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.
Four cyclic peptides, diandrine A–D ( 1 – 4 ), were isolated from the MeOH extract of Formosan Drymaria diandra. Their structures were elucidated by chemical and spectroscopic analyses as cyclo(‐Gly1‐Pro2‐Trp3‐Pro4‐Tyr5‐Phe6‐), cyclo(‐Gly1‐Pro2‐Leu3‐Pro4‐Leu5‐Trp6‐Ser7‐Ser8‐), cyclo(Gly1‐Gly2‐Pro3‐Tyr4‐Trp5‐Pro6‐), and cyclo(Gly1‐Gly2‐Pro3‐Tyr4‐Trp5‐Pro6‐), respectively. Compounds 3 and 4 were stable conformational isomers. Cyclopeptide 1 showed a selective inhibitory effect on collagen‐induced platelet aggregation with an IC50 value of 44.2 μM .  相似文献   

2.
tert‐Butoxy­carbonyl­glycyl‐dehydro­alanyl‐glycine methyl ester (systematic name: methyl {2‐[(tert‐butoxycarbonylamino)­acetamido]prop‐2‐enamido}acetate) (Boc0‐Gly1‐ΔAla2‐Gly3‐OMe), C13H21N3O6, has been structurally characterized by single‐crystal X‐ray diffraction and by density functional theory (DFT) calculations at the B3LYP/6–311+G** level. The peptide chain in both the solid‐state and calculated structures adopts neither β nor γ turns. All amino acid residues in the tripeptide sequence are linked trans to each other. The bond lengths and valence angles of the amino acid units in the crystal structure and gas phase are comparable. However, the conformation of the third glycyl residue (Gly3) is different in the crystalline state and in the gas phase. It is stabilized in the calculated structure by an additional intra­molecular short contact between Gly3 NH and methyl ester COMe groups.  相似文献   

3.
The α,β‐dehydro­phenyl­alanine residues influence the conformation of the title penta­peptide Boc0–Gly1–ΔZPhe2–Gly3–ΔEPhe4–l ‐Phe5p‐NA ethanol solvate, C42H43N7O9·C2H5OH. The first unsaturated phenyl­alanyl (ΔZPhe2) and the third glycyl (Gly3) residues form a type I β turn, while the second unsaturated phenyl­alanyl (ΔEPhe4) and the last phenyl­alanyl (l ‐Phe5) residues are part of a type II β turn. All the amino acids in the peptide are linked trans to one another. The crystal structure is stabilized by intra‐ and inter­molecular hydrogen bonds.  相似文献   

4.
The vapor pressures p the excess partial molar enthalpies of glycerol H Gly E the densities d and the thermal expansivities p of aqueous glycerol were measured at 25°C. From the vapor pressure data, the excess chemical potential of H2O µ W E was calculated, assuming that the partial pressure of glycerol p Gly is negligibly small. The excess chemical potential of glycerol µ Gly E was estimated by applying the Gibbs–Duhem relation and these data were used to calculate the excess partial molar entropies S Gly E . From the density data, the excess partial molar volumes of glycerol V Gly E and from the thermal expansivity data, the normalized cross fluctuations SV, introduced by us earlier, were evaluated. While the detailed manner in which glycerol modifies the molecular arrangement of H2O in its immediate vicinity is yet to be elucidated, the hydrogen bond probability in the bulk H2O away from solute molecules is reduced gradually as the glycerol composition increases to the point where putative presence of icelike patches is no longer possible. Thereupon, a qualitatively different mixing scheme seems to set in.  相似文献   

5.
Comparison of the crystal structures of two pentadehydropeptides containing ΔPhe residues, namely (Z,Z)‐N‐(tert‐butoxycarbonyl)glycyl‐α,β‐phenylalanylglycyl‐α,β‐phenylalanylglycine (or Boc0–Gly1–ΔZPhe2–Gly3–ΔZPhe4–Gly5–OH) methanol solvate, C29H33N5O8·CH4O, (I), and (E,E)‐N‐(tert‐butoxycarbonyl)glycyl‐α,β‐phenylalanylglycyl‐α,β‐phenylalanylglycine (or Boc0–Gly1–ΔEPhe2–Gly3–ΔEPhe4–Gly5–OH), C29H33N5O8, (II), indicates that the ΔZPhe residue is a more effective inducer of folded structures than the ΔEPhe residue. The values of the torsion angles ϕ and ψ show the presence of two type‐III′β‐turns at the ΔZPhe residues and one type‐II β‐turn at the ΔEPhe residue. All amino acids are linked trans to each other in both peptides. β‐Turns present in the peptides are stabilized by intramolecular 4→1 hydrogen bonds. Molecules in both structures form two‐dimensional hydrogen‐bond networks parallel to the (100) plane.  相似文献   

6.
The sequence and conformational effects on the gas-phase acidities of peptides have been studied by using two pairs of isomeric cysteine-polyglycine peptides, CysGly3,4NH2 and Gly3,4CysNH2. The extended Cooks kinetic method was employed to determine the gas-phase acidities using a triple quadrupole mass spectrometer with an electrospray ionization source. The ion activation was achieved via collision-induced dissociation experiments. The deprotonation enthalpies (Δacid H) were determined to be 323.9 ± 2.5 kcal/mol (CysGly3NH2), 319.2 ± 2.3 kcal/mol (CysGly4NH2), 333.8 ± 2.1 kcal/mol (Gly3CysNH2), and 321.9 ± 2.8 kcal/mol (Gly4CysNH2), respectively. The corresponding deprotonation entropies (Δacid S) of the peptides were estimated. The gas-phase acidities (Δacid G) were derived to be 318.4 ± 2.5 kcal/mol (CysGly3NH2), 314.9 ± 2.3 kcal/mol (CysGly4NH2), 327.5 ± 2.1 kcal/mol (Gly3CysNH2), and 317.4 ± 2.8 kcal/mol (Gly4CysNH2), respectively. Conformations and energetic information of the neutral and anionic peptides were calculated through simulated annealing (Tripos), geometry optimization (AM1), and single point energy calculations (B3LYP/6-31+G(d)), respectively. Both neutral and deprotonated peptides adopt many possible conformations of similar energies. All neutral peptides are mainly random coils. The two C-cysteine anionic peptides, Gly3,4(Cys-H)NH2, are also random coils. The two N-cysteine anionic peptides, (Cys-H)Gly3,4NH2, may exist in both random coils and stretched helices. The two N-cysteine peptides, CysGly3NH2 and CysGly4NH2, are significantly more acidic than the corresponding C-terminal cysteine ones, Gly3CysNH2 and Gly4CysNH2. The stronger acidities of the former may come from the greater stability of the thiolate anion resulting from the interaction with the helix-macrodipole, in addition to the hydrogen bonding interactions.  相似文献   

7.
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   

8.
An extensive exploration of Met-enkephalin in its zwitterionic form has been carried out, in order to characterize the different low-energy conformational domains accessible to this pentapeptide. The study builds on previous studies carried out in our lab for the neutral molecule, which provided the initial geometries from which the conformational space of the charged molecule could be scanned. The initial conformations were subjected to a series of high- and lowtemperature molecular dynamics simulations. Snapshots along each trajectory were taken, minimized, and used as starting points in further MD trajectories until no lower-energy conformers could be characterized. The CHARMm force field was used throughout the study for this purpose. The same search strategy was used in these studies simulating two different environmental conditions, a distance-dependent dielectric ? = r and a high constant dielectric ? = 80. In the low dielectric environment, the formation of the salt bridge dominates the structure. In the high dielectric environment, the screening of the electrostatic interactions results in weaker intramolecular interactions. In both cases, the Gly2–Gly3 β-turn-type structures are preferred over the Gly3–Phe4 turns, in marked contrast to what is found for the neutral molecule. The lowest-energy structures from both environmental conditions were reoptimized in the presence of a cluster of explicit water molecules. Reoptimization of the structures considering explicit water structures did not result in significant conformational changes for the structures characterized with the ε = r or ε = 80 environments.  相似文献   

9.
IR data of the neuropeptide substance P ( 1 ) and its synthetic segments des-(Arg1-Gln6)-substance P ( 6 ), des-(Arg1-Pro4)-substance P ( 4 ), des-(Arg1-Lys3)-substance P ( 3 ), and des-Arg1-substance P ( 2 ) indicate predominant β-structures in the solid state and α-helical structures in CF3CH2OH (amide I band shape analysis). In MeOH, the spectra of 1 suggest a partly helical structure. On membranes prepared from 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine, a C-terminal α-helix consisting of 8 or 9 peptide bonds appears to be induced (IR attenuated total reflection studies). Its perpendicular orientation on the membrane is suggested by the dichroic ratios of the amide-I and -II bands. This study is consistent with our CD experiments and lends support to the membrane structure of 1 predicted from the estimated amphiphilic moment, hydrophobic-association constant, and helix length.  相似文献   

10.
NBO analysis and hybrid density functional theory–based method (B3LYP/6-311+G**) was used to study the anomeric effects (AE), dipole–dipole interactions, and steric repulsion effects on the conformational properties of 2,3,5,6-tetrahalo-1,4-dioxane [halo = F (1), Cl (2), Br (3)], 2,3,5,6-tetrahalo-1,4-dithiane [halo = F (4), Cl (5), Br (6)], and 2,3,5,6-etrahalo-1,4-diselenane [halo = F (7), Cl (8), Br (9)]. B3LYP/6-311+G** results revealed a strong axial preference in compounds 13. Gibbs free energy difference (G eqG ax) values (e.g., ΔG eq-ax) between the axial and equatorial conformations of compound 1 to compound 3 are 8.19, 3.86, and 3.13 kcal mol?1, respectively, as calculated by the B3LYP/6-311+G** level of theory. On the other hand, the NBO analysis of donor–acceptor (bond–antibond) interactions revealed that the AE for compounds 1–3 are ?12.26, ?16.46, and ?18.11 kcal mol?1, respectively. Contrary to the increase of the AE values from compound 1 to compound 3, the increase of the steric repulsions (e.g., 1,3-syn-axial repulsions) could fairly explain the decrease of the axial conformation stability in compounds 13 compared to their equatorial conformations. Further, the correlations between the AE, structural parameters, and conformational behavior of compounds 4–9 have been investigated.  相似文献   

11.
Abstract

The conformation of lactal (1) in D2O and DMSO-d6 has been investigated by employing NMR techniques and molecular mechanics calculations. The glucal ring shows a 4 H5 conformation in deuterium oxide and DMSO-d6. In contrast, the glucal ring of the hexa-O-acetyl derivative (2) in CDCl3 can be described as an equilibrium between the 4 H5 and 4 H5 forms. The disaccharide 1 has a restricted flexibility with the Φ angle oscillating about -105 ± 30° and the Ψ angle varying about -145 ± 30°. Compound 2 shows a similar conformational behaviour.  相似文献   

12.
The Cinchona alkaloid analogs (+)- and (?)- 5 with a quinuclidine-2-methanol residue attached to C(2) of a 9,9′-spirobifluorene moiety were prepared as a racemic mixture by reacting lithiated 2-bromo-9,9′-spirobifluorene 7 with (2-ethoxycarbonyl)quinuclidine (±)- 6 to give ketone (±)- 8 , followed by diastereoselective reduction with diisobutylaluminum hydride (DIBAL-H). The absolute configuration at C(9) and C(8), i.e., at the methanol bridge and the adjacent quinuclidine C-atom, in the two enantiomers of 5 is identical to the configuration at the corresponding centers in (?)-quinine ( 1 ) and (+)-quinidine ( 2 ), respectively. For the optical resolution of (±)- 5 , a chiral stationary phase for HPLC was prepared by covalently bonding quinine via a thiol spacer to a silica-gel surface. The enantiomer separation was accomplished at an α value of 1.61 with (±)- 5 being eluted last, in agreement with 1H-NMR studies in CDCl3 which showed that (+)- 5 underwent a more stable host-guest association with quinine than (?)- 5 . 1H{1H} Nuclear Overhauser effect (NOE) difference spectroscopical analysis of the host-guest associations with quinine in CDCl3, combined with computer-model examinations, allowed the assignment of the absolute configurations as (+)-(8R,9S)- 5 and (?)-(8S,9R)- 5 . A detailed conformational analysis displayed excellent agreement between the results of computational methods (Monte Carlo multiple minimum simulations, analyses of the total energy as a function of the flexible dihedral angles in the molecule) and 1H{1H}-NOE difference spectroscopical data. It was found that (?)- 5 and (+)- 5 differ significantly in their conformational preference from their natural counterparts quinine ( 1 ) and quinidine ( 2 ). Whereas the natural alkaloids prefer the ‘open’ conformation, with the quinuclidine N-atom pointing away from the quinoline ring, analog (±)- 5 adopts preferentially (by ca. 4 kcal mol?1) a ‘closed’ conformation, in which the quinuclidine N-atom points into the cleft of the 9,9′-spirobifluorene moiety. Since the basic quinuclidine N-atom in the ‘closed’ conformation is sterically shielded from forming strong H-bonds, the new Cinchona alkaloid analogs form less stable host-guest associations via H-bonding than quinine or quinidine.  相似文献   

13.
Abstract

The poly-Diels-Alder addition between the new bisdiene 1,4-bis(5-methoxy-2-oxazolyl)benzene (4) and N,N′-hexamethylene-bis[2-(2,5-dihydro-2,5-dioxo-pyrrole-1-yl) acetamide] (7) is described. The structure of the resulting polyadduct 12 was proved by 1H NMR spectroscopy with the aid of the low-molecular-weight model compounds 1,4-bis(1,3-dihydro-7-hydroxy-1,3-dioxo-2-phenyl-pyrrolo[3,4-c] pyridine-4-yl)benzene (9) and N,N'-hexamethylene-bis[2-(1, 3-dihydro-7-hydroxy-6-methyl-1,3-dioxo-4-phenyl-pyrrolo [3,4-c]pyridine-2-yl)acetamide] (11). The reaction proceeds via the aromatization of the primarily formed cycloadducts. Polyadduct 12 shows a number average degree of polymerization Pn of about 11 – 12 (Mn = 8500 ? 9200 g/mol), calculated from 1H NMR endgroup signals.  相似文献   

14.
Abstract

Stereoselective electron-transfer between optically active Λ-[Co(EDDS)]? and [Co(diamine)3]2+ (diamine = racemic-1,2-diaminopropane (pn); R-1,2-diaminopropane (R-pn)) have been investigated in aqueous and in DMSO solutions. In aqueous solution, the reaction between Λ-[Co(EEDS)]? and [Co(pn)3]2+ produces four conformational isomers of the [Co(pn)3]3+ complex which have the Λ-forms in excess. Their optical purities are increased ob3>lelob2> lel2ob> lel3 . In contrast, the reaction in DMSO results in more δ isomers than Λ and the sequence of optical purities are in reversed order.

The reaction between Λ-[Co(EDDS)]? and [Co(R-pn)3]2+ in aqueous solution produces two conformational isomers of [Co(R-pn)3]3+; their absolute configurations are identified as δ-lel2ob and Λ-ob3 . In DMSO, in contrast, the reaction produces only one isomer which is identified as δ-lel2ob.  相似文献   

15.
An organic–inorganic coplanar-shaped hexa-CuII sandwiched hybrid, (H2en){[Cu(en)2]2[Cu2(en)2Cu4(H2O)2](B-α-AsW9O34)2}?·?5H2O (en?=?ethylenediamine) (1), has been synthesized via hydrothermal method and further characterized by IR spectroscopy, thermogravimetric analysis, and X-ray single-crystal diffraction. Single-crystal X-ray diffraction analysis shows that 1 contains an organic–inorganic hybrid polyoxoanion {Cu2(en)2Cu4(H2O)2(B-α-AsW9O34)2}6?, which can be described as a coplanar-shaped hexa-CuII cluster sandwiched by two trivacant [B-α-AsW9O34]9? fragments. Complex 1 represents a rare organic–inorganic coplanar-shaped hexa-CuII cluster sandwiched arsenotungstate. Magnetic susceptibility measurements indicate that 1 demonstrates ferromagnetic coupling interactions within the CuII centers.  相似文献   

16.
The synthesis and structural characterization of a tetrazine-based catecholamide (CAM) ligand, N,N′-bis(N″-(aminoethyl)-2,3-bis(hydroxy)benzamide)-1,2,4,5-tetrazine-3,6-diamine (5), were investigated. All compounds were characterized by 1H NMR spectroscopy, 13C NMR spectroscopy, and FTIR spectroscopy. The protonation equilibria of 5 and complexation capacities (log βpqr) of Cd2+, Co2+, and Cu2+ complexes of 5 were evaluated through potentiometric titration and spectrophotometric titration, respectively. Species independent pM value (=?log [M]free) was used to compare metal affinities with the final sequence Cu2+ > Cd2+ > Co2+. Results show that 5 has potential for heavy metal removal.  相似文献   

17.
The complexes [(η5-C9H6R)Re(CO)3] [R = nBu (8), tBu (9), CH(CH2)4 (10), Ph (11), Bz (12), 4-methoxyphenyl (13), 4-chlorophenyl (14)] were synthesized by refluxing substituted indenyl ligands [C9H7R] [R = nBu (1), tBu (2), CH(CH2)4 (3), Ph (4), Bz (5), 4-methoxyphenyl (6), 4-chlorophenyl (7)], and Re2(CO)10 in decalin. The molecular structures of 9, 10, 12, and 13 were determined by X-ray diffraction analysis. These four crystals have similar molecular structures of the mononuclear carbonyl complex. In each of these molecules, Re is η5-coordinated to the five-membered ring of the indenyl group. Complexes 814 have catalytic activity for Friedel-Crafts reactions of aromatic compounds with a variety of alkylation and acylation reagents. Compared with traditional catalysts, these mononuclear metal carbonyl complexes have obvious advantages such as high activities, mild reaction conditions, high selectivity, and environmentally friendly chemistry.  相似文献   

18.
The observation that per-2,6-O-methyl-3-O-benzoyl--cyclodextrin (1) displays some unusual conformational behaviour in solution has led to a detailed investigation by (dynamic) NMR spectroscopy of the equilibration process that occurs in solutions of per-2,3-O-benzoyl--cyclodextrin (3) and some related compounds (7–9) between conformational isomers with averagedC 6 andC 3 molecular symmetries in certain organic solvents such as benzene, dichloromethane, and chloroform. The solvent dependence of the conformational equilibrium is also reflected in a spread of values for the specific optical rotations for 3 from +9° in 1,1,2,2-tetrachloroethane, where there is a degenerate equilibrium between species withC 3 molecular symmetry, to +92° in acetone where a species with averagedC 6 symmetry is present.This paper is dedicated to the memory of the late Dr C. J. Pedersen.  相似文献   

19.
苯丙炔酸和苯丙烯酸酯苯并菲盘状液晶的合成及介晶性   总被引:2,自引:0,他引:2  
合成了7个含苯丙炔酸和苯丙烯酸酯链的苯并菲盘状液晶化合物C18H6(OCnH2n+1)5O2CR' [R'=C≡CC6H5, n=4~9 (1a1f); R'=CH=CHC6H5, n=6 (1g)]. 该系列化合物结构通过1H NMR, IR和元素分析表征. 液晶性通过差示扫描量热法和偏光显微镜进行了研究, 结果显示: 化合物均为六方柱状相热致型液晶; 含苯丙炔酸酯链苯并菲盘状液晶化合物1a1f, 随着烷氧链的增长, 清亮点呈现逐渐下降的趋势; 对于含苯丙烯酸酯链苯并菲化合物1g与具有同样软链长度的炔基酯链苯并菲化合物1c比较, 具有更低的熔点和结晶点, 而它们的清亮点几乎一致, 因而化合物1g有更宽的介晶性温度范围.  相似文献   

20.

Abstract  

Treatment of [Cp*MoCl4], 1 (Cp* = η5-C5Me5), with [LiBH4.thf] in toluene at −40 °C, followed by thermolysis with [(thf)Li{CH(PPh2–BH3)2}] results in the formation of a new class of phosphido bridged molybdaborane [(Cp*Mo)2B4H7(μ-PPh2)], 2 which has been characterized crystallographically. In addition, the above reaction also produces known [(Cp*Mo)2B5H9], 3 and an unusual molybdaborane [(Cp*Mo)2B5H8(O i Pr)], 4 ( i Pr = –CH(CH3)2). All the new compounds have been characterized in solution by 1H, 11B, 13C, 31P NMR spectroscopy and the structural types were unambiguously established by X-ray crystallographic analysis of compounds 2 and 4.  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