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.
An unprecedented new natural product named nocarsin A (1), 5H‐4a,6,7a‐triazacyclopenta[cd]indene‐5,7(6H)‐dione (1), together with seven known compounds lumichrome (2), cyclo (L ‐Leu‐L ‐Tyr) (3), cyclo (L ‐Ala‐L ‐Ile) (4), cyclo (L ‐Ala‐L ‐Leu) (5), cyclo (L ‐Val‐L ‐Ala) (6), 5‐methyluracil (7) and uracil (8), was isolated from Nocardia alba sp.nov (YIM 30243T), which was isolated from a soil sample collected from Yunnan Province, P. R. China. NMR techniques including COSY, HSQC, ROESY, and HMBC were used to elucidate the structures of these compounds. We report the unambiguous assignments of the 1H and 13C NMR spectra of the new compound nocarsin A (1). Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

2.
From the roots of Pseudostellaria heterophylla, three cyclopeptides and three amides were isolated, besides heterophyllin A and B. Their structures were determined as cyclo (Ala‐Gly‐Pro‐Val‐Tyr‐) (heterophyllin J; 1 ), cyclo (Ala‐Gly‐Pro‐Tyr‐Leu‐) (pseudostellarin A; 2 ), cyclo (Gly‐Gly‐Gly‐Pro‐Pro‐Phe‐Gly‐Ile‐) (pseudostellarin B; 3 ), methyl γ‐hydroxypyroglutamate ( 4 ), methyl pyroglutamate ( 5 ), and pyroglutamic acid ( 6 ) on the basis of spectral data, especially 2D‐NMR data. Among them, compounds 1 and 4 are new compounds.  相似文献   

3.
Two new cyclic tetrapeptides, cyclo(l ‐Val‐l ‐Leu‐l ‐Val‐l ‐Ile) ( 1 ) and cyclo(l ‐Leu‐l ‐Leu‐l ‐Ala‐l ‐Ala) ( 2 ), and 15 known compounds, cyclo(Gly‐l ‐Leu‐Gly‐l ‐Leu) ( 3 ), cyclo(l ‐Ser‐l ‐Phe) ( 4 ), cyclo(l ‐Leu‐l ‐Ile) ( 5 ), cyclo(l ‐Tyr‐l ‐Phe) ( 6 ), cyclo(Gly‐l ‐Trp) ( 7 ), cyclo(l ‐Leu‐l ‐Tyr) ( 8 ), cyclo(Gly‐l ‐Phe) ( 9 ), cyclo(l ‐Phe‐trans‐4‐hydroxy‐l ‐Pro) ( 10 ), cyclo(l ‐Leu‐l ‐Leu) ( 11 ), cyclo(l ‐Val‐l ‐Phe) ( 12 ), cyclo(l ‐Val‐l ‐Leu) ( 13 ), cyclo(l ‐Ile‐l ‐Ile) ( 14 ), cyclo(l ‐Tyr‐l ‐Tyr) ( 15 ), turnagainolide A ( 16 ), and bacimethrin ( 17 ) were isolated from the fermentation broth of Streptomyces rutgersensis T009 obtained from Elaphodus davidianus excrement. Their structures were identified on the basis of spectroscopic analysis. Meanwhile, the absolute configurations of the amino acid residues of compounds 1 and 2 were determined by advanced Marfey method. Compound 3 was obtained from a natural source for the first time. The X‐ray single crystal diffraction data of bacimethrin ( 17 ) were also reported for the first time. Compounds 1  –  17 exhibited no antimicrobial activities with the MICs > 100 μg/ml.  相似文献   

4.
刘勉  叶蕴华 《中国化学》2002,20(11):1347-1353
IntroductionCyclicpeptides ,whichareconstrainedconforma tionallyandmoreresistanttoproteasedigestionsthantheirlinearprecursors ,havebeenofgreatinterestassynthetictargetsbothaspotentialdrugleadsandasmodelsforcon formationalanalysis .1 4 Currentmethodsforsynt…  相似文献   

5.
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   

6.
A new cyclopeptide, clausenain I ( 1 ), has been isolated by a multi‐step chromatography procedure from Clausena anisum‐olens. Its structure was elucidated as cyclo (‐Gly1‐Ile2‐Ile3‐Val4‐Leu5‐Ile6‐Ile7‐Leu8‐Leu9‐) by extensive 2D‐NMR spectroscopic methods and chemical evidence. It is the first time that a natural cyclic peptide has been isolated from the genus Clausena.  相似文献   

7.
Two new cyclopeptides, named arenariphilin A ( 1 ) and arenariphilin B ( 2 ), were isolated from the whole plants of Arenaria oreophila. Their structures were determined as cyclo‐(Thr‐Gly) ( 1 ) and cyclo‐(Ser1‐Gly ‐Ser2‐Ile ‐Phe1‐Phe2) ( 2 ) on the basis of spectral data, especially by 2D‐NMR.  相似文献   

8.
Four new 9,10‐secocycloartane (=9,19‐cyclo‐9,10‐secolanostane) triterpenoidal saponins, named huangqiyenins G–J ( 1 – 4 , resp.), were isolated from Astragalus membranaceus Bunge leaves. The acid hydrolysis of 1 – 4 with 1M aqueous HCl yielded D ‐glucose, which was identified by GC analysis after treatment with L ‐cysteine methyl ester hydrochloride. The structures of 1 – 4 were established by detailed spectroscopic analysis as (3β,6α,10α,16β,24E)‐3,6‐bis(acetyloxy)‐10,16‐dihydroxy‐12‐oxo‐9,19‐cyclo‐9,10‐secolanosta‐9(11),24‐dien‐26‐yl β‐D ‐glucopyranoside ( 1 ), (3β,6a,10α,24E)‐3,6‐bis(acetyloxy)‐10‐hydroxy‐12,16‐dioxo‐9,19‐cyclo‐9,10‐secolanosta‐9(11),24‐dien‐26‐yl β‐D ‐glucopyranoside ( 2 ), (3β,6α,9α,10α,16β,24E)‐3,6‐bis(acetyloxy)‐9,10,16‐trihydroxy‐9,19‐cyclo‐9,10‐secolanosta‐11,24‐dien‐26‐yl β‐D ‐glucopyranoside ( 3 ), and (3β,6α,10α,24E)‐3,6‐bis(acetyloxy)‐10‐hydroxy‐16‐oxo‐9,19‐cyclo‐9,10‐secolanosta‐9(11),24‐dien‐26‐yl β‐D ‐glucopyranoside ( 4 ).  相似文献   

9.
Reaction of [U(TrenTIPS)] [ 1 , TrenTIPS=N(CH2CH2NSiiPr3)3] with 0.25 equivalents of P4 reproducibly affords the unprecedented actinide inverted sandwich cyclo‐P5 complex [{U(TrenTIPS)}2(μ‐η55‐cyclo‐P5)] ( 2 ). All prior examples of cyclo‐P5 are stabilized by d‐block metals, so 2 shows that cyclo‐P5 does not require d‐block ions to be prepared. Although cyclo‐P5 is isolobal to cyclopentadienyl, which usually bonds to metals via σ‐ and π‐interactions with minimal δ‐bonding, theoretical calculations suggest the principal bonding in the U(P5)U unit is polarized δ‐bonding. Surprisingly, the characterization data are overall consistent with charge transfer from uranium to the cyclo‐P5 unit to give a cyclo‐P5 charge state that approximates to a dianionic formulation. This is ascribed to the larger size and superior acceptor character of cyclo‐P5 compared to cyclopentadienyl, the strongly reducing nature of uranium(III), and the availability of uranium δ‐symmetry 5f orbitals.  相似文献   

10.
Rate constants were determined for the reactions of OH radicals with halogenated cyclobutanes cyclo‐CF2CF2CHFCH2? (k1), trans‐cyclo‐CF2CF2CHClCHF? (k2), cyclo‐CF2CFClCH2CH2? (k3), trans‐cyclo‐CF2CFClCHClCH2? (k4), and cis‐cyclo‐CF2CFClCHClCH2? (k5) by using a relative rate method. OH radicals were prepared by photolysis of ozone at a UV wavelength (254 nm) in 200 Torr of a sample reference H2O? O3? O2? He gas mixture in an 11.5‐dm3 temperature‐controlled reaction chamber. Rate constants of k1 = (5.52 ± 1.32) × 10?13 exp[–(1050 ± 70)/T], k2 = (3.37 ± 0.88) × 10?13 exp[–(850 ± 80)/T], k3 = (9.54 ± 4.34) × 10?13 exp[–(1000 ± 140)/T], k4 = (5.47 ± 0.90) × 10?13 exp[–(720 ± 50)/T], and k5 = (5.21 ± 0.88) × 10?13 exp[–(630 ± 50)/T] cm3 molecule?1 s?1 were obtained at 253–328 K. The errors reported are ± 2 standard deviations, and represent precision only. Potential systematic errors associated with uncertainties in the reference rate constants could add an additional 10%–15% uncertainty to the uncertainty of k1k5. The reactivity trends of these OH radical reactions were analyzed by using a collision theory–based kinetic equation. The rate constants k1k5 as well as those of related halogenated cyclobutane analogues were found to be strongly correlated with their C? H bond dissociation enthalpies. We consider the dominant tropospheric loss process for the halogenated cyclobutanes studied here to be by reaction with the OH radicals, and atmospheric lifetimes of 3.2, 2.5, 1.5, 0.9, and 0.7 years are calculated for cyclo‐CF2CF2CHFCH2? , trans‐cyclo‐CF2CF2CHClCHF? , cyclo‐CF2CFClCH2CH2? , trans‐cyclo‐CF2CFClCHClCH2? , and cis‐cyclo‐CF2CFClCHClCH2? , respectively, by scaling from the lifetime of CH3CCl3. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 532–542, 2009  相似文献   

11.
The enantioselective condensing reagent 4,6‐dimethoxy‐1,3,5‐triazine (DMT)/strychnine/BF$\rm{{_{4}^{-}}}$ was obtained by treatment of 2‐chloro‐4,6‐dimethoxy‐1,3,5‐triazine (CDMT) with strychnine tetrafluoroborate. The reagent was useful under typical conditions of solid‐phase peptide synthesis (SPPS) with enantiomerically homogeneous substrates. By SPPS, desired dipeptides were obtained in 84–94% yield using 4 equiv. of racemic Fmoc‐Ala, Fmoc‐Phe, and/or Fmoc‐Tyr for 1 equiv. of Wang resin loaded with Gly, Ala, Leu, Phe, Glu(tBu), and/or Pro, respectively. For all three Fmoc‐protected amino acids, the configuration of the enantiomer preferred under SPPS conditions was independent of the structure of the acylated component and identical to that established in condensations proceeding in solution. In all cases, the enantiomer ratios L /D (er) were in a similar range, and varied from 9 : 92 to 2 : 98 for alanine, and from 90 : 10 to 100 : 0 for aromatic amino acids. The synthesis of Ac‐L ‐Lys(Ac)‐D ‐Ala‐D ‐Ala‐OH from racemic Fmoc‐Ala gave an L /D ratio of 10 : 90 for the esterification of Wang resin, and 0 : 100 for the formation of peptide bonds.  相似文献   

12.
Water‐soluble cationic alkynylplatinum(II) 2,6‐bis(benzimidazol‐2′‐yl)pyridine (bzimpy) complexes have been demonstrated to undergo supramolecular assembly with anionic polyelectrolytes in aqueous buffer solution. Metal–metal‐to‐ligand charge transfer (MMLCT) absorptions and triplet MMLCT (3MMLCT) emissions have been found in UV/Vis absorption and emission spectra of the electrostatic assembly of the complexes with non‐conjugated polyelectrolytes, driven by Pt???Pt and π–π interactions among the complex molecules. Interestingly, the two‐component ensemble formed by [Pt(bzimpy‐Et){C?CC6H4(CH2NMe3‐4)}]Cl2 ( 1 ) with para‐linked conjugated polyelectrolyte (CPE), PPE‐SO3?, shows significantly different photophysical properties from that of the ensemble formed by 1 with meta‐linked CPE, mPPE‐Ala. The helical conformation of mPPE‐Ala allows the formation of strong mPPE‐Ala– 1 aggregates with Pt???Pt, electrostatic, and π–π interactions, as revealed by the large Stern–Volmer constant at low concentrations of 1 . Together with the reasonably large Förster radius, large HOMO–LUMO gap and high triplet state energy of mPPE‐Ala to minimize both photo‐induced charge transfer (PCT) and Dexter triplet energy back‐transfer (TEBT) quenching of the emission of 1 , efficient Förster resonance energy transfer (FRET) from mPPE‐Ala to aggregated 1 molecules and strong 3MMLCT emission have been found, while the less strong PPE‐SO3?– 1 aggregates and probably more efficient PCT and Dexter TEBT quenching would account for the lack of 3MMLCT emission in the PPE‐SO3?– 1 ensemble.  相似文献   

13.
A kind of new lanthanocene complex with an ansa carbonous‐bridged cyclopentadienyl/aromatic heterocycle ligand was prepared and characterized. Based on the data of elemental analyses, MS and IR, they were presumed to be solvent‐free complexes (cyclo‐C4H3SCMe2C5H4)2LnCl [Ln = Er (1), Dd ( 2 ), Y ( 3 ), Sm ( 4 )]. These complexes were effective for the polymerization of methyl methacrylate in the presence of co‐catalyst. When AlEt3 and NaH (nanometric) were used as different co‐catalysts, the lanthanocene complexes 1–4 showed different catalytic behavior. These differences resulted from the formation of different active species. The catalyst system (cyclo‐C4H3SCMe2C5H4)2LnCl/NaH (nanometric) showed high catalytic activity (yield ≥ 95% and Mη > 105) in a short time at the ambient temperature. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

14.
Abstract

Strain HT88 was isolated from the fresh stems of Mallotus nudiflorus L, and it was identified as Nocardiopsis sp. by analyzing its morphology and the 16S rRNA sequence. The extracts of fermented HT88 showed potent antimicrobial activities. Bioassay guided separation of extracts led to eight proline (or hydroxyproline, Hyp)-containing cyclic dipeptides. Their structures were determined by 1D and 2D NMR spectroscopy and ESI mass spectrometry and further comparison with existing 1H and 13C NMR, melting points and specific rotation data. The eight 2,5-diketopiperazines (DKPs) were identified as cyclo(L-Pro-L-Leu) (1), cyclo(Pro-Leu) (2), cyclo(L-trans-Hyp-L-Leu) (3), cyclo(D-trans-Hyp-D-Leu) (4), and cyclo(D-Pro-L-Phe) (5), cyclo(L-Pro-L-Phe) (6), and cyclo(D-cis-Hyp-L-Phe) (7), cyclo(L-trans-Hyp-L-Phe) (8), respectively. Up to date, this is the first isolation of four pairs of proline based DKPs from Nocardiopsis sp.  相似文献   

15.
Quantum mechanics/molecular mechanics calculations in tyrosine ammonia lyase (TAL) ruled out the hypothetical Friedel–Crafts (FC) route for ammonia elimination from L ‐tyrosine due to the high energy of FC intermediates. The calculated pathway from the zwitterionic L ‐tyrosine‐binding state (0.0 kcal mol?1) to the product‐binding state ((E)‐coumarate+H2N? MIO; ?24.0 kcal mol?1; MIO=3,5‐dihydro‐5‐methylidene‐4H‐imidazol‐4‐one) involves an intermediate (IS, ?19.9 kcal mol?1), which has a covalent bond between the N atom of the substrate and MIO, as well as two transition states (TS1 and TS2). TS1 (14.4 kcal mol?1) corresponds to a proton transfer from the substrate to the N1 atom of MIO by Tyr300? OH. Thus, a tandem nucleophilic activation of the substrate and electrophilic activation of MIO happens. TS2 (5.2 kcal mol?1) indicates a concerted C? N bond breaking of the N‐MIO intermediate and deprotonation of the pro‐S β position by Tyr60. Calculations elucidate the role of enzymic bases (Tyr60 and Tyr300) and other catalytically relevant residues (Asn203, Arg303, and Asn333, Asn435), which are fully conserved in the amino acid sequences and in 3D structures of all known MIO‐containing ammonia lyases and 2,3‐aminomutases.  相似文献   

16.
The (3R,5S,6E,8S,10R)‐11‐amino‐3,5,8,10‐tetramethylundec‐6‐enoic acid (ATUA; 1 ), which was designed as a βII′‐turn mimic according to the concepts of allylic strain and 2,4‐dimethylpentane units, was incorporated into a cyclic RGD peptide. The three‐dimensional structure of cyclo(‐RGD‐ATUA‐) (=cyclo(‐Arg‐Gly‐Asp‐ATUA‐)) 4 in H2O was determined by NMR techniques, distance geometry calculations and molecular‐dynamics simulations. The RGD sequence of 4 shows high conformational flexibility but some preference for an extended conformation. The structural features of the RGD sequence of 4 were compared with the RGD moiety of cyclo(‐RGDfV‐) (=cyclo(‐Arg‐Gly‐Asp‐D ‐Phe‐Val‐)). In contrast to cyclo(‐RGDfV‐), which is a highly active αvβ3 antagonist and selective against αIIbβ3, cyclo(‐RGD‐ATUA‐) shows a lower activity and selectivity. The structure of the ATUA residue in the cyclic peptide resembles a βII′‐turn‐like conformation. Its middle part, adjacent to the C?C bond, strongly prefers the designed and desired structure.  相似文献   

17.
Three new sesquiterpenes, schisansphenins A ( 1 ) and B ( 2 ) and (?)‐γ‐cuparenol ( 3 ), were isolated from an acetone extract of the fruits of Schisandra sphenanthera. The known compound 4 was isolated for the frist time from a natural source. The structures of the isolated compounds were elucidated through extensive spectroscopic analyses, particularly 2D‐NMR experiments (1H,1H‐COSY, HMQC, HMBC, and NOESY). A plausible biogenetic pathway for schisansphenin B ( 2 ) is proposed. Compounds 2 and 3 significantly reduced activation of NF‐AT and NF‐κB in the luciferase‐reporter assay.  相似文献   

18.
Hpn, one of Helicobacter pylori′s nickel‐accessory proteins, is an amazingly peculiar protein: Almost half of its sequence consists of polyhistidyl (poly‐His) residues. Herein, we try to understand the origin of this naturally occurring sequence, thereby shedding some light on the bioinorganic chemistry of Hpn′s numerous poly‐His repeats. By using potentiometric, mass spectrometric, and various spectroscopic techniques, we studied the NiII‐ and CuII complexes of the wild‐type Ac‐THHHHYHGG‐NH2 fragment of Hpn and of its six analogues, in which consecutive residues (His or Tyr) were replaced by Ala (Ala‐substitution or Ala‐scan approaches), thereby resulting in Ac‐TAHHHYHGG‐NH2, Ac‐THAHHYHGG‐NH2, Ac‐THHAHYHGG‐NH2, Ac‐THHHAYHGG‐NH2, Ac‐THHHHAHGG‐NH2, and Ac‐THHHHYAGG‐NH2 peptides. We found that the His4 residue is critical for both NiII‐ and CuII‐ion binding and the effectiveness of binding varies even if the substituted amino acid does not take part in the direct binding interactions.  相似文献   

19.
We report here a new‐skeleton tricyclic diterpenoid, neorogioldiol ( 8 ), along with new prenylbisabolanes, rogioldiol D ( 6 ) and O11,15‐cyclo‐14‐bromo‐14,15‐dihydrorogiol‐3,11‐diol ( 5 ), and their putative biogenetic precursor, (−)‐geranyllinalool ( 7 ), isolated from the red seaweed Laurencia microcladia, which has colonized a small tract of the Tuscany coast called Il Rogiolo. In a case study of the assignment of the absolute configuration for molecules composed of chiral halves that are connected by single bonds, the absolute configuration of neorogioldiol ( 8 ) was based on a) the assumption of the identity of the cyclohexane moiety with co‐occurring (2S,3R,6S)‐rogiolal ( 4 ) and b) NMR‐derived relative configurations for the bicyclic moiety, and c) the combination of these two pieces of information by molecular‐mechanics‐aided conformational analysis, in agreement with NOE data.  相似文献   

20.
Two new series of Boc‐N‐α,δ‐/δ,α‐ and β,δ‐/δ,β‐hybrid peptides containing repeats of L ‐Ala‐δ5‐Caa/δ5‐Caa‐L ‐Ala and β3‐Caa‐δ5‐Caa/δ5‐Caa‐β3‐Caa (L ‐Ala = L ‐alanine, Caa = C‐linked carbo amino acid derived from D ‐xylose) have been differentiated by both positive and negative ion electrospray ionization (ESI) ion trap tandem mass spectrometry (MS/MS). MSn spectra of protonated isomeric peptides produce characteristic fragmentation involving the peptide backbone, the Boc‐group, and the side chain. The dipeptide positional isomers are differentiated by the collision‐induced dissociation (CID) of the protonated peptides. The loss of 2‐methylprop‐1‐ene is more pronounced for Boc‐NH‐L ‐Ala‐δ‐Caa‐OCH3 (1), whereas it is totally absent for its positional isomer Boc‐NH‐δ‐Caa‐L ‐Ala‐OCH3 (7), instead it shows significant loss of t‐butanol. On the other hand, second isomeric pair shows significant loss of t‐butanol and loss of acetone for Boc‐NH‐δ‐Caa‐β‐Caa‐OCH3 (18), whereas these are insignificant for its positional isomer Boc‐NH‐β‐Caa‐δ‐Caa‐OCH3 (13). The tetra‐ and hexapeptide positional isomers also show significant differences in MS2 and MS3 CID spectra. It is observed that ‘b’ ions are abundant when oxazolone structures are formed through five‐membered cyclic transition state and cyclization process for larger ‘b’ ions led to its insignificant abundance. However, b1+ ion is formed in case of δ,α‐dipeptide that may have a six‐membered substituted piperidone ion structure. Furthermore, ESI negative ion MS/MS has also been found to be useful for differentiating these isomeric peptide acids. Thus, the results of MS/MS of pairs of di‐, tetra‐, and hexapeptide positional isomers provide peptide sequencing information and distinguish the positional isomers. Copyright © 2010 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