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 major reactant ion in conventional ion mobility spectrometry (IMS) is the hydronium ion, H3O+ which is produced in the usual ionization sources such as corona discharge or radioactive sources. Using the hydronium reactant ion, mostly the analytes with proton affinity higher than that of water are ionized. A broader range of compounds can be detected by IMS if other alternative ionization channels, such as charge transfer from NO+, are employed. In this work we introduce a simple and novel method for producing NO+ as the major reactant ion in IMS. This was achieved by adding neutral NO to the corona discharge ionization source. The neutral NO was prepared via an additional discharge in an air stream, flowing into the corona discharge source. A curtain plate was mounted in front of the corona discharge to prevent the influence of the analyte on the production of NO+. Using this technique, the reactant ion could easily and quickly switch between the H3O+ and NO+. The performance of the new source was evaluated by recording ion mobility spectra of test compounds with both H3O+ and NO+ reactant ions.  相似文献   

2.
The behavior of potassium tetrachloropalladate(II) in media simulating biological fluids has been studied. In aqueous solutions of NaCl, the aquation rate is higher than the rate of chloro ligand introduction into the internal coordination sphere of palladium. In HCl solutions, on the contrary, the process of palladium chloro complex formation predominates. The latter is apparently due to protonation of water molecules composing aqua complexes. By means of the ZINDO/1 method, the substitution of ligands – water molecules and hydronium ion – in planar complexes of palladium(II) by chloride ion has been investigated. All complexes containing H2O and H3O+ ligands, other than [Pd(H2O)4]2+, have intramolecular hydrogen bonds. In [Pd(H2O)3(H3O)]3+ and trans-[Pd(H2O)2(H3O)Cl]2+, a “non-classic” symmetric hydrogen bond O ··· H ··· O is established (ZINDO/1, RHF/STO-6G*). By the first three steps the substitution of hydronium ion in the internal sphere of palladium atom is more favorable thermodynamically, compared to water molecules. Logarithms of stepwise stability constants of palladium(II) chloride complexes correlate linearly to enthalpies (ZINDO/1, PM3) of water substitution by chloride ion.  相似文献   

3.
19F, 1H NMR and DSC methods are used to study the ion mobility and phase transitions in hexafluoro complex compounds of tantalum(V), niobium(V), and titanium(IV) with the tetramethylammonium cation. A transformation of the NMR spectra observed in a temperature range 77(130)-450 K is found to be associated with a change in the type of ion motions in the anion and cation sublattices of the studied compounds. In a temperature range 170-450 K the main types of ion motions are isotropic reorientations of TaF6, NbF6, TiF6 octahedra and tetramethylammonium ions. Two endothermic effects with the maxima at 232.5 K and 256.5 K for [N(CH3)4]TaF6 and 235 K and 250 K for [N(CH3)4]NbF6, which are observed in the DSC curve in the range 170-400 K correspond to phase transitions that have almost no effect on the NMR spectral parameters the a temperature range 230-260 K. For the [N(CH3)4]2TiF6 compound, an endothermic effect at 422 K is observed, corresponding to the phase transition from the rhombohedral to the cubic modification.  相似文献   

4.
Complexes between crown ethers and quaternary ammonium cations have been studied by electrospray ionisation mass spectrometry (ESI-MS). The ESI-MS method has been shown to allow observation of not only stable inclusion complexes between large crown ethers and tetramethylammonium cation (e.g. [DB30C10 + (CH3)4N]+ ion) but also of unstable inclusion complexes between smaller crown ethers and quaternary ammonium cations which are difficult to observe by other methods, namely [18C6 + (CH3)4N]+ ion. Stability of the complexes between crown ethers containing aromatic ring and tetramethylammonium cation is enhanced by cation-Π interactions. The molecule of 18C6 does not contain aromatic rings, thus [18C6 + (CH3)4N]+ ion exists due to the formation of C–H···O hydrogen bonds. Such a complex is quite unusual, since C–H···O hydrogen bonds are very weak and usually coexist with other strong interactions.  相似文献   

5.
Metastable transitions arising from the loss of C2H2 and HF from the [C7H6F]+ ion have been investigated. Under standard operating conditions, the intensity ratio of the metastable peaks was approximately independent of the precursor of the [C7H6F]+ ion, indicating fragmentation from a common structure such as the symmetrical fluorotropylium ion. The variation of the intensity ratio with several instrumental parameters suggests that I(C2H2 loss)/I(HF loss) rises as the internal energy of the [C7H6F]+ ion falls. Possible interference from discrimination effects at the β-slit when comparing intensity ratios for first and second field free region are discussed.  相似文献   

6.
We examine theoretically the three channels that are associated with the detachment of a single water molecule from the aqueous clusters of the alkaline earth dications, [M(H2O) n ]2+, M = Mg, Ca, Sr, Ba, n ≤ 6. These are the unimolecular water loss (M2+(H2O) n?1 + H2O) and the two hydrolysis channels resulting the loss of hydronium ([MOH(H2O) n?2]+ + H3O+) and Zundel ([MOH(H2O) n?3]+ + H3O+(H2O)) cations. Minimum energy paths (MEPs) corresponding to those three channels were constructed at the Møller–Plesset second order perturbation (MP2) level of theory with basis sets of double- and triple-ζ quality. We furthermore investigated the water and hydronium loss channels from the mono-hydroxide water clusters with up to four water molecules, [MOH(H2O) n ]+, 1 ≤ n ≤ 4. Our results indicate the preference of the hydronium loss and possibly the Zundel-cation loss channels for the smallest size clusters, whereas the unimolecular water loss channel is preferred for the larger ones as well as the mono-hydroxide clusters. Although the charge separation (hydronium and Zundel-cation loss) channels produce more stable products when compared to the ones for the unimolecular water loss, they also require the surmounting of high-energy barriers, a fact that makes the experimental observation of fragments related to these hydrolysis channels difficult.  相似文献   

7.
Metastable abundance ratios have been measured involving four decomposition reactions of C3H6 radical cations formed from a variety of precursors. The ratios are quite similar in accord with extensive isomerization to a propene structure prior to fragmentation. Small, yet constant differences are observed for those C3H6 ions which have been shown to be formed as cyclic ions by ion cyclotron resonance studies. The differences are interpreted to reflect internal energy variations, which result because the initially formed ions have two different structures. The abundance ratios are shown to depend on ionizing energy, repeller voltage and accelerating voltage, but are independent of the degrees-of-freedom in the precursor as well as the number of steps necessary to produce the [C3H6] Despite small variations in metastable ratios, the classification of various [C3H6] ions can be achieved under a variety of conditions which affect the internal energy of the decomposing ions.  相似文献   

8.
Density functional theory (DFT) using SVWN5, B3LYP, B3P86, O3LYP, B3PW91, B1LYP, B971, MPW1PW91, PBE1PBE, BHandH, and BHandHLYP density functionals was employed to study the structural characteristics of the Y(H2O) 8 3+ yttrium aqua ion. The nonlocal hybrid GGA functionals show worse predictive ability in structural calculations of the Y(H2O) 8 3+ aqua ion compared to the relatively simple combined functional BHandH and to the simplest SVWN5 functional in LSDA theory.  相似文献   

9.
The effects of Nd(III)/Am(III) complexation with sulfate were studied by 1) re-examining existing data for the Am–SO4 system using more, advanced aqueous electrolyte models valid to high concentration to obtain reliable thermodynamic data for SO 4 2– complexes or ion interactions with Nd3+ and Am3+ and 2) conducting experimental solubility studies of NdPO4(c), an analog phase of AmPO 4 (c), a possibly important phase in high level nuclear wastes, in the presence of SO 4 2– to test the newly developed thermodynamic model and show the possible influence of sulfate in a repository environment. The data showed that the increase in the solubility of NdPO 4 (c) resulted primarily from the increase in ionic strength. Slightly higher observed Nd concentrations in the presence of sulfate, as compared with concentrations predicted at the experimental ionic strengths, resulted from the weak complexes or ion interactions involving Nd 3+ –SO 4 2– . The Pitzer ion interaction parameters, applicable to 0.5m sulfate, were obtained for Am 3+ –SO 4 2– from a reinterpretation of known solvent extraction data. These parameters are also consistent with literature data for Am 3+ /Na+ exchange and solvent extraction in the presence of sulfate. When used for the analogous Nd 3+ –SO 4 2– system to predict NdPO 4 (c) solubility in the presence of sulfate, they provided excellent agreement between the predicted and the observed solubilities, indicating that they can be reliably used to determine Nd 3+ or Am 3+ ion interactions with SO 4 2– in all ground waters where SO 4 2– is less than 0.5m  相似文献   

10.
The preparation and construction of tungsten bronze electrodes, responsive to sodium ions, are reported. The electrodes were made from tungsten bronze powder (Na0.4WO3), compressed into pellets with a resin binder and sealed in epoxy resin.The electrodes were found to respond almost instantenously to changes in sodium ion concentration and to be relatively stable with time. The pH dependence decreases as the sodium ion concentration increases, indicating suitability for solutions of sea water concentration.Unfortunately the electrodes show a response to potassium ions of apparently equal magnitude to the sodium response between 0.001 mol kg−1 and 0.1 mol kg−1. Preliminary tests indicate that they also respond to lithium, calcium, aluminium, cupric and tetramethylammonium ions at concentrations of 0.1 mol kg−1. At high sodium ion concentrations ( 1 mol kg−1) the effect of cross sensitivity is negligible for these ions. No attempts are made in this short paper to evaluate selectivity coefficients. Some data demonstrating the low but linear response to sodium ion are given. Tests carried out in DNA—saline-bearing solutions suggest that the possible suitability for marine or biological applications.  相似文献   

11.
Extensive computations were performed on aqueous clusters of monovalent sodium cation [Na+(H2O) n ; (n = 1–20)] using MP2/cc-pVTZ and density functional theory. The structure, energy, and coordination number (CN) preference of a large number of competing conformations of different complexes have been explored. For complexes up to n = 12, the CN 4 is most preferred while 5, 6 CNs are favored in case of larger complexes containing up to 20 water molecules. These results are in very good agreement with experimental observations. The strength of hydrogen bonding among the waters coordinated to the Na+ ion is found to play a major role in the stability of the complexes. The varying preferences for CN of Na+ ion were explored by screening two important databases: Protein Databank and Cambridge Structural Database. A linear correlation is observed between the M (Metal)–O distance and the charge on metal ion in complex with the increase in CN of metal ion.  相似文献   

12.
Results of mass spectrometric studies are reported for the collisional dissociation of Group XI (Cu, Ag, Au) metal ion complexes with fatty acids (palmitic, oleic, linoleic and α-linolenic) and glycerolipids. Remarkably, the formation of M2H+ ions (M = Cu, Ag) is observed as a dissociation product of the ion complexes containing more than one metal cation and only if the lipid in the complex contains a double bond. Ag2H+ is formed as the main dissociation channel for all three of the fatty acids containing double bonds that were investigated while Cu2H+ is formed with one of the fatty acids and, although abundant, is not the dominant dissociation channel. Also, Cu(I) and Ag(I) ion complexes were observed with glycerolipids (including triacylglycerols and glycerophospholipids) containing either saturated or unsaturated fatty acid substituents. Interestingly, Ag2H+ ion is formed in a major fragmentation channel with the lipids that are able to form the complex with two metal cations (triacylglycerols and glycerophosphoglycerols), while lipids containing a fixed positive charge (glycerophospocholines) complex only with a single metal cation. The formation of Ag2H+ ion is a significant dissociation channel from the complex ion [Ag2(L–H)]+ where L = Glycerophospholipid (GP) (18:1/18:1). Cu(I) also forms complexes of two metal cations with glycerophospholipids but these do not produce Cu2H+ upon dissociation. Rather organic fragments, not containing Cu(I), are formed, perhaps due to different interactions of these metal cations with lipids resulting from the much smaller ionic radius of Cu(I) compared to Ag(I).  相似文献   

13.
The existence of gas‐phase electrostatic ion–ion interactions between protonated sites on peptides ([Glu] Fibrinopeptide B, Angiotensin I and [Asn1, Val5]‐Angiotensin II) and attaching anions (ClO4? and HSO4?) derived from strong inorganic acids has been confirmed by CID MS/MS. Evidence for ion–ion interactions comes especially from the product ions formed during the first dissociation step, where, in addition to the expected loss of the anion or neutral acid, other product ions are also observed that require covalent bond cleavage (i.e. H2O loss when several carboxylate groups are present, or NH3 loss when only one carboxylate group is present). For [[Glu] Fibrinopeptide B + HSO4]?, under CID, H2O water loss was found to require less energy than H2SO4 departure. This indicates that the interaction between HSO4? and the peptide is stronger than the covalent bond holding the hydroxyl group, and must be an ion–ion interaction. The strength and stability of this type of ion‐pairing interaction are highly dependent on the accessibility of additional mobile charges to the site. Positive mobile charges such as protons from the peptide can be transferred to the attaching anion to possibly form a neutral that may depart from the complex. Alternatively, an ion–ion interaction can be disrupted by a competing proximal additional negatively charged site of the peptide that can potentially form a salt bridge with the positively charged site and thereby facilitate the attaching anion's departure. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

14.
Conclusions The 2H+-biphenylenonium ion is formed when biphenylene is dissolved in the system HSO3F-SbF5-SO2ClF, which undergoes degenerate rearrangement as the result of one of the hydrogen atoms of the CH2 group shifting to the 3 position.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 5, pp. 1199–1201, May, 1978.  相似文献   

15.
Metastable ion peak shapes, dimensions and relative abundances have been measured for the three fragmentations [C3H6]+· → [C3H4]+· + H2, [C3H6]+· → [C3H5]+ + H· and [C3H6]+· → [C3H3]+ + H2 + H·. [C3H6]+· ions were derived from propene, cyclopropane, tetrahydrofuran, cyclohexanone, 2-methyl but-1-ene and cis-pent-2-ene. Activation energies for these fragmentations have been evaluated. Three daughter ion dissociations ([C3H5]+ → [C3H3]+ + H2, [C3H5]+ → [C3H4]+· + H· and [C3H4]+· → [C3H3]+ + H·) have been similarly examined. Ion structures have been determined and the metastable energy releases have been correlated with the thermochemical data. It is concluded that the molecular ions of propene and cyclopropane become structurally indistinguishable prior to fragmentation and that differences in their metastable ion characteristics can be ascribed wholly to internal energy differences; the latter can be correlated with the photoelectron spectra of the isomers. The pathway for the consecutive fragmentation which generates the metastable ion peak (m/e 42 → m/e.39) has been shown to be It is likewise concluded that fragmentating [C3H6]+· ions generated from the various precursor molecules are also structurally indistinguishable and cannot be classified with either molecular ion of the isomeric C3H6 hydrocarbons.  相似文献   

16.
H3O+ and OH?, formed by the self‐ionization of two coordinating water molecules during the crystal growing of a host molecule [1,3,5‐tris(hydroxymethyl)2,4,6‐triethylbenzene ( 1 )], could be effectively stabilized by hydrogen‐bonding interactions with the preorganized hydroxy groups of three molecules of 1. The binding motifs observed in the complex ( 1 )3?H3O+?HO? show remarkable similarity to those postulated for the hydrated hydronium and hydroxide ion complexes, which play important roles in various chemical, biological, and atmospheric processes, but their molecular structures are still not fully understood and remain a subject of intensive research.  相似文献   

17.
Potentiometric sensors with plasticized polymer membranes based on organic ion exchangers, tetraalkylammonium dodecyl sulfates (benzyldimethyldodecylammonium, benzyldimethyltetradecylammonium, dimethyldistearylammonium), have been proposed for the determination of quaternary ammonium salts in model solutions and KATAPAV technical solutions. The thermal stability, composition, and solubility product have been estimated. It has been shown that ion associates are stable to 60?C70°C, K S varies in the range from 2 × 10?8 to 5 × 10?10. The basic electrochemical parameters of the sensors have been determined as well, such as linearity ranges of the electrode function (5 × 10?5 (5 × 10?6)?1 × 10?2 (1 × 10?3) M) and slopes of the electrode functions (47?C59 mV/pc), response time (60?C90 sec), potential drift (2?C3 mV/day), operation period (3?C4 months), limits of detection for tetramethylammonium salts (1 × 10?5?4 × 10?7 M).  相似文献   

18.
The oxidation of hydrazoic acid in perchloric acid in the absence of added chloride under pseudo first-order conditions ([HN3] » [AuCl 4 ? ]) is first order in [Au(III)]. Michaelis–Menten type of dependence (linear plots of k obs ?1 vs [HN3]?1) is observed with respect to [HN3]. The k obs is independent of ionic strength and the plot between k obs ?1 and [H+] is linear. The inner-sphere mechanism is consistent with the formation of an axial complex (K = 25 dm3 mol?1) between AuCl3(HO)? ion and HN3 prior to its rate determining decomposition (k = 0.0182 s?1). It is inferred that the free radicals N 3 ? do not oxidise Au(II). The reaction becomes outer-sphere in the presence of added Cl? ions which are inferred to form a cage around the hydronium ion surrounding the AuCl 4 ? ions. The penetration of N 3 ? through the cage is rate controlling and within the cage, the electron transfer from N 3 ? ion to AuCl 4 ? is fast. The value of the rate determining constant k 2 is 0.547 dm3 mol?1 s?1 and the equilibrium constant K Cl for the cage formation is 5 dm3 mol?1 at 25 °C. It is calculated that the minimum HN3 concentration required before the reaction exhibits zero-order dependence in HN3 is 0.31 mol dm?3 when [H+] = 0.18 mol dm?3 at 25 °C.  相似文献   

19.
Solvation and association interactions in solutions of LiBF4/DMCC (DMCC for N,N-dimethylcarbamoyl chloride) and LiBF4/DMCC–DME (DME for 1,2-dimethoxyethane) have been studied as a function of concentration of lithium tetrafluoroborate by infrared and Raman spectroscopy. Strong interactions between Li+ and solvent molecules or BF4 anions are observed. The apparent solvation numbers of Li+ in LiBF4/DMCC solutions were deduced. Band-fitting to the B–F stretching band of BF4 anion permits detailed assess of the ion pairing. Based on the calculations of density function theory, optimal structures of Li+(DMCC)n (n = 1–3) were suggested. It is found that the lithium ion was preferentially solvated by DME in DMCC–DME binary solvents. This finding is supported by quantum chemistry calculations.  相似文献   

20.
A 1H NMR spectroscopic study of tetra-n-butylammonium halides (TBAX: X = Cl, Br or I) in CDCl3 solutions was conducted. Complexation studies of TBAX salts with different host molecules using 1H NMR in CDCl3 have previously revealed that the reference residual CHCl3 proton signal had been shifted downfield. The aim of the study was to quantify the extent of these chemical shift changes with TBAX salts. Linear concentration–chemical shift relationships in each case were obtained from the resulting titration plots obtained from the addition of the TBAX salts alone to CDCl3. Interactions in the solid state as determined by X-ray crystallography support the solution-state investigations indicating halide ion–chloroform proton interactions.  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