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.
Two cyclooctapeptides, cycloreticulin A, cyclo(Pro1-Gly2-Asp3-Ile4-Ser5-Ile6-Tyr7-Tyr8) (1) and cycloreticulin B, cyclo(Pro1-Mso2-Tyr3-Gly4-Thr5-Val6-Ala7-Val8) (2), have been isolated from the methanol extract of the seeds of Annona reticulata L. The sequences were elucidated on the basis of the MS/MS fragmentation using a QTOF mass spectrometer equipped with an ESI source, chemical degradation and extensive 2D-NMR. The solid state conformation of cycloreticulin A, carried out by X-ray study, is characterised by the presence of two β-turns (types II and III) and an inversed γ-turn. Its solution structure appeared quite similar to the crystal one. The cyclic backbone solution structure of cycloreticulin B, close to that of the cyclooctapeptide squamin A, from which its sequence only differs by a Val8/Ile8 substitution, involves three β-turns, two of type I and one of type III, being similar to the crystal structure of squamin A.  相似文献   

2.
The oxidation of the 28 VE cyclo‐E6 triple‐decker complexes [(CpRMo)2(μ,η66‐E6)] (E=P, CpR=Cp( 2 a ), Cp*( 2 b ), CpBn( 2 c )=C5(CH2Ph)5; E=As, CpR=Cp*( 3 )) by Cu+ or Ag+ leads to cationic 27 VE complexes that retain their general triple‐decker geometry in the solid state. The obtained products have been characterized by cyclic voltammetry (CV), EPR, Evans NMR, multinuclear NMR spectroscopy, MS, and structural analysis by single‐crystal X‐ray diffraction. The cyclo‐E6 middle decks of the oxidized complexes are distorted to a quinoid ( 2 a ) or bisallylic ( 2 b , 2 c , 3 ) geometry. DFT calculations of 2 a , 2 b , and 3 persistently result in the bisallylic distortion as the minimum geometry and show that the oxidation leads to a depopulation of the σ‐system of the cyclo‐E6 ligands in 2 a – 3 . Among the starting complexes, 2 c is reported for the first time including its preparation and full characterization.  相似文献   

3.
Nitrogen‐rich heterocyclic bases and oxygen‐rich acids react to produce energetic salts with potential application in the field of composite explosives and propellants. In this study, 12 salts formed by the reaction of the bases 4‐amino‐1,2,4‐trizole (A), 1‐amino‐1,2,4‐trizole (B), and 5‐aminotetrazole (C), upon reaction with the acids HNO3 (I), HN(NO2)2 (II), HClO4 (III), and HC(NO2)3 (IV), are studied using DFT calculations at the B97‐D/6‐311++G** level of theory. For the reactions with the same base, those of HClO4 are the most exothermic and spontaneous, and the most negative ΔrGm in the formation reaction also corresponds to the highest decomposition temperature of the resulting salt. The ability of anions and cations to form hydrogen bonds decreases in the order NO3?>N(NO2)2?>ClO4?>C(NO2)3?, and C+>B+>A+. In particular, those different cation abilities are mainly due to their different conformations and charge distributions. For the salts with the same anion, the larger total hydrogen‐bond energy (EH,tot) leads to a higher melting point. The order of cations and anions on charge transfer (q), second‐order perturbation energy (E2), and binding energy (Eb) are the same to that of EH,tot, so larger q leads to larger E2, Eb, and EH,tot. All salts have similar frontier orbitals distributions, and their HOMO and LUMO are derived from the anion and the cation, respectively. The molecular orbital shapes are kept as the ions form a salt. To produce energetic salts, 5‐aminotetrazole and HClO4 are the preferred base and acid, respectively.  相似文献   

4.
Molecular dynamics has been used with a Lennard-Jones (6–12) potential in order to study the decay behavior of neutral Argon clusters containing between 12 and 14 atoms. The clusters were heated to temperatures well above their melting points and then tracked in time via molecular dynamics until evaporation of one or more atoms was observed. In each simulation, the mode of evaporation, energy released during evaporation, and cluster lifetime were recorded. Results from roughly 2000 simulation histories were combined in order to compute statistically significant values of cluster half-lives and decay energies. It was found that cluster half-life decreases with increasing energy and that for a given value of excess energy (defined asE=(E tot ?E gnd)/n), the 13 atom cluster is more stable against decay than clusters containing either 12 or 14 atoms. The dominant decay mechanism for all clusters was determined to be single atom emission.  相似文献   

5.
A number of force fields of the molecular mechanics type have been tested for their ability to represent as an energy minimum, the observed crystal structure for three cyclic hexapeptides, cyclo-(-Ala-Ala-Gly-Gly-Ala-Gly-), cyclo-(-Ala-Ala-Gly-Gly-Ala-Gly-), and cyclo-(-D-Ala-D-Ala-Gly-Gly-Gly-Gly-). The most effective force field tested was that recently proposed by Kollman and co-workers, notwithstanding its use of “united” atoms for CH, CH2, and CH3 groups. Fields proposed by Levitt, and adaptations of that of Scheraga and co-workers, were also effective. Force fields in which hydrogens bonded to electronegative atoms were not specified explicitly were less accurate in representation.  相似文献   

6.
Simulated annealing (SA) is a popular global minimizer that can conveniently be applied to complex macromolecular systems. Thus, a molecular dynamics or a Monte Carlo simulation starts at high temperature, which is decreased gradually, and the system is expected to reach the low-energy region on the potential energy surface of the molecule. However, in many cases this process is not efficient. Alternatively, the low-energy region can be reached more effectively by minimizing the energy of selected molecular structures generated along the simulation pathway. The efficiency of SA to locate energy-minimized structures within 5 kcal/mol above the global energy minimum is studied as applied to three peptide models with increasing geometrical restrictions: (1) The linear pentapeptide Leu-enkephalin described by the ECEPP potential, (2) a cyclic hexapeptide described by the GROMOS force field energy EGRO alone, and (3) the same cyclic peptide with EGRO combined with a restraining potential based on 31 proton–proton restraints obtained from nuclear magnetic resonance (NMR) experiments. The efficiency of SA is compared to that of the Monte Carlo minimization (MCM) method of Li and Scheraga, and to our local torsional deformations (LTD) method for the conformational search of cyclic molecules. The results for the linear peptide show that SA provides a relatively weak guidance towards the most stable energy region; as expected, this guidance increases for the cyclic peptide and the cyclic peptide with NMR restraints. However, in general, MCM and LTD are significantly more efficient than SA as generators of low-energy minimized structures. This suggests that LTD might provide a better search tool than SA in structure determination of protein regions for which a relatively small number of restraints are provided by NMR. ©1999 John Wiley & Sons, Inc. J Comput Chem 20: 1659–1670, 1999  相似文献   

7.
Energy minimization plays an important role in structure determination and analysis of proteins, peptides, and other organic molecules; therefore, development of efficient minimization algorithms is important. Recently, Morales and Nocedal developed hybrid methods for large-scale unconstrained optimization that interlace iterations of the limited-memory BFGS method (L-BFGS) and the Hessian-free Newton method (Computat Opt Appl 2002, 21, 143-154). We test the performance of this approach as compared to those of the L-BFGS algorithm of Liu and Nocedal and the truncated Newton (TN) with automatic preconditioner of Nash, as applied to the protein bovine pancreatic trypsin inhibitor (BPTI) and a loop of the protein ribonuclease A. These systems are described by the all-atom AMBER force field with a dielectric constant epsilon = 1 and a distance-dependent dielectric function epsilon = 2r, where r is the distance between two atoms. It is shown that for the optimal parameters the hybrid approach is typically two times more efficient in terms of CPU time and function/gradient calculations than the two other methods. The advantage of the hybrid approach increases as the electrostatic interactions become stronger, that is, in going from epsilon = 2r to epsilon = 1, which leads to a more rugged and probably more nonlinear potential energy surface. However, no general rule that defines the optimal parameters has been found and their determination requires a relatively large number of trial-and-error calculations for each problem.  相似文献   

8.
Salts of 3d, 4d, and 5d metals in the presence of the ligands 1,1,1,-tris(diphenylphosphinomethyl)ethane (triphos) or tris (2-diphenylphosphinoethyl) amine (np3) react with white phosphorus P4 (or yellow As4) to produce several mononuclear sandwich and dinuclear triple-decker sandwich complexes, which contain the unprecedented cyclo-triphosphorus (or cyclo-triarsenic) unit acting as a trihapto-ligand. In these complexes the metal atoms are bonded to the there phosphorus atoms of the phosphane ligand and to the three atoms of the cyclo-P3 or cyclo-As3 unit. The complexes are diamagnetic or have μeff-values corresponding to one or two unpaired electrons. The cyclo-P3 ligand is coordinatively unsaturated as proved by the fact that the mononuclear sandwich compounds may form Lewis-base adducts with electron-acceptor fragments. Reaction of the complexes (np)3M (M = Ni, Pd) with white P4 leads to formation of diamagnetic compounds [(np3)M(η1-P4)], in which the metal atom is bonded to the three phosphorus atoms of the np3-ligand and in addition to one P atom of the intact P4 molecule, which behaves as a monohapto-ligand. This article contains a review of the syntheses and structures of these complexes as well as a unified, albeit qualitative, approach to their bonding and properties.  相似文献   

9.
By combining Hartree–Fock results for nonrelativistic ground-state energies of N-electron atoms with analytic expressions for the large-dimension limit, we have obtained a simple renormalization procedure. For neutral atoms, this yields energies typically threefold more accurate than the Hartree–Fock approximation. Here, we examine the dependence on Z and N of the renormalized energies E(N, Z) for atoms and cations over the range Z, N = 2 → 290. We find that this gives for large Z = N an expansion of the same form as the Thomas–Fermi statistical model, E → Z7/2(C0 + C1Z?1/3 + C2Z?2/3 + C3Z?3/3 + ?), with similar values of the coefficients for the three leading terms. Use of the renormalized large-D limit enables us to derive three further terms. This provides an analogous expansion for the correlation energy of the form δE δZ4/3(δC3 + δC5Z?2/3 + δC6Z?3/3 + ?); comparison with accurate values of δE available for the range Z ? 36 indicates the mean error is only about 10%. Oscillatory terms in E and δE are also evaluated. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
Structure and bonding in triple-decker cationic complexes [(η5-Cp)Fe(μ,η:η5-E5) Fe(η5-Cp)]+ (1: E = CH, 2: E = P, 3: E = As) and [(η5-Cp)Fe(μ,η:η5-Cp)Fe(η5-E5)]+ (E = P, As) are examined by density functional theory (DFT) calculations at the B3LYP/6-31+G* level. These species exhibit the lowest energy when all the three ligands are eclipsed. In the complexes with bifacially coordinated cyclo-E5, the perfectly eclipsed D5h sandwich structure a is found to be a potential minimum. The energy difference between the fully eclipsed and the staggered conformations b and c are within 1.0, 2.1, and 6.3 kcal/mol, respectively, for E = CH, P, and As. The isomeric species with monofacially coordinated cyclo-E5 (E =P, As), [(η5 -Cp)Fe(μ,η :η5-Cp)Fe(η5-E5)]+ are predicted to be about 30 and 60 kcal/mol higher in energy , respectively, for E = P and As. The calculations predict that the bifacially coordinated cyclo-E5 (E =P, As) undergoes significant ring expansion leading to ``loosening of bonds' as observed experimentally. The consequent loss of aromaticity in the central cyclo-E5 indicates that significant π-electron density from the ring can be directed towards bonding with the iron centers on both sides. The diffuse nature of the π-orbitals of cyclo-P5 and cyclo-As5 can lead to better overlap with the iron d-orbitals and result in stronger bonding. This is reflected in the bond order values of 0.377 and 0.372 for the Fe-P and Fe-As bonds in 2a and 3a, respectively. The natural population analysis reveals that the Fe atom that is coordinated to a cyclo-E5 (E = P, As) possesses a negative charge of −0.23 to −0.38 units due to transfer of electron density from the inorganic ring to the metal center.  相似文献   

11.
We proposed a complete calculation scheme for attributing the total energy by the Hartree–Fock theory to atoms (EA) and the region between two atoms (EAB). It was pointed out that the conventional method using the Fock matrix includes a large amount of mutual contamination in both EA and EAB. The new scheme was derived from the basic expression of the total energy. Calculated results by the new scheme satisfy the theoretical requirements. The scaling effect on partitioned energies was also examined. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 71: 35–46, 1999  相似文献   

12.
Numerous experimental data for the cyclization of free radicals C·H2(CH2)nCH=CH2 cyclo-[(CH2)n+1CH(C·H2)], and C·H2(CH2)nCH=CHR cyclo-[(CH2)n+1C·HCHR] were analyzed in the framework of the parabolic model. The activation energy of thermoneutral (H e = 0) cyclization E e0 decreases linearly with an increase in the energy of cycle strain E rsc: E e0(n) (kJ mol–1) = 85.5 – 0.44E rsc(n) (n is the number of atoms in the cycle). The activation entropy of cyclization S # also depends on the cycle size: the larger the cycle, the lower S #. A linear dependence of S # on the difference between the entropies of formation S° of cyclic hydrocarbon and the corresponding paraffin was found: S # = 1.00[S°(cycle) – S°(CnH2n+2)]. The E e0 values coincide for cyclization reactions with the formation of the six-membered cycle and the bimolecular addition of alkyl radicals to olefins.  相似文献   

13.
A double-headed trypsin inhibitor (MCI-1) was isolated and purified from the seeds of Momordica charantia Linn. Cucurbitaceae, by using the trypsin-sepharose-4B affinity chromatography and CM-Sephadex-C50 ion exchange chromatography. It is composed of 77 amino acid residues: Asp8 Thr1 Ser4 Glu8 Pro2 Gly6 Ala4 Cys14 Val2 Met4 Ile8 Leu1 Phe1 His3 Lys4 Arg7. The amino acid sequence of MCI-1 was determined by sequencing the cyanogen bromide, tryptic and staphylococcus aureus V8 proteolytic peptides, then aligned by overlapped sequences. The result shows that MCI-1 contains 7 pairs of disulfide bonds, its sequence showed the high homology with those of “Bowman-Birk” inhibitors. About 50% trypsin inhibitory activity still remained after MCI-1 was cleavaged with cyanogen bromide.  相似文献   

14.
Five uranium complexes with UO2L′(solv) formula (L′: prepared by condensation reaction between 2-hydroxyacetophenon S-pentyl isothiosemicarbazone (H2L) and 5-bromosalicylaldehyde (Sal); solv: ethanol (1), 2-propanol (2), 2-butanol (3), ethylene glycol (4), 1,2-propanediol (5)) were synthesized through template reactions between H2L, Sal, and UO2(CH3COO)2?2H2O for 1 and recrystallization of 1 in appropriate solvents for the other complexes. The compounds are characterized by melting point, elemental analyses, FT-IR, 1H NMR, 13C NMR, TGA, and X-ray crystallography. Molecular structures of the obtained crystals, determined by X-ray diffraction analysis, showed that the complexes have distorted pentagonal bipyramidal geometry. In all complexes, the bianionic tetradentate ligand (N2O2) with phenolic oxygens (O3, O4), thioamidic and azomethine nitrogen donor atoms is coordinated to the metal center in equatorial positions and the solv molecules occupied the fifth equatorial site and finally linear UO2 is located in axial position. The thermal behavior of the complexes was studied with TGA and DTG data and the results revealed three weight loss stages. The Coats–Redfern method is used for all degradation steps to determine kinetic parameters.  相似文献   

15.
Syntheses and Crystal Structures of [μ‐(Me3SiCH2Sb)5–Sb1,Sb3–{W(CO)5}2] and [{(Me3Si)2CHSb}3Fe(CO)4] – Two Cyclic Complexes with Antimony Ligands cyclo‐(Me3SiCH2Sb)5 reacts with [(THF)W(CO)5] (THF = tetrahydrofuran) to form cyclo‐[μ‐(Me3SiCH2Sb)5–Sb1,Sb3–{W(CO)5}2] ( 1 ). The heterocycle cyclo‐ [{(Me3Si)2CHSb}3Fe(CO)4] ( 2 ) is formed by an insertion reaction of cyclo‐[(Me3Si)2CHSb]3 and [Fe2(CO)9]. The crystal structures of 1 and 2 are reported.  相似文献   

16.
The total energies and one-electron energies for first- and second-row atoms were calculated by using the Hartree–Fock and the Hartree–Fock-Slater Hamiltonian with Xα orbitals, uiexp); α was parametrized from EHFS exp) = Eexp. The EHF exp) total energies are always higher than the Hartree–Fock energies for the atoms. The relation of the calculated ionization potential to the experimental ionization potential depends on the α used to define ui(α), αexp, or αHF.  相似文献   

17.
A search of the published chemical and engineering literature found enthalpy of solution data for an additional 104 and 49 organic compounds dissolved in benzene and acetonitrile, respectively. Standard thermodynamic relationships were used to convert the experimental enthalpy of solution data, ΔHsolv, to enthalpies of solvation, ΔHsolv. Updated Abraham model correlations were derived for describing gas-to-benzene and gas-to-acetonitrile enthalpies of solvation by combining the 104 and 49 additional values to existing benzene and acetonitrile ΔHsolv databases. The updated Abraham model correlations for benzene and acetonitrile described the observed ΔHsolv values to within overall standard deviations of less than 3.4 kJ mol?1.  相似文献   

18.
The compound [(μ‐dppp)(AuCl)2], previously reported to associate intermolecularly in a chain (catena) structure through AuI–AuI interactions (3.316Å), was obtained from gold(III) precursors in a cyclo form with shortened intramolecular AuI—AuI contacts at 3.237Å and a puckered AuPCCCPAu seven‐membered ring. DFT calculations using a large relativistic basis to account for the d10–d10 interaction reproduce the observed molecular structure in the crystal of this “linkage isomer”, including the conspicuous distortion at one of the gold atoms. The chelate complex [(dppp)PtCl2] was crystallized and structurally characterized as the dichloromethane solvate.  相似文献   

19.
In this work, we add different strength of external electric field (Eext) along molecule axis (Z‐axis) to investigate the electric field induced effect on HArF structure. The H‐Ar bond is the shortest at Eext = ?189 × 10?4 and the Ar‐F bond show shortest value at Eext = 185 × 10?4 au. Furthermore, the wiberg bond index analyses show that with the variation of HArF structure, the covalent bond H‐Ar shows downtrend (ranging from0.79 to 0.69) and ionic bond Ar‐F shows uptrend (ranging from 0.04 to 0.17). Interestingly, the natural bond orbital analyses show that the charges of F atom range from ?0.961 to ?0.771 and the charges of H atoms range from 0.402 to 0.246. Due to weakened charge transfer, the first hyperpolarizability (βtot) can be modulated from 4078 to 1087 au. On the other hand, make our results more useful to experimentalists, the frequency‐dependent first hyperpolarizabilities were investigated by the coupled perturbed Hartree‐Fork method. We hope that this work may offer a new idea for application of noble‐gas hydrides. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
Two novel phosphinic amides, (C6H5)2P(O)(NH?cyclo?C7H13) (I) and (C6H5)2P(O)(NH?cyclo?C6H11) (II) were synthesized and characterized by spectroscopic methods and X-ray crystallography. Both compounds crystallize in the orthorhombic chiral space group P212121 and in both structures, the N—H···O hydrogen bonds lead to one-dimensional arrangements along the a axis. The molecular geometries and vibrational frequencies of I and II were investigated with quantum chemical calculations at the B3LYP/6–311G** level of theory. Furthermore, the hydrogen bonds were studied by means of the Bader theory of atoms in molecules (AIM) and natural bond orbital (NBO) analysis.  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