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 46 毫秒
1.
A unified mechanism of the interaction of vibrational relaxation and dissociation of polyatomic molecules working in a wide temperature range (from 2000 to 10000 K) is proposed, which is described by a double step-ladder model. Relaxation due to collisions with the transfer of small and large portions of energy is taken into account. The transfer efficiency of the portions of thermal energy in the high-temperature decomposition upon the collisions of CO2 molecules with atomic and molecular partners is determined. The reaction rate constant of high-temperature dissociation of carbon dioxide is calculated. The data presented in the article suggest a new method for elucidating the mechanism of energy exchange in the absence of vibrational and translational equilibrium and at ultrahigh temperatures when the dissociation takes place during the time of several collisions.  相似文献   

2.
The current research is constructed for considering the chemical ionization and dissociation of perindopril in the positive mode of corona discharge ion mobility spectrometry. Four product ion peaks are observed in the ion mobility spectrum of perindopril erbumine at the cell temperature of 473 K. These peaks are assigned through the obtained intensity variation analysis in the ion mobility spectra over the elapsed time accompanied by the calculations backed by the validated density functional theory (DFT). In this regard, the most stable ionic species associated with each peak and the corresponding reliable generation pathways are found by the well-confirmed meta hybrid density functional method, M06-2X. The peaks are assigned to the protonated perindopril and its dissociation products, including counter ion and the related fragment ions. However, the structures of the neutral perindopril in the gas phase are thoroughly assessed to find a more stable one. The predicted chemical ionization products by the theory are in excellent agreement with our presented experiment here. Theoretical evaluations demonstrated that the production of a fragment by dissociation process occurs when perindopril gets a proton from the ionization region. Also, without protons, there is no dissociation process. Therefore, our mechanism investigated here is the proton transfer one. All possible sites of perindopril are considered theoretically for protonation along with their possible reactions. In addition to the computed PES, the assigned ions for obtained spectra are confirmed by the computed equilibrium constants and rate constants. Our theoretical results show that the peak of the main fragment is for M-CH3CH2OH produced by a reaction pathway involving no barrier. This study opens new perspectives in interpreting large molecules spectra for future studies.  相似文献   

3.
4.
The one-photon IR excitation and subsequent UV dissociation of ammonia molecules selective with respect to nitrogen isotopes were studied. The selectivity of vibrational excitation is achieved by tuning CO2 laser radiation to resonance with 14NH3 or 15NH3 molecules. The dependences of the yield of dissociation for each isotopic component and the selectivity on the buffer gas (N2, O2, Ar) pressure, the partial pressure of ammonia, and the time of delay between IR and UV laser pulses were established. At low pressures (67–270 Pa) of the isotopic mixture with a 15N concentration of 4.8%, the dissociation selectivity for 15N was 17. The mechanisms responsible for the selectivity of IR + UV-initiated dissociation are discussed. The phenomenological model has been developed that takes into consideration the effect of the interisotopic V-V exchange and V-T relaxation on the formation of the yield and selectivity of the two-stage IR+UV dissociation of ammonia.  相似文献   

5.
A prominent dissociation path for electrospray generated tryptic peptide ions is the dissociation of the peptide bond linking the second and third residues from the ammo-terminus. The formation of the resulting b2 and y n−2 fragments has been rationalized by specific facile mechanisms. An examination of spectral libraries shows that this path predominates in diprotonated peptides composed of 12 or fewer residues, with the notable exception of peptides containing glutamine or glutamic acid at the N-terminus. To elucidate the mechanism by which these amino acids affect peptide fragmentation, we synthesized peptides of varying size and composition and examined their MS/MS spectra as a function of collision voltage in a triple quadrupole mass spectrometer. Loss of water from N-terminal glutamic acid and glutamine is observed at a lower voltage than any other fragmentation, leading to cyclization of the terminal residue. This cyclization results in the conversion of the terminal amine group to an imide, which has a lower proton affinity. As a result, the second proton is not localized at the N-terminus but is readily transferred to other sites, leading to fragmentation near the center of the peptide. Further confirmation was obtained by examining peptides with N-terminal pyroglutamic acid and N-acetyl peptides. Peptides with N-terminal proline maintain the trend of forming b2 and y n−2 because their ring contains an imine rather than imide and has sufficient proton affinity to retain the proton at the N-terminus.  相似文献   

6.
Kinetic and thermodynamic isotope effects on the unimolecular dissociation of proton bound dimers were studied in the gas phase using mass spectrometry techniques. In addition proton transfer reactions were investigated using equilibrium techniques in conjunction with a theoretical study. Normal isotope effects were observed for all of the amine systems studied. The effect of label position, extent of labeling, size and structure of the proton bound dimers have been discussed with respect to (i) the kinetic and thermodynamic isotope effect on the dissociation reaction, (ii) the kinetic energy release on the dissociation reaction, (iii) the thermodynamic isotope effect on the proton exchange reaction between the labeled and unlabeled amines, and (iv) the effective temperatures and the excess energies of the metastable proton bound dimers. Other compound classes (CH3OH, (CH3)2O, CH3CN and (CH3)2CO) were studied and discussed in the same way, though not as thoroughly. All the systems show normal isotope effects, except for the proton bound dimer of CH3CN and CD3CN, which showed an inverse isotope effect.  相似文献   

7.
Abstract

Phosphates of tetravalent elements are practically important for ion exchange, catalysis and conductivity. This study deals with a number of hydrates of titanium, tin and zirconium phosphates. PMR data show that the structure of water molecules in hydrates is slightly distorted, and at temperatures higher than 160 K water has high translation mobility. NMR 31P proves HPO2 4 dissociation to be growing with increase of temperature. Energetic parameters of this process are determined. Close values of anion dissociation enthalpy (0, 16/2/Ev) and obtained activation energy of conductivity for di- and monohydrates (0,17/2/Ev) show tunnel pattern of proton transfer along H-bond direction, This type of correlation was not observed in anhydrous compounds. That can be explained by impossibility of anion proton tunneling because of H-bond weakening. Proton conductivity of acid phosphates was studied. Ten-fold decrease of conductivity at room temperature with the loss of each water molecule proves H2O participation in proton transport. Mechanism of this process is discussed with the use of NMR data. Dependence of water mobility and conductivity level on the degree of crystallinity is also discussed. With the help of NMR-data processes of ion exchange in tin and zirconium acid phosphates, as well as the state of developed salt forms were studied. Presence of lithium with high mobility in Li2Sn(PO4)2 ·nF2O was established.  相似文献   

8.
The rate of H/D exchange among methyl group protons in a series of substituted 3-hydroxypyridines, 5-hydroxypyrimidines, and their N-oxides has been shown to increase with increasing acidity of the medium. The most reactive form of these molecules is the cationic form at pH<2. The rate of H/D exchange of CH3 group protons in 3-hydroxypyridine derivatives has also been found to be several orders of magnitude lower than the rates of exchange for methyl-substituted 5-hydroxypyrimidine and its N-oxide. Effective rate constants for methyl group proton exchange have been estimated. In the case of methyl-substituted 5-hydroxypyrimidine N-oxide derivatives it has been established that the rate of proton exchange is greater for an ortho-methyl group than for a methyl group in the para-position relative to the N-oxide site.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 1, pp. 93–96, January, 1991.  相似文献   

9.
Two distinct proton exchange reactions occur in metastable gaseous benzylbenzenium ions, generated by isobutane chemical ionization of diphenylmethane and four deuterium-labelled analogues. Whereas the proton ring-walk at the benzenium moiety is fast giving rise to a completely random intraannular proton exchange, the interannular proton exchange is surprisingly slow and competes with the elimination of benzene. A kinetic isotope effect of kH/kD= 5 has been determined for the interannular proton transfer, and a particularly high energy barrier of 50–75 kJ mol?1 has been estimated. These observations are attributed to steric restrictions of the ring-to-ring proton transfer in benzylbenzenium ions and contrasted to the fast interannular proton exchange in the higher homologues.  相似文献   

10.
We investigate the role played by the coordination state of pre-existing water wires during the dissociation of moderately strong acids by means of first-principles molecular dynamics calculations. By preparing 2,4,6-tricyanophenol (calc. pKa~0.5) in two different initial states, we are able to observe sequential as well as concerted trajectories of dissociation: On one hand, equilibrium dissociation takes place on a ~50 ps timescale; proton conduction occurs through three-coordinated water wires in this case, by means of sequential Grotthus hopping. On the other hand, by preparing 2,4,6-tricyanophenol in a hydration state inherited from that of equilibrated phenol (calc. pKa=7.6), the moderately strong acid finds itself in a presolvated state from which dissociation can take place on a ~1 ps timescale. In this case, concerted dissociation trajectories are observed, which consist of proton translocation through two intervening, four-coordinated, water molecules in 0.1-1.0 ps. The present results suggest that, in general, the mechanism of proton translocation depends on how the excess proton is injected into a hydrogen bond network. In particular, if the initial conditions favour proton release to a fourfold H-bonded water molecule, proton translocation by as much as 6-8 A? can take place on a sub-picosecond timescale.  相似文献   

11.
Metal—proton and metal—metal exchange reactions have been studied by PMR for thiophenol, 2-methylthiophenol, 2,6-dimethylthiophenol, benzyl mercaptan and their C6H5HG, (C6H5)3Sn and (C6H5)3Pb derivatives in chlorobenzene and pyridine solutions. In chlorobenzene the metal—metal exchange has been found to proceed in many cases at a greater rate than the metal—proton type, the exchange mobility of hydrogen and organometallic groups in chlorobenzene increasing in the order (C6H5)3Sn<H<(C6H5)3Pb<C6H5Hg. In the case of the (C6H5)3Sn and (C6H5)3Pb groups, pyridine accelerates the metal—proton exchange to a greater extent than the metal—metal exchange.The influence of various factors on the exchange reactions has been studied Analysis of the experimental findings and literature data has led to the conclusion that most probably the mechanism of the exchange reactions involves an associative pathway, the ease of exchange being mainly determined by the ability of the migrating group to form a cyclic transition state with delocalized bonds. The data on the exchange equilibria of the organometallic derivatives of 2-methylthiophenol and 2,6-dimethylthiophenol with thiophenol and its derivatives demonstrate that the C6H5HgS, (C6H5)3SnS and (C6H5)3PbS groups have equal steric requirements when involved in non-bonded interactions with o-methyl substituents.  相似文献   

12.
《Chemical physics》1986,101(1):45-54
The mechanism of the rection between the diatomic molecule and a solid non-corrugated surface has been studied, employing the classical trajectory method. The study was performed in the energy range 1–6 eV, where all four types of collisions, inelastic, reaction (dissociative trapping), adsorption and dissociation, take place. To study mass effects three kinds of diatomic molecules were considered: the light—light (LL) mass combination represented by H2, the light—heavy (LH) mass combination representedby HCl and HI and the heavy—heavy(HH) mass combination represented by Cl2 and I2. The main findings are: (a) The reactive process for the LL and HH systems is governed by one reaction mechanism, and those for the LH systems are governed by two different mechanisms. (b) Whereas the LL and HH systems tend to dissociate upon collision with the surface, once threshold for dissociation is reached no dissociation is encountered for the LH systems, even for energies of tens of electron volts. (c) The gas/3-solid-phase reactive process differs significantly from that encountered in the gas phase in that the important gas-phase quasicollinear arrangements play a negligible role in the gas—solid phase.  相似文献   

13.
Azaphilones represent numerous groups of wild fungal secondary metabolites that exhibit exceptional tendency to bind to nitrogen atoms in various molecules, especially those containing the amine group. Nitrogenized analogues of mitorubrin azaphilones, natural secondary metabolites of Hypoxylon fragiforme fungus, have been detected in the fungal methanol extract in very low concentrations. Positive electrospray ionization interfaced with high‐resolution mass spectrometry was applied for confirmation of the elemental composition of protonated species. Collision‐induced dissociation (CID) experiments have been performed, and fragmentation mechanisms have been proposed. Additional information regarding both secondary metabolite analogue families has been reached by application of gas‐phase proton/deuterium (H/D) exchanges performed in the collision cell of a triple quadrupole mass spectrometer. An incomplete H/D exchange with one proton less than expected was observed for both protonated mitorubrin azaphilones and their nitrogenized analogues. By means of the density functional theory, an appropriate explanation of this behavior was provided, and it revealed some information concerning gas‐phase H/D exchange mechanism and protonation sites. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

14.
The transfer of protons from dithiophosphoric acids R2PSSH to substituted pyridines and proton exchange between the SH groups of the same acids and NH of protonated pyridines in chloroform solution were studied. The results obtained are compared with the analogous data for monothiocarboxylic acids RCOSH. The mechanism of SH-NH exchange was established; its slow step, just as for RCOSH, is transfer of a proton in the ion pairs. In the transition from RCOSH to R2PSSH, an increase in the thermal effect of the deprotonation of the acids by amines and the activation energy E of SH-NH proton exchange is observed, which is due to the greater stability of the R2PSS anions in comparison with RCOS on account of the greater delocalization of charge on the two sulfur atoms. With increasing excess of acid (above two-fold), the values of E are increased, since the excess acid molecules more effectively solvate the ion pairs than the original molecular complexes.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 21, No. 6, pp. 676–681, November–December, 1985.  相似文献   

15.
The kinetics of the metal exchange reactions between open‐chain Gd(DTPA)2? and Gd(DTPA‐BMA), macrocyclic Gd(DOTA)? and Gd(HP‐DO3A) complexes, and Cu2+ ions were investigated in the presence of endogenous citrate, phosphate, carbonate and histidinate ligands in the pH range 6–8 in NaCl (0.15 M ) at 25 °C. The rates of the exchange reactions of Gd(DTPA)2? and Gd(DTPA‐BMA) are independent of the Cu2+ concentration in the presence of citrate and the reactions occur via the dissociation of Gd3+ complexes catalyzed by the citrate ions. The HCO3?/CO32? and H2PO4? ions also catalyze the dissociation of complexes. The rates of the dissociation of Gd(DTPA‐BMA), catalyzed by the endogenous ligands, are about two orders of magnitude higher than those of the Gd(DTPA)2?. In fact near to physiological conditions the bicarbonate and carbonate ions show the largest catalytic effect, that significantly increase the dissociation rate of Gd(DTPA‐BMA) and make the higher pH values (when the carbonate ion concentration is higher) a risk‐factor for the dissociation of complexes in body fluids. The exchange reactions of Gd(DOTA)? and Gd(HP‐DO3A) with Cu2+ occur through the proton assisted dissociation of complexes in the pH range 3.5–5 and the endogenous ligands do not affect the dissociation rates of complexes. More insights into the interaction scheme between Gd(DTPA‐BMA) and Gd(DTPA)2? and endogenous ligands have been obtained by acquiring the 13C NMR spectra of the corresponding diamagnetic Y(III)‐complexes, indicating the increase of the rates of the intramolecular rearrangements in the presence of carbonate and citrate ions. The herein reported results may have implications in the understanding of the etiology of nephrogenic systemic fibrosis, a rare disease that has been associated to the administration of Gd‐containing agents to patients with impaired renal function.  相似文献   

16.
The emergence of ceria (CeO2) as an efficient catalyst for the selective hydrogenation of alkynes has attracted great attention. Intensive research effort has been devoted to understanding the underlying catalytic mechanism, in particular the H2–CeO2 interaction. Herein, we show that the adsorption of propyne (C3H4) on ceria, another key aspect in the hydrogenation of propyne to propene, strongly depends on the degree of reduction of the ceria surface and hydroxylation of the surface, as well as the presence of water. The dissociation of propyne and the formation of methylacetylide (CH3CC‐) have been identified through the combination of infrared reflection absorption spectroscopy (IRAS) and DFT calculations. We demonstrate that propyne undergoes heterolytic dissociation on the reduced ceria surface by forming a methylacetylide ion on the oxygen vacancy site and transferring a proton to the nearby oxygen site (OH group), while a water molecule that competes with the chemisorbed methylacetylide at the vacancy site assists the homolytic dissociation pathway by rebounding the methylacetylide to the nearby oxygen site.  相似文献   

17.
A series of substituted 3-aroyldithiocarbazates has been synthesized and studied. The corresponding acid dissociation constants have been determined potentiometrically. Semiempirical PM3 molecular orbital calculations suggest the existence of several tautomeric forms of the compounds. Geometrical parameters, proton affinities, and static reactivity indices have been examined. Structural properties and protonation sites are well described by calculations. The strong correlations between the pK a values and the Hammett constants as well as the N(3) calculated proton affinities indicate that the N(3) atom is the most probable protonation site. The thermodynamics of the protonation process are mainly controlled by HOMO-LUMO rather than electrostatic interactions. According to PM3 results, 3-aroyldithiocarbazic acid should be quite stable in the gas phase, while a mechanism for its decomposition in solution is proposed.  相似文献   

18.
Hydrogen‐release by photoexcitation, excited‐state‐hydrogen‐transfer (ESHT), is one of the important photochemical processes that occur in aromatic acids and is responsible for photoprotection of biomolecules. The mechanism is described by conversion of the initial state to a charge‐separated state along the O(N)‐H bond elongation, leading to dissociation. Thus ESHT is not a simple H‐atom transfer in which a proton and a 1s electron move together. Here we show that the electron‐transfer and the proton‐motion are decoupled in gas‐phase ESHT. We monitor electron and proton transfer independently by picosecond time‐resolved near‐infrared and infrared spectroscopy for isolated phenol–(ammonia)5, a benchmark molecular cluster. Electron transfer from phenol to ammonia occurred in less than 3 picoseconds, while the overall H‐atom transfer took 15 picoseconds. The observed electron‐proton decoupling will allow for a deeper understanding and control of of photochemistry in biomolecules.  相似文献   

19.
The mechanism of cyclopalladation of azobenzene (L) with PdCl2 in N,N-dimethylformamide (dmf) was studied computationally, using DFT (B3LYP) methods supplemented with a continuum solvation model. Since the exact nature of the reacting complex is unknown, several candidates were considered. These were: (1) a mononuclear adduct with two ligand molecules, L-PdCl2-L, (2) a mononuclear adduct with one ligand and one solvent molecule, L-PdCl2-dmf, (3) a dinuclear adduct with a double chloride bridge, [L-PdCl-(μ-Cl)]2, and (4) a coordinatively unsaturated complex with an agostic interaction, L-PdCl2. The reaction profile initiating from L-PdCl2-dmf, which displays an agostic intermediate produced after displacement of the dmf molecule by the activating C-H bond, has the lowest barrier (20.4 kcal/mol in the step with the proton transfer to the O(dmf) atom). In all other reaction pathways, the proton migration is to a chlorine atom and is associated with remarkably high barriers. The results are related to previous experimental and other computational findings. While none of the reaction profiles includes explicit dissociation of the ligand, the proton transfer was found to occur only after the ligand is almost completely displaced from the coordinating shell. It was concluded that the transition state corresponds to 14-electron coordination of Pd and that ease of a ligand dissociation is an important, but not necessarily decisive, factor for cyclopalladation.  相似文献   

20.
In an unforced flowing NaCl solution in bulk, gravitational or electro convection supplies ions from bulk toward the membrane surface through a boundary layer. In a boundary layer formed on an anion exchange membrane, the convection converts to migration and diffusion and carries an electric current. In a boundary layer formed on a cation exchange membrane, the convection converts to migration and carry an electric current. In a forced flowing solution in bulk, the boundary layer thickness is reduced and gravitation or electro convection is disappeared. An electric current is carried by diffusion and migration on the anion exchange membrane and by migration on the cation exchange membrane. Ion transport in a boundary layer on the cation exchange membrane immersed in a NaCl solution is more restricted comparing to the phenomenon on the anion exchange membrane. This is due to lower counter-ion mobility in the boundary layer and the restricted water dissociation reaction in the membrane. The water dissociation reaction is generated in an ion exchange membrane and promoted due to the increased forward reaction rate constant. However, the current efficiency for the water dissociation reaction is generally low. The intensity of the water dissociation is more suppressed in the strong acid cation exchange membrane comparing to the phenomenon in the strong base anion exchange membrane due to lower forward reaction rate constant in the cation exchange membrane. In the strong acid cation exchange membrane, the intensity of electric potential is larger than the values in the strong base anion exchange membrane. Accordingly, the stronger repulsive force is developed between ion exchange groups (SO 3 ? groups) and co-ions (OH? ions) in the cation exchange membrane, and the water dissociation reaction is suppressed. In the strong base anion exchange membrane, the repulsive force between ion exchange groups (N+(CH3)3 groups) and co-ions (H+ ions) is relatively low, and the water dissociation reaction is not suppressed. Violent water dissociation is generated in metallic hydroxides precipitated on the desalting surface of the cation exchange membrane. This phenomenon is caused by a catalytic effect of metallic hydroxides. Such violent water dissociation does not occur on the anion exchange membrane.  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