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 structure of the title compound, C6H18N22+.2Cl?, has been determined and has a centre of symmetry. The mol­ecule has strong intermolecular hydrogen bonding between each Cl? and an N—H bond [Cl?N = 3.012 (3) Å].  相似文献   

2.
Negative ion formation in CF2Cl2, CF3Cl and CFCl3 under low-energy electron impact has been investigated using a trochoidal monochromat The ions observed are F?, Cl?, FCl?, Cl2?, CFCl2? from CF2Cl2; F?, Cl?, FCl?, CF2Cl Quoting available thermochemical data, it can be shown that most of the observed negative ions arise from dissociative attachment processes. Appearance The extremely high yield of Cl? in CFCl3, which is observed at ε = 0.0 eV, will be discussed with regard to the lifetime of this molecule i  相似文献   

3.
The success of secondary ion mass spectrometry (SIMS) analyses depends largely on the ionization probability of the analyzed elements. The chemical state of the surface changes with the chemical nature and the concentration of implanted ions. The positive ionization probability can be enhanced by bombarding the surface with electronegative elements. In view of such an enhancement of the positive secondary ion yield, we present SIMS analyses carried out with O?, F?, Cl?, Br? and I? primary ion beams. Useful yields were experimentally determined for metal (Al, Ni, Cu, Ag and Ta) and semiconductor samples (Si, Ge, InP and GaAs). For metal samples, an enhancement of the useful yield under halogen bombardment, compared with O? bombardment, was observed for Ni, Cu and Ag under F? bombardment (enhancement of up to two orders of magnitude). For semiconductors, lower useful yields are obtained under halogen bombardment as compared with O? bombardment. The observed results are discussed in terms of the surface concentration of the implanted primary ion species and their electronegativity. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

4.
In the structure Ba12F19Cl5 [hexagonal space group P6 2m] the two chlorides on the sites Cl(1) and Cl(2) can partially be replaced by bromide ions. Single crystals of the type Ba12F19ClδBr5–δ with a chloride to bromide ratio up to 2 : 3 could be obtained by cooling a flux of 75 mol% BaF2 and 25 mol% BaX2 with X = Cl, Br. The crystal quality decreases with increasing bromide concentration. Structural parameters of five selected single crystals with different chloride/bromide ratio were studied by single crystal X-ray diffraction methods. The refined total Cl?/Br? population ratio in the crystals is close to the one of the flux. The lattice parameters and interatomic distances change in various ways, when the smaller chloride ion is replaced by the bigger bromide ion. The refinements show a statistical disorder on the halide sites with preferential bromide substitution on site Cl(1).  相似文献   

5.
Why is silicon hypervalent and carbon not? Or why is [Cl? CH3? Cl]? labile with a tendency to localize one of its axial C? Cl bonds and to largely break the other one, while the isostructural and isoelectronic [Cl? SiH3? Cl]? forms a stable pentavalent species with a delocalized structure featuring two equivalent Si? Cl bonds? Various hypotheses have been developed over the years focusing on electronic and steric factors. Here, we present the so‐called ball‐in‐a‐box model, which tackles hypervalence from a new perspective. This model reveals the key role of steric factors and provides a simple way of understanding the above phenomena in terms of different atom sizes. Our bonding analyses are supported by computation experiments in which we probe, among other things, the shape of the SN2 potential‐energy surface of Cl? attacking a carbon atom in the series of substrates CH3Cl, .CH2Cl, ..CHCl, and ...CCl. Our findings for ClCH3Cl? and ClSiH3Cl? are generalized to other Group 14 central atoms (Ge, Sn, and Pb) and axial substituents (F).  相似文献   

6.
In the synthesis of Pt/C catalysts via H2PtCl6 adsorption onto a carbon support, NH4Cl can be formed catalytically during the reduction of the precursor with H2 at 250°C. This compound favors the sintering of metal particles. This effect is likely due to the weakening of metal-support bonding because of NH4Cl adsorption on the Pt surface. The sources of nitrogen and chlorine atoms are basic surface sites of the support, which contain nitrogen atoms in their structure and adsorb Cl? ions from the precursor solution. This effect is typical of active carbons, whose surface contains chemically bound nitrogen as amino groups, and weakens as the Pt/N atomic ratio in the supported catalyst precursors is increased.  相似文献   

7.
Low molecular weight polyisobutylenes (PIB) with chlorine, olefin and succinic acid end‐groups were studied using direct analysis in real time mass spectrometry (DART‐MS). To facilitate the adduct ion formation under DART conditions, NH4Cl as an auxiliary reagent was deposited onto the PIB surface. It was found that chlorinated adduct ions of olefin and chlorine telechelic PIBs, i.e. [M + Cl]? up to m/z 1100, and the deprotonated polyisobutylene succinic acid [M? H]? were formed as observed in the negative ion mode. In the positive ion mode formation of [M + NH4]+, adduct ions were detected. In the tandem mass (MS/MS) spectra of [M + Cl]?, product ions were absent, suggesting a simple dissociation of the precursor [M + Cl]? into a Cl? ion and a neutral M without fragmentation of the PIB backbones. However, structurally important product ions were produced from the corresponding [M + NH4]+ ions, allowing us to obtain valuable information on the arm‐length distributions of the PIBs containing aromatic initiator moiety. In addition, a model was developed to interpret the oligomer distributions and the number average molecular weights observed in DART‐MS for PIBs and other polymers of low molecular weight. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
Low-energy reactive collisions between the negative molecular ion of a tetrachlorodibenzo-p-dioxin (TCDD) and oxygen inside the collision cell of a triple-stage quadrupole mass spectrometer produce a substitution ion [M ? Cl + O]?, a phenoxide ion [C6H4-nO2Cln], [M ? HCl], and Cl? by which 1,2,3,4-, 1,2,3,6/1,2,3,7- and 2,3,7,8-TCDD isomers can be distinguished either directly or on the basis of intensity ratios. The collision conditions have an important effect on the relative abundances. Energy- and pressure-resolved curves show that the ions formed by a collisionally activated reaction (CAR) process, i.e. [M ? Cl + O]? and [C6H4-n,O2Cln], are favoured by a high pressure of oxygen (3-6 mTorr) (1 Torr = 133.3 Pa) and a low collision energy (0.1-7 eV), whereas the ions formed by a collisionally activated dissociation (CAD) process, i.e. [M ? HCl] and Cl?, are favoured by high pressure and high energy. By choosing a relatively low collision energy (5 eV) and high pressure (4 mTorr), the CAR and CAD ions can be clearly detected.  相似文献   

9.
Quantum-chemical calculations of the structures and thermodynamic stabilities of isomers of molecular clusters of lower Mo chlorides Mo13Cl24, Mo13Cl26, and Mo13Cl30, as well as of the Mo13Cl 26 ? anion of cuboctahedral or icosahedral structure, have been performed by the DFT B3LYP method with the use of the Lanl2dz basis set for the Mo atoms and the 631G(d) basis set for the Cl atoms. The highest symmetry polyhedra are considered: Mo13Cl30 and Mo13Cl24 have a structure of centered icosahedron and cuboctahedron with bridging Cl atoms sitting atop the midpoints of 30 and 24 edges, respectively. In both cases, calculations lead to a rapid increase in the Mo-Mo distance to the values that are evidence of the absence of metal-metal bonds; hence, the cluster structure, in its classical meaning, is not retained. Both the neutral cluster Mo13Cl26 and the Mo13Cl 26 ? anion—a symmetric icosahedron, in which Cl ions sit atop the centers of 20 faces and 6 Cl atoms are terminal, and a cuboctahedron, in which 14 Cl atoms are located on the three- and fourfold axes over the centers of molybdenum faces and 12 terminal Cl atoms are located on twofold axes. For both the neutral cluster and the anion, stable structures have been obtained. In both cases, the total energies are lower for the initial icosahedral configuration.  相似文献   

10.
The equilibrium geometrical parameters and frequencies of the normal vibrations of Ln2Cl 7 ? ions (Ln = La and Lu) and the enthalpies of the dissociation reactions Ln2Cl 7 ? → Ln2Cl6 + Cl? and Ln2Cl 7 ? → LnCl3 + LnCl 4 ? were calculated at the MP2 and MP4 levels (with regard to single, double, triple, and quadruple perturbations). The basis set superposition errors were eliminated by using the counterpoise approach. The potential energy surface of the Ln2Cl 7 ? ions was found to reach a minimum for a configuration with three bridging and four terminal Cl atoms (symmetry C 2). The terminal fragments LnCl2 show almost free inherent rotation. The lanthanide compression of the interatomic Ln-Cl distances differed for the bridging and terminal Ln-Cl bonds. The calculated enthalpies of dissociation of the Ln2Cl 7 ? ions were compared with data from high-temperature mass spectrometry.  相似文献   

11.
The Crystal Structure of the 1:1 Addition Compound between Antimony Trichloride and Diphenylammonium Chloride, SbCl3 · (C6H5)2NH2+Cl? The 1:1 addition compound between antimony trichloride and diphenylammoniumchloride SbCl3 · (C6H5)2NH2+Cl? crystallizes in the monoclinic space group P21/n with a = 5.668(8), b = 20.480(12), c = 14.448(17) Å, β = 110.4(1)° and Z = 4 formula units. Chains of SbCl3 molecules and anion cation chains are bridged by Cl ions and form square tubes. The coordination of the Sb atoms by Cl atoms by Cl atoms and Cl ions is distorted octahedral. Mean distances are Sb? Cl = 2.37 Å for Sb? Cl (3×), 3.09 Å for Sb…Cl? (2×) and 3.42 Å for Sb…Cl (1×). The Sb…Cl? contacts and hydrogen bonds NH…Cl? at 3.15 Å generate tetrahedral coordination of the Cl ions.  相似文献   

12.
The structures of the title compounds, [CuCl(C3H5N3S)4]Cl·H2O, (I), and [CuCl(C4H7N3S)4]Cl, (II), comprise square‐pyramidal Cu centres with four N‐bound organic ligands filling the base positions, a Cl atom in the apical position and a Cl? as a free counter‐ion. The cation and free chloride ion in (II) have fourfold crystallographic symmetry. Hydro­gen‐bonding associations from the 2‐amino H atoms dominate both structures, with the principal acceptors being the chlorides, although in (I), the N4 atoms are also involved. Furthermore, (I) is a hydrate, with the water mol­ecule participating in the hydrogen‐bonding network.  相似文献   

13.
The anionization of molecules supplied from the gas phase onto a negatively charged [Cl]? or [NO3]? ion donating surface has been investigated. The charged surface was prepared by exposing an aqueous solution of LiCl (or LiNO3) and polyethylene oxide to a high external field as is done in negative ion field desorption mass spectrometry. The ionization of some monosaccharides and adenosine by [Cl]? and [NO3]? attachment and of some acids by proton abstraction is reported.  相似文献   

14.
(PPh4)2[V2Cl9][VCl5] · CH2Cl2. Synthesis, I.R. Spectrum, and Crystal Structure The title compound was obtained by addition of CCl4 to a solution of PPh4Cl and excess VCl4 in CH2Cl2. It forms black crystals which are light and moisture sensitive. The i.r. spectrum is reported. The crystal structure was determined by X-ray diffraction (3044 independent reflexions, R = 0.063). Crystal data: triclinic, space group P1 , Z = 2, a = 1186, b 1325, c = 1995 pm, α 97.5, β 105.6°, γ 93.4°. The structure consists of PPh4+ cations, V2Cl9? and VCl5? anions and CH2Cl2 molecules. The V2Cl9? ions consist of face-sharing octahedra with a long V…?V distance of 333 pm; the VCl5? ions form nearly ideal trigonal bipyramids with V? Cl bonds of 228 pm (axial) and 218 pm (equatorial). Both anions deviate only marginally from D3h symmetry. Half of the cations is arranged to (PPh4+)2 pairs about inversion centers.  相似文献   

15.
Two diquaternary ammonium chloride salts have been used to examine the roles of solvent and counterion in determination of the degree of ion pairing in solution and the resultant charge state distributions in electrospray ionization mass spectrometry (ESI-MS). Three series of solvents, that is, alcohol, polar aprotic, and chlorinated solvents, have been employed to test the influence of solvent polarity and other parameters on the desorption behavior of diquaternary ammonium ions observed in ESI-MS. Solvents of higher polarity were found to yield gas-phase ions of higher charge states, in accordance with their reduced tendency toward ion pairing in solution. Counterion effects were investigated via the following approaches: (1) increase the diquaternary ammonium salt concentration; (2) increase the concentration of an external electrolyte that contained the common counterion Cl?; (3) replace Cl? with trifluoroacetate (TFAc ?); (4) increase the concentration of an external electrolyte that contained TFAc?. These experiments indicate that variation of the specific counterion employed alters the degree of influence that the counterion exerts (via ion pairing) on electrospray ionization mass spectra. Increasing amounts of trifluoroacetate ions in a variety of solvent systems invariably led to a progressive shift of the observed ESI-MS charge states of diquaternary ammonium ions toward lower values.  相似文献   

16.
Thermodynamic data, ΔH n-1, n o and ΔS n-1, n o, for clustering reactions of halide ions X?(X = F, Cl, Br, and I) with N2Owere measured with a pulsed electron beam high-pressure mass spectrometer. In contrast to the fact that CO2 forms a covalent bond with the fluoride ion to yield the fluoroformate ion, FCO2 ?, the interaction between F? and N2O is mainly electrostatic. It was found that the cluster ions F? (N2O)n complete the first shell at n = 6, thus forming an octahedral structure. The difference between F—CO2 ? and F? ... N2O is discussed in terms of Coulombic, exchange, and charge-transfer interactions. The X? (N2O)2 clusters (X = Cl, Br and I) are found to be of C2h symmetry, while F? (N2O)2 is of a twisted form and is slightly asymmetric due to a slight participation of covalency (charge transfer) in the core ion F? ... N2O.  相似文献   

17.
《中国化学快报》2020,31(10):2725-2729
The porous g-C3N4 (PCN) nanosheets are successfully synthesized and further modified with nano-sized Ag by a simple wet-chemical process. Interestingly, the Ag-modified porous g-C3N4 (Ag-PCN) nanosheets exhibit competitive fluorescence detection performance of chloride ion (Cl) in aqueous solution. Under the optimized conditions, the concentration of Cl could be quantitative analyzed with the Ag-PCN in a wide detection range from 0.5 mmol/L to 0.1 mol/L, with a low detection limitation of 0.06 mmol/L. It is confirmed that the fluorescence of PCN could be effectively decayed by the photoinduced charge transfer via the adsorbed Cl for trapping holes, mainly by means of the time-resolved fluorescence and surface photovoltage spectra. The porous structure and modified Ag promote the adsorption of Cl on resulting Ag-PCN, leading to excellent fluorescence detection for Cl. This work provides a feasible route to develop a fluorescence detection of Cl with g-C3N4 nanosheets in environment water.  相似文献   

18.
Given the biological importance of organic cations, the facilitated transport of organic ion pairs could find many applications. Calix[6]arene tris(thio)ureas, which possess a cavity that can accommodate primary ammonium ions, can not only act as carriers for Cl?/NO3? antiport but can also perform the cotransport of PrNH3Cl. Transport was monitored by fluorescence spectroscopy and the presence of the different species inside the vesicles was characterized by 1H and 35Cl NMR experiments involving shift reagents. The cotransport of PrNH3Cl was also observed by receptors deprived of a cavity, but the presence of the cavity conveys an advantage, as the cotransport by calix[6]arenes was observed to be more efficient than the Cl?/NO3? antiport, which is not the case with receptors without a cavity. The role played by the cavity was further highlighted by the disappearance of this advantage when using a bulky ammonium ion, which cannot be complexed within the cavity.  相似文献   

19.
Developing high‐performance all‐solid‐state batteries is contingent on finding solid electrolyte materials with high ionic conductivity and ductility. Here we report new halide‐rich solid solution phases in the argyrodite Li6PS5Cl family, Li6?xPS5?xCl1+x, and combine electrochemical impedance spectroscopy, neutron diffraction, and 7Li NMR MAS and PFG spectroscopy to show that increasing the Cl?/S2? ratio has a systematic, and remarkable impact on Li‐ion diffusivity in the lattice. The phase at the limit of the solid solution regime, Li5.5PS4.5Cl1.5, exhibits a cold‐pressed conductivity of 9.4±0.1 mS cm?1 at 298 K (and 12.0±0.2 mS cm?1 on sintering)—almost four‐fold greater than Li6PS5Cl under identical processing conditions and comparable to metastable superionic Li7P3S11. Weakened interactions between the mobile Li‐ions and surrounding framework anions incurred by substitution of divalent S2? for monovalent Cl? play a major role in enhancing Li+‐ion diffusivity, along with increased site disorder and a higher lithium vacancy population.  相似文献   

20.
The title compound, {(C7H15N2Cl)2[Cd3Cl10]·4H2O}n, consists of 1‐chloromethyl‐1,4‐diazoniabicyclo[2.2.2]octane dications, one‐dimensional inorganic chains of {[Cd3Cl10]4−} anions and uncoordinated water molecules. Each of the two independent CdII ions, one with site symmetry 2/m and the other with site symmetry m, is octahedrally coordinated by chloride ions (two with site symmetry m and one with site symmetry 2), giving rise to novel polymeric zigzag chains of corner‐sharing Cd‐centred octahedra parallel to the c axis. The organic cations, bisected by mirror planes that contain the two N atoms, three methylene C atoms and the Cl atom, are ordered. Hydrogen bonds (O—H...Cl and O—H...O) between the water molecules (both with O atoms in a mirror plane) and the chloride anions of neighbouring chloridocadmate chains form a three‐dimensional supramolecular network.  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