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.
Two new halogenated eremophilane‐type sesquiterpenoids, (11S)‐ and (11R)‐12‐chloronootkaton‐11‐ol ( 1 and 2 , resp.), together with five known sesquiterpenoids, nootkatone ( 3 ), 7‐epiteucrenone B ( 4 ), oxyphyllenodiol A ( 5 ), oxyphyllenodiol B ( 6 ), and alpinenone ( 7 ), were isolated from the EtOH extract of the fruits of Alpinia oxyphylla Miq . The structures were elucidated based on the analyses of their spectroscopic data.  相似文献   

2.
Seven new sesquiterpenoids, namely eupatochinilides Ⅰ-Ⅶ (1-7), together with eight known compounds, euponin (8), mollisorin A (9), niveusin B (10), 8β-(4'-acetoxy-tiglyloxy)-3β-hydroxy-6Hβ,7Hα-germacra-1(10)E,4E,11(13)-trien-6,12-olide (11), eupalinifide B (12), 8β-(4'-hydroxytigloyloxy)-5-desoxy-8-desacyleuparotin (13), 3-deacetyeupalinin A (14), and 15-hydroxyleptocarpin (15), were isolated from the ethanolic extract of the whole plant of Eupatorium chinense L. Their structures and stereochemistry were established by spectroscopic methods and GIAO based ^13C NMR chemical shift calculations.  相似文献   

3.
Eight new metabolites, including five new sesquiterpenoids, 10,11‐epoxyguaian‐10‐ol ( 1 ), 10,11‐epoxyguaian‐13‐ol ( 2 ), a new backbone sesquiterpene rearranged from guaiane ( 3 ), two 1,5 : 1,10‐disecoguaianes, 4 and 5 , two new dihydroisocoumarins, 7‐chloromellein‐4‐ol ( 6 ) and 7‐chloromellein‐5‐ol ( 7 ), and one new tetralone, 7‐chloroscytalone ( 8 ), were isolated from the mutant strain G‐444 of Tubercularia sp. TF5, an endophytic fungus of Taxus mairei, along with ten known compounds, 3,4‐dihydro‐4,8‐dihydroxy‐2H‐naphthalen‐1‐one ( 9 ), (3R,4S)‐4‐hydroxymellein ( 10 ), 5‐formylmellein ( 11 ), 5‐carboxymellein ( 12 ), sporogen‐AO1 ( 13 ), tuberculariols A ( 14 ) and B ( 15 ), hymatoxin E ( 16 ), 4‐oxo‐4H‐pyran‐3‐acetic acid ( 17 ), and penicillic acid ( 18 ). Their structures were elucidated by spectroscopic analyses including HR‐ESI‐MS, 1D‐ and 2D‐NMR (HMQC, HMBC, 1H,1H‐COSY and NOESY). The antimicrobial activities of 1 – 8 were evaluated, but none showed any substantial effect.  相似文献   

4.
A series of 7‐fluorinated 7‐deazapurine 2′‐deoxyribonucleosides related to 2′‐deoxyadenosine, 2′‐deoxyxanthosine, and 2′‐deoxyisoguanosine as well as intermediates 4b – 7b, 8, 9b, 10b , and 17b were synthesized. The 7‐fluoro substituent was introduced in 2,6‐dichloro‐7‐deaza‐9H‐purine ( 11a ) with Selectfluor (Scheme 1). Apart from 2,6‐dichloro‐7‐fluoro‐7‐deaza‐9H‐purine ( 11b ), the 7‐chloro compound 11c was formed as by‐product. The mixture 11b / 11c was used for the glycosylation reaction; the separation of the 7‐fluoro from the 7‐chloro compound was performed on the level of the unprotected nucleosides. Other halogen substituents were introduced with N‐halogenosuccinimides ( 11a → 11c – 11e ). Nucleobase‐anion glycosylation afforded the nucleoside intermediates 13a – 13e (Scheme 2). The 7‐fluoro‐ and the 7‐chloro‐7‐deaza‐2′‐deoxyxanthosines, 5b and 5c , respectively, were obtained from the corresponding MeO compounds 17b and 17c , or 18 (Scheme 6). The 2′‐deoxyisoguanosine derivative 4b was prepared from 2‐chloro‐7‐fluoro‐7‐deaza‐2′‐deoxyadenosine 6b via a photochemically induced nucleophilic displacement reaction (Scheme 5). The pKa values of the halogenated nucleosides were determined (Table 3). 13C‐NMR Chemical‐shift dependencies of C(7), C(5), and C(8) were related to the electronegativity of the 7‐halogen substituents (Fig. 3). In aqueous solution, 7‐halogenated 2′‐deoxyribonucleosides show an approximately 70% S population (Fig. 2 and Table 1).  相似文献   

5.
The preparation of (2S,3S)‐ and (2R,3S)‐2‐fluoro and of (3S)‐2,2‐difluoro‐3‐amino carboxylic acid derivatives, 1 – 3 , from alanine, valine, leucine, threonine, and β3h‐alanine (Schemes 1 and 2, Table) is described. The stereochemical course of (diethylamino)sulfur trifluoride (DAST) reactions with N,N‐dibenzyl‐2‐amino‐3‐hydroxy and 3‐amino‐2‐hydroxy carboxylic acid esters is discussed (Fig. 1). The fluoro‐β‐amino acid residues have been incorporated into pyrimidinones ( 11 – 13 ; Fig. 2) and into cyclic β‐tri‐ and β‐tetrapeptides 17 – 19 and 21 – 23 (Scheme 3) with rigid skeletons, so that reliable structural data (bond lengths, bond angles, and Karplus parameters) can be obtained. β‐Hexapeptides Boc[(2S)‐β3hXaa(αF)]6OBn and Boc[β3hXaa(α,αF2)]6‐OBn, 24 – 26 , with the side chains of Ala, Val, and Leu, have been synthesized (Scheme 4), and their CD spectra (Fig. 3) are discussed. Most compounds and many intermediates are fully characterized by IR‐ and 1H‐, 13C‐ and 19F‐NMR spectroscopy, by MS spectrometry, and by elemental analyses, [α]D and melting‐point values.  相似文献   

6.
The four new sesquiterpenoids 1 – 4 , and the new 2‐(2‐phenylethyl)‐4H‐chromen‐4‐one (=2‐(2‐phenylethyl)‐4H‐1‐benzopyran‐4‐one) derivative 5 , together with the two known sesquiterpenoids 6 and 7 , the five known chromenones 8 – 12 , and 1‐hydroxy‐1,5‐diphenylpentan‐3‐one ( 13 ), were isolated from a 70% MeOH extract of Aquilaria malaccensis agarwood chips. Their structures were elucidated on the basis of comprehensive spectral analyses and comparison with literature data.  相似文献   

7.
Two new diterpenoids, pedunculatic acid A (= (4R,5α,7α)‐7‐ethoxy‐9β,13β‐dioxyabiet‐8(14)‐en‐18‐oic acid; 1 ) and pedunculatic acid B (= (4S,5α,12β)‐8β,14β‐epoxy‐12‐hydroxy‐11‐oxototaran‐19‐oic acid; 2 ), together with three known sesquiterpenoids, were isolated from the Chinese medicinal herb Callicarpa pedunculata R. Brown . Their structures were elucidated by spectroscopic analyses, including 1D‐ and 2D‐NMR, and by high‐resolution mass spectrometry.  相似文献   

8.
Addition of various amines to the 3,3‐bis(trifluoromethyl)acrylamides 10a and 10b gave the tripeptides 11a – 11f , mostly as mixtures of epimers (Scheme 3). The crystalline tripeptide 11f 2 was found to be the N‐terminal (2‐hydroxyethoxy)‐substituted (R,S,S)‐ester HOCH2CH2O‐D ‐Val(F6)‐MeLeu‐Ala‐OtBu by X‐ray crystallography. The C‐terminal‐protected tripeptide 11f 2 was condensed with the N‐terminus octapeptide 2b to the depsipeptide 12a which was thermally rearranged to the undecapeptide 13a (Scheme 4). The condensation of the epimeric tripeptide 11f 1 with the octapeptide 2b gave the undecapeptide 13b directly. The undecapeptides 13a and 13b were fully deprotected and cyclized to the [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐D ‐valine]]‐ and [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐L ‐valine]]cyclosporins 14a and 14b , respectively (Scheme 5). Rate differences observed for the thermal rearrangements of 12a to 13a and of 12b to 13b are discussed.  相似文献   

9.
Two new guaiane sesquiterpenoids, (1β,5β)‐1‐hydroxyguaia‐4(15),11(13)‐dieno‐12,5‐lactone ( 1 ) and 1,5‐epoxy‐4‐hydroxyguai‐11(13)‐en‐12‐oic acid ( 2 ), together with five known compounds, rupestonic acid ( 3 ), strobilactone A ( 4 ), antiquorin ( 5 ), isosclerone ( 6 ), and 5‐hydroxy‐2′,3′,4′,7,8‐pentamethoxyflavone ( 7 ), were isolated from a 95% EtOH extract of Artemisia rupestris. Compounds 1 and 2 are rare examples of guaiane sesquiterpenoids, incorporating a 12,5‐lactone group or featuring a 1,5‐epoxy ring, respectively. The structures of 1 and 2 were identified by various spectroscopic methods. Compounds 1, 4 , and 5 exhibited moderate cytotoxic activities against the human lung cancer 95‐D cell line with IC50 values of 11.3, 19.8, and 34.5 μM , respectively.  相似文献   

10.
Two new compounds, (6S,13S)‐6‐{[β‐D ‐glucopyranosyl‐(1→4)‐α‐L ‐rhamnopyranosyl]oxy}cleroda‐3,14‐dien‐13‐ol ( 1 ) and kadsuric acid 3‐methyl ester ( 2 ), together with nine known compounds, (6S,13E)‐6‐{[β‐D ‐glucopyranosyl‐(1→4)‐α‐L ‐rhamnopyranosyl]oxy}cleroda‐3,13‐dien‐15‐ol ( 3 ), (6S,13S)‐6‐[6‐O‐acetyl‐β‐D ‐glucopyranosyl‐(1→4)‐α‐L ‐rhamnopyranosyl]oxy}‐13‐{[α‐L ‐rhamnopyranosyl‐(1→4)‐β‐D ‐fucopyranosyl]oxy}cleroda‐3,14‐diene ( 4 ), (6S,13S)‐6‐{[6‐Oβ‐D ‐glucopyranosyl‐(1→4)‐α‐L ‐rhamnopyranosyl]oxy}‐13‐{[α‐L ‐rhamnopyranosyl‐(1→4)‐β‐D ‐fucopyranosyl]oxy}cleroda‐3,14‐diene ( 5 ), 15‐hydroxydehydroabietic acid ( 6 ), 15‐hydroxylabd‐8(17)‐en‐19‐oic acid ( 7 ), junicedric acid ( 8 ), (4β)‐kaur‐16‐en‐18‐oic acid ( 9 ), (4β)‐16‐hydroxykauran‐18‐oic acid ( 10 ), and (4β,16β)‐16‐hydroxykauran‐18‐oic acid ( 11 ) were isolated from the fronds of Dicranopteris linearis or D. ampla. Their structures were established by extensive 1D‐ and 2D‐NMR spectroscopy. Compounds 1 and 3 – 8 showed no anti‐HIV activities.  相似文献   

11.
Two series of territrem B analogs, i.e., 5 – 10 , containing both the 2‐en‐1‐one‐A‐ring and the aromatic‐E‐ring pharmacophores were designed and synthesized from jujubogenin ( 4a ). The anti‐acetylcholinesterase (anti‐AChE), anti‐caspase‐3, and other biological activities of these territrem‐B analogs and their intermediates were assessed. Compound 9b, 22a , and 24f were shown to be weak inhibitors of AChE. None of the synthesized compounds exhibited significant inhibitory activity on caspase‐3. On the other hand, compounds 22e, 24a, 7b , and 8a showed mild cytotoxicity on cultured KB cells, with IC50 values of 2.0, 3.5, 6.5, and 14 μM , respectively. In addition, compounds 23b and 5f were active against injury arising from oxygen‐glucose deprivation.  相似文献   

12.
Nine new sesquiterpenes, i.e., dendronobilins A–I ( 1 – 9 ), with copacamphane‐type ( 1 ), picrotoxane‐type ( 2 – 6 ), muurolene‐type ( 7 ), alloaromadendrane‐type ( 8 ), and cyclocopacamphane‐type ( 9 ) skeletons, were isolated from the 60% EtOH extract of the stems of Dendrobium nobile. Their structures were established as (1R,2R,4S,5S,6S,8S,9R)‐2,8‐dihydroxycopacamphan‐15‐one ( 1 ), (2β,3β,4β,5β)‐2,4,11‐trihydroxypicrotoxano‐3(15)‐lactone ( 2 ), (2β,3β,5β,9α,11β)‐2,11‐epoxy‐9,11,13‐trihydroxypicrotoxano‐3(15)‐lactone ( 3 ), (2β,3β,5β,12R*)‐2,11,13‐trihydroxypicrotoxano‐3(15)‐lactone ( 4 ), (2β,3β,5β,12S*)‐2,11,13‐trihydroxypicrotoxano‐3(15)‐lactone ( 5 ), (2β,3β,5β,9α)‐9,10‐cyclo‐2,11,13‐trihydroxypicrotoxano‐3(15)‐lactone ( 6 ), (9β,10α)‐muurol‐4‐ene‐9,10,11‐triol ( 7 ), (10α)‐alloaromadendrane‐10,12,14‐triol ( 8 ), and (5β)‐cyclocopacamphane‐5,12,15‐triol ( 9 ) on the basis of spectroscopic analysis. The absolute configuration of compound 1 was tentatively assigned as (1R,2R,4S,5S,6S,8S,9R) according to its CD spectrum and the octant rule. Compounds 1 and 4 – 9 were inactive in our preliminary in vitro immunomodulatory bioassay.  相似文献   

13.
The reactions of 4N‐ethyl‐2‐[1‐(pyrrol‐2‐yl)methylidene(hydrazine carbothioamide ( 4 EL1 ) and 4N‐ethyl‐2[1‐(pyrrol‐2‐yl)ethylidene(hydrazine carbothioamide ( 4 EL2 ) with Group 12 metal halides afforded complexes of types [M(L)2X2] (M = Zn, Cd; L = 4 EL1, 4 EL2; X = Cl, Br, I; 1 – 6 , 14 – 19 ) and [M(L)X2] (M = Hg; L = 4 EL1, 4 EL2; X = Cl, Br, I; 7 – 9 , 20 – 22 ). In addition, reaction of 4 EL1 with salts of CuII, NiII, PdII and PtII afforded compounds of type [M(4 EL1–H)2] ( 10 – 13 ). The new compounds were characterized by elemental analysis, FAB mass spectrometry, IR and electronic spectroscopy and, for sufficiently soluble compounds, 1H, 13C and, when appropriate, 113Cd or 199Hg NMR spectrometry. The spectral data suggest that in their complexes with Group 12 metal cations, both thiosemicarbazones are neutral and S‐monodentate; and for [Zn(4 EL1)2I2] ( 3 ), [Cd(4 EL1)2Br2] ( 5 ) and [Hg(4 EL1)Cl2]2 ( 7 ) this was confirmed by X‐ray diffractometry. By contrast, in its complexes with CuII and Group 10 metal cations, 4 EL1 is monodeprotonated and S,N‐bidentate, as was confirmed by X‐ray diffractometry for [Ni(4 EL1–H)2] ( 11 ) and [Pd(4 EL1–H)2] ( 12 ).  相似文献   

14.
Two new eremophilane‐type sesquiterpenoids eremophil‐6‐en‐11‐ol ( 1 ) and (7α,9α,10α)‐9,10‐epoxy‐eremophilan‐11‐ol ( 2 ), together with a known eremophilane‐type (6α,8α)‐6,8‐dihydroxyeremophil‐7(11)‐en‐12‐oic acid 12,8‐lactone ( 3 ) were isolated from the rhizomes of Ligularia veitchiana. The structures of 1 and 2 were established by spectral analysis including 1H‐ and 13C‐NMR, HSQC, HMBC, and HR‐ESI‐MS data. The compounds 1 and 3 were assessed against lung‐cancer (A549) and stomach‐cancer (BCG823) cell lines by the MTT method. The results showed that 1 exhibited significant inhibiting activities on the growth of these cancer cells with IC50 values between 1–100 μg/ml, whereas compound 3 had no effect on the same cell lines.  相似文献   

15.
Di(tert‐butyl)diazomethane ( 4 ) is a nucleophilic 1,3‐dipole with strong steric hindrance at one terminus. In its reaction with 2,3‐bis(trifluoromethyl)fumaronitrile ((E)‐ BTE ), a highly electrophilic tetra‐acceptor‐substituted ethene, an imino‐substituted cyclopentene 9 is formed as a 1 : 2 product. The open‐chain zwitterion 10 , assumed as intermediate, adds the second molecule of (E)‐ BTE . The 19F‐ and 13C‐NMR spectra allow the structural assignment of two diastereoisomers, 9A and 9B . The zwitterion 10 can also be intercepted by dimethyl 2,3‐dicyanofumarate ( 11 ) and furnishes diastereoisomeric cyclopentenes 12A and 12B ; an X‐ray‐analysis of 12B confirms the ‘mixed’ 1 : 1 : 1 product. Competing is an (E)‐ BTE ‐catalyzed decomposition of 4 to give 2,3,4,4‐tetramethylpent‐1‐ene ( 7 )+N2; the reaction of (E)‐ BTE with a trace of water appears to be responsible for the chain initiation. The H2SO4‐catalyzed decomposition of diazoalkane 4 , indeed, produced the alkene 7 in high yield. The attack on the hindered diazoalkane 4 by 11 is slower than that by (E)‐ BTE ; the zwitterionic intermediate 21 undergoes cyclization and furnishes the tetrasubstituted furan 22 . In fumaronitrile, electrophilicity and steric demand are diminished, and a 1,3‐cycloaddition produces the 4,5‐dihydro‐1H‐pyrazole derivative 25 . The reaction of 4 with dimethyl acetylenedicarboxylate leads to pyrazole 29 +isobutene.  相似文献   

16.
Gambogic acid (GA, 1 ), the most prominent member of Garcinia natural products, has been reported to be a promising anti‐tumor agent. Previous studies have suggested that the planar B ring and the unique 4‐oxa‐tricyclo[4.3.1.03,7]dec‐2‐one caged motif were essential for anti‐tumor activity. To further explore the structure‐activity relationship (SAR) of caged Garcinia xanthones, two new series of B‐ring modified caged GA analogues 13a – 13e and 15a – 15e were synthesized utilizing a Claisen/Diel‐Alder cascade reaction. Subsequently, these compounds were evaluated for their in vitro anti‐tumor activities against A549, MCF‐7, SMMC‐7721 and BGC‐823 cancer cell lines by MTT assay. Among them, 13b – 13e exhibited micromolar inhibition against several cancer cell lines, being approximately 2–4 fold less potent in comparison to GA. SAR analysis revealed that the peripheral gem‐dimethyl groups are essential for maintaining anti‐tumor activity and substituent group on C1 position of B‐ring has a significant effect on potency, while modifications at C‐2, C‐3 and C‐4 positions are relatively tolerated. These findings will enhance our understanding of the SAR of Garcinia xanthones and lead to the development of simplified analogues as potential anti‐tumor agents.  相似文献   

17.
The neutral hexacoordinate silicon(IV) complex 6 (SiO2N4 skeleton) and the neutral pentacoordinate silicon(IV) complexes 7 – 11 (SiO2N2C skeletons) were synthesized from Si(NCO)4 and RSi(NCO)3 (R=Me, Ph), respectively. The compounds were structurally characterized by solid‐state NMR spectroscopy ( 6 – 11 ), solution NMR spectroscopy ( 6 and 10 ), and single‐crystal X‐ray diffraction ( 8 and 11 were studied as the solvates 8? CH3CN and 11? C5H12 ? 0.5 CH3CN, respectively). The silicon(IV) complexes 6 (octahedral Si‐coordination polyhedron) and 7 – 11 (trigonal‐bipyramidal Si‐coordination polyhedra) each contain two bidentate ligands derived from an α‐amino acid: (S)‐alanine, (S)‐phenylalanine, or (S)‐tert‐leucine. The deprotonated amino acids act as monoanionic ( 6 ) or as mono‐ and dianionic ligands ( 7 – 11 ). The experimental investigations were complemented by computational studies of the stereoisomers of 6 and 7 .  相似文献   

18.
A total of 21 natural products, 1 – 21 , were isolated from a supercritical CO2 extract of the rhizomes of Petasites hybridus. Thereby, seven new eremophilane (= (1S,4aR,7R,8aR)‐decahydro‐1,8a‐dimethyl‐7‐(1‐methylethyl)naphthalene) sesquiterpenes, compounds 4, 5, 9, 11, 12, 15 , and 17 , were identified. The new constituent 9‐hydroxyisobakkenolide ( 15 ) is the first representative of a group of compounds closely related to the well‐known, but rare, bakkenolides. Tsoongianolide B ( 18 ) and its degradation product ligularenolide ( 19 ) were found as new Petasites constituents as well. The known eremophilanolide 2 was isolated from a plant source for the first time and the oxofuranopetasin 16 was isolated for the first time from the rhizomes of P. hybridus, together with eight other known compounds. The C(8)‐epimeric 2‐[(tigloyl)oxy]eremophilanolides 3 and 8 could clearly be differentiated. All structures were established by extensive 1D‐ and 2D‐NMR experiments (Tables 1–3), and confirmed by in‐depth GC/MS and HPLC/MS experiments.  相似文献   

19.
Four C(8),C(12)‐diastereoisomers, (8S,12S)‐isoandrographolide ( 1 ), (8S,12R)‐isoandrographolide ( 2 ), (8R,12R)‐isoandrographolide ( 3 ), and (8R,12S)‐isoandrographolide ( 4 ) were isolated from the aerial parts of Andrographis paniculata. The structures of the new compounds 1 – 3 were established on the basis of the spectroscopic data including UV, IR, NMR, HR‐ESI‐MS, and X‐ray diffraction analysis.  相似文献   

20.
A series of side chain reactions starting from the 6‐ and 7‐styryl‐substituted 1,3‐dimethyllumazines 1 and 21 as well as from the 6‐ and 7‐[2‐(methoxycarbonyl)ethenyl]‐substituted 1,3‐dimethyllumazine 2 and 22 were performed first by addition of Br2 to the C?C bond forming the 1′,2′‐dibromo derivatives 3, 4, 24 , and 26 in high yields (Schemes 1 and 3) (lumazine=pteridine‐2,4(1H,3H)‐dione). Treatment of 3 with various nucleophiles gave rise to an unexpected tele‐substitution in 7‐position and elimination of the Br‐atoms generating 7‐alkoxy‐ (see 5 and 6 ), 7‐hydroxy‐ (see 7 ) and 7‐amino‐6‐styryl‐1,3‐dimethyllumazines (see 8 – 11 ) (Scheme 1). On the other hand, 4 underwent, with dilute DBU (1,8‐diazabicyclo[5.4.0]undec‐2‐ene), a normal HBr elimination in the side chain leading to 18 , whereas treatment with MeONa afforded a more severe structural change to 19 . Similarly, 24 and 26 reacted to 27, 32 , and 33 under mild conditions, whereas in boiling NaOMe/MeOH, 24 gave 7‐(2‐dimethoxy‐2‐phenylethyl)‐1,3‐dimethyllumazine ( 30 ) which was hydrolyzed to give 31 (Scheme 3). From the reactions of 4 and 24 with DBU resulted the dark violet substance 20 and 25 , respectively, in which DBU was added to the side chain (Scheme 2). The styryl derivatives 1 and 21 could be converted, by a Sharpless dihydroxylation reaction, into the corresponding stereoisomeric 6‐ and 7‐(1,2‐dihydroxy‐2‐phenylethyl)‐1,3‐dimethyllumazines 34 – 37 (Scheme 4). The dihydroxy compounds 34 and 35 were also acetylated to 38 and 39 which, on catalytic reduction followed by formylation, yielded the diastereoisomer mixtures 40 and 41 . Deacetylation to 42 and 45 allowed the chromatographic separation of the diastereoisomers resulting in the isolation of 43 and 44 as well as 46 and 47 , respectively. Introduction of a 6‐ or 7‐ethynyl side chains proceeded well by a Sonogashira reaction with 6‐ ( 48 ) or 7‐chloro‐1,3‐dimethyllumazine ( 55 ) yielding 49 – 51 and 56 – 58 (Scheme 5). The direction of H2O addition to the triple bond is depending on the substituents since the 6‐ ( 49 ) and 7‐(phenylethynyl)‐1,3‐dimethyllumazine ( 56 ) showed attack at the 2′‐position yielding 53 and 60 , in contrast to the 6‐ ( 51 ) and 7‐ethynyl‐1,3‐dimethyllumazine ( 58 ) favoring attack at C(1′) and formation of 6‐ ( 52 ) and 7‐acetyl‐1,3‐dimethyllumazine ( 59 ).  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