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.
The 2′-deoxyribofuranose analog of the naturally occurring antibiotics SF-2140 and neosidomycin were prepared by the direct glycosylation of the sodium salts of the appropriate indole derivatives, with 1-chloro-2- deoxy-3,5-di-O-p-toluoyl-α-D-erythropentofuranose ( 5 ). Thus, treatment of the sodium salt of 4-methoxy-1H- indol-3-ylacetonitrile ( 4a ) with 5 provided the blocked nucleoside, 4-methoxy-1-(2-deoxy-3,5-di-O-p-toluoyl-β- D-erythropentofuranosyl)-1H-indol-3-ylacetonitrile ( 6a ), which was treated with sodium methoxide to yield the SF-2140 analog, 4-methoxy-1-(2-deoxy-β-D-erythropentofuranosyl)-1H-indol-3- ylacetonitrile ( 7a ). The neosidomycin analog ( 8 ) was prepared by treatment of the sodium salt of 1H-indol-3-ylacetonitrile ( 4b ) with 5 to obtain the blocked intermediate 1-(2-deoxy-3,5-di-O-p-toluoyl-β-D-erythropentofuranosyl) ?1H-indol-3-ylace-tonitrile ( 6b ) followed by sodium methoxide treatment to give 1-(2-deoxy-β-D-erythropentofuranosyl)-1H- indol-3-ylacetonitrile ( 7b ) and finally conversion of the nitrile function of 7b to provide 1-(2-deoxy-β-D- erythropentofuranosyl)-1H-indol-3-ylacetamide ( 8 ). In a similar manner, indole ( 9a ) and several other substituted indoles including 1H-indole-4-carbonitrile ( 9b ), 4-nitro-1H-indole ( 9c ), 4-chloro-1H-indole-2-carboxamide ( 9d ) and 4-chloro-1H-indole-2-carbonitrile ( 9e ) were each glycosylated and deprotected to provide 1-(2-deoxy-β-D-erythropentofuranosyl)-1H-indole ( 11a ), 1-(2-deoxy-β-D-erythropentofuranosyl)-1H-indole-4- carbonitrile ( 11b ), 4-nitro-1-(2-deoxy-β-D-erythropentofuranosyl)-1H-indole ( 11c ), 4-chloro-1-(2-deoxy-β-D- erythropentofuranosyl)-1H-indole-2-carboxamide ( 11d ) and 4-chloro-1-(2-deoxy-β-D-erythropentofuranosyl)- 1H-indole-2-carbonitrile ( 11e ), respectively. The 2′-deoxyadenosine analog in the indole ring system was prepared for the first time by reduction of the nitro group of 11c using palladium on carbon thus providing 4-amino-1-(2-deoxy-β-D-erythropentofuranosyl)- 1H-indole ( 16 , 1,3,7-trideaza-2′-deoxyadenosine).  相似文献   

2.
The base-catalyzed cyclization of some N-propargylamides to oxazoles has been studied in the presence of sodium hydride and potassium carbonate. The α-arylsubstituted propargylamides 1c-d (Ar = p-OMeC6H4 ( 1c ), C6H5 ( 1d ), and p-O2NC6H4 ( 1e )) cyclized with markedly higher rates ( 1e > 1d > 1c ) than the unsubstituted and α-methyl substituted propargylamides 1a and 1b . A 1H nmr spectroscopic experiment demonstrated the presence of an allenic intermediate in the potassium carbonate-catalyzed ring closure of 1e . The observed rank order of reactivities correlates well with the acidities of the respective propargylic hydrogens of the amides and with the ability of the ring closed intermediates to stabilize an oxazole anion. The results demonstrate that the base-catalyzed formation of oxazoles from propargylamides may proceed via an allenic intermediate.  相似文献   

3.
The cycloaddition reaction of cyclic imidates, 2‐benzyl‐5,6‐dihydro‐4H‐1,3‐oxazines 1a , 1b , 1c , 1d , 1e , 1f , with dimethyl acetylenedicarboxylate 2 , trimethyl ethylenetricarboxylate 4 , or dimethyl 2‐(methoxymethylene)malonate 6 afforded new fused heterocyclic compounds, such as methyl (6‐oxo‐3,4‐dihydro‐2H‐pyrrolo[2,1‐b]‐1,3‐oxazin‐7‐ylidene)acetates 3a , 3b , 3c , 3d , 3e , 3f (71–79%), dimethyl 2‐(6‐oxo‐3,4,6,7‐tetrahydro‐2H‐pyrrolo[2,1‐b]‐1,3‐oxazin‐7‐yl)malonates 5b , 5c , 5d , 5e , 5f (43–71%), or methyl 6‐oxo‐3,4‐dihydro‐2H,6H‐pyrido[2,1‐b]‐1,3‐oxazine‐7‐carboxylates 7a , 7b , 7c , 7d , 7e , 7f (32–59%), respectively. In these reactions, 1a , 1b , 1c , 1d , 1e , 1f (cyclic imidates, iminoethers) functioned as their N,C‐tautomers (enaminoethers) 2 to α,β‐unsaturated esters 2 , 4, and 6 to give annulation products 3 , 5 , and 7 following to the elimination of methanol, respectively. J. Heterocyclic Chem., (2011).  相似文献   

4.
The unexplored carbon rich cationic closo carboranes, C3Bn?3Hn+1 (n=5, 6, 7, 10, 12) are investigated theoretically. The position isomers were calculated at the B3LYP/6‐31G* level, and the charge distribution in the cluster is estimated by NBO analysis. The criterion of ring‐cap orbital overlap compatibility along with the number of B? C, C? C, and B? B bonds help in explaining the stability order in each category. The most stable isomer is the one with maximum ring‐cap orbital overlap and largest number of B? C bonds. The order of relative stability among the trigonal bipyramid is 1c > 1b > 1a ′, where the stability is proportional to the number of CH caps over the small three‐membered ring. The C3B3H6+ isomer with the one allyl C3 group ( 2b ) is more favorable than the one with a cyclopropenyl group ( 2a ). Among the C3B4H7+ isomers the stability order is 3e > 3d > 3c > 3b > 3a , which mostly depends on the ring‐cap orbital overlap. In the bicapped square antiprism (4) where there is large number of isomers, the order follows the rule of ring cap compatibility and the number of B? C bonds. The order of 5e > 5d > 5c > 5b > 5a obtained from the calculations is in perfect agreement with the above sited rules. Equations (1) – (5) devised for estimating the stability of isomers of C3Bn?3Hn+ indicate an increase in stability with cage size. The mono‐positive charge of the isomers is distributed throughout the cage, making them suitable candidates as weakly electrophillic cations. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 1542–1551, 2001  相似文献   

5.
The 5,6,7,8,9,10‐hexahydro‐2‐methylthiopyrimido[4,5‐b]quinolines 4a , 4b , 4c , 4d , 5a , 5b , 5c , 5d and their oxidized forms 6a , 6b , 6c , 6d , 7a , 7b , 7c , 7d were obtained from the reaction of 6‐amino‐2‐(methylthio)pyrimidin‐4(3H)‐one 2 or 6‐amino‐3‐methyl‐2‐(methylthio)pyrimidin‐4(3H)‐one 3 and α,β‐unsaturated ketones 1a , 1b , 1c , 1d using BF3.OEt2 as catalyst and p‐chloranil as oxidizing agent. Some of the new compounds were evaluated in the US National Cancer Institute (NCI), where compound 5a presented remarkable activity against 46 cancer cell lines, with the most important GI50 values ranging from 0.72 to 18.4 μM from in vitro assays.  相似文献   

6.
The monomeric octa-aza bis-α-diimine macrocyclic complex [CoII(C10H20N8)(H2O)](ClO4)2 I, undergoes various reactions on the macrocyclic ligand. Reaction of complex I with triethylamine in double molar proportions, followed by slow aerial oxidation, produces a molecular dimeric complex [CoII(C10H14N8)]2, III, and a novel Co(I) complex [CoI(C10H19N8)], IV. Complex III is a staggered cofacial dimer with a cobalt-cobalt bond length 2.86(1) Å. The macrocyclic ligand of the complex contains an a-diimine function in each five-membered chelate ring, and a three-atom N-C-N? delocalized system in each six-membered chelate ring. Complex IV has the 5-5-6-6 chelate arrangement because one α-diimine moiety is rearranged to a syn-anti configuration. In the structure, the two fused six-membered chelate rings are fully conjugated and the two fused five-membered rings are saturated. However, when complex I reacts with excess triethylamine under the similar conditions, a dimeric complex of another type, [CoII(C10Hl6N8)]2, II, was generated, in which one N-N bond of the macrocyclic ligand is broken. Complex IV can be isolated also from the reaction of complex I with excess hydrazine, followed by slow aerial oxidation. When hydrazine in double molar proportions was used, complex [CoI(C10H17N8)(NHNH)] V, which contains a coordinated diazene ligand, was obtained. Only one six-membered chelate ring of complex V is deprotonated and oxidized to form a three-atom N-C-N? delocalized system. The structures of octa-aza complexes I-V are determined by X-ray crystallography: I, orthorhombic, C mca, a = 11.646(4), b = 17.049(3), c = 10.706(3) Å, Z = 4, R = 0.045, Rw = 0.047, based on 1024 reflections with I > 2σ(I); II, monoclinic, P 21/c, a = 9.814(3), b = 22.583(6). c = 14.632(9) Å, β = 98.90(5)°, Z = 4, R = 0.085, Rw = 0.101, based on 2033 reflections with I > 2σ(I); III, tetragonal, P 4/nmm, a = 15.614(3), c = 6.498(2) Å, Z = 4, R = 0.081, Rw = 0.115, based on 340 reflections with I > 2σ(I); IV, orthorhombic, P bca, a = 8.484(1), b = 16.662(3), c = 18.760(2) Å, Z = 8, R = 0.029, Rw = 0.024, based on 1441 reflections with I > 2σ(I); V, monoclinic, P 21/m, a = 7.892(3), b = 11.713(6), c = 9.326(4) Å, β = 108.03(3), Z = 2, R = 0.047, Rw = 0.056, based on 948 reflections with I > 2σ(I).  相似文献   

7.
The kinetics and product studies of oxidation of eight olefins 1 - 8 by ClO2 in H2O in the pH range 3-7 are described. The reaction is faster as the pH decreases. At pH < 4, ClO2 reacts equimolarly with olefins to yield isomeric mixtures of chlorohydrines and 1,2-dioxygenated products, following the equation: The order of reactivity is: (E)-stilbene > indene > β-methylstyrene > acenaphthylene > α-methylstyrene > styrene > cyclohexene > allylbenzene. A multi-stage radical-cation mechanism is proposed, in which an initial reversible protonation: is followed by an electron-transfer stage (rate-determining): The cation-radical thus produced, adds rapidly an additional ClO2 to form dioxygenated products. The chlorohydrines most likely arise from HClO additions to the olefinic double bonds, which, in turn, generate from dismutation of 2 HClO2 into HClO + H+ + ClO.  相似文献   

8.
Several 3-alkoxysubstituted pyrazolo[3,4-d]pyrimidine ribonucleosides structurally related to adenosine, inosine and guanosine have been prepared by the direct glycosylation of preformed aglycon precursor containing a 3-alkoxy substituent. Ring closure of 5(3)-amino-3(5)-ethoxypyrazole-4-carboxamide ( 6b ) with either formamide or potassium ethyl xanthate gave 3-ethoxyallopurinol ( 7b ) and 3-ethoxy-6-thioxopyrazolo[3,4-d]-pyrimidin-4(5H,7H)-one ( 10 ), respectively. Methylation of 10 gave the corresponding 6-methylthio derivative 15 . Similar ring annulation of 5(3)-methoxypyrazole-4-carboxamide ( 6a ) with formamide afforded 3-methoxyallopurinol ( 7a ). Treatment of 5(3)-amino-3(5)-methoxypyrazole-4-carbonitrile ( 5a ) with formamidine acetate furnished 4-amino-3-methoxypyrazolo[3,4-d]pyrimidine ( 4 ). High-temperature glycosylation of 7b with 1-O-acetyl-2,3,5-tri-O-benzoyl-D-ribofuranose in the presence of boron trifluoride etherate gave a 2:1 mixture of N-1 and N-2 glycosyl blocked nucleosides 11b and 13b . Deprotection of 11b and 13b with sodium methoxide gave 3-ethoxy-1-β-D-ribofuranosylpyrazolo[3,4-d]pyrimidin-4(5H)-one ( 12b ) and the corresponding N-2 glycosyl isomer 14b , respectively. Similar glycosylation of either 4 or 7a , and subsequent debenzoylation gave exclusively 4-amino-3-methoxy-1-β-D-ribofuranosylpyrazolo[3,4-d]pyrimidine ( 9 ) and 3-methoxy-1-β-D-ribofuranosylpyrazolo[3,4-d]pyrimidin-4-(5H)-one ( 12a ), respectively. The structural assignment of 12a was made on the basis of single-crystal X-ray analysis. Application of this general glycosylation procedure to 15 gave the corresponding N-1 glycosyl derivative 16 as the sole product, which on debenzoylation afforded 3-ethoxy-6-(methylthio)-1-(3-D-ribofuranosylpyrazolo[3,4-d]pyrimidin-4(5H)-one ( 17 ). Oxidation of 16 and subsequent ammonolysis furnished the guanosine analog 6-arnino-3-ethoxy-1-β-D-ribofuranosylpyrazolo[3,4-d]-pyrimidin-4(5H)-one ( 19 ). Similarly, starting from 3-methoxy-4,6-bis(methylthio)pyrazolo[3,4-d]pyrimidine ( 20 ), 6-amino-3-methoxy-1-β-D-ribofuranosylpyrazolo[3,4-d]pyrimidin-4(5H)-one ( 23 ) was prepared.  相似文献   

9.
2′-Deoxyribofuranosyl and arabinofuranosyl nucleosides of certain purine-6-sulfenamides, sulfinamides and sulfonamides have been prepared by sequential amination and controlled oxidation of the corresponding 6-thiopurine nucleosides, and evaluated for antiviral and antitumor activities in mice. Amination of 2′-deoxy-6-thioinosine ( 4a ) and 9-β-D-arabinofuranosyl-6-thiopurine ( 4c ) with chloramine solution gave the corresponding 6-sulfenamides 5a and 5c , respectively, which on selective oxidation with 3-chloroperoxybenzoic acid (MCPBA) gave diastereomeric 9-(2-deoxy-β-D-erythro-pentofuranosyl)purine-6-sulfinamide ( 6a ) and 9-β-D-arabinofuranosylpurine-6-sulfinamide ( 6c ), respectively. However, oxidation of 5a and 5c with excess of MCPBA gave the corresponding 6-sulfonamide derivatives 7a and 7c , respectively. Similar amination of 2′-deoxy-6-thioguanosine ( 4b ), ara-6-thioguanine ( 4d ) and α-2′-deoxy-6-thioguanosine ( 8 ) gave the respective 6-sulfenamide derivatives 5b, 5d and 9 . Controlled oxidation of 5b, 5d and 9 gave (R,S)-2-amino-9-(2-deoxy-β-D-erythro-pentofuranosyl)purine-6-sulfinamide ( 6b ), (R,S)-2-amino-9-β-D-arabinofuranosylpurine-6-sulfinamide ( 6d ) and the α-anomer of ( 6b) (10 ), respectively. The diastereomeric mixture of (R,S )-10 was partially resolved and the structure of S -10 was assigned by single-crystal X-ray diffraction analysis. Oxidation of 5b, 5d and 9 with excess of MCPBA afforded the respective 6-sulfonamide derivatives 7b, 7d and 11 . Nucleosides 5c and 7c were significantly active against Friend leukemia virus in mice, whereas 6c was somewhat less active. Of the 20 nucleosides evaluated, 12 exhibited biologically significant anti-L1210 activity in mice. Nucleosides 6b and 7a at 173 mg/kg/day × 1 showed a T/C of 153, whereas 7d at 800 mg/kg/day × 1 showed a T/C of 153 against L1210 leukemia. The α-nucleoside 9 at 480 mg/kg/day × 1 gave a T/C of 172. A single treatment with 6b, 7a, 7d and 9 reduced the body burdens of viable L1210 cells by more than 99.2%. The antileukemic activity of these novel nucleosides tended to parallel solubility.  相似文献   

10.
Butanolic extracts of the Mediterranean sponge Aplysina (= Verongia) cavernicola have given, by reverse-phase HPLC, the antibacterial quinols (±)-3-bromoverongiaquinol (= (±)-3-bromo-1-hydroxy-4-oxo-2,5-cyclohexadine-1-acetamide; 1d) and (±)-3-bromo-5-chloroverongiaquinol (= (±)-3-bromo-5-chloro-1-hydroxy-4-oxo-2,5-cyclohexadine-1-acetamide; 1c ) besides the products of their formal cyclization 5-chlorohexadiene-1-acetamide; 1c ) besides the products of their formal cyclization 5-chlorocavernicolin (= 5-cloro-3,3a,7,7aβ-tetrahydro-3aβ-hydroxy-2,6(1H)-indoledione; 6) , the C(7)-epimerizing 7β-bromo-5-chlorocavernicolin (=7 β-bromo-5-chloro-3,3a,7,7aβ-tetrahydro-3aβ-hydroxy-2,6(1H)-indoledione; 4a and 7α-bromo-5-chlorocavernicolin (4b) , and the C(7)-epimerizing 5-bromo-7β-chlorocavernicolin ( = 5-bromo-7β-chloro-3,3a,7,7aβ-tetrahydro-3aβ-hydroxy-2,6(1H)-indoledione; 5a) and 5-bromo-7α-chlorocavernicolin (5b) . The latter four were isolated as mixtures of C(7)-epimerizing monoacetates 4a′/4b′ and 5a′/5b′. Both 1 and 1c proved to be racemic from NMR examination of their esterification products with (–)-methyl-oxyacetic acid, whilst 6 had a ca. 6% enantiomeric purity as shown by a 1H-NMR study of its monoacetate 6′ in the presence of a chiral shift reagent. These chiroptical data of the first chiral quinols from the Verongida and of 6 suggest phenol oxidative routes from tyrosine precursors for their formation. In view of their bioactivities, 1d and 1c have been synthesized from (p-hydroxyphenyl)acetic acid byt phenol oxidative routes.  相似文献   

11.
Two α,β-unsaturated thiolactones, 5-(2-propynyl)-2(5H)-thiophenone (5) and 3,5-di(2-propenyl)?2(5H)-thiophenone (6) , were newly synthesized. Irradiation (λ = 300 nm) of 5 in MeOH containing cyclopentene afforded a 3:1 mixture of diasteroisomeric methyl 7-thiatricyclo[6.4.0.02,6]dodec-10-ene-12-carboxylates (8a/8b) , while irradiation of 6 in MeOH saturated with 2-methylpropene gives a 3:2 mixture of diastereoisomeric methyl 3,3,9-trimethyl-5-thiatricyclo[6.2.1.02,6]undecane-1-carboxylates ( 10a / 10b ).  相似文献   

12.
The reactions of 2, 2′-dithiodianiline 1 with 2-alkyl-1, 3-diketones 2a-d have been employed in order to synthesize 2-acyl-2H-1, 4-benzothiazines. In the cases of 2a, b the expected 2-acyl-2H-1, 4-benzothiazines, i.e. 3a, b were obtained, whereas the reactions of 1 with the 1, 3-diketones 2c, d afforded the α-ketosulfide 12 and the 1, 4-benzothiazine 17 , respectively. The products 3a, b underwent the hydrolytic C2-C3 bond cleavage of the thiazine nucleus to give the α-ketosulfides 6 and 11 , respectively. Such an hydrolytic process explains the formation of the compound 12 in the reaction of 1 with 2c . The formation of 17 in the case of 2d is considered to be formed through a rearrangement involving the 1, 3-sulfur shift of the preformed 1, 4-benzothiazine 3d .  相似文献   

13.
Three new Zn‐phosphonates based on 5‐phosphonoisophthalic acid, (HO2C)2C6H3PO3H2 (H4L), [Zn2(H2O)(O2C)2C6H3PO3] · H2O ( 1 ), Zn2(H2O)2(O2C)2C6H3PO3 ( 2 ), and KZn[H(O2C)2C6H3PO3] ( 3 ) have been hydrothermally synthesized and characterized by single‐crystal X‐ray diffraction ( 1 : triclinic, , a = 742.49(3) pm, b = 846.37(4) pm, c = 992.84(4) pm, α = 80.936(2)°, β = 81.574(2)°, γ = 73.139(3)°, V = 586.28(4) · 106 pm3, R1 = 0.0583, wR2 = 0.1347 (for I > 2σ(I)); 2 : monoclinic, P21/m, a = 464.78(9) pm, b = 1329.2(3) pm, c = 974.5(3) pm, β = 95.80(3)°, V = 599.0(2) · 106 pm3, R1 = 0.0395, wR2 = 0.1086 (for I > 2σ(I)); 3 : monoclinic, P21/c, a = 501.9(1) pm, b = 2489.9(5) pm, c = 946.2(5) pm, β = 105.38(3)°, V = 1140.0(7) · 106 pm3, R1 = 0.0365, wR2 = 0.0848 (for I > 2σ(I))). The title compounds 1 and 2 have the same chemical composition but exhibit different structures and are therefore polymorphs. Thus, in compound 1 , isolated ZnO4‐tetrahedra, and in 2 , infinite double‐chains of corner‐sharing ZnO6 polyhedra are observed. In, KZn[H(O2C)2C6H3PO3] ( 3 ) K+‐ions have been incorporated into the structure leading to the formation of a bimetallic inorganic‐organic hybrid compound.  相似文献   

14.
Using the readily accessible chiral auxiliaries 1 – 3 the sulfonamide-shielded O-silylated esters 5 underwent π-face-selective α-acetoxylation on successive treatment with Pb(OAc)4 and NEt3 HF to give after recrystallization α-acetoxy ester 6 in 55–67% yields and in 95–100% d.e. Starting from conjugated enoates addition of RCu and subsequent acetoxylation 10 → 11 → 12 yielded α,β-bifunctionalized esters 12 with >95% configurational control at both Cα and Cβ. Nondestructive removal of the auxiliary ( 6 → 7 , 6 → 8 and 12 → 13 ) gave either α-hydroxycarboxylic acids or terminal α-glycols in high enantiomeric purity. The prepared glycols 8c and 13a are key intermediates for previously reported syntheses of the natural products 16 and 17 , respectively.  相似文献   

15.
1-Unsubstituted dihydro-6-methyl-2(1H)-pyrimidinethiones undergo an aminolysis in dialkylformamides or methylformamide resp., at higher temperature, and then are rearranged to 4-dialkylaminodihydro-2(1H)-pyridinethiones6 a, b or the 4-methylamino compound6 c. 1-Alkyl- and 1-aryldihydro-6-methyl-2(1H)-pyrimidinethiones1 b, c and tautom, methylenecompounds2b,c resp., react at the same conditions withDMF not only to 4-alkylamino-or 4-arylaminodihydro-2(1H)-pyridinethiones6 c, d but also to 4-dimethylaminodihydro-2(1H)-pyridinethiones6 a. From 3-substituted dihydro-6-methyl-2(1H)-pyrimidinethiones only the 3-aryl compound12 b is converted byDMF to the corresponding dihydro-4-dimethylamino-1-phenyl-2(1H)-pyridinethione14. Also 4-alkylamino- and 4-arylaminodihydro-2(1H)-pyridones18a, b are formed by heating inDMF at 230°C from 1-alkyl- and 1-aryldihydro-2(1H)-pyrimidinones15b, c and of methylene compounds16b, c resp. 2-Methylimino- and 2-phenyliminodihydro-1,3-thiazinethiones19a, b react inDMF viaDimrothrearrangement to the corresponding 1-alkyl- and 1-phenyl-dihydro-2(1H)-pyrimidinethiones1 b, 2 b, 1 c, 2 c and further to 4-alkylamino-and 4-arylaminodihydro-2(1H)-pyridinethiones6 c, d.  相似文献   

16.
Methyl- and phenylhydrazines react with 2-(diphenylacetyl)-1,3-indandione ( 1 ) to yield respectively the 1-(methylhydrazone) and the 1-(phenylhydrazone) of 2-(diphenylacetyl)-1,3-indandione ( 2a and 2b ). In comparison, acetic and benzoic acid hydrazides react with 1 to give respectively the α-(acetylhydrazone) and the α-(benzoylhydrazone) of 2-(diphenylacetyl)-1,3-indandione ( 3a and 3b ). Cyclization of 2a and 2b gives 2,3-disubstituted indeno[1,2-c]pyrazol-4(2H)-ones ( 7a and 7b ). Cyclization of 3a and 3b , followed by methylation, gives 1-methyl- and 2-methyl-3-(diphenylmethyl)indeno[1,2-c]pyrazol-4(1 and 2H)-ones ( 9a and 9b ). 2-Isovaleryl-1,3-indandione reacts with phenylhydrazine to give directly 3-isobutyl-1-phenylindeno[1,2-c]-pyrazol-4(1H)-one ( 10 ).  相似文献   

17.
Cycloadditions of the α,β-unsaturated-acyl cyanides 1–3 with (Z)-or (E)-1-bromo-2-ethoxyethene ( 4 ) may be performed at moderate temperatures and provide in good yields the 3-bromo-2-ethoxy-3,4-dihydro-2H-pyran-6-carbonitriles 5–7 , respectively (Scheme 1). Diastereoisomeric pairs of products result at room temperature merely from the ‘endo’- and ‘exo’-transition states; more complex mixtures appear above 60° as a consequence of (Z)/(E)-isomerization of 4 . The relative stability of the anomers of 5 and 6 is explored by treatment with BF3·Et2O. Acid alcoholysis (MeOH or EtOH) of 5 leads to acetals 9a , b of 4-bromo-5-oxopentanoate. Alkyl (2Z,4E)-5-ethoxypenta-2,4-dienoates 12 , 17 , and 20 , are formed in alcoholic alkoxide solutions from 5 , 6 , and 7 , respectively, which is compatible with the intermediacy of 2-alkoxy-2H-pyrans and their valence tautomers, α,β-unsaturatedacyl cyanides. Methoxide addition to the CN group competes with dehydrobromination in case of 5 ; it leads to 3-bromo-3,4-dihydro-2H-pyran-6-carboximidate 13 (ca. 50% at ?20°) which can be hydrolyzed to the methyl carboxylate 14 . DBU (1,8-diazabicyclo[5,4,0]undec-7-ene) in benzene converts 5 to 6-ethoxy-2-oxohexa-3,5-dienenitrile ( 11 ), the ring-opening product of an obviously unstable 2-ethoxy-2H-pyran; the same reagent dehydrobrominates 6 to 2-ethoxy-4-methyl-2H-pyran-6-carbonitrile ( 15 ). HBr Elimination from 7 takes place with great ease in presence of pyridine, or even during chromatography on alumina, and leads to the stable ethyl 6-cyano-2-ethoxy-2H-pyran-4-carboxylate ( 18 ); this dimerizes at room temperature to give a 1:3 mixture of tricyclic adducts ‘endo’- 21 and ‘exo’- 21 . The structure of the latter is established by an X-ray crystallographic analysis.  相似文献   

18.
Three new pregnane glycosides, cynatroside A ( 1 ), cynatroside B ( 2 ), and cynatroside C ( 3 ), isolated from the roots of Cynanchum atratum (Asclepiadaceae), were characterized as 7β‐{[Oα‐L ‐cymaropyranosyl‐(1→4)‐Oβ‐D ‐digitoxopyranosyl‐(1→4)‐β‐D ‐oleandropyranosyl]oxy}‐3,4,4a,4b,5,6,7,8,10,10a‐decahydro‐6α‐hydroxy‐4b‐ methyl‐2‐(2‐methyl‐3‐furyl)phenanthren‐1(2H)‐one ( 1 ), 7β‐{[Oβ‐D ‐cymaropyranosyl‐(1→4)‐Oα‐L ‐diginopyranosyl‐(1→4)‐β‐D ‐cymaropyranosyl]oxy}‐3,4,4a,4b,5,6,7,8,10,10a‐decahydro‐2,6α‐dihydroxy‐4b‐methyl‐2‐(2‐methyl‐3‐furyl)phenanthren‐1(2H)‐one ( 2 ), and 7β‐{[Oα‐L ‐cymaropyranosyl‐(1→4)‐Oβ‐D ‐digitoxopyranosyl‐(1→4)‐β‐L ‐cymaropyranosyl]oxy}‐3,4,4a,4b,5,6,7,8,10,10a‐decahydro‐2,6α‐dihydroxy‐4b‐methyl‐2‐(2‐methyl‐3‐furyl)phenanthren‐1(2H)‐one ( 3 ), respectively. In addition, ten known constituents were identified, i.e., cynascyroside D ( 4 ), glaucoside C ( 5 ), glaucoside D ( 6 ), atratoside A ( 7 ), 2,4‐dihydroxyacetophenone ( 8 ), 4‐hydroxyacetophenone ( 9 ), syringic acid ( 10 ), azelaic acid ( 11 ), suberic acid ( 12 ), and succinic acid ( 13 ). Among these compounds, 1 – 4 significantly inhibit acetylcholinesterase activity.  相似文献   

19.
The reaction of ethyl 1H-benzimidazole-2-acetate (1) with methyl or ethyl isocyantes 2a,b resulted in excellent yields of the respective 2-methyl- or 2-ethylpyrimido[1,6-a]benzimidazole-1,3(2H,5H)-diones 3a,b , while the reaction of 1 with phenyl isocyanate (2c) gave, unexpectedly, ethyl 2-(1-phenylcarbamoyl-1H,3H-benzimidazol-2-ylidene)-2-phenylcarbamoylacetate (4). Alkylation of 3 with trimethyl or triethyl phosphates 5a,b led to the 5-methyl or 5-ethyl derivatives 6a-d . Chlorination of 6 with sulfuryl chloride afforded the 4-chloro derivatives 7a-d.  相似文献   

20.
The 6‐aza‐nido‐decaboranes RNB9H11 ( 1a—d ; R = H, Ph, 4‐C6H4Me, 4‐C6H4Cl) act as 1, 2‐hydroboration agents via their 9‐BH vertex, giving products RNB9H10R′. The boranes 1a, b and 3‐hexyne yield the 9‐(1‐ethyl‐1‐butenyl)‐6‐aza‐nido‐decaboranes 2a, b (R′ = CEt = CHEt). 2, 3‐Dimethyl‐2‐butene is hydroborated by 1a—d under formation of the 9‐(1, 1, 2‐trimethylpropyl)‐6‐aza‐nido‐decaboranes 3a—d (R′ = —CMe2 —CHMe2). With the boranes 1a—c and (trimethylsilyl)ethene, a 85:15 mixture of the products (RNB9H10)CH2CH2(SiMe3)( 4a—c ) and their chiral isomers (RNB9H10)CH(SiMe3)CH3 ( 5a—c ) is obtained. The action of BH3(SMe2) on the mixtures 4b/5b or 4c/5c results in a closure of the nido‐NB9 skeleton of 4b or 4c , respectively, with a closo‐NB11 skeleton of the products RNB11H10R′ ( 6b or 6c;R′ = CH2CH2(SiMe3)); R′ is found in position 7 of 6b, c . All products of the type 2—6 are characterised by NMR.  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