lead has to be stated, too, as three hydrogen atoms from three different hydroborate anions are attached to the Pb2+ cations (d(Pb–H) = 258–270 pm) completing their first‐sphere coordination number to six. These three oxygen and three hydrogen ligands are arranged as quite irregular polyhedron leaving enough space for a stereochemical lone‐pair activity (6sp) at each Pb2+ cation. Since additional intercalating water of hydration is present as well, both classical H–Oδ ··· +δH–O‐ and unconventional B–Hδ ··· +δH–O hydrogen bonds play a significant role in the stabilization of the entire crystal structure.  相似文献   

20.
Quantification of hygroscopicity of ionic liquids (ILs) is of great importance in both fundamental studies and practical applications of ILs. This study demonstrates an electrochemical method for effectively quantifying the hygroscopicity of ILs through electrochemically monitoring water contents absorbed into ILs. The measurements of water content absorbed into the ILs are performed with square wave voltammetry (SWV) based on the water‐induced enhancement of diffusion of solution‐dissolved potassium ferricyanide (K3Fe(CN)6) redox probe. For demonstration, two kinds of ILs with different hygroscopicity (i.e., hydrophilic Bmim+Gly? and hydrophobic Bmim+PF6?) are employed in this study. The dissolution of K3Fe(CN)6 redox probe into ILs is found to have little effect on the hygroscopicity of ILs. The hygroscopicity of ILs is thus able to be quantified by monitoring water content absorbed into ILs as a function of time when ILs are stored at room temperature and standard atmospheric pressure under 55 % relative humidity (RH). Under the conditions employed in this study, the hygroscopicity of Bmim+Gly? and Bmim+PF6? is determined to be 1.33 M per hour and 0.05 M per hour, respectively, which are almost consistent with those measured with Karl Fischer titration under the same conditions. The electrochemical method demonstrated in this study is experimentally simply and environmentally benign and may be potentially extended for general quantification of hygroscopicity of ILs.  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