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 296 毫秒
1.
The negative ion mass spectra of phosphonitrile chlorides (PNCl2)n (n≥3) are studied. Since this series of compounds give very intense negative [M]? and [M? Cl]? ions, they can be used as good reference standards for negative ion mass spectrometry.  相似文献   

2.
The stabilities of alkali halide cluster ions [M(MX)n]+ (M ? Li, Na, K, Rb, Cs; X ? F, Cl, Br, I) have been studied by measuring the fragment ion yields following dissociation of the ions in the second field free region of a ZAB-2F mass spectrometer. Extractable cluster ions were observed for certain values of n. It was found that the stabilities of the neutral fragment species formed are also of importance in determining the fragmentation rates. Possible configurations of M and X in the stable ions are discussed.  相似文献   

3.
Linear ion-trap multiple-stage mass spectrometric approach (MS n ) towards nearly complete structural elucidation of triacylglycerol (TAG) including (1) assignment the fatty acid substituents on the glycerol backbone and (2) location of the double bond(s) on the unsaturated fatty acyl groups is reported. The characterization is established by the findings that MS2 on the [M+Li]+ ions of TAG yields more abundant ions reflecting losses of the outer fatty acid substituents either as free acids (i.e., [M+Li-R1CO2H]+ and [M+Li-R3CO2H]+ ions) or as lithium salts (i.e., [M+Li-R1CO2Li]+ and [M+Li-R3CO2Li]+ ions) than the ions reflecting the similar losses of the inner fatty acid substituent (i.e., [M+Li-R2CO2Li]+ and [M+Li-R2CO2Li]+ ions). Further dissociation (MS3) of [M+Li-R n CO2H]+ (n=1, 2, or 3) gives rise to the ion series locating the double bonds along the fatty acid chain. These ions arise from charge-remote fragmentations involving β-cleavage with γ-H shift, analogous to those seen for the unsaturated long-chain fatty acids characterized as initiated ions. Significant differences in abundances in the ion pairs reflecting the additional losses of the fatty acid moieties, respectively, were also seen in the MS3 spectra of the [M+Li-R n CO2H]+ and [M+Li-R n CO2Li]+ ions, leading to confirmation of the fatty acid substituents on the glycerol backbone. MS n on the [M+Na]+ and [M+NH4]+ adduct ions also affords location of fatty acid substituents on the glycerol backbone, but not the position of the double bond(s) along the fatty acid chain. Unique ions from internal losses of the glycerol residues were seen in the MS3 spectra of [M+Alk-R n CO2H]+ (n=1, 2, 3) and of [M+Alk-R n CO2Alk]+ (Alk=Li, Na, NH4; n=1, 3). They are signature ions for glycerides and the pathways leading to their formation may involve rearrangements.  相似文献   

4.
Thirteen of the salts of the alkali metals (Li, Na, K, Rb, Cs) with acetic, 2,2-dimethylpropionic, trifluoroacetic and heptafluorobutyric acid have been found to be sufficiently volatile to give mass spectra under normal electron impact conditions. The metal containing ions observed include (M=metal): [M]+, [MO]+, [MCO2]+, [M2], [M2O], [M2CO2] and the cluster ions [Mn (carboxylate)n-1]+ for n = 2–8.  相似文献   

5.
Herein we report density functional calculations of homolytic and heterolytic dissociation energies of the diatomic alkalimetal halides MX (M=Li, Na, K, Rb, and Cs and X=F, Cl, Br, I, and At) and their corresponding microsolvated structures MX?(H2O)n (n=1 to 4). Our results show that the homolytic dissociation energy of the MX?(H2O)n species increases with the number of water molecules involved in the microsolvated salts. On the other hand, the heterolytic dissociation energy follows exactly the opposite trend. As a result, while for the isolated diatomic alkalimetal halides, homolytic dissociation is always favored over heterolytic dissociation, the latter is preferred for CsF and CsCl already for n=2, and for n=4 it is the preferential mode of dissociation for more than half of the species studied.  相似文献   

6.
The positive and negative FAB mass spectra of a series of alkoxy- and chloro-silanes Xm(CH3)3-mSi(CH2)nR [m = 1 or 3, n = 3, 10 or 17, X = Cl or OMe or OEt, R = Me, NH2, glycidoxy, COOMe, NHCO(CH2)7COOMe or NHCO(CH2)10CH2OAc] were recorded in NBA and NPOE matrices. The chlorosilanes underwent rapid hydrolysis into silanols which condense to form siloxanes, the process being complete in NBA and partial in NPOE, yielding siloxane-based fragment ions in the positive spectra and silyloxyanions in the negative spectra. The alkoxysilanes were more resistant to hydrolysis, affording abundant [MH – HX]+ ions (X = OMe or OEt) in their positive FAB spectra and moderate to high intensity [M – H]? ions in the negative mode, the latter undergoing characteristic sequential loss of C2H4, EtOH and C2H4. Significant variations were observed in the positive spectra of all the silanes with change of matrix.  相似文献   

7.
Liquid secondary ionization mass spectra of solutions of alkali chlorides in glycerol were studied as a function of salt concentration. The experimental abundances of glycerol ions and of Cs+(CsCl) n cluster ions were successfully reproduced by assuming that most of the randomly distributed ions pair up with counterions shortly after impact. Further, it is considered that clustering (or proton transfer) reactions occur mainly between an ion that survives the pairing process and ion pairs (or basic analytes) in the immediate vicinity; however, some mixing undoubtedly occurs in the later stages of the desorption process. At the density of the original matrix, the range of proton transfer is calculated to be 5–15 Å, and that of clustering approximately 25% shorter. These reaction distances are inversely correlated with the internal energy of the ejected ions. In general, liquid secondary ionization mass spectra of alkali chloride solutions can be seen to result from competitive ion-ion recombination reactions in the decaying matrix. Finally, from the abundances of cluster ions containing [glycerol - H]? ions, it is estimated that approximately 1% of the glycerol molecules in the ejected volume are ionized in the collision cascade.  相似文献   

8.
Thermogravimetric (t.g.) and differential scanning calorimetric (d.s.c.) data have been used to study metal–amino acid interactions in adducts of general formula MnCl2 · ngly (gly = glycine, n = 0.7, 2.0, 4.0 and 5.0). All the prepared adducts exhibit only a one step mass loss associated with the release of glycine molecules, except for the 0.7gly adduct, which exhibits two glycine mass loss steps. From d.s.c. data, the enthalpy values associated with the glycine mass loss can be calculated: MnCl2 · 0.7gly = 409 and 399 kJ mol–1, MnCl2 · 2.0gly = 216 kJ mol–1, MnCl2 · 4.0gly = 326 kJ mol–1 and MnCl2 · 5.0gly = 423 kJ mol–1, respectively. The enthalpy associated with the ligand loss, plotted as function of the number of ligands for the n = 2.0, 4.0 and 5.0 adducts, gave a linear correlation, fitting the equation: H (ligand loss)/kJ mol–1 = 67 × (number of ligands, n) + 76. A similar result was achieved when the enthalpy associated with the ligand loss was plotted as a function of the a(COO) bands associated with the coordination through the carboxylate group, 1571, 1575 and 1577 cm–1, respectively, for the n = 2.0, 4.0 and 5.0 adducts, giving the equation H (ligand loss) /kJ mol–1 = 33.5 × a(COO) /cm–1 – 52418.5. This simple equation provides evidence for the enthalpy associated with the ligand loss being very closely related to the electronic density associated with the metal–amino acid bonds.  相似文献   

9.
Calculations in the framework of the density functional theory are performed to study the lowest‐energy isomers of coinage metal fluoride and chloride clusters (MnFn, MnCln, M = Cu, Ag, or Au, n = 1–6). For all calculated species starting from the trimers the most stable structures are found to be cyclic arrangements. However, planar rings are favored in the case of metal fluorides whereas metal chlorides prefer nonplanar cycles. Calculated bond lengths and infrared frequencies are compared with the available experimental data. The nature of the bonding, involving both covalent and ionic contributions, is characterized. The stability and the fragmentation are also investigated. Trimers are found to be particularly stable when considering the Gibbs free energies. © 2012 Wiley Periodicals, Inc.  相似文献   

10.
This paper overviews three living cationic polymerization systems (for styrene, p-methoxystyrene, and isobutyl vinyl ether) that are, in common, featured by: (i) specifically in nonpolar solvents, the use of the hydrogen halide/metal halide initiating systems (HX/MXn; X: I, Br, Cl; MXn: ZnX2, SnCl4), which generate a living growing carbocation stabilized by a nucleophilic counteranion (X…MXn); (ii) specifically in polar solvents, the use of externally added ammonium salts (nBu4N+Y; Y: I, Br, Cl), which permit the generation of living species from HX/MXn by providing nucleophilic halogen anions Y, either the same as or different from the halogen X in HX.  相似文献   

11.
N-Acetylcysteine and nine N-acetylcysteine conjugates of synthetic origin were characterized by positive- and negative-ion plasma desorption mass Spectrometry. For sample preparation the electrospray technique and the nitrocellulose spin deposition technique were applied. The fragmentation of these compounds, which are best seen as S-substituted desaminoglycylcysteine dipeptides, shows a similar behaviour to that of linear peptides. In the positive-ion mass spectra intense protonated molecular ion peaks are observed. In addition, several sequence-specific fragment ions (A+, B+, [Y + 2H]+, Z+), immonium ions (I+) and a diagnostic fragment ion for mercap-turic acids (RM+) are detected. The negative-ion mass spectra exhibit deprotonated molecular ions and in contrast only one fragment ion corresponding to side-chain specific cleavage ([RXS]?) representing the xenobiotic moiety. In the case of a low alkali metal concentration on the target, cluster molecular ions of the [nM + H]+ or [nM - H]? ion type (n = 1-3) are observed. The analysis of an equimolar mixture of eight N-acetylcysteine conjugates shows different quasi-molecular ion yields for the positive- and negative-ion spectra.  相似文献   

12.
Poly(propylene glycol) [α-hydro-ω-hydroxypoly(oxypropylene)] of number-average molar mass n ≈ 2000 g · mol−1 (PPG2000) was cyclised with high conversion (ca. 75%) by reaction with dichloromethane in the presence of powdered KOH. The cyclic product was separated from chain extended polymer by preparative GPC, giving an overall yield of polymer (n ≈ 2000 g · mol−1, narrow molar mass distribution) in excess of 50%. Characterisation by analytical GPC and 13C NMR spectroscopy confirmed cyclisation. DEPT and 1H-coupled NMR spectra were used to show that the links in cyclic poly(oxypropylene) were 77% single acetal, 12% double acetal and 11% triple acetal (or higher). This complexity probably results from competitive reaction with water introduced with KOH.  相似文献   

13.
In the 70 e V electron impact mass spectra of a series of alkyldiphenylphosphine oxides (R?2PO, R = Me, Et, n-Pr, i-Pr, n-Bu, i-Bu, t-Bu, neopentyl, n-decyl), molecular ions of low abundance are observed and [M + H]+ ions are formed to a small extent at high sample pressures. The major ions include [?2PO]+, [?2POH]+; [?2CH2PO]+ and [?2CH2POH]+ which are formed by rearrangement and cleavage processes. The chemical ionization mass spectra obtained with methane and isobutane reagents consist of [M + H]+ ions. The proton affinity of R?2PO was found to be 219 ± 2.5 kcal mol?1.  相似文献   

14.
We investigate surface-enhanced Raman scattering (SERS) spectra of pyridine–Agn (n = 2–8) complexes by density functional theory (DFT) and time-dependent DFT (TDDFT) methods. In simulated normal Raman scattering (NRS) spectra, profiles of pyridine–Agn (n = 2–8) complexes are analogical with that of isolated pyridine. Nevertheless, calculated pre-SERS spectra are strongly dependent on electronic transition states of new complexes. Wavelengths at 335 nm, 394.8 nm, 316.9 nm and 342.6 nm, which are nearly resonant with pure charge transfer excitation states, are adopted as incident light when simulating pre-SERS spectra for pyridine–Agn (n = 2–8) complexes, respectively. We obtain enhancement factors from 103 to 105 in pre-SERS spectra compared with corresponding NRS spectra. The obvious increase in Raman intensities mainly result from charge transfer resonance Raman enhancement. A charge difference densities (CDDs) methodology is adopted in describing chemical enhancement mechanism. This methodology aims at visualizing charge transfer from Agn (n = 2–8) clusters to pyridine on resonant electronic transition, which is one of the most direct evidences for chemical enhancement mechanism.  相似文献   

15.
The coordination chemistry of 1,10-phenanthroline (phen) and sodium 1,3-benzothiazole-2-thiolate (SBT) with selected s-block elements has been investigated. Four metal complexes were prepared and their structures were characterized using a variety of analytical techniques including infrared, UV–visible, 1H NMR and 13C NMR spectroscopies, elemental analysis, mass spectrometry and single-crystal X-ray diffraction. The reactions of phen and SBT with M(X)n (M = K(I), Cs(I), Mg(II), Sr(II); X = OH, CO3, Cl; n = 1, 2) in MeOH–H2O yielded one-dimensional chains of both potassium [K2(phen)2(BT)2(H2O)4]n ( 1 ) and caesium [Cs(phen)(BS)(H2O)]n ( 2 ) (where BT = 1,3-benzthiazole-2-thiolate and BS = 1,3-benzthiazole-2-sulfinothiolate) and mononuclear complexes of both magnesium {[Mg(phen)(H2O)4](BT)2·phen} ( 3 ) and strontium {[Sr(phen)2(H2O)4](BT)2} ( 4 ). In these complexes, phen binds via an N,N′ chelate pocket, while the monoanonic BT ligands either coordinate in a bidentate fashion (in the case of 1 ) or remain uncoordinated (in the case of 3 and 4 ). In complex 2 , SBT ligand was oxidized in situ into a new BS ligand. The sulfinothiolate oxygen atoms in BS coordinate with caesium in a tridentate fashion. Complexes 1 – 4 were evaluated against urease for enzyme inhibition. The complexes displayed significant inhibition with IC50 values in the range 10.8–45.8 μM. In order to examine the structure–activity relationship, the complexes were docked at the active site of urease. Docking results clearly demonstrate the binding of each complex within the active site of the enzyme.  相似文献   

16.
Negative chemical ionization mass spectrometry is used as a probe to examine reactions between hydrocarbon radicals and metal complexes in the gas phase. The methane negative chemical ionization mass spectra of 27 complexes of cobalt(II ), nickel(II ) and copper(II ) in the presence of O4, O2N2 and N4 donor atom sets are characterized by two dominant series of adduct ions of the form [M + CnH2n]? and [M + CnH2n+1]? at m/z values above the molecular ion, [M]?. Insertion of the CH radical into the ligand followed by radical/radical recombination and electron capture is proposed as the major mechanism leading to the formation of [M + CnH2n]? adduct ions. A second pathway involves ligand substitution by CnH2n+1 radicals concomitant with H elimination and electron capture. Oxidative addition at the metal followed by ionization is suggested as the principal pathway for the formation of [M + CnH2n+1]? adduct ions.  相似文献   

17.
The 70 e V-electron impact mass spectra of the C7–C10 n-alkynes have been determined as well as the metastable ion spectra of the molecular ions and the [CS2]+ and [N2O]+ charge exchange mass spectra of the C7-C9 n-alkynes. The metastable ion mass spectra provide only a limited opportunity to distinguish between isomers; however, the 70-eV EI mass spectra of isomeric compounds permit a ready distinction between isomers. The [CS2]+ charge exchange mass spectra of isomeric compounds also show substantial differences. The [N2O]+ charge exchange mass spectra do not show the enhancement of β-fission fragments observed in field ionization experiments, despite representing ions of similar internal energy, and it is concluded that field dissociation is responsible for the β-fission fragments in the field ionization experiments.  相似文献   

18.
Nanorods of vanadium oxide doped with alkali metal ions M x V2O5 · nH2O (M = Na, K, Rb, Cs, x = 0.31–0.44) have been obtained under hydrothermal conditions. The particles are 30–80 nm in diameter and a few micrometers in length. The chemical state of atoms and their concentration ratios have been studied by XPS. It has been shown that vanadium atoms are in two oxidation states V5+ and V4+ and the concentration of vanadium(IV) ions directly depends on the alkali metal. The X-ray photoelectron spectra of the valence bands of M x V2O5 · nH2O (M = Na, K, Rb, Cs) nanorods have been measured and interpreted.  相似文献   

19.
The reactions of Co(II), Ni(II), and Cu(II) chlorides and bromides and their metallic powders with tetrazol-1-yl-tris(hydroxymethyl)methane (L) afforded new complexes ML2Hal2 · mH2O(M = Co(II) or Ni(II), Hal = Cl; M = Cu(II), Hal = Cl or Br, m = 0; and M = Co(II) or Ni(II), Hal = Br, m = 2), MLnCl2 (M = Co(II) or Ni(II), n = 2 or 4; M = Cu(II), n = 2), and MLnBr2 · mH2O (M = Ni(II), n = 2, m = 2; M = Cu(II), n = 2, m = 0). The compositions and structures of the synthesized complexes were determined by elemental analysis, IR spectroscopy (50–4000 cm−1), and X-ray diffraction analysis. The introduction of a bulky substituent into position 1 of the tetrazole cycle was shown to exert almost no effect on the coordination mode but affected the composition and structure of the complexes.  相似文献   

20.
After exploring the potential energy surfaces of MmCE2p (E=S−Te, M=Li−Cs, m=2, 3 and p=m-2) and MnCE3q (E=S−Te, M=Li−Cs, n=1, 2, q=n-2) combinations, we introduce 38 new global minima containing a planar hypercoordinate carbon atom (24 with a planar tetracoordinate carbon and 14 with a planar pentacoordinate carbon). These exotic clusters result from the decoration of V-shaped CE22− and Y-shaped CE32− dianions, respectively, with alkali counterions. All these 38 systems fulfill the geometrical and electronic criteria to be considered as true planar hypercoordinate carbon systems. Chemical bonding analyses indicate that carbon is covalently bonded to chalcogens and ionically connected to alkali metals.  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