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.
We report the use of unimolecular dissociation by infrared radiation for gaseous multiphoton energy transfer to determine relative activation energy (Ea,laser) for dissociation of peptide sequence ions. The sequence ions of interest are mass-isolated; the entire ion cloud is then irradiated with a continuous wave CO2 laser, and the first order rate constant, kd, is determined for each of a series of laser powers. Provided these conditions are met, a plot of the natural logarithm of kd versus the natural logarithm of laser power yields a straight line, whose slope provides a measure of Ea,laser. This method reproduces the Ea values from blackbody radiative dissociation (BIRD) for the comparatively large, singly and doubly protonated bradykinin ions (nominally y 9 and y 9 2+ ). The comparatively small sequence ion systems produce Ea,laser values that are systematic underestimates of theoretical barriers calculated with density functional theory (DFT). However, the relative Ea,laser values are in qualitative agreement with the mobile proton model and available theory. Additionally, novel protonated cyclic-dipeptide (diketopiperazine) fragmentation reactions are analyzed with DFT. FT-ICR MS provides access to sequence ions generated by electron capture dissociation, infrared multiphoton dissociation, and collisional activation methods (i.e., b n , y m , c n , z m ions).  相似文献   

2.
Adsorption of pyridine on Nin‐clusters (with n = 2,3,4) is studied by quantum chemical calculations at B3LYP/LANL2DZ and B3LYP/6‐311G** levels. First, Nin‐clusters are investigated for accessible structure and electronic states. The lowest electronic state with four unpaired electrons is predicted for Ni4‐cluster based on geometry and electronic structure, showing that the cluster stability nicely depends on number of unpaired electrons. Correction for basis set superposition error of metal‐metal bond is appreciable and has increasing effect on cluster binding energy. Next, adsorption of pyridine in planar and vertical adsorption modes is investigated on rhombus Ni4‐cluster. The vertical mode is found (at B3LYP/6‐311G** level) as the most favorable adsorption mode. Adsorption energy (ΔEads) depends on cluster size; adsorption on Ni4‐cluster is most favorable with ΔEads = ?207.33 kJ/mol. The natural bond orbital analysis reveals the charge transfer in adsorbate/metal‐cluster. Results of investigations for the Ni2‐ and Ni3‐cluster are also presented. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
The thermodynamic stabilities of P2, P4, and three P8 cage structure were investigated through high‐precision CBS‐Q calculations. The CBS‐Q values for the bond energy of P2 (ΔEo: +115.7 kcal mol−1) and the formation of P4 from P2 (Δ Eo:‐56.6 kcal mol−1) were in excellent agreement with the experimental values (Eo: +117 and ‐56.4 kcal mol−1 respectively). Among the P8 cages, the cubane structure was the least stable (Δ Eo +37 kcal vs. 2×P4). The most stable P8 isomer adopts a cuneane structure resembling S4N4, and is more stable than white phosphorus at T = 0 K (Δ Eo −3.3 kcal mol−1), but still unstable under standard conditions for entropic reasons (Δ Go of +8.1 kcal mol−1 vs. 2×P4). The CBS‐Q energies represent significant revisions (6–20 kcal mol−1) of previous computational predictions obtained by high‐level single method calculations. © 2005 Wiley Periodicals, Inc. Heteroatom Chem 16:453–457, 2005; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20119  相似文献   

4.
Two p‐phenylenevinylene (PV) trimers, containing 3′‐methylbutyloxyl (in MBOPV3) and 2′‐ethylhexyloxyl (in EHOPV3) side chains, are used as model compounds of PV‐based conjugated polymers (PPV) with the purpose of clarifying the origin of fast (picosecond time) components observed in the fluorescence decays of poly[2‐methoxy‐5‐(2′‐ethylhexyloxy)‐p‐phenylenevinylene] (MEH‐PPV). The fluorescence decays of MBOPV3 and EHOPV3 reveal the presence of similar fast components, which are assigned to excited‐state conformational relaxation of the initial population of non‐planar trimer conformers to lower‐energy, more planar conformers. The rate constant of conformational relaxation kCR is dependent on solvent viscosity and temperature, according to the empirical relationship kCR=o?exp(?αEη/RT), where o is the frequency factor, ηo is the pre‐exponential coefficient of viscosity, Eη is the activation energy of viscous flow. The empirical parameter α, relating the solvent microscopic friction involved in the conformational change to the macroscopic solvent friction (α=1), depends on the side chain. The fast component in the fluorescence decays of MEH‐PPV polymers (PPVs), is assigned to resonance energy transfer from short to longer polymer segments. The present results call for revising this assignment/interpretation to account for the occurrence of conformational relaxation, concurrently with energy transfer, in PPVs.  相似文献   

5.
The energy transfer of highly vibrationally excited isomers of dimethylnaphthalene and 2‐ethylnaphthalene in collisions with krypton were investigated using crossed molecular beam/time‐of‐flight mass spectrometer/time‐sliced velocity map ion imaging techniques at a collision energy of approximately 300 cm?1. Angular‐resolved energy‐transfer distribution functions were obtained directly from the images of inelastic scattering. The results show that alkyl‐substituted naphthalenes transfer more vibrational energy to translational energy than unsubstituted naphthalene. Alkylation enhances the V→T energy transfer in the range ?ΔEd=?100~?1500 cm?1 by approximately a factor of 2. However, the maximum values of V→T energy transfer for alkyl‐substituted naphthalenes are about 1500~2000 cm?1, which is similar to that of naphthalene. The lack of rotation‐like wide‐angle motion of the aromatic ring and no enhancement in very large V→T energy transfer, like supercollisions, indicates that very large V→T energy transfer requires special vibrational motions. This transfer cannot be achieved by the low‐frequency vibrational motions of alkyl groups.  相似文献   

6.
In an attempt to synthesize a homologous series of peptide peresters, we investigated the reaction of the oxazol‐5(4H)‐ones of Pht‐(Aib)n‐OH (n=2–8) and tert‐butyl hydroperoxide in the presence of 4‐(dimethylamino)pyridine. Unexpectedly, the major product isolated in each case proved to be the peptide dialkyl peroxide. This novel class of peptide derivatives was characterized by FT‐IR, 1H‐NMR, MS, cyclic voltammetry, and X‐ray diffraction. On the basis of the experimental data, a plausible mechanism is proposed for this reaction.  相似文献   

7.
Dissociation constants (pKa) of trazodone hydrochloride (TZD⋅HCl) in EtOH/H2O media containing 0, 10, 20, 30, 40, 50, 60, 70, and 80% (v/v) EtOH at 288.15, 298.15, 308.15, and 318.15 K were determined by potentiometric techniques. At any temperature, pKa decreased as the solvent was enriched with EtOH. The dissociation and transfer thermodynamic parameters were calculated, and the results showed that a non‐spontaneous free‐energy change (ΔdissGo>0) and unfavorable enthalpy (ΔdissHo>0) and entropy (ΔdissSo<0) changes occurred on dissociation of trazodone hydrochloride. The free‐energy change or pKa varied nonlinearly with the reciprocal dielectric constant, indicating the inadequacy of the electrostatic approach. The dissociation equilibria are discussed on the basis of the standard thermodynamics of transfer, solvent basicity, and solute‐solvent interactions. The values of ΔtransGo and ΔtransHo increased negatively with increasing EtOH content, revealing a favorable transfer of trazodone hydrochloride from H2O to EtOH/H2O mixtures and preferential solvation of H+ and trazodone (TZD). Also, ΔtransSo values were negative and reached a minimum, in the H2O‐rich zone that has frequently been related to the initial promotion and subsequent collapse of the lattice structure of water. The pKa or ΔdissGo values correlated well with the Dimroth‐Reichardt polarity parameter ET(30), indicating that the physicochemical properties of the solute in binary H2O/organic solvent mixtures are better correlated with a microscopic parameter than the macroscopic one. Also, it is suggested that preferential solvation plays a significant role in influencing the solvent dependence of dissociation of trazodone hydrochloride. The solute‐solvent interactions were clarified on the basis of the linear free‐energy relationships of Kamlet and Taft. The best multiparametric fit to the Kamlet‐Taft equation was evaluated for each thermodynamic parameter. Therefore, these parameters in any EtOH/H2O mixture up to 80% were accurately derived by means of the obtained equations.  相似文献   

8.
The reaction of a P4 butterfly complex with yellow arsenic yields the largest mixed PnAsm ligand complexes synthesized to date. [{Cp′′′Fe(CO)2}2(μ,η1:1‐P4)] reacts with As4 to yield [{Cp′′′Fe}2(μ,η4:4‐PnAs4‐n)] and [Cp′′′Fe(η5‐PnAs5‐n)]. Mass spectrometry together with NMR spectroscopy and X‐ray crystallography give clear evidence about the arrangement of the E positions within the cyclo‐E5 and E4 moieties of the products. Moreover, the results of DFT calculations agree well with the experimental determined outcomes. By coordinating the E4 complex [{Cp′′′Fe}2(μ,η4:4‐PnAs4‐n)] with CuCl, a rearrangement of the E positions occurs in favor with a preferred phosphorus coordination towards copper atoms in the resulting 1D polymeric chain.  相似文献   

9.
We report non‐chiral amino acid residues cis‐ and trans‐1,4‐diaminocyclohexane‐1‐carboxylic acid (cyclo‐ornithine, cO) that exhibit unprecedented stereospecific control of backbone dissociations of singly charged peptide cations and hydrogen‐rich cation radicals produced by electron‐transfer dissociation. Upon collision‐induced dissociation (CID) in the slow heating regime, peptide cations containing trans‐cO residues undergo facile backbone cleavages of amide bonds C‐terminal to trans‐cO. By contrast, peptides with cis‐cO residues undergo dissociations at several amide bonds along the peptide ion backbone. Diastereoisomeric cO‐containing peptides thus provide remarkably distinct tandem mass spectra. The stereospecific effect in CID of the trans‐cO residue is explained by syn‐facially directed proton transfer from the 4‐ammonium group at cO to the C‐terminal amide followed by neighboring group participation in the cleavage of the CO―NH bond, analogous to the aspartic acid and ornithine effects. Backbone dissociations of diastereoisomeric cO‐containing peptide ions generate distinct [bn]+‐type fragment ions that were characterized by CID‐MS3 spectra. Stereospecific control is also reported for electron‐transfer dissociation of cis‐ and trans‐cO containing doubly charged peptide ions. The stereospecific effect upon electron transfer is related to the different conformations of doubly charged peptide ions that affect the electron attachment sites and ensuing N―Cα bond dissociations.  相似文献   

10.
The average downward energy transfer (〈Δ Edown〉) is obtained for highly vibrationally excited acetyl chloride with Ne and C2H4 bath gases at ca. 870 K. Data are obtained by the technique of very low-pressure pyrolysis (VLPP). Fitting these data by solution of the appropriate reaction-diffusion integrodifferential master equation yields the gas/gas collisional energy transfer parameters: 〈Δ Edown〉 values are 220 ± 10 cm?1 (Ne bath gas) and 330 ± 20 cm?1 (C2H4). These energy transfer quantities are much less than those predicted by statistical theories, or those observed for similar sized molecules such as CH3CH2Cl. These results are explained by the qualitative predictions of the biased random walk model wherein the fundamental mechanism of energy transfer is the multiple interactions between the bath gas and the individual atoms of the reactant molecule, during the course of the collision event. The charge distribution of acetyl chloride decreases the number of such interactions, thereby reducing the amount of energy transferred per collision.  相似文献   

11.
High oxidation potential perfluorinated zinc phthalocyanines (ZnFnPcs) are synthesised and their spectroscopic, redox, and light‐induced electron‐transfer properties investigated systematically by forming donor–acceptor dyads through metal–ligand axial coordination of fullerene (C60) derivatives. Absorption and fluorescence spectral studies reveal efficient binding of the pyridine‐ (Py) and phenylimidazole‐functionalised fullerene (C60Im) derivatives to the zinc centre of the FnPcs. The determined binding constants, K, in o‐dichlorobenzene for the 1:1 complexes are in the order of 104 to 105 M ?1; nearly an order of magnitude higher than that observed for the dyad formed from zinc phthalocyanine (ZnPc) lacking fluorine substituents. The geometry and electronic structure of the dyads are determined by using the B3LYP/6‐31G* method. The HOMO and LUMO levels are located on the Pc and C60 entities, respectively; this suggests the formation of ZnFnPc.+–C60Im.? and ZnFnPc.+–C60Py.? (n=0, 8 or 16) intra‐supramolecular charge‐separated states during electron transfer. Electrochemical studies on the ZnPc–C60 dyads enable accurate determination of their oxidation and reduction potentials and the energy of the charge‐separated states. The energy of the charge‐separated state for dyads composed of ZnFnPc is higher than that of normal ZnPc–C60 dyads and reveals their significance in harvesting higher amounts of light energy. Evidence for charge separation in the dyads is secured from femtosecond transient absorption studies in nonpolar toluene. Kinetic evaluation of the cation and anion radical ion peaks reveals ultrafast charge separation and charge recombination in dyads composed of perfluorinated phthalocyanine and fullerene; this implies their significance in solar‐energy harvesting and optoelectronic device building applications.  相似文献   

12.
Fully unprotected peptide o‐aminoanilides can be efficiently activated by NaNO2 in aqueous solution to furnish peptide thioesters for use in native chemical ligation. This finding enables the convergent synthesis of proteins from readily synthesizable peptide o‐aminoanilides as a new type of crypto‐thioesters. The practicality of this approach is shown by the synthesis of histone H2B from five peptide segments. Purification or solubilization tags, which are sometimes needed to improve the efficiency of protein chemical synthesis, can be incorporated into the o‐aminoanilide moiety, as demonstrated in the preparation of the cyclic protein lactocyclicin Q.  相似文献   

13.
The synthesis and electrochemical properties of ferrocene conjugates are presented for the purpose of investigating adenosine 5′‐[γ‐ferrocenoylalkyl] triphosphate ( 1 a – 4 a , ferrocene (Fc)–ATP) as co‐substrates for phosphorylation reactions. Compounds 1 a – 4 a were synthesized, purified by HPLC, and characterized by NMR spectroscopy and mass spectrometry. In solution, all Fc–ATP bioconjugates exhibit a reversible one‐electron redox process with a half‐wave potential (E1/2) in the 390–430 mV range, peak separations (ΔEp) in the 40–70 mV range, and the peak current ratio (ipa/ipc) near unity. The peptide‐modified surface Glu‐Gly‐Ile‐Tyr‐Asp‐Val‐Pro was used to study the sarcoma‐related protein (Src) kinase activity by employing the Fc–ATP bioconjugates as co‐substrates. Subsequent kinase‐catalyzed transfer of the γ‐Fc‐phosphate group to the tyrosine residues of the surface‐bound peptides was characterized by a formal potential (Eo) ≈390 mV (vs. Ag/AgCl). The Fc‐coverage, estimated by time‐of‐flight secondary‐ion mass spectrometry (TOF‐SIMS) and cyclic voltammetry (CV), suggested validity of Fc–ATP conjugates as kinase co‐substrates. Depending on the length of the alkyl spacer of the Fc–ATP conjugate, different current densities were obtained, pointing to a direct correlation between the two. Molecular modeling revealed that the structural constraint imposed by the short alkyl spacer ( 1 a ) causes a steric congestion and negatively affects the outcome of phosphorylation reaction. An optimal analytical response was obtained with the Fc–ATP conjugates with linker lengths longer than six CH2 groups.  相似文献   

14.
A series of tributyltin(IV) complexes of 2‐[(E)‐2‐(3‐formyl‐4‐hydroxyphenyl)‐1‐diazenyl]benzoic acid and 4‐[((E)‐1‐{2‐hydroxy‐5‐[(E)‐2‐(2‐carboxyphenyl)‐1‐diazenyl]phenyl}methylidene)amino]aryls have been investigated by electrospray mass spectrometry (ESI‐MS) and tandem mass spectrometry (MSn) techniques. The assignments are facilitated by agreement between observed and calculated isotopic patterns and MSn studies. Single‐crystal X‐ray crystallography of (Bu3Sn[O2CC6H4{N?N(C6H3‐4‐OH(C(H)?NC6H4OCH3‐4))}‐o])n reveals a polymeric structure. Toxicity studies of the tributyltin(IV) complexes of the 4‐[((E)‐1‐{2‐hydroxy‐5‐[(E)‐2‐(2‐carboxyphenyl)‐1‐diazenyl]phenyl}methylidene)amino]aryls on the second larval instar of the Aedes aegypti and Anopheles stephensi mosquito larvae are also reported. The LC50 values indicate that the complexes are effective larvicides, which range from a low of 0.36 ppm to a high of 0.69 ppm against the Ae. aegypti larvae and between 0.82 and 1.17 ppm against the An. stephensi larvae. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

15.
The Conder and Young (CY) and the peak maximum (PM) methods were used to estimate the retention time of n-alkane probes on chemithermomechanical pulp (CTMP) wood fibers treated with a low molecular weight grade phenol-formaldehyde resin (PFR). Thermodynamic functions (ΔHao, ΔGao, and ΔSao) and the London dispersive component of the surface energy were derived from these retention times. Treated wood fibers show a high energy surface due to the presence of the thermoset resin on their surface. Values of ΔHao obtained from the CY method were higher than those obtained with the PM method at relatively high temperatures and with relatively low molecular weight alkanes. The results from the two methods were identical at low temperature (293 K) and with the relatively high molecular weight alkane n-undecane.  相似文献   

16.
Cationic emulsions of triblock copolymer particles comprising a poly(n‐butyl acrylate) (PnBA) central block and polystyrene (PS) outer blocks were synthesized by activator generated by electron transfer (AGET) atom transfer radical polymerization (ATRP). Difunctional ATRP initiator, ethylene bis(2‐bromoisobutyrate) (EBBiB), was used as initiator to synthesize the ABA type poly(styrene‐bn‐butyl acrylate‐b‐styrene) (PS‐PnBA‐PS) triblock copolymer. The effects of ligand and cationic surfactant on polymerizations were also discussed. Gel permeation chromatography (GPC) was used to characterize the molecular weight (Mn) and molecular weight distribution (MWD) of the resultant triblock copolymers. Particle size and particle size distribution of resulted latexes were characterized by dynamic light scattering (DLS). The resultant latexes showed good colloidal stability with average particle size around 100–300 nm in diameter. Glass transition temperature (Tg) of copolymers was studied by differential scanning calorimetry (DSC). © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 611–620  相似文献   

17.
Unnatural metal‐chelating amino acids bearing aminodiacetate side‐chains have been introduced into two hexapeptides to obtain efficient lanthanide‐binding peptides. The synthesis of the enantiopure Fmoc‐Adan(tBu)2‐OH synthons is described with overall yields of 32 and 50 % for n=2 and n=3 side‐chain carbon atoms, respectively. The two peptides AcWAdanPGAdanGNH2 ( P n ) were synthesized from the protected synthons by standard solid‐phase peptide synthesis. Studies of the lanthanide complexes of the two peptides P n by luminescence titrations, mass spectrometry, circular dichroism, and solution NMR spectroscopy demonstrate that the Adan chain length has a dramatic effect on the complexation properties. Indeed, the flexible compound P3 forms a mononuclear complex of moderate stability (β11=109.9), which tends to transform into a binuclear species in the presence of excess of the metal ion. Interestingly, the more compact peptide P2 provides stable Ln3+ complexes with the exclusive formation of the mononuclear Ln P2 adduct. The stability constant of Tb P2 is two orders of magnitude higher (β11=1012.1) than that measured for P3 . The 800 MHz NMR spectrum of the La3+ complex of P2 evidences a well‐defined type II β‐turn as well as a hydrophobic Trp(indole)–Pro interaction. These interactions exemplify the non‐innocent character of the peptide spacer in the complex La P2 as well as the role of a peptide secondary structure in the stabilization of metal complexes.  相似文献   

18.
We review the processes which have been observed from collisions between alkali-halide clusters and solid surfaces. Soft impact of nanocrystalline NanF n?1 + clusters against solid surfaces causes them to cleave along the lowest energy (100) plane. At higher collision energies (Ei>1 eV/atom), an evaporative cascade occurs which is characteristic of a transformation of the nanocrystal to a molten state. Efficient F? transfer from the cluster to the surface can occur for the larger clusters (>60 atoms) scattering from Si(111), in direct competition with the cleaving channel at low energies. In this regime, strong bonds can form between the F? and silicon surface. The reaction probability increases with cluster size indicating that an impact-initiated shock wave is needed to enhance the reactive process.  相似文献   

19.
20.
Positive singly charged ionic liquid aggregates [(Cnmim)m+1(BF4)m]+ (mim = 3‐methylimidazolium; n = 2, 4, 8 and 10) and [(C4mim)m+1(A)m]+ (A = Cl, BF4, PF6, CF3SO3 and (CF3SO2)2N) were investigated by electrospray ionisation mass spectrometry and energy‐variable collision induced dissociation. The electrospray ionisation mass spectra (ESI‐MS) showed the formation of an aggregate with extra stability for m = 4 for all the ionic liquids with the exception of [C4mim][CF3SO3]. ESI‐MS‐MS and breakdown curves of aggregate ions showed that their dissociation occurred by loss of neutral species ([Cnmim][A])a with a ≥ 1. Variable‐energy collision induced dissociation of each aggregate from m = 1 to m = 8 for all the ionic liquids studied enabled the determination of Ecm, 1/2 values, whose variation with m showed that the monomers were always kinetically much more stable than the larger aggregates, independently of the nature of cation and anion. The centre‐of‐mass energy values correlate well with literature data on ionic volumes and interaction and hydrogen bond energies. Copyright © 2008 John Wiley & Sons, Ltd.  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