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 359 毫秒
1.
The BF3-catalyzed cyclization of 3-acetyl-1-aryl-2-pentene-1,4-diones 1a-e in the presence of water in boiling tetrahydrofuran gave bis(3-acetyl-5-aryl-2-furyl)methanes 2a-e in 26-79% yields along with a small amount of 3-acetyl-5-aryl-2-methylfurans 3a-e. The exact structure of 2a was determined by X-ray crystallography. The use of a half volume of the solvent for the reaction of 1a resulted in the formation of 2,4-bis(3-acetyl-5-phenyl-2-furfuryl)-3-acetyl-5-phenylfuran (4) together with 2a and 3a. A similar reaction of 1a was carried out in the presence of 3-acetyl-5-(4-methylphenyl)-2-methylfuran (3d) to afford 4-(3-acetyl-5-phenyl-2-furfuryl)-3-acetyl-5-(4-methylphenyl)-2-methylfuran (5) in 49% yield. The BF3-catalyzed reaction of 1a with 2,4-pentanedione in dry tetrahydrofuran at 23°C gave 3-(3-acetyl-5-phenyl-2-furfuryl)-4-hydroxy-3-penten-2-one (6a) and 3-(3-acetyl-2-methyl-4-phenyl-5-furyl)-4-hydroxy-3-penten-2-one (7a) in 66 and 24% yields, respectively. The product distribution depended on the reaction temperature. A similar reaction of 1b-e also yielded the corresponding trisubstituted furans 6b-e and tetrasubstituted furans 7b-e in good yields. These results suggested the presence of the furfuryl carbocation intermediate A during the reaction. The one-pot synthesis of 6a and 7a was also achieved by a similar reaction using phenylglyoxal. The deoxygenation of 1a with triphenylphosphine gave 3a in 88% yield, while 1a was treated with concentrated hydrochloric acid to yield 3-acetyl-2-chloromethyl-5-phenylfuran (8) which was quantitatively transformed in ethanol into 3-acetyl-2-ethoxymethyl-5-phenylfuran (9) and in water into 3-acetyl-5-phenylfurfuryl alcohol (10), respectively. In addition, the Diels-Alder reaction of cyclopantadiene with 1a gave the corresponding [4+2] cycloaddition products 11 and 12.  相似文献   

2.
2-Aminobenzylamine was reacted with corresponding aromatic or heteroaromatic aldehydes to give 1,2,3,4-tetrahydroquinazolines 1 the oxidation of which with H2O2-tungstate in methanol led to the formation of the corresponding 1,2,3,4-tetrahydroquinazolin-1-ols 2. A one-pot procedure involving the treatment of the in situ formed quinazoline 1 with H2O2-tungstate again led to the formation of 2. Compounds 2 react with 2 equiv of aryl isocyanate in toluene at room temperature to produce compounds 3. The probable mechanism of the ring-expanding carbamoylation of quinazolin-1-ols 2 to 6-oxa-5,8-diaza-benzocycloheptenes 3 is discussed.  相似文献   

3.
(Z)-5-(2-(1H-Indol-3-yl)-2-oxoethylidene)-3-phenyl-2-thioxothiazolidin-4-one (7a-q) derivatives have been synthesized by the condensation reaction of 3-phenyl-2-thioxothiazolidin-4-ones (3a-h) with suitably substituted 2-(1H-indol-3-yl)-2-oxoacetaldehyde (6a-d) under microwave condition. The thioxothiazolidine-4-ones were prepared from the corresponding aromatic amines (1a-e) and di-(carboxymethyl)-trithiocarbonyl (2). The aldehydes (6a-h) were synthesized from the corresponding acid chlorides (5a-d) using HSnBu3.  相似文献   

4.
Reduction of 3,4-cis-diacetyl-1,2,3,4-tetramethyl-1-cyclobutene 1 with NaBH4 yielding 2-hydroxyoxolanes 2a and 2b with complete diastereoselectivity at the anomeric carbon atom suggests that a highly stereoselective intramolecular hemiketalization process leading to the formation of the 2-hydroxyoxolane unit is much faster than the attack of the second molecule of the nucleophile on the second carbonyl group. Further reduction of 2a at 0 °C gives selectively the meso-diol 3a. Reaction of 1 with MeLi or MeMgBr also involves the participation of the adjacent carbonyl, thereby yielding hydroxyoxolane 5a selectively. A mechanistic rationale is proposed on the basis of the relative energies of all isomeric hemiketals, 2a-d and 5a and b, (calculated by the PM3 method) and the relative stabilities of the conformers of 1 (calculated by the MM2 method).  相似文献   

5.
The reaction of a rhodanine derivative (=(Z)-5-benzylidene-3-phenyl-2-thioxo-1,3-thiazolidin-4-one; 1) with (S)-2-methyloxirane (2) in the presence of SiO2 in dry CH2Cl2 for 10 days led to two diastereoisomeric spirocyclic 1,3-oxathiolanes 3 and 4 with the Me group at C(2) (Scheme 2). The analogous reaction of 1 with (R)-2-phenyloxirane (5) afforded also two diastereoisomeric spirocyclic 1,3-oxathiolanes 6 and 7 bearing the Ph group at C(3) (Scheme 3). The structures of 3, 4, 6, and 7 were confirmed by X-ray crystallography (Figs. 1 and 2). These results show that oxiranes react selectively with the thiocarbonyl group (CS) in 1. Furthermore, the nucleophilic attack of the thiocarbonyl S-atom at the SiO2-activated oxirane ring proceeds with high regio- and stereoselectivity via an SN2-type mechanism.  相似文献   

6.
Starting from trimethylsilyl enol ether of 1-acetyl-1,3,5-cycloheptatriene, the title 1,1-dimethyl-, 1,1-diethyl-, and 1,1-dipropyl-1H-azulenium cations 6-8 were synthesized in five steps. The order of pKR+ values of these cations was found to be 7>8>6. A comparison of the values between 1,1-dialkyl- and 1,1-spiroalkylated 1H-azulenium cations with the same number of carbon atoms at the 1-position provided the results of 7>1 and 8<3. The cation 8 shows a relatively lower pKR+ value to those of 3 and 7 probably due to its slightly bulkier propyl groups from which solvation stabilization of 8 under the conditions suffers. An intermolecular charge-transfer interaction between the cations and dibenzo-24-crown-8 was also studied.  相似文献   

7.
Muscarone analogues are compounds that have been proposed for the prevention or treatment of senile dementias. ARL-16607 (I), ARL-15467 (II), ARL-14995 (III) and YM-796 (IV) are spiromuscar-3-one derivatives and behave as muscarinic M1 agonists, with different binding selectivity and efficacy at the M1 receptors. In this work, we have elucidated the solid-state structures of I-III and compared the results obtained with the data available in the literature for IV.The solid-state arrangements of I-IV have then been used to input a series of theoretical calculations. For each molecule, eight conformations have been modeled and the obtained structures (1-32) have been submitted to a series of molecular dynamics/simulated annealing and molecular mechanics calculations aimed to explore the conformational freedom of the spiromuscar-3-one moiety. Some hints about the reactivity of I-IV have been obtained by performing Hartree-Fock, density functional theory and semiempirical quantum mechanics calculations. These studies analyzed the properties of the frontier orbitals and of the molecular electrostatic potential of I-IV.The information gained has been used to explain the better efficacy and poorer selectivity shown by III. Our results suggest that the behavior of III is due to its smaller size, the features of its molecular surface, the shape of its electrostatic potential and the orientation of its reactive domains.  相似文献   

8.
Anirban Kar 《Tetrahedron》2005,61(22):5297-5302
Starting from citraconic anhydride (13), a simple multistep (9-10 steps) synthesis of naturally occurring butyrolactones maculalactone A (3), maculalactone B (1), maculalactone C (2) and nostoclide I (4) have been described with good overall yields via dibenzylmaleic anhydride (20) and benzylisopropylmaleic anhydride (27). The two anhydrides 20 and 27 were prepared by SN2′ coupling reactions of appropriate Grignard reagents with dimethyl bromomethylfumarate (14), LiOH-induced hydrolysis of esters to acids, bromination of carbon-carbon double bond, in situ dehydration followed by dehydrobromination and chemoselective allylic substitution of bromoatom in disubstituted anhydrides 19 and 26 with appropriate Grignard reagents. The NaBH4 reduction of these anhydrides 20 and 27 furnished the desired lactones 21 and 29, respectively. The lactone 21 on Knoevenagel condensation with benzaldehyde, furnished maculalactone B (1), which on isomerization gave maculalactone C (2). Selective catalytic hydrogenation of 1 gave maculalactone A (3). The conversion of lactone 29 to nostoclide I (4) is known.  相似文献   

9.
1-Fluoroindan-1-carboxylic acid (FICA) (1) was designed and synthesized as its methyl ester (FICA Me ester) (4) in order to develop an efficient chiral derivatizing agent (CDA) which excels α-methoxy-α-(trifluoromethyl)phenylacetic acid (MTPA) in capability. FICA Me ester (4) was prepared by fluorination of methyl 1-hydroxyindan-1-carboxylate (3) with (diethylamino)sulfur trifluoride (DAST) and derived to the esters of racemic secondary alcohols by ester exchange reaction. The resulting ΔδF value was large in the case of 2-butyl ester of FICA (5a), whereas not detectable in the case of the corresponding MTPA ester (6a). The magnitude of the ΔδH values was similar to that of MTPA esters. The diastereomers of (R)-(−)-8-phenylmenthyl ester of FICA (5i) was separated and their 1H NMR analyses revealed that the concept of the modified Mosher's method was successfully applied to 5i.  相似文献   

10.
The (2-methoxyphenyl)piperazine pharmacophore, a part of the WAY 100635 structure, has been functionalized with phosphinoarylbenzylamide or phosphinoarylbenzylamine chelator groups using propylene or hexylene alkyl chains as linkers (L2-L4). These heterofunctionalized phosphines bearing an arylpiperazine moiety have been used to stabilize rhenium tricarbonyl complexes of the type [Re(CO)3Br(κ2-L)] (4, L = L2; 5, L = L3; 6, L = L4), which have been fully characterized, including by X-ray crystallographic analysis in the case of compounds 4 and 5. These monomeric complexes are six-coordinate, displaying a distorted octahedral coordination geometry with a facial arrangement of the carbonyl groups. The other three remaining positions are occupied by a bromide and by the bidentate heterofunctionalized phosphine, which coordinates through the phosphorus and the oxygen atom or through the phosphorus and the nitrogen atom in 4 and 5, respectively. The 99mTc complexes (3a-6a) were also prepared and their characterization established by comparative HPLC, using the Re complexes as surrogates. The in vitro binding affinity for the 5HT1A receptor subtype and the selectivity against the 5HT2A receptors for the rhenium complexes were determined. Compound 3 is the only one which presents a reasonable affinity and selectively towards 5HT1A (IC50 = 20 nM) and 5HT2A (IC50 = 4680 nM) receptors, respectively. When the spacer length between the chelate unit and receptor binding domain increased and/or the amide group in the chelator was replaced by a secondary amine unacceptable affinity values for 5HT1A receptors (IC50 = 200-1100 nM) and lost of selectivity were observed.  相似文献   

11.
A series of group 12 metal coordination polymers with 1,2-bis(diphenylphosphino)ethane dioxide (dppeO2), {[ZnCl2(μ-dppeO2)]·CH2Cl2}n (1·CH2Cl2), [ZnBr2(μ-dppeO2)]n (2), [CdI2(μ-dppeO2)]n (4), [(HgI2)2(μ-dppeO2)]n (5), [Zn(SCN)(μ-SCN)(μ-dppeO2)]n (6), and [Cd(NO3)(μ-SCN)(μ-dppeO2)]n (7), have been synthesized and structurally characterized. The structures of the compounds are all based on an infinite 1D chain constructed by four-coordinate metal ions and dppeO2 ligands adopting the trans bridging coordination fashion. In the coordination polymers 1, 2 and 4, the halide ions act as terminal ligands, leading to discrete 1D chains with alternative MX2 and dppeO2 repeating units. The mercury compound 5 features a unique square-wave-like inorganic chain –[Hg(1)–I–Hg(2)–I]–, and the 1D HgI2(μ-dppeO2) chains are further linked by HgI2 bridges to form a 3D network. In the thiocyanate-containing compounds 6 and 7, the 1D chains are linked by one (6) or two (7) bridging SCN ions to result in 2D layered structures. Solid-state emission spectra of the coordination polymers show different variations compared to the free dppeO2 ligand, such as enhancement (1, 2, 6 and 7), shift (3 and 4) and quenching (5) upon metal coordination.  相似文献   

12.
β-CF3-α,β-diphenylvinyl sulfide 3a was prepared stereoselectively in 77% yield from the reaction of 2 with phenyllithium at room temperature for 5 h. Oxidation of 3a with MCPBA afforded the corresponding vinyl sulfone 4a, in which (E)-4a can be crystallized in a mixture of CH2Cl2 and hexane. The addition-elimination reaction of (E)-4a with phenyllithium having substituents on the benzene ring provided 5a-j in 51-82% yields stereospecifically. Similarly, the treatment of (E)-4a with p-chloroethoxyphenyllithium in the presence of 12-crown-4 (20 mol %) at −10 °C, followed by slowly warming to room temperature, resulted in the formation of the corresponding panomifene precursor 6 in 82% yield.  相似文献   

13.
The P63 (a=2ap, b=2bp, c=cp) crystal structure reported for BaAl2O4 at room temperature has been carefully re-investigated by a combined transmission electron microscopy and neutron powder diffraction study. It is shown that the poor fit of this P63 (a=2ap, b=2bp, c=cp) structure model for BaAl2O4 to neutron powder diffraction data is primarily due to the failure to take into account coherent scattering between different domains related by enantiomorphic twinning of the P6322 parent sub-structure. Fast Fourier transformation of [0 0 1] lattice images from small localized real space regions (∼10 nm in diameter) are used to show that the P63 (a=2ap, b=2bp, c=cp) crystal structure reported for BaAl2O4 is not correct on the local scale. The correct local symmetry of the very small nano-domains is most likely orthorhombic or monoclinic.  相似文献   

14.
The reactions of Ar2TeO (Ar = 4-MeO-C6H4) with 2-, 3- and 4-pyridine carboxylic acids (LH) afforded different organotelluroxane structural types depending on the stoichiometry of the reactants and the conditions of the reaction. Ar2Te(L)OH (1a-1c) are formed in a 1:1 reaction of Ar2TeO with LH in the presence of water. On the other hand a 1:2 reaction under anhydrous conditions leads to the formation of Ar2TeL2 (2a-2c). A 2:2 reaction under anhydrous conditions affords the ditelluroxanes Ar2Te(L)OTe(L)Ar2 (3a-3c) while tritelluroxanes Ar2Te(L)OTeAr2OTe(L)Ar2 (4a-4c) are formed in 3:2 reactions. Interestingly, 3a-3c are formed in the reaction of 2a-2c with Ar2TeO. The former can be hydrolyzed to 1a-1c while the latter upon reaction with Ar2TeO lead to the formation of the tritelluroxanes 4a-4c. Attempts to metalate 2a with PdCl2(MeCN)2 leads to a transfer of the carboxylate ligand to palladium affording Ar2TeCl2 and PdL2. X-ray crystal structures of representative examples of the family of 1, 2 and 3 reveal interesting supramolecular structures and the formation of a novel [TeO]2 structural unit. The latter results from intermolecular secondary Te?O interactions.  相似文献   

15.
Two cassane diterpenoids, pulcharrin G (1) and 6β-cinnamoyl-7β-hydroxy-voucapen-5-α-ol (2), the constituents of Caesalpinia pulcherrima, were treated with BF3·OEt2 to furnish two olefinic products 3 and 4, respectively. The products were formed by elimination of water and migration of a methyl group from C-4 to C-5. The cytotoxic and antimicrobial activities of 3 and 4 were examined.  相似文献   

16.
The effective syntheses of the enantiomerically pure C1-C17 2 and C18-C25 3 fragments as promising synthetic intermediates of bafilomycin A1, 1 have been achieved.  相似文献   

17.
Novel vitamin D receptor (VDR) antagonists, 24,24-dimethyl-1α-hydroxyvitamin D3-26,23-lactones (8 and 9) and their C2α functionalized analogues (8a-c and 9a-c) were efficiently synthesized and their biological activities were evaluated. The construction of vitamin D3 triene skeleton was achieved by palladium-catalyzed alkenylative cyclization of A-ring precursor enyne (22 and 22a-c) with CD-ring bromoolefin having a 24,24-dimethyl-α-methylene-γ-lactone unit on the side chain (13 and 14). The CD-ring precursors 13 and 14 were prepared by using chromium-mediated allylation of the aldehyde 10 derived from vitamin D2. On the other hand, the A-ring enyne having 2α-(3-hydroxypropyl) group (22b) was newly synthesized from epoxide 15 using regio- and stereoselective alkylation methodology. The potency of the antagonistic activity of the newly designed analogues (8 and 9) increased up to 12 times that of TEI-9647 (2). Furthermore, introduction of the three motifs, that is, a methyl (8a and 9a), an ω-hydroxypropyl (8b and 9b) or an ω-hydroxypropoxyl group (8c and 9c) into the C2α position of 8 and 9, respectively, resulted in remarkable enhancement, up to 89 times, of the antagonistic activity on VDR.  相似文献   

18.
Phase transitions in the elpasolite-type K3AlF6 complex fluoride were investigated using differential scanning calorimetry, electron diffraction and X-ray powder diffraction. Three phase transitions were identified with critical temperatures , and . The α-K3AlF6 phase is stable below T1 and crystallizes in a monoclinic unit cell with a=18.8588(2)Å, b=34.0278(2)Å, c=18.9231(1)Å, β=90.453(1)° (a=2accc, b=4bc, c=ac+2cc; ac, bc, cc—the basic lattice vectors of the face-centered cubic elpasolite structure) and space group I2/a or Ia. The intermediate β phase exists only in very narrow temperature interval between T1 and T2. The γ polymorph is stable in the T2<T<T3 temperature range and has an orthorhombic unit cell with a=36.1229(6)Å, b=17.1114(3)Å, c=12.0502(3)Å (a=3ac−3cc, b=2bc, c=ac+cc) at 250 °C and space group Fddd. Above T3 the cubic δ polymorph forms with ac=8.5786(4)Å at 400 °C and space group . The similarity between the K3AlF6 and K3MoO3F3 compounds is discussed.  相似文献   

19.
The reactions of ketones 1a-o, nitromethane 2, and a stoichiometric amount of piperidine 3a or ethylenediamine 3b in the presence of mercaptan 6a in THF or CH3CN solution give high yields of β-nitrosulfides 7a-o. The latter can be oxidized by 8a (m-CPBA or m-CPBA/AcOH) at 0°C, 8b (H2O2/AcOH), or 8c (H2O2) at room temperature, thus generating β-nitroalkylsulfoxides 9a-o, which then undergo elimination to produce medium to high yields of 2,2-disubstituted-1-nitroalkenes 5a-o, when refluxed in a solution of ClCH2CH2Cl (1,2-dichloroethane). After preparation from 1a-o, 2, 3, and 6a, 7a-o were oxidized with 8a, 8b, or 8c in a mixture of CH3CN and ClCH2CH2Cl to generate β-nitrosulfoxides 9a-o, which then underwent elimination under refluxing under one-pot conditions. Compounds 14 and 15g were also prepared using 13, 2, 3b, and 6, in a similar manner.  相似文献   

20.
The synthesis and the characterization of some new aluminum complexes with bidentate 2-pyrazol-1-yl-ethenolate ligands are described. 2-(3,5-Disubstituted pyrazol-1-yl)-1-phenylethanones, 1-PhC(O)CH2-3,5-R2C3HN2 (1a, R = Me; 1b, R = But), were prepared by solventless reaction of 3,5-dimethyl pyrazole or 3,5-di-tert-butyl pyrazole with PhC(O)CH2Br. Reaction of 1a or 1b with (R1 = Me, Et) yielded N,O-chelate alkylaluminum complexes (2a, R = R1 = Me; 2b, R = But, R1 = Me; 2c, R = Me, R1 = Et). Compound 1a was readily lithiated with LiBun in thf or toluene to give lithiated species 3. Treatment of 3 with 0.5 equiv of MeAlCl2 or AlCl3 yielded five-coordinated aluminum complexes [XAl(OC(Ph)CH{(3,5-Me2C3HN2)-1})2] (4, X = Me; 5, X = Cl). Reaction of 5 with an equiv of LiHBEt3 generated [Al(OC(Ph)CH{(3,5-Me2C3HN2)-1})3] (6). Complex 6 was also obtained by reaction of 3 with 1/3 equiv of AlCl3. Treatment of 5 with 2 equiv of AlMe3 yielded complex 2a, whereas with an equiv of AlMe3 afforded a mixture of 2a and [Me(Cl)AlOC(Ph)CH{(3,5-Me2C3HN2)-1}] (7). Compounds 1a, 1b, 2a-2c and 4-6 were characterized by elemental analyses, NMR and IR (for 1a and 1b) spectroscopy. The structures of complexes 2a and 5 were determined by single crystal X-ray diffraction techniques. Both 2a and 5 are monomeric in the solid state. The coordination geometries of the aluminum atoms are a distorted tetrahedron for 2a or a distorted trigonal bipyramid for 5.  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