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 46 毫秒
1.
Identification and synthesis of new depsides isolated from oakmoss (Evernia Prunastri (L .) ACH .) Lecanoric acid 5 and 4 new depsides have been isolated from oakmoss (Evernia Prunastri (L .) ACH .) extracts by means of silicagel column chromatography. Various spectral methods including 13C-NMR. were used in determining the structure of 4,2″-O-methylgyrophoric acid ( 1 ), lecanoric acid ( 5 ), 2′-O-methylevernic acid ( 2 ), 3′-methylevernic acid ( 3 ) and methyl 3′-methyllecanorate ( 4 ).  相似文献   

2.
Isolation and identification of evernine in ‘Oakmosse’ (Evernia prunastri (L. ) Ach.). A new depside, evernin ( 1 ), has been found in the lichen Evernia Prunastri (L. ) Ach. The determination of the structure and the spectral parameters of 1 , especially the 13C-NMR. spectrum, are described.  相似文献   

3.
Volatiles constituents of “Oakmoss”(Evernia Prunastri (L .) ACH .) The chemical consititution of different fractions of two oakmoss oleoresines were investigated by GC./MS. using SCOT. and capillary columns. Besides many known compounds, over 52 (sesquiterpenes, aliphatic and aromatic hydrocarbons) have not yet been reported as constituents of oakmoss.  相似文献   

4.
Synthesis of ‘D -Isothreonine’ and ‘L -Alloisothreonine’ Starting from L -Alanine Starting from L -alanine, ‘D -isothreonine’ ( = (2R, 3S)-3-amino-2-hydroxybutanoic acid) and ‘L -alloisothreonine’ ( = (2S, 3S)-3-amino-2-hydroxybutanoic acid) were synthesized.  相似文献   

5.
Three new fragrant compounds formerly isolated from the absolute oil of cassie (Acacia farnesiana WILLD .) have been synthesized through stereospecific reactions, and their respective structures thus confirmed. These unusual C11 constituents are cis-3-methyl-dec-3-enoic acid (Ia), trans-3-methyl-dec-4-enoic acid (IIa) and cis-3-methyl-dec-3-en-1-ol (Ib). The related ‘non natural’ stereomers, namely, trans-3-methyl-dec-3-enoic acid (X) and cis-3-methyl-dec-4-enoic acid (XIX), have also been synthesized, as well as the 3-methyl-decen-1-ols IIb, XVII and XX.  相似文献   

6.
Due to their inherent liability towards highly acidic conditions previously considered to be a prerequisite for data acquisition, betaxanthin structure dereplication by NMR spectroscopy has been scarcely reported and was, hitherto, exclusively based on 1H‐NMR data interpretation. Applying only slightly acidic conditions, we herein report the first 13C‐NMR data of two betaxanthins, i.e., indicaxanthin ( 1 ), isolated from yellow‐orange cactus pear fruits (Opuntia ficus‐indica [L.] Mill . cv. ‘Gialla’), and of miraxanthin V ( 2 ) from yellow Swiss chard petioles (Beta vulgaris L. ssp. cicla [L.] Alef . cv. ‘Bright Lights’), as derived by gHSQC‐ and gHMQC‐NMR experiments and inverse detection.  相似文献   

7.
Steroselective Alkylation at C(α) of Serine, Clyceric Acid, Threonine, and Tartaric Acid Involving Heterocyclic Enolates with Evocyelic Double Bonds The chiral, non-racemic title acids are converted to methyl dioxolane-(cf. 13 ), oxazoline-( 4 ) and oxazolidinecarboxylates (cf. 9 ). Deprotonation by Li(i-Pr) 2N at dry-ice temperature gives solutions of the lithium enolates A–D With exocyclic enolate double bonds. These are stable crough with respect to β-elimination (Scheme 1) to be alkylated with or without cosolvents such as HMPA or DMPU The products are formed in good to excellent yields and, with the exception of the tartrate-derived acetonlde (see Scheme 2), with diastereoselectivities above 90%. While the tartrate-and threonine-derived enolates ( A and B , resp.) are chiral due to the second stereogenic center of the precursors, the serine- and glyceric-acid-derived enolates ( A and B , resp.) are chiral due to the second sterogenic center of the precursors, the serine-nd glyceric-acid-derived enolates are non-racemic due to a tert butyl-substituted (pivalaldehyde-derived) acetal center ( C and D , resp.). The products of alkylation can be hydrolyzed to give α-branched tartaric acid (Scheme 2), allothreonine (Scheme 3), serine (Scheme 4), and glyceric-acid derivatives (Scheme 5) with quaternary stereogenic centers. The configurations of the products are determined by NOE-NMR measurements and by chemical correlation. These show that the dioxolane-derived enolates A and D are alkylated preferentially from that face of the ring which is already substituted (‘syn’-attack), while the dihydrooxazol-and oxazolidine-derived enolates B and C are alkylated from the opposite face (‘anti’-attack). The ‘syn’-attack is postulated to arise from strong folding of the heterocyclic ring due to electronic repulsion between the enolate π-system and non-bonding electron pairs on the heteroatoms (see Scheme 6).  相似文献   

8.
(S)‐5‐Benzoyloxymethyl‐3‐[(E)‐(dimethylamino)methylidene]tetrahydrofuran‐2‐one (6), prepared in 5 steps from L‐glutamic acid (1), was used as precursor in a one step ‘ring switching’ synthesis of (S)‐2‐hydroxy‐3‐heteroaryl‐l‐propyl benzoates 13‐18, 23, 24. In the reaction of 6 with 2‐aminopyridine (21) and 2‐amino‐4,6‐dimethylpyrimidine (22) the corresponding dimethylamine substitution products (25, 26) were obtained.  相似文献   

9.
Isolation and Identification of Furfuric Acid a New Depsidone from the Lichen Pseudevernia furfuracea (L.) Ach. The structure of furfuric acid, ( 1 ), a new depsidone, isolated from the lichen Pseudevernia furfuracea (L.) Ach. has been established as 4-formyl-3,8-dihydroxy-9-(2,4-dihydroxy-5-methoxycarbonyl-3,6-dimethylbenzyl)-1,6-dimethyl-11-oxo-11H-dibenzo[b,e][1,4]dioxepin-7-carboxylic acid by spectral and chemical methods.  相似文献   

10.
Pd‐Catalyzed amination of 3,5‐dibromo‐ and 3,5‐dichloropyridine ( 1a and 1b , resp.) with linear polyamines 2 leads to the formation of a new family of pyridine‐containing macrocycles 3 with an ‘exo’‐oriented pyridine N‐atom (Schemes 1 and 2). The dependence of the macrocycle yield on the nature of the halogen atom, the length of the polyamine chain and C/N atom ratio, and the composition of the catalytic system is studied. The synthesis of mono‐ and bis(5‐halopyridin‐3‐yl)‐substituted polyamines 4, 5, 8, 9 , and of 3,5‐bis(polyamino)‐substituted pyridines 6 is described (Schemes 3 and 4), and the use of these compounds as intermediates on the way to the macrocycles 7, 16 , and 18 with larger cavity (‘cyclodimers’ and ‘cyclotrimers’) is demonstrated (Schemes 510).  相似文献   

11.
Crystal and solution structures of the enantiomerically pure and the racemic pairs of (η3-allyl) {2-[2′-(diphenylphosphino)phenyl]-4,5-dihydro-4-phenyloxazole}palladium(II) hexafluorophosphates ( 1 , and rac- 1 , resp.) and tetraphenylborates ( 2 , and rac- 2 , resp.) as well as (η3-allyl){2-[2′-(diphenylphosphino)phenyl]-4,5-dihydro-4-isopropyloxazole}palladium(II) tetraphenylborate ( 3 ) were characterized by X-ray crystallography and 1H-NMR spectroscopy. In the solid state, rac- 1 and rac- 2 proved to be disordered with both diastereoisomeric complexes in the crystal. The complexes 2 and 3 exist only in the ‘exo’ form. The X-ray structures show that the [PdII3-allyl)] moiety may adopt different configurations between a nearly symmetrical three-electron PdII3-allyl) system and an asymmetrical allyl group with a η1- and a η2-bonding to the metal center. The [PdII3-allyl)] system of rac- 1 and of ‘endorac- 2 is closer to the former, and that of 2 , ‘exo’-rac- 2 , and 3 closer to the later geometry. The 1H-NMR spectra of the hexafluorophosphates 1 and rac- 1 show two sets of signals of the allylic protons in an ‘exo’/‘endo’ ratio of 2:3. The tetraphenylborates 2, rac- 2 , and 3 give only one set of broad signals of the allylic protons.  相似文献   

12.
Herein we disclosed the use of a remote ‘imidazole’‐based precatalyst [(para‐cymene)RuII(L)Cl]+, C‐1 where L=2‐(4‐substituted‐phenyl)‐1H‐imidazo[4,5‐f][1,10] phenanthroline) for the selective oxidation of a variety of alkyl arenes/heteroarenes and alcohols to their corresponding aldehydes or ketones in presence of tert‐butyl hydroperoxide (TBHP). The remote ‘imidazole’ moiety present in the complex facilitates the activation of oxidant and subsequent generation of active species via the release of para‐cymene from C‐1 , which in‐turn was less effective without the ‘imidazole’ moiety. The mechanistic features of C‐1 promoted oxidation of alkyl arenes were also assessed from spectroscopic, kinetic, and few control experiments. The substrate scope for C‐1 promoted oxidation reaction was assessed based on the selective oxidation of 27‐different alkyl arenes/heteroarenes and 25 different alcohols to their corresponding aldehydes/ketones in moderate to good yields.  相似文献   

13.
Rigorous regio- and stereospecific labeling experiments are performed to demonstrate the operation of the previously suggested operation of ‘isotopically sensitive branching’ in FeI-mediated C? H bond activation. For the hexane-1,6-diol/Fe+-complex, it is shown that dehydrogenation involves specifically the central C(3)/C(4) position, and the study of the stereospecifically labeled D ,L - and meso-[3,4-D2]-isotompomers 1e and 1f demonstrates that dehydrogenation proceedes via two competing pathways (i.e. ‘anti’- vs. ‘syn’-route). The contribution of these routes to the product formation is – due to a kinetic isotope effect – controlled by the relative configuration at the labeled positions C(3)/C(4). For the D ,L -form 1e , we estimate a ratio of 49:1 in favor of the ‘anti’-route; due to an isotope effect, this ratio drops to 4.3:1 for the meso-form 1f .  相似文献   

14.
A screen‐printed silver strip with a built‐in three‐in‐one electrode (SPAgE) configuration of Ag‐working, Ag‐counter and Ag/AgxO (silver oxides) pseudoreference electrodes has been developed for sensitive and selective electrochemical flow injection analysis (FIA) of aluminum chlorohydrate (ACH) present in antiperspirants, through the free Cl? ion liberated from ACH in aqueous medium, as a redox signal at Ag‐working electrode in pH 6 phosphate buffer solution (PBS). The solution phase and instrumental parameters were systematically optimized. The calibration graph was linear in the window 1–200 ppm concentration of ACH and the lowest detection limit (S/N=3) was 295 ppb with a slope of 0.0989 μA/ppm and regression coefficient of 0.998. Calculated relative standard deviation (RSD) values for the detection of 5 and 50 ppm ACH by this method are 2.21 % and 2.16 %, respectively. Four different antiperspirant deodorants real samples with and without ACH content were successfully analyzed and the detected values obtained were found to be in good agreement with the product labeled values.  相似文献   

15.
The position of the ‘Inclined W theory’, i.e., the linear correlation of the properties (Pi) of the lanthanide and the actinide ions with their ground state total orbital angular quantum numbers (L), Pi = wiL + ki, where wi and ki are the slopes and intercepts of the least squares lines, is made clear. The position of gadolinium is ambivalent, and it can be a common point for the symmetric ‘Inclined W plots’ or it may either belong to the second or the third tetrad depending on the system investigated. The propionate and the isobutyrate systems are used to elucidate the ‘Inclined W concept’. The ‘Inclined W parameters’ for the solvent extraction of the lanthanides with PMBP ligand have been calculated.  相似文献   

16.
The orientation of the cycloaddition of diazomethane on unsaturated branchedchain sugars has been studied. For 3-C-cyanomethylidene-3-deoxy-1,2-O-isopropylidene-α-D-glycero-tetrofuranose the orientation was ‘normal’ and did not depend on the configuration at the double bond. The same situation prevailed with derivatives of 3-deoxy-1,2:5,6-di-O-isopropylidene-3-C-methylidene-α-D-xylo-hexofuranose. For the 3-C-acylmethylidene- and the 3-C-cyanomethylidene-3-deoxy-1,2:5,6-di-O-isopropylidene-α-D-ribo-hexofuranoses, the trans-(H–C(3′)–C(2))-isomer gave the ‘normal’ cycloadduct whereas the cis-isomer gave predominantly the αabnormal spiro-pyrazoline. This observation represents the first instance where the regioselectivity of a cycloaddition reaction is affected by the geometrical isomerism of the dipolarophile. The most probable explanation of the phenomenon is the conformational perturbation about the C(4)--C(5) bond of the unsaturated sugars induced by a change in the configuration at C(3). The consequence of that ‘conformational transmission’ of a difference in configuration at C(3) is that the steric crowding on the cis- than in the trans-isomer. Several novel examples of a new series of C-glycosylidenic derivatives, the spiro-pyrazolines, are described.  相似文献   

17.
Six unsaturated γ-lactones, (Z)-5-octen-4-olide ( 1 ), (Z)-5-decen-4-olide ( 2 ).(Z)-6-nonen-4-olide ( 3 ), (Z)-6-dodecen-4-olide ( 4 ), (Z, Z)-6,9-dodecadien-4-olide ( 5 ), and tuberolide ( 6 ) have been identified for the first time in tuberose absolute (from Polianthes tuberosa L.). All structures were corroborated by synthesis and all, except 3 and 4 , are new.
  • 1 The name ‘tuberolactone’ has been suggested for (Z, Z)-2,7-decadien-5-olide [1]. We propose the name ‘tuberolide’ for the bicyclic lactone 6 . (IUPAC name (1R*,5S*,Z)-6-(2′-pentenyl)-2-oxabicyclo[3.3.0]octan-3-one).
  • An improved method for the stereoselective synthesis of (±)-cis-bicyclo [4.3.0]-non-3-en-7-one ( 23 ) by an AlCl3-catalyzed Diels-Alder reaction is reported.  相似文献   

    18.
    In the context of Eschenmoser's work on pyranosyl‐RNA (‘p‐RNA’), we investigated the synthesis and base‐pairing properties of the 5‐methylisocytidine derivative. The previously determined clear‐cut restrictions of base‐pairing modes of p‐RNA had led to the expectation that a 5‐methylisocytosine β‐D ‐ribopyranosyl (= D ‐pr(MeisoC)) based (4′ → 2′)‐oligonucleotide would pair inter alia with D ‐pr(isoG) and L ‐pr(G) based oligonucleotides (D ‐pr and L ‐pr = pyranose form of D ‐ and L ‐ribose, resp.). Remarkably, we could not observe pairing with the D ‐pr(isoG) oligonucleotide but only with the L ‐pr(G) oligonucleotide. Our interpretation concludes that this – at first hand surprising – observation is caused by a change in the nucleosidic torsion angle specific for isoC.  相似文献   

    19.
    From the carbolithiation of 1‐(cyclopenta‐2,4‐dien‐1‐ylidene)‐N,N‐dimethylmethanamine (=6‐(dimethylamino)fulvene; 3 ) and different lithiated azaindoles 2 (1‐methyl‐7‐azaindol‐2‐yl, 1‐[(diethylamino)methyl]‐7‐azaindol‐2‐yl, and 1‐(methoxymethyl)‐7‐azaindol‐2‐yl), the corresponding lithium cyclopentadienide intermediates 4a – 4c were formed (7‐azaindole=1H‐pyrrolo[2,3‐b]pyridine). The latter underwent a transmetallation reaction with TiCl4 resulting in the (dimethylamino)‐functionalised ‘titanocenes’ 5a – 5c . When the ‘titanocenes’ 5a – 5c were tested against LLC‐PK cells, the IC50 values obtained were of 8.8, 12, and 87 μM , respectively. The most cytotoxic ‘titanocene’, 5a , with an IC50 value of 8.8 μM is nearly as cytotoxic as cis‐platin, which showed an IC50 value of 3.3 μM when tested on the epithelial pig kidney LLC‐PK cell line, and ca. 200 times better than ‘titanocene dichloride’ itself.  相似文献   

    20.
    Novel three‐residue helix‐turn secondary structures, nucleated by a helix at the N terminus, were generated in peptides that have ‘β‐Caa‐L ‐Ala‐L ‐Ala,’ ‘β‐Caa‐L ‐Ala‐γ‐Caa,’ and ‘β‐Caa‐L ‐Ala‐δ‐Caa’ (in which βCaa is C‐linked carbo‐β‐amino acid, γCaa is C‐linked carbo‐γ‐amino acid, and δ‐Caa is C‐linked carbo‐δ‐amino acid) at the C terminus. These turn structures are stabilized by 12‐, 14‐, and 15‐membered (mr) hydrogen bonding between NH(i)/CO(i+2) (i+2 is the last residue in the peptide) along with a 7‐mr hydrogen bond between CO(i)/NH(i+2). In addition, a series of α/β‐peptides were designed and synthesized with alternating glycine (Gly) and (S)‐β‐Caa to study the influence of an achiral α‐residue on the helix and helix‐turn structures. In contrast to previous results, the three ‘β–α–β’ residues at the C terminus (α‐residue being Gly) are stabilized by only a 13‐mr forward hydrogen bond, which resembles an α‐turn. Extensive NMR spectroscopic and molecular dynamics (MD) studies were performed to support these observations. The influence of chirality and side chain is also discussed.  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