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 15 毫秒
1.
Neutralization-reionization experiments were performed on beams of [H2]+˙ ions of different, known vibrational energy content using a variety of neutralization target gases (Xe, H2, Ne) and reionization gases (He, O2). The recovery of [H2]+˙ ions was found to be only weakly dependent on the vibrational energy of the original [H2]+˙ ions. The ion kinetic energy spectra of H+ fragments from the neutralization-reionization experiments were independent of the collision gas; the processes by which they were generated were identified.  相似文献   

2.
The octahedral cationic CoIII complexes [Co(ida)(bigH)2]+, [Co(glyO)(bigH)2]2+, [Co(α-alanO)(bigH)2]2+, [Co(β-alanO)BigH)2]2+ and Co(PhbigH)3]3+, where idaH2=iminodiacetic acid, glyOH=glycine, α-alanOH=α-alanine, β-alanOH=β-alanine and PhbigH=phenylbiguanide, were studied by thin-layer chromatography on silica gel in solutions of different electrolytes (NaCl, NaBr, NaI, Na2SO4, Na2S2O3, NaNO2 and NaNO3). Among other factors, the movement of the cationic complexes was found to be dependent on the surface tension and equivalent conductance of the developer electrolyte.  相似文献   

3.
The effect of acidity and equilibrium chloride ion concentration on the interaction of PdCl4 2- with cystine (H2CySS) in hydrochloric acid solutions was studied. Pd(H4CySS)Cl3+ complex was found to form at [Cl] = 1.0, 0.5, or 0.25 mol/Linthe [H+] range from 0.10 to 1.00 mol/L; the relevant equilibrium constant was determined. Monodentate coordination of cystine to palladium(II) through the sulfur atom was proposed on the basis of analysis of conditional stability constants as functions of [Cl] and [H+].  相似文献   

4.
The complex species formed in aqueous solutions (25 °C, I=3.0 mol⋅dm−3 KCl ionic medium) between the V(III) cation and the ligands 6-methylpicolinic acid (MePic, HL), salicylic acid (H2Sal, H2L) and phthalic acid (H2Phtha, H2L) have been studied by potentiometric and spectrophotometric measurements. Application of the least-squares computer program LETAGROP to the experimental emf(H) data, taking into account the hydrolytic species and hydrolysis constants of V(III), indicates that under the employed experimental conditions the complexes [VL]2+, [V(OH)L]+, [V(OH)2L], [V(OH)3L], [VL2]+, [VL3] and [V2OL4] form in the vanadium(III)–MePic system. Were observed the complexes [VL]+, [VL2], [V(OH)L2]2− and [VL3]3− in the vanadium(III)–H2Sal system, and the species [VHL]2+, [VL]+, [V(OH)L], [VHL2], [VL2], [V(OH)L2]2−, [V(OH)2L2]3− and [VL3]3− in the vanadium(III)–H2Phtha system. The stability constants of these complexes were determined by potentiometric measurements, and spectrophotometric measurements were made in order to perform a qualitative characterization of the complexes formed in aqueous solution.  相似文献   

5.
Charge exchange of neutral C3F6 by a variety of atomic and molecular ions in the 1 to 25 eV range of collision energies is used to characterize the energies associated with formation of [C3F6]+˙. The internal energy of the nascent [C3F6]+˙ ion, assessed by observing the degree to which it fragments, increases with the recombination energy of the charge-exchange reagent. The existence of excited states of the reagent ions is identified from the fragmentation behaviour of [C3F6]+˙ in the cases of [CS2]+˙, NO+, O2+˙, [NH3]+˙ and possibly [CH4]+˙. In addition, the data confirm that the [C3F6]+˙ parent ion fragments from both the ground state and a long-lived isolated electronic state. The latter is populated by near-resonant charge transfer. Translational excitation contributes relatively little to internal excitation of the charge-exchanged product ion and even less in the case of the isolated state.  相似文献   

6.
The molecular ions of isomeric octanes retain their structural identity, while their alkyl fragments [CnH2n+1]+ (n = 3 to 7) isomerise to common structures prior to decomposition. Structures for [C6H13]+ and [C7H15]+ ions are proposed.  相似文献   

7.
In ab initio calculations, we determined the most probable routes of decomposition of the [CF3Cl]+, [CF2Cl2]+, [CFCl3]+, [CCl4]+ molecular ions of freons and [C3H8]+ ions of hydrocarbons formed by collision of neutral molecules with protons with energies of the order of 10 keV. The calculated potential surface sections are compared on a qualitative level with the probability of various ion fragments in experiments on fragmentation. The role of the charge transfer dynamics between the proton and the molecule is discussed.  相似文献   

8.
The kinetics of the electron-transfer reactions between promazine (ptz) and [Co(en)2(H2O)2]3+ in CF3SO3H solution ([CoIII] = (2–6) × 10−3 m, [ptz] = 2.5 × 10−4 m, [H+] = 0.02 − 0.05 m, I = 0.1 m (H+, K+, CF3SO 3 ), T = 288–308 K) and [Co(edta)] in aqueous HCl ([CoIII] = (1 − 4) × 10−3 m, [ptz] = 1 × 10−4 m, [H+] = 0.1 − 0.5 m, I = 1.0 m (H+, Na+, Cl), T = 313 − 333 K) were studied under the condition of excess CoIII using u.v.–vis. spectroscopy. The reactions produce a CoII species and a stable cationic radical. A linear dependence of the pseudo-first-order rate constant (k obs) on [CoIII] with a non-zero intercept was established for both redox processes. The rate of reaction with the [Co(en)2(H2O)2]3+ ion was found to be independent of [H+]. In the case of the [Co(edta)] ion, the k obs dependence on [H+] was linear and the increasing [H+] accelerates the rate of the outer-sphere electron-transfer reaction. The activation parameters were calculated as follows: ΔH = 105 ± 4 kJ mol−1, ΔS = 93 ± 11 J K−1mol−1 for [Co(en)2(H2O)2]3+; ΔH = 67 ± 9 kJ mol−1, ΔS = − 54 ± 28 J K−1mol−1 for [Co(edta)].  相似文献   

9.
Two solid complexes, fac–[Cr(gly)3] and [Cr(gly)2(OH)]2, (where gly is glycinato ligand) were prepared and their acid-catalysed aquation products were identified. The structure of [Cr(gly)3] was solved by X-ray diffraction, revealing a cationic 3D sublattice with perchlorate anions inside its cavities. Acid-catalysed aquation of [Cr(gly)3] and [Cr(gly)2(OH)]2 leads to the same inert product, [Cr(gly)2(H2O)2]+, in a two-stages process. At the first stage, intermediate complexes, [Cr(gly)2(O–glyH)(H2O)]+ and [Cr(gly)2(H2O)–OH–Cr(gly)2(H2O)]+, are formed respectively. Kinetics of the first aquation stage of [Cr(gly)3] were studied in HClO4 solutions. The dependencies of the pseudo first-order rate constants on [H+] are as follows: k obs1H = k 0 + k 1 K p1[H+], where k 0 and k 1 are rate constants for the chelate-ring opening via spontaneous and acid-catalysed reaction paths, respectively, and K p1 is the protonation constant. The proposed mechanism assumes formation of the reactive intermediate as a result of proton addition to the coordinated carboxylate group of the didentate ligand. Some kinetic studies on the second reaction stage, the one-end bonded glycine liberation, were also done. The obtained results were analogous to those for stage I. In this case, the proposed reactive species are intermediates, protonated at the carboxylate group of the monodentate glycine. Base hydrolysis of two complexes, [Cr(gly)2(O–gly)(OH)] and [Cr(gly)2(OH)2], was studied in 0.2–1.0 M NaOH. The pseudo first-order rate constants, k obsOH, were [OH] independent in the case of [Cr(gly)2(O–gly)(OH)], whereas those for [Cr(gly)2(OH)2] linearly depended on [OH]. The reaction mechanisms were proposed, where the OH -catalysed reaction path was rationalized in terms of formation of the reactive conjugate base, [Cr(gly)2(OH)(O)]2−, as a result of OH ligand deprotonation. Activation parameters were determined and discussed.  相似文献   

10.
An ion–neutral complex is a non-covalently bonded aggregate of an ion with one or more neutral molecules in which at least one of the partners rotates freely (or nearly so) in all directions. A density-of-states model is described, which calculates the proportion of ion–neutral complex formation that ought to accompany simple bond cleavages of molecular ions. Application of this model to the published mass spectrum of acetamide predicts the occurrence of ions that have not hitherto been reported. Relative intensities on the order of 0.1 (where the abundance of the most intense fragment ion = 1) ere predicted for [M – HO]+ and [M – CH4]+˙ ions, which have the same nominal masses as the prominent [M – NH3]+˙ and [M – NH2]+ fragments. High-resolution mass spectrometric experiments confirm the presence of the predicted fragment ions. The [M – HO]+ and [M – CH4]+˙ fragments were observed with relative abundances of 0.02 and 0.04, respectively. Differences between theory and experiment may be ascribed to effects of competing distonic ion pathways.  相似文献   

11.
The gas phase structures of the [M–H] cations and anions of glycine have been studied by using a combination of ab initio calculations (at the MP2(FC)/6–31+G1 level of theory) and tandem mass spectrometry (MS/MS). It was found that the ab initio stability order for the anions is [H2NCH2CO2] > [H2NCHCO2H] > [HNCH2CO2H]. In contrast, the cations exhibit different behaviour, whereas [H2NCHCO2H]+ is predicted to be a stable structure, [H2NCH2CO2]+ spontaneously fragments to the ion–molecule complex [H2NCH2+ ⋯ (OCO)] and the singlet [HNCH2CO2H]+ isomer is predicted to undergo a skeletal rearrangement to form [CH2NHCO2H]+. MS/MS spectra of [M–H]+ cations of various glycine isotopomers were obtained via: (i) collisional activation of electron impact generated cations and (ii) charge reversal of anions formed via HO negative ion chemical ionization. The resulting spectra were significantly different, suggesting different structures were involved. Neutralization–reionization experiments were performed on [M–H] anions in order to gain insights into the structures of the intermediate radicals.  相似文献   

12.
Rearrangement of the molecular ions of tetracyclone, tetraphenylquinone, tetraphenylthiophene dioxide and pentaphenylcyclopentadienol prior to decomposition, rather than formation of a symmetrical [C4Ar4]+˙ ion, is considered as an alternative method for the production of [C2Ar2]+˙. The p-fluoro substituent is used to elicit information about the positional origins of the [C2Ar2]+˙ fragments. Operation of a double focusing mass spectrometer in the defocused mode reveals that the integrity of the molecular ion in different systems is affected to varying degrees, ranging from total lack of rearrangement to virtually complete scrambling.  相似文献   

13.
Five novel 2,3-naphtho crown ether group 10 metal bis(dithiolate) complexes, [Na(N15C5)2]2[Pd(mnt)2] (1), [Na(N15C5)]2[Pd(i-mnt)2] (2) and [K(N18C6)]2[M(i-mnt)2] (3 5) (where mnt = 1,2-dicyanoethylene-1,2-dithiolate, i-mnt = 1,1-dicyanoethylene-2,2-dithiolate and M = Ni, Pd, Pt for complexes 35, respectively), have been synthesized and characterized by elemental analysis, FT-IR, UV–Visible spectra and single crystal X-ray diffraction. X-ray diffraction analyses reveal that complexes 1 and 2 have different structural features while complexes 35 are structurally isomorphous. Complex 1 consists of two [Na(N15C5)2]+ sandwich complex cations and one [Pd(mnt)2]2− anion, affording a zero-dimensional structure. For 2, the [Na(N15C5)]+ mono-capped complex cations act as the bridges linking the [Pd(i-mnt)2]2− anions into a 1D infinite chain through Na–N interactions and SȮFC and SȮFπ interactions are observed in the resulting chain. Complexes 35 all consist of two [K(N18C6)]+ complex cations and one [M(i-mnt)2]2− (M = Ni, Pd or Pt) anion and the complex molecules are linked into␣1D␣chains by the bridging K–O(ether) interactions between the adjacent [K(N18C6)]+ units. What’s novel is that the resulting chains are assembled into novel 2D networks through interchain π–π stacking interactions between the neighboring naphthylene moieties of N18C6. The stack model of naphthylene group in complexes 35 is discussed.  相似文献   

14.
An assembly strategy for metal nanoclusters using electrostatic interactions with weak interactions, such as C?H???π and π???π interactions in which cationic [Ag26Au(2‐EBT)18(PPh3)6]+ and anionic [Ag24Au(2‐EBT)18]? nanoclusters gather and assemble in an unusual alternating array stacking structure is presented. [Ag26Au(2‐EBT)18(PPh3)6]+ [Ag24Au(2‐EBT)18]? is a new compound type, a double nanocluster ion compound (DNIC). A single nanocluster ion compound (SNIC) [PPh4]+ [Ag24Au(2‐EBT)18]? was also synthesized, having a k‐vector‐differential crystallographic arrangement. [PPh4]+ [Ag24Au(2,4‐DMBT)18]? adopts a different assembly mode from both [Ag26Au(2‐EBT)18(PPh3)6]+ [Ag24Au(2‐EBT)18]? and [PPh4]+ [Ag24Au(2‐EBT)18]?. Thus, the striking packing differences of [Ag26Au(2‐EBT)18(PPh3)6]+ [Ag24Au(2‐EBT)18]?, [PPh4]+ [Ag24Au(2‐EBT)18]? and the existing [PPh4]+ [Ag24Au(2,4‐DMBT)18]? from each other indicate the notable influence of ligands and counterions on the self‐assembly of nanoclusters.  相似文献   

15.
The iron(III) dimeric complex [Fe2(CN)10]4− is reduced to the iron(III)iron(II) species [Fe2(CN)10]5− by iodide ion, the equilibrium constant being strongly dependent upon the nature of the alkali metal cation, reduction being favoured in the sequence: Cs+>NH 4 + ≥K+>Na+>Li+. The reaction kinetics are autocatalytic in character, the catalytic species being the mixed valence dimer. The rates of reactions are also strongly catalysed by alkali metal cations, in the same sequence as for the equilibrium constants. The reaction mechanism involves the formation of I 2 as a reactive intermediate which can be oxidised by both [Fe2(CN)10]4− and [Fe2(CN)10]5−.  相似文献   

16.
The following chromium(III) complexes with serine (Ser) and aspartic acid (Asp) were obtained and characterized in solution: [Cr(ox)2(Aa)]2− (where Aa = Ser or Asp), [Cr(AspH−1)2] and [Cr(ox)(Ser)2]. In acidic solutions, [Cr(ox)2(Aa)]2− undergoes acid-catalysed aquation to cis-[Cr(ox)2(H2O)2] and the appropriate amino acid. [Cr(ox)(Ser)2] undergoes consecutive acid-catalysed Ser liberation to give [Cr(ox)(H2O)4]+, and the [Cr(Asp)2] ion is converted into [Cr(Asp)(H2O)4]2+. Kinetics of these reactions were studied under isolation conditions. The determined rate expressions for all the reactions are of the form: k obs = a + b[H+]. Reaction mechanisms are proposed, and the meaning of the determined parameters has been established. Evidence for the formation of an intermediate with O-monodentate amino acid is given. The effect of the R-substituent at the α-carbon atom of the amino acid on the complex reactivity is discussed.  相似文献   

17.
Summary Fragmentation patterns of the essential amino acids (AAs) as their silyl derivatives have been obtained with the aid of ion trap detection (ITD). Three derivatizing reagents, hexamethyldisilazane+trifluoroacetic acid (HMDS+TFAA),bis-(trimethylsilyl)trifluoroacetamide (BSTFA) andN-methyl-N-(t-butyldimethylsilyl)trifluoroacetamide (MTBSTFA) were used. Simple and multiple derivatives obtained with each reagent have been investigated, with regard to their sensitivity and selectivity. Our study performed in the concentration range of 5-2000 ng amino acids has shown that, contrary to literature data, thirteen of the twenty-two AAs investigated including the TBDMS derivatives give rise to more than one peak when eluted. As a result of ion/molecule interaction the very informative ions of high masses, ([M]+, [M+TMS/(TBDMS)]+, [M+1]+) are formed with considerable intensities. The fragments [M-CH3]+, [M-C4H9]+, [M-(CH3)2Si]+, [M-TMS/(TBDMS)COO]+, [M-TBDMSOH]+, [M-TBDMSO]+, [M-TBDMSNH]+ and numerous others could be utilized for identification purposes. Presented at Balaton Symposium on High Performance Separation Methods, Siófok, Hungary, September 1–3, 1999  相似文献   

18.
Photoion-photoion coincidence spectra of benzene and benzene-d6 photoionized by He(II) light and synchrotron radiation show the existence of six major and eight minor charge-separation reactions of the [C6H6]2+ ion. Three main groups of ion pairs are related to [C3H3]+ + [C3H3]+, [C2H3]+ + [C4H3]+ and [CH3]+ + [C5H3]+, with appearance energies of 32.2 ± 0.5 eV, 31.3 ± 0.5 eV and 28.4 ± 0.3 eV. The kinetic energy release is the same for all pairs within a group, irrespective of hydrogen number, but differs from group to group. Results are interpreted in terms of fast, direct charge separation of [C6H6]2+, and subsequent hydrogen loss by the singly charged fragments.  相似文献   

19.
A spectroelectrochemical sensor was developed for [Re(dmpe)3]+ as a nonradioactive analog for [Tc(dmpe)3]+. The sensor consists of an optically transparent electrode (OTE) coated with a thin film of sulfonated polystyrene‐block‐poly(ethylene‐ran‐butylene)‐block‐polystyrene (SSEBS). Colorless [Re(dmpe)3]+ was reversibly oxidized to [Re(dmpe)3]2+ (λmax=530 nm). [Re(dmpe)3]+ preconcentrated by ion‐exchange into the SSEBS film, resulting in a 20‐fold increase in peak current compared to a bare OTE after 1 h of exposure to aqueous [Re(dmpe)3]+ solution. Detection of [Re(dmpe)3]+ at concentrations down to 2×10?6 M was accomplished by electrochemical modulation of the complex and monitoring absorbance by attenuated total reflectance (ATR).  相似文献   

20.
The mass spectra of 1,2-dichloro-3,4-bis(dichloromethylene)cyclobutene (IV) and of hexachloropentafulvene (II) have been studied. Compound IV cannot be an intermediate in the formation of II from octachloro-1,2-dimethylenecyclobutane (III) under electron-impact, as previously suggested. In the mass spectra of II and IV the species [C6]+ and [C5]+ occur, obviously through cleavage of the semicyclic C-C bond. The mass spectrum of pentachlorofulvalene (VI) shows strikingly that successive elimination of an even number of CI atoms is preferred over that of an odd number of CI atoms; probably corresponding C-CI bonds in the two rings are broken simultaneously. Amongst the fragments, the species [C10]+ and [C7]+ and possibly also [C8]+ and [C9]+ have been observed.  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