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.
Four isobutyl amides were isolated from the fruits of white pepper (Piper nigrum L.) and identified to be N-isobutyl-13-(3,4-methylenedioxyphenyl)-2E,4E,12E-tridecatrienamide (3, guineensine), N-isobutyl-2E,4E,8Z-eicosatrienamide (5), N-isobutyl-2E,4E-octadecadienamide (6) and N-isobutyl-2E,4E-decadienamide (7, pellitorine).  相似文献   

2.
From the fresh roots of Heliopsis longipes three new minor alkamides: longipinamide A (N-isobutyl-8,10-diynoic-3Z-undecenamide), longipenamide A (N-isobutyl-syn-8,9-dihydroxy-2E,6Z-decadienamide) and longipenamide B (N-isobutyl-syn-6,9-dihydroxy-2E,7E-decadienamide); three known alkamides: affinin (spilanthol, N-isobutyl-2E,6Z,8E-decatrienamide), N-isobutyl-2E,6Z-decadienamide and N-isobutyl-2E-decenamide; and 11 other known compounds were isolated. The structures of the three new minor alkamides were established by 1D and 2D NMR spectroscopy including 1H, 13C, DEPT, COSY, HSQC, and HMBC experiments, as well as by EI and FAB+ mass spectrometry. To our knowledge, this is the first report of the isolation of linear dihydroxyalkamides as natural products.  相似文献   

3.
A main pungent amide, spilanthol (1), and three alkamides, (2E)-N-(2-methylbutyl)-2-undecene-8,10-diynamide (2), (2E,7Z)-N-isobutyl-2,7-tridecadiene-10,12-diynamide (3), and (7Z)-N-isobutyl-7-tridecene-10,12-diynamide (4) were isolated from the flower heads of Spilanthes acmella L. var. oleracea Clarke. Their structures were established by spectroscopic methods. Compounds 2 and 4 were new and 3 was found for the first time in Spilanthes species. Chemotaxonomic aspects are discussed.  相似文献   

4.
Bioassay-guided fractionation of a methanol extract of the brown alga, Ishige sinicola, led to the isolation of five algicidal compounds. Their structures were determined to be α-monoglycerides of eicosa-5Z,8Z,11Z,14Z-tetraenoic (arachidonic) acid, octadeca-6Z,9Z,12Z,15Z-tetraenoic acid, linoleic acid and oleic acid, and 1-O-palmitoyl-3-O-(6-sulfo-α-D-quinovopyranosyl)-sn-glycerol on the basis of spectroscopic data and a comparison with the data in the literature. These glycerolipids showed moderate-to-high cell lysis activity against the red tide microalgal species, Heterosigma akashiwo, Karenia mikimotoi and Alexandrium catenella, at a concentration of 20 μg/mL.  相似文献   

5.
The methanol extract of Ehretia dicksonii provided (10E,12Z,15Z)-9-hydroxy-10,12,15-octadecatrienoic acid methyl ester (1) which was isolated as an anti-inflammatory compound. Compound 1 suppressed 12-O-tetradecanoyl-phorbol-13-acetate (TPA)-induced inflammation on mouse ears at a dose of 500 μg (the inhibitory effect (IE) was 43%). Linolenic acid methyl ester did not inhibit this inflammation at the same dose. However, the related compounds of 1, (9Z,11E)-13-hydroxy-9,11-octadecadienoic acid (5) and (9Z,11E)- 13-oxo-9,11-octadecadienoic acid (6), showed potent activity (IE500 μg of 63% and 79%, respectively). Compounds 1, 4 ((9Z,12Z,14E)-16-hydroxy-9,12,14-octadecatrienoic acid), 5 and 6 also showed inhibitory activity toward soybean lipoxygenase at a concentration of 10 μg/ml.  相似文献   

6.
[2–14C]-(2Z,4E)-γ-Ionylideneethanol and [2–14C]-(2Z,4E)-γ-ionylideneacetic acid were converted by Cercospora cruenta to [2–14C]-(2Z,4E)-1′,4′-dihydroxy-γ-ionylideneacetic acid and [2-14C]-(2Z,4E)-4′-hydroxy-γ-ionylideneacetic acid, which are intermediates of ABA biosynthesis in C. cruenta.  相似文献   

7.
Several compounds having the basic α-ionylideneacetic acid structure were tested in Cercospora rosicola resuspensions. At 100 μm, all the compounds inhibited abscisic acid (ABA) biosynthesis. Time studies with unlabelled and deuterated (2Z,4E)- and (2E,4E)-α-ionylideneacetic acids showed rapid conversions into both (2Z,4E)- and (2E,4E)-4′-keto-α-ionylideneacetic acids as major products. Incorporation of the label into ABA was specific for the 2Z,4E-isomer. Minor products, identified by GC-MS, were (2Z,4E)- and (2E,4E)-4′-hydroxy-α-ionylideneacetic acids and (2Z,4E)-1′-hydroxy-α-ionylideneacetic acid. The conversion to (2Z,4E)-l′-hydroxy-α-ionylideneacetic acid has not been previously reported and was specific for the 2Z,4E-isomer. A time study for the conversion of methyl esters of [2H3]-(2Z,4E)- and [2H3]-(2E,4E)-4′-keto-α-ionylideneacetates showed a slow introduction of the l′-hydroxyl group and specificity for 2Z,4E-isomer. Conversion of the ethyl esters of (2Z,4E)- and (2E,4E)-l′-hydroxy-α-ionylideneacetates into the ethyl esters of both ABA and (2E,4E)-ABA demonstrated that ABA can be formed by oxidation of the 4′-position after the insertion of the 1′-hydroxy group. The ethyl 1′-hydroxy acids were also isomerized to the corresponding ethyl (2Z,4E)- and ethyl (2E,4E)-3′-hydroxy-β-ionylideneacetates. Ethyl (2Z,4E)-1′-hydroxy acid also gave small amounts of ethyl l′,4′-trans-diol of ABA. These results suggest that ABA may be formed through a (2Z,4E)-1′-hydroxy-α-ionylidene-type intermediate in addition to the previously proposed route through (2Z,4E)-4′-keto-α-ionylideneacetic acid.  相似文献   

8.
(3E, 5Z)-3,5-Dodecadienyl acetate, the female sex pheromone of Phtheochroa cranaodes, was regio and stereo-selectively synthesized from 1-octyne and (E)-4-bromo-3-buten-1-ol by using Pd(PPh3)4, CuI and piperidine to afford the enyne (5). Further elaboration afforded the target pheromone. The synthetic pheromone was identified with the natural product by its MS and IR, data GLC retention time and biological activity.  相似文献   

9.
Cynasibirolide A ( 1 ), one new humulanolide sesquiterpene, together with four known analogs, asteriscanolide ( 2 ), (1S,8S)-8-hydroxyhumula-2Z,6E,9E-trien-1,12-olide ( 3 ), (1S,7R)-8-oxohumula-2Z,9E-dien-1,12-olide ( 4 ), and (+)-6,7,9,10-tetrahydroasteriscunolide ( 5 ) were isolated from the roots and rhizomes of Cynanchum acutum subsp. sibiricum. Their structures and configurations were elucidated by spectroscopic methods, including 2D-NMR techniques, and the structure of 1 was confirmed by single-crystal X-ray diffraction. All compounds were evaluated for their anti-complementary activity in vitro, and compound 3 exhibited anti-complement effect with CH50 value of 0.45 mM.  相似文献   

10.
Farnesyl and α-ionylideneethyl compounds with tertiary and quaternary amine functional groups were synthesized and their effects on abscisic acid (ABA) biosynthesis of Cercospora rosicola observed. The trimethylammonium compounds were lethal at 10 μm and inhibitory at 10 μm, but lesser amounts of α-ionylideneethyltrimethylammonium iodide enhanced ABA biosynthesis. N,N-Dimethylfarnesylamine had little effect on ABA biosynthesis. N,N-Dimethyl (2Z,4E)- and (2E,4E)-α-ionylideneethylamines inhibited ABA biosynthesis at 100 μm but had no or little effect at lower concentrations. Farnesol and farnesylpyrophosphate (FPP) enhanced ABA biosynthesis. FPP appears to be both a precursor and an inducer and farnesol is an inducer of ABA biosynthesis. N,N-Dimethyl (2Z,4E)- and (2E,4E)-α-ionylideneethylamines were converted to N,N-dimethyl (2Z,4E)- and (2E,4E)-4′-keto-α-ionylideneethylamines, respectively. These conversions are analogous to those reported for α-ionone and α-ionylideneacetic acids and show that basic as well as acidic and neutral compounds with α-ionone type rings can undergo oxidation at the 4′-position. α-Ionylideneacetic acids inhibited growth of C. rosicola and the dimethylamines enhanced growth. Complete feedback inhibition was obtained with 400 μm of ABA.  相似文献   

11.
Two phenolic amides were isolated from the fruits of white pepper (Piper nigrum L.) and identified to be N-trans-feruloyl tyramine (2a) and N-5-(4-hydroxyphenyl)-2E, 4E-pentadienoyl piperidine (6a) on the basis of chemical and specrtal evidence. Both compounds were synthesized.  相似文献   

12.
The epimeric diterpenes (+)‐(1S,3E,7E,11S,12S)‐verticilla‐3,7‐dien‐12‐ol ( 1 ), isolated from Bursera suntui, and (+)‐(1S,3E,7E,11S,12R)‐verticilla‐3,7‐dien‐12‐ol ( 2 ), isolated from Bursera kerberi, gave the same Wagner‐Meerwein rearrangement product (?)‐(1E,4Z,8Z,11S,12R)‐phomacta‐1,(15)4,8‐triene ( 3 ). The Et2O:BF3‐induced transformations evidence that verticillenes and phomactanes, both containing the bicyclo[9.3.1]pentadecane skeleton, are biogenetically related through the verticillen‐12‐yl cation ( A + ), which also is a key intermediate in the biosynthetic pathways to generate antitumor taxanes. Molecular modeling using the Monte Carlo protocol, followed by density functional theory (DFT) geometry optimization employing the hybrid functionals B3LYP and B3PW91, both with the DGDZVP basis set, secured the configuration of 3 as followed from the good agreement between the calculated and experimental vibrational circular dichroism spectra. Similar DFT calculations allowed determining the absolute configuration of (+)‐(1R,4R,5R,8S,9S,11S,12R,15R)‐1,15:4,5:8,9‐triepoxyphomactane ( 9 ), which surprisingly derives from epoxidation of the second minimum energy conformer of 3 .  相似文献   

13.
海南特有植物华石斛(Dendrobium sinense)的化学成分和药理活性研究报道较少。为了深入研究华石斛的化学成分,该研究采用MCI小孔树脂柱色谱、硅胶柱色谱和Sephadex LH-20柱色谱等多种现代分离纯化技术,从其全草乙醇提取物的乙酸乙酯部位中分离纯化了10个化合物。结果表明:对分离纯化得到的10个化合物鉴定为罗汉松脂素(1)、4,5-二羟基-2,3-二甲氧基-9,10-二氢菲(2)、华石斛素C(3)、对甲氧基苯乙醇(4)、顺式对羟基肉桂酸乙酯(5)、对羟基苯丙酸乙酯(6)、丁香醛(7)、3-羟基苯甲醛(8)、3,9-dihydroxy-megastigma-5-ene(9)和(9Z,12Z)-9,12-二烯十八碳酸甲酯(10)。其中,首次从华石斛中分离得到的有化合物1、4-6、8-10。通过体外抑制乙酰胆碱酯酶活性实验发现,化合物2、4和化合物 9能够使乙酰胆碱酯酶活力下降。  相似文献   

14.
The mallo prenol isolated from the leaves of Mallotus japonicus was elucidated to be a mixture of (2Z,6Z, 10Z, 14Z, 18Z, 22Z, 26E, 30E, 34E)-3,7,11,15,19,23,27,31,35,39-decamethyl-2,6,10,14,18,22,26,30,34,38-tetracontadecaen-1-ol and its C45- and C55-homologues and not the previously reported structure. The malloprenols were demonstrated to be biosynthesized by successive cis condensation of isoprene residues with (2E, 6E, 10E)-geranylgeranyl pyrophosphate.  相似文献   

15.
We examined the biosynthetic pathway of abscisic acid (ABA) after isopentenyl diphosphate in a fungus, Cercospora cruenta. All oxygen atoms at C-1, -1, -1′, and -4′ of ABA produced by this fungus were labeled with 18O from 18O2. The fungus did not produce the 9Z-carotenoid possessing γ-ring that is likely a precursor for the carotenoid pathway, but produced new sesquiterpenoids, 2E,4E-γ-ionylideneethane and 2Z,4E-γ-ionylideneethane, along with 2E,4E,6E-allofarnesene. The fungus converted these sesquiterpenoids labeled with 13C to ABA, and the incorporation ratio of 2Z,4E-γ-ionylideneethane was higher than that of 2E,4E-γ-ionylideneethane. From these results, we concluded that C. cruenta biosynthesized ABA by the direct pathway via oxidation of ionylideneethane with molecular oxygen following cyclization of allofarnesene. This direct pathway via ionylideneethane in the fungus is consistent with that in Botrytis cinerea, except for the positions of double bonds in the rings of biosynthetic intermediates, suggesting that the pathway is common among ABA-producing fungi.  相似文献   

16.
Buczek A  Wałęsa R  Broda MA 《Biopolymers》2012,97(7):518-528
The tendency to adopt β‐turn conformation by model dipeptides with α,β‐dehydrophenylalanine (ΔPhe) residue in the gas phase and in solution is investigated by theoretical methods. We pay special attention to a dependence of conformational properties on the side‐chain configuration of dehydro residue and the influence of N‐methylation on β‐turn stability. An extensive computational study of the conformational preferences of Z and E isomers of dipeptides Ac‐Gly‐(E/Z)‐ΔPhe‐NHMe ( 1a / 1b ) and Ac‐Gly‐(E/Z)‐ΔPhe‐NMe2 ( 2a / 2b ) by B3LYP/6‐311++G(d,p) and MP2/6‐311++G(d,p) methods is reported. It is shown that, in agreement with experimental data, Ac‐Gly‐(Z)‐ΔPhe‐NHMe has a great tendency to adopt β‐turn conformation. In the gas phase the type II β‐turn is preferred, whereas in the polar environment, the type I. On the other hand, dehydro residue in Ac‐Gly‐(E)‐ΔPhe‐NHMe has a preference to adopt extended conformations in all environments. N‐methylation of C‐terminal amide group, which prevents the formation of 1←4 intramolecular hydrogen bond, change dramatically the conformational properties of studied dehydropeptides. Especially, the tendency to adopt β‐turn conformations is much weaker for the N‐methylated Z isomer (Ac‐Gly‐(Z)‐ΔPhe‐NMe2), both in vacuo and in the polar environment. On the contrary, N‐methylated E isomer (Ac‐Gly‐(E)‐ΔPhe‐NMe2) can easier adopt β‐turn conformation, but the backbone torsion angles (?1, ψ1, ?2, ψ2) are off the limits for common β‐turn types. © 2012 Wiley Periodicals, Inc. Biopolymers 97:518–528, 2012.  相似文献   

17.
(±)-(2Z,4E)-α-Ionylideneacetic acid (2) was enantioselectively oxidized to (?)-(l′S)-(2Z,4E)-4′-hydroxy-α-ionylideneacetic acid (3), (+)-(1′R)-(2Z,4E)-4′-oxo-α-ionylideneacetic acid (4) and (+)-abscisic acid (ABA) (1) by Cercospora cruenta IFO 6164, which can produce (+)-ABA and (+)-4′-oxo-α-acid 4. This metabolism was confirmed by the incorporation of radioactivity from (±)-(2-14C)-(2Z,4E)-α-acid 2 into three metabolites. (?)-4′-Hydroxy-α-acid 3 was a diastereoisomeric mixture consisting of major 1′,4′-trance-4′-hydroxy-α-acid 3a and minor 1′,4′-cis-4′-hydroxy-α-acid 3b. These structures, 3a and 3b, were confirmed by 13C-NMR and 1H-NMR analysis. Also, the enantioselectivity of the microbial oxidation was reexamined by using optically pure α-acid (+)-2 and (?)-2, as the substrates.  相似文献   

18.
Synthesis of “reversed” methylenecyclopropane analogues of nucleoside phosphonates 6a,7a, 6b, and 7b is described. 1-Bromo-1-bromomethylcyclopropane 8 was converted to the bromocyclopropyl phosphonate 9 by Michaelis-Arbuzov reaction with triisopropyl phosphite. Base-catalyzed β-elimination and deacetylation gave the key Z- and E-hydroxymethylcyclopropyl phosphonates 10 and 11 separated by chromatography. The Mitsunobu type of alkylation of 10 or 11 with adenine or 2-amino-6-chloropurine afforded phosphonates 12a, 12b, 13a, and 13b. Acid hydrolysis furnished the adenine and guanine analogues 6a, 7a, 6b, and 7b. The E and Z configuration was assigned on the basis of NOE experiments with phosphonates 6b and 7b. All Z- and E-isomers were also distinguished by different chemical shifts of CH2O or CH2N (H4 or H4′). Significant differences of the chemical shifts of the cyclopropane C3(3’) carbons and coupling constants 3JP,C2(2’) or 3JP,C3(3’) selective for the Z- or E-isomers were also noted. Phosphonates 6a, 7a, 6b, and 7b are devoid of significant antiviral activity.  相似文献   

19.
The crystal and molecular structures of the ligand bpenH2 (N,N′-bis(2′-pyridinecarboxamide)-1,2-ethane) and its deprotonated dimeric cobalt(III) complex fac-[Co2(bpen)3]·12H2O have been determined by single-crystal X-ray diffraction methods. Crystal data: (a) bpenH2, C14H14N4O2, orthorhombic, space group Pccn, a=9.638(1), b= 15.288(1), c = 8.684(1) Å, Z=4; (b) Co2(bpen)3· 12H2O, C42H60N12O18Co2, triclinic, space group P1, a = 11.128(3), b = 14.316(5), c = 16.466(4) Å, α= 92.02(2)°, β = 95.21(2)°, γ = 99.30(2)°, Z = 2.The structures were refined to R 0.034 and 0.053 for 1064 and 7748 independent reflexions, respectively. The bpenH2 molecule has a space group imposed centre of symmetry, with the amide group adopting a trans configuration in the closely planar picolinamide moiety. The cobalt complex is dimeric in which three bpen ligands, acting each as a bis(N2-bidentate), bridge the two metal atoms. Each cobalt atom is octahedral with CoNpy 1.944(3) and Co Nam 1.933(3) Å. The Co··Co separation is 5.493(1) Å. The symmetry of the dimeric molecule is D3 which is consistent with that indicated from solution NMR studies.  相似文献   

20.
5-Oxo-(7E,9E,11Z,14Z)-eicosatetraenoic acid (5-oxo-ETE) has been identified as a non-enzymatic hydrolysis product of leukotriene A4 (LTA4) in addition to 5,12-dihydroxy-(6E,8E,10E,14Z)-eicosatetraenoic acids (5,12-diHETEs) and 5,6-dihydroxy-(7E,9E,11Z,14Z)-eicosatetraenoic acids (5,6-diHETEs). The amount of 5-oxo-ETE detected in the mixture of the hydrolysis products of LTA4 was found to be pH-dependent. After incubation of LTA4 in aqueous medium, the ratio of 5-oxo-ETE to 5,12-diHETE was 1:6 at pH 7.5, and 1:1 at pH 9.5. 5-Oxo-ETE was isolated from the alkaline hydrolysis products of LTA4 in order to evaluate its effects on human polymorphonuclear (PMN) leukocytes. 5-Oxo-ETE induced a rapid and dose-dependent mobilization of calcium in PMN leukocytes with an EC50 of 250 nM, as compared to values of 3.5 nM for leukotriene B4 (LTB4) and >500 nM for 5(S)-hydroxy-(6E,8Z,11Z,14Z)-eicosatetraenoic acid (5-HETE). Pretreatment of the cells with LTB4 totally abolished the calcium response induced by 5-oxo-ETE. In contrast, the preincubation with 5-oxo-ETE did not affect the calcium mobilization induced by LTB4. The calcium response induced by 5-oxo-ETE was totally inhibited by the specific LTB4 receptor antagonist LY223982. These data demonstrate that 5-oxo-ETE can induce calcium mobilization in PMN leukocyte via the LTB4 receptor in contrast to the closely related analog 5-oxo-(6E,8Z,11Z,14Z)-eicosatetraenoic acid which is known to activate human neutrophils by a mechanism independent of the receptor for LTB4.  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