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.
Six alkyl alcohols were studied using thermospray mass Spectrometry. Whereas the dominant ion in the spectrum up to a repeller potential of 120 V was [M + NH4]+, above that potential [M + H]+ and fragment ions appeared. The fragments observed were largely due to hydrogen release from alkyl ions ([CnH2n+1]+ – H2 → [CnH2n-1]+) and loss of water or some other stable molecule from the same species. The results are compared with those from ionization of the same alcohols under electron impact and photoionization conditions and with results obtained for methanol under thermospray conditions.  相似文献   

2.
Analysis of the isobutane chemical ionization mass spectra of hexenols, cyclohexenols and various syn/anti pairs of bicyclic and tricyclic homoallylic alcohols shows that: (i) the spectra of the allylic alcohols are dominated by [M + H – H2O]+ and [M + C4H9–H2O]+ ions and contain traces of [M + H]+ ions; (ii) [M + H]+ ions are prominent in the spectra of acyclic and certain cyclic homoallylic alcohols; and (iii) [M + H]+ ions dominate the spectra of other acyclic unsaturated alcohols. The [M + H]+ ions may result from either: (a) protonation of the hydroxyl group, followed by a very rapid intramolecular proton transfer from the protonated hydroxyl group to the carbon–carbon double bond or internal solvation of the protonated hydroxyl group by the carbon–carbon double bond; and/or (b) direct protonation of the carbon–carbon double bond with significant internal solvation of the resulting carbocation by the hydroxyl group, which may lead to carbon–oxygen bond formation to give a protonated cyclic ether. The consequences of placing various geometric constraints on the possible intramolecular interactions between the hydroxyl group and the carbon–carbon double bond in unsaturated alcohols are explored.  相似文献   

3.
The reaction between the dihydride of decacarbonyltriosmium [H2Os3(CO)10] and phenyl arsine oxide (PAO) in benzene yields only one product [Os3(O)9(μ-H){μ-PhAs(O)OAsPh}] (1), which is characterized by high resolution mass spectrometry (HRMS), Fast Atomic Bombardment Mass Spectrometry (FAB)+, IR, 1H and 13C NMR, and single crystal X-ray diffraction. The solid state X-ray diffraction study of compound (1) shows that the molecule is polycyclic and has an osmium triangle with a bridging hydride bonded to a PhAs(O)-O-AsPh ligand.  相似文献   

4.
The enantiomeric differentiation of a series of chiral β‐amino alcohols (A) is attempted, for the first time, by applying the kinetic method using L‐proline, L‐tryptophan, 4‐iodo‐L‐phenylalanine or 3, 5‐diiodo‐L‐tyrosine as the chiral references (Ref) and Cu2+ or Ni2+ ion (M) as the central metal ion. The trimeric diastereomeric adduct ions, [M+(Ref)2+A‐H]+, formed under electrospray ionization conditions, are subjected for collision‐induced dissociation (CID) experiments. The products ions, formed by the loss of either a reference or an analyte, detected in the CID spectra are evaluated for the enantiomeric differentiation. All the references showed enantiomeric differentiation and the Rchiral values are better for the aromatic alcohols than for aliphatic alcohols. Notably, the Rchiral values of the aliphatic amino alcohols enhanced when Ni2+ is used as the central metal ion. The experimental results are well supported by computational studies carried out on the diastereomeric dimeric complexes. The computational data of amino alcohols is correlated with that of amino acids to understand the structural interaction of amino alcohols with reference molecule and central metal ion and their role on the stabilization of the dimeric complexes. Application of flow injection MS/MS method is also demonstrated for the enantiomeric differentiation of the amino alcohols. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

5.
The experimental determination of ionization and appearance energies is discussed, together with the calculation of heats of formation of ions. Results are presented for the ions [CnH2n+1]+, [CnH2n+2N]+ and [CnH2n+1O]+. The low temperature (~350K), low energy (12.1eV) mass spectra of some alkanes, amines and alcohols are presented and discussed.  相似文献   

6.
cis-[CrIII(phen)2(H2O)2]3+ and cis-[CrIII(bipy)2(H2O)2]3+ (phen = 1,10-phenanthroline and bipy = 2,2-bipyridine) were readily oxidized by either PbO2 or PhIO to form the chromium(V) complexes [CrV(phen)2(O)2]+ and [CrV(bipy)2(O)2]+ respectively, which were characterized by elemental analysis, i.r. and e.s.r. spectroscopy.  相似文献   

7.
The synthesis of two new polyamines containing 2-pyridyl and 6-methyl-(2-pyridyl) groups is described. The equilibria between H+ and Co2+ and the new ligand 1,9-di(2-pyridyl)-2,5,8-triazanonane (dptn) as well as the protonation of the hydroxo complexes of 1,6-di(2-pyridyl)-2,5-diazahexane-Co(II) (Co(dpdh) and 1-(6-methyl-2-pyridyl-6-(2-pyridyl)-2,5-diazahexane-Co(II) (Co(mdpdh)) have been studied in aqueous solution using the pH method. The coordination ability of the pyridine containing ligand dptn is compared with the chelating tendency of the analogous aliphatic amine (tetren). In spite of the lower basicity of the pyridine derivative the stability constants of its Co(II) complex is higher by a factor of thirty. The absorption spectra give evidence for a pseudooctahedral geometry of Co(dpdh) (H2O) and Co(dpdh)(H2O)(OH)+. Oxygen-uptake measurements indicate the formation of binuclear peroxo species. The potentiometric equilibrium data indicate the presence of dibridged species (dpdh)Co(O2, OH)Co(dpdh)3+ and (mdpdh)Co(O2, OH)Co-(mdpdh)3+. The kinetics of the rapid O2-uptake was measured over a wide pH range on a stopped-flow apparatus. For Co(dpdh)2+ and Co(mdpdh)2+ we found a second order rate constant independent of pH up to pH 9, but in more alkaline solutions it increases and reaches an upper limit around pH 12.3. The data could be fitted by a rate law of the form k1 = (k1[H+] + k1 KH) ([H+] + KH)?1. This variation with pH was explained by a rapid equilibrium Co(dpdh) (H2O) ? Co(dpdh)(H2O)(OH)+ + H+(KH). The enhanced rate constants of the hydroxo species must arise from a rate determining H2O replacement by O2, dominated by Co-OH2 bond breaking and the expected ability of an OH? group to labilize neighboring H2O molecules. The protonation constant of the hydroxo complex obtained by equilibrium measurements (pKH = 11.19 ± 0.03) was in good agreement with that derived from kinetic data (11.12 ± 0.04). The hydrolysis of Co(dptn)(H2O)2+ influences the rate of O2-incorporation in a different way. In this system retardation occurs as a result of hydrolysis ascribed to the slower leaving of OH? compared to H2O. This was expected if a mechanism with rate determining H2O replacements by O2 holds.  相似文献   

8.
Solvent-mediated crystal-to-crystal transformations of [Au6Ag3Cu3(H2O)3(d -pen)6(tdme)2]3+ (d -[ 1 (H2O)3]3+; pen2−= penicillaminate, tdme=1,1,1-tris(diphenylphosphinomethyl)ethane) to form unique supramolecular species are reported. Soaking crystals of d -[ 1 (H2O)3]3+ in aqueous Na2bdc (bdc2−=1,4-benzenedicarboxylate) yielded crystals containing d -[ 1 (bdc)(H2O)2]+ due to the replacement of a terminal aqua ligand in d -[ 1 (H2O)3]3+ by a monodentate bdc2− ligand. When γ-cyclodextrin (γ-CD) was added to aqueous Na2bdc, d -[ 1 (H2O)3]3+ was transformed to d -[ 1 (bdc@γ-CD)(H2O)2]+, where a γ-CD ring was threaded by a bdc2− molecule to construct a pseudorotaxane structure. While the use of dicarboxylates with an aliphatic carbon chain instead of bdc2− afforded analogous pseudorotaxanes, such pseudorotaxane species were not formed when crystals of [Au6Ag3Cu3(H2O)3(l -pen)6(tdme)2]3+ (l -[ 1 (H2O)3]3+) enantiomeric to d -[ 1 (H2O)3]3+ were soaked in aqueous Na2bdc and γ-CD, affording only crystals containing l -[ 1 (bdc)(H2O)2]+.  相似文献   

9.
Loss of an alkyl group X? from acetylenic alcohols HC?C? CX(OH)(CH3) and gas phase protonation of HC?C? CO? CH3 are both shown to yield stable HC?C? \documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm C}\limits^{\rm + } $\end{document}(OH)(CH3) ions. Ions of this structure are unique among all other [C4H5O]+ isomers by having m/z 43 [C2H3O]+ as base peak in both the metastable ion and collisional activation spectra. It is concluded that the composite metastable peak for formation of m/z 43 corresponds to two distinct reaction profiles which lead to the same product ion, CH3\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm C}\limits^{\rm + } $\end{document}?O, and neutral, HC?CH. It is further shown that the [C4H5O]+ ions from related alcohols (like HC?C? CH(OH)(CH3)) which have an α-H atom available for isomerization into energy rich allenyl type molecular ions, consist of a second stable structure, H2C?\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm C}\limits^{\rm + } $\end{document}? C(OH)?CH2.  相似文献   

10.
Ring cleavage α to the oxygen function leads to [C3H5O]+ and [C4H7O]+ ions in the mass spectra of 4-arylcyclohexanols and their methyl ethers, respectively. Cleavage α to the aryl group gives rise to [C3H7O]+ (from the alcohols), [C4H9O]+ (from the ethers) and [ArC3H4]+ (from both) ions. The competition between the two ring cleavages explains the effect of the substituents of the aryl groups on the relative abundances of the resulting ions.  相似文献   

11.
The oxidation of primary alcohols by sodium N-chloroethylcarbamate in acid solution, results in the formation of corresponding aldehydes. The reaction is first order with respect to the oxidant and alcohol. The rate increases with an increase in acidity. The oxidation of α,α-dideuterioethanol exhibited a primary kinetic isotope, kH/kD = 2.11 at 298 K. The value of solvent isotope effect k(H2O)/k(D2O) = 2.23 at 298 K. Addition of ethyl carbamate does not affect the rate. (EtOC(OH)NHCl)+ has been postulated as the reactive species. Plots of (log k2 + Ho) against (Ho + log[H+]) are linear with the slope, ?, having values from 1.78–1.87. This suggested a proton abstraction by water in the rate-determining step. The rates of oxidation of alcohols bearing both electron-withdrawing and electron-donating groups are more than that of methanol. A concerted mechanism involving transfer of a hydride ion from the C? H bond of the alcohol tothe oxidant and removal of a proton from the O? H group by a water molecule has been proposed.  相似文献   

12.
For identification of ginsenoside enantiomers, electrospray ionization mass spectrometry (ESI‐MS) was used to generate silver complexes of the type [ginsenoside + Ag]+. Collision induced dissociation of the silver‐ginsenoside complexes produced fragment ions by dehydration, allowing differentiation of ginsenoside enantiomers by the intensity of [M + Ag ? H2O]+ ion. In the meanwhile, an approach based on the distinct profiles of enantiomer‐selective fragment ion intensity varied with collision energy was introduced to refine the identification and quantitation of ginsenoside enantiomers. Five pairs of enantiomeric ginsenosides were distinguished and quantified on the basis of the distribution of fragment ion [M + Ag ? H2O]+. This method was also extended to the identification of other type of ginsenoside isomers such as ginsenoside Rb2 and Rb3. For demonstrating the practicability of this novel approach, it was utilized to analyze the molar ratio of 20‐(S) and 20‐(R) type enantiomeric ginsenosides in enantiomer mixture in red ginseng extract. The generation of characteristic fragment ion [M + Ag ? H2O]+ likely results from the reduction of potential energy barrier of dehydration because of the catalysis of silver ion. The mechanism of enantiomer identification of ginsenosides was discussed from the aspects of computational modeling and internal energy. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

13.
Molecular geometries and heats of formation have been calculated, using MINDO/3, for mass spectral fragment pairs (A+ + B) derived from formamide. There are five stable isomeric forms of the molecular ion: [H2NC(OH)]+, (H3NCO)+, [HNC(OH2)]+, [NCH(OH2)]+, and (NCOH3)+ (in order of increasing
, but no isomer (H2NCHO)+. There are three isomeric forms of (M—H)+: (H2NCO)+ (HNCOH)+, and (NCOH2)+: the only stable form of (M—2H)+ is (NCOH)+. Other (A/B)+ fragment pairs calculated are (CO/NH3)+, (HCO/NH2)+, (H2O/HCN)+, (H2O/HNC)+ [HO. + (HCNH)+], and [HO. + (H2NC)+]. The structure of the doubly charged ion M2+ is also reported.  相似文献   

14.
The temperature dependence of the isobutane chemical ionization (CI.) mass spectra of 54 open-chain, cyclic and unsaturated C5- to C10-alcohols was studied at temperatures ranging from 60 to 250°, and enthalpy changes were calculated for the corresponding main reactions of typical alcohols. The CI. reactivity is controlled by the temperature and the substrate structure as usual, and in addition, by the molecular size. The combination of thermal, structural and substrate-size effects leads to the following main conclusions. At low-reactivity conditions, i.e. at 150° or less, the alcohols with less than 11 C-atoms give four distinct types of spectra, with (M – OH)+ usually as the base peak. The characteristic ions are MC4H9+ and (M – H)+ for primary, MH+ and (MC4H9 – H2O)+ for secondary, (MC4H9 – H2O)+ for tertiary and allyl-type alcohols. Configurational assignments of stereoisomeric alcohols are also possible, by means of steric compression and shielding effects. The MH+/(M – OH)+ ratio in the spectra of epimeric methylcyclohexanols is at least 3 to 4 times higher for the isomers with mainly axial OH-group conformation compared to the equatorial isomers. Stereospecific (M - H)+ ions are apparently formed from trans-2-methylcyclopentanol and endo-norbornan-2-ol by a favorable abstraction of the unshielded H(α)-atoms versus normal behavior of the other epimers. While the spectra recorded at 200° show almost exclusively (M – OH)+ ions, those at 250° give nevertheless some C-skeleton information through the temperature dependent decomposition of the (M – OH)+ ions.  相似文献   

15.
Reactions of oxygenated cobalt(II) complexes. IX. Oxidative properties of tetrakis(ethylenediamine)-μ-peroxo-μ-hydroxo-dicobalt(III)
  • 1 VIII s. [1].
  • [(en)2Co(O2, OH)Co(en)2]3+ ( a ) reacts with I? in acidic aqueous solution according to: CoIII(O2, OH)CoIII + 21? + 5H+ ? 2CoIII + 3H2O + I2. Using I? in excess first order rate constants are obtained which, to a first approximation, are independent of [I?]. Comparison with kinetic data of deoxygenation of [(en)2Co(O2, OH)Co(en)2]3+ under analogous conditions suggests that both reactions have the same rate determining step. The singly bridged species [(en)2(H2O)CoO2Co(H2O) (en)2]4+ is shown to be the reactive intermediate in the iodide oxidation (Schema 2).  相似文献   

    16.
    Diaquacobinamide (H2O)2Cbi2+ or its conjugate base hydroxyaquacobinamide (OH(H2O)Cbi+)) can bind up to two cyanide ions, making dicyanocobinamide. This transition is accompanied by a significant change in color, previously exploited for cyanide determination. The reagent OH(H2O)Cbi+ is used in excess; when trace amounts of cyanide are added, CN(H2O)Cbi+ should be formed. But the spectral absorption of CN(H2O)Cbi+ is virtually the same as that of OH(H2O)Cbi+. It has been inexplicable how trace amounts of cyanide are sensitively measured by this reaction. It is shown here that even with excess OH(H2O)Cbi+, (CN)2Cbi is formed first due to kinetic reasons; this only slowly forms CN(H2O)Cbi+. This understanding implies that CN(H2O)Cbi+ will itself be a better reagent.  相似文献   

    17.
    Reaction of [Ru(trpy)Cl3] with quinolin-8-ol (HQ) yields [Ru(trpy)(Q)Cl]. Treatment of [Ru(trpy)(Q)Cl] with Ag+ in Me2CO–H2O (3:1) and MeCN gives [Ru(trpy)- (Q)(H2O)]+ and [Ru(trpy)(Q)(MeCN)]+, respectively, which were isolated as their perchlorate salts. A similar reaction in EtOH, in the presence of NaN3, yields [Ru(trpy)(Q)(N3)]. All complexes are diamagnetic (low-spin, d6, S = 0) and show many intense m.l.c.t. transitions in the visible region. They display a reversible RuII-RuIII oxidation in the -0.13-0.48 V versus s.c.e. range, followed by an irreversible RuIII-RuIV oxidation in the 0.46–1.08V versus s.c.e. range and three trpy-based reductions on the negative side of s.c.e. Chemical oxidation of [RuII(trpy)(Q)Cl] by Ce4+ gives [Ru(trpy)-(Q)Cl]+ which shows intense l.m.c.t. transitions in the visible region together with a weak ligand field transition in the lower energy region. The complex is one-electron paramagnetic (low-spin, d5, S=1/2) and shows a rhombic e.s.r. spectrum in MeCN–PhMe (1:1) solution at 77K. Chemical oxidation of [Ru(trpy)(Q)-(H2O)]+ results in the formation of a -oxo dimer, [{Ru(trpy)(Q)}2O]2+.  相似文献   

    18.
    The reactivity of ClMn(H2O)+ towards small organic compounds (L) was examined in a Fourier transform ion cyclotron resonance (FT‐ICR) mass spectrometer. The organic compounds studied are aliphatic and aromatic alcohols, aliphatic amines, ketones, an epoxide, an ether, a thiol and a phosphine. All the reactions lead to the formation of the ClMn(H2O)(L)+ complex, which dissociates by loss of the H2O molecule. In general, the reactions were found to occur with high efficiencies (>85%), indicating them to be exothermic. Electron transfer was also observed between ClMn(H2O)+ and compounds with low ionization energies (IE), to form the molecular ion (L+?) of the analyte. Based on these observations, the IE of ClMn(H2O)+ is approximated to be 8.1 ± 0.1 eV. Thus, the utility of ClMn(H2O)+ as a chemical ionization reagent in mass spectrometry is expected to be limited to organic compounds with IEs greater than 8 eV. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

    19.
    The decomposition reactions of [C2H5O]+ ions produced by dissociative electron-impact ionization of 2-propanol have been studied, using 13C and deuterium labeling coupled with metastable intensity studies. In addition, the fragmentation reactions following protonation of appropriately labeled acetaldehydes and ethylene oxides with [H3]+ or [D3]+ have been investigated. In both studies particular attention has been paid to the reactions leading to [CHO]+, [C2H3]+ and [H3O]+. In both the electron-impact-induced reactions and the chemical ionization systems the fragmentation of [C2H5O]+ to both [H3O]+ and [C2H3]+ proceeds by a single mechanism. For each case the reaction involves a mechanism in which the hydrogen originally bonded to oxygen is retained in the oxygen containing fragment while the four hydrogens originally bonded to carbon become indistinguishable. The fragmentation of [C2H5O]+ to produce [CHO]+ proceeds by a number of mechanisms. The lowest energy route involves complete retention of the α carbon and hydrogen while a higher energy route proceeds by a mechanism in which the carbons and the attached hydrogens become indistinguishable. A third distinct mechanism, observed in the electron-impact spectra only, proceeds with retention of the hydroxylic hydrogen in the product ion. Detailed fragmentation mechanisms are proposed to explain the results. It is suggested that the [C2H5O]+ ions formed by protonation of acetaldehyde or ionization of 2-propanol are produced initially with the structure [CH3CH?\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm O}\limits^ + $\end{document}H] (a), but isomerize to [CH2?CH? \documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm O}\limits^ + $\end{document}H2] (e) prior to decomposition to [C2H3]+ or [H3O]+. The results indicate that the isomerization ae does not proceed directly, possibly because it is symmetry forbidden, but by two consecutive [1,2] hydrogen shifts. A more general study of the electron-impact mass spectrum of 2-propanol has been made and the fragmentation reactions proceeding from the molecular ion have been identified.  相似文献   

    20.
    Gaseous oxygen and nitrogen bases, both singly and as binary mixtures, have been introduced into ion mobility spectrometers to study the appearance of protonated molecules, and proton-bound dimers and trimers. At ambient temperature it was possible to simultaneously observe, following the introduction of molecule A, comparable intensities of peaks ascribable to the reactant ion (H2O)nH+, the protonated molecule AH+ and AH+ · H2O, and the symmetrical proton bound dimer A2H+. Mass spectral identification confirmed the identifications and also showed that the majority of the protonated molecules were hydrated and that the proton-bound dimers were hydrated to a much lesser extent. No significant peaks ascribable to proton-bound trimers were obtained no matter how high the sample concentration. Binary mixtures containing molecules A and B, in some cases gave not only the peaks unique to the individual compounds but also peaks due to asymmetrical proton bound dimers AHB+. Such ions were always present in the spectra of mixtures of oxygen bases but were not observed for several mixtures of oxygen and nitrogen bases. The dimers, which were not observable, notable for their low hydrogen bond strengths, must have decomposed in their passage from the ion source to the detector, i.e. in a time less than ∼5 ms. When the temperature was lowered to −20 °C, trimers, both homogeneous and mixed, were observed with mixtures of alcohols. The importance of hydrogen bond energy, and hence operating temperature, in determining the degree of solvation of the ions that will be observed in an ion mobility spectrometer is stressed. The possibility is discussed that a displacement reaction involving ambient water plays a role in the dissociation.  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