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 23 毫秒
1.
Abstract

New dinaphthosulfoxide aza crowns were prepared from the corresponding dinaphthosulfoxide diester and diamines under reflux conditions (conventional heating) and microwave (MW) irradiation. These reaction routes were used for the synthesis of a series of aza crowns, such as dinaphthosulfide, dinaphthosulfone, dibenzosulfide, and dibenzosulfoxide. The synthesis of aza crowns under MW irradiation in comparison with conventional heating has several advantages, such as shorter reaction times, simpler reaction conditions, and the use of ecofriendly solvents, higher yields, and regioselectivity. In the MW synthesis, macrocyclization is preferred to oligomerization, and thus, the yields of byproducts are low.

Supplemental materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements to view the free supplemental file.

GRAPHICAL ABSTRACT   相似文献   

2.
Dinaphthosulfone aza macrocycles were synthesized from the reaction of diamines and dinaphthosulfone diester (1,1′-sulfoxobis-(2-naphthoxy(2-methyl acetate))) in methanol in the presence of catalytic amounts of para toluenesulfonic acid (p-TsOH). Dinaphthosulfone diester (1,1′-sulfoxobis-(2-naphthoxy(2-methyl acetate))) was synthesized from the corresponding dinaphthosulfide diester and hydrogen peroxide in formic acid at room temperature. Dinaphthosulfide diester was prepared from initial dinaphtholsulfide diol (1,1′-thio bis(2-hydroxy naphthalene)) and methylchloroacetate. 1H NMR spectroscopy showed the unusual splittings for these dinaphthosulfone aza macrocycles, and this finding could be proposed as the role of tetrahedral structure of sulfone functional group, hydrogen bonding in the cavity and size of macrocycle.  相似文献   

3.
Abstract

New dinaphthosulfide aza oxa thia crowns from 1-naphthol were prepared. At first, 2,2′-thiobis(1-hydroxynaphthalene) (a dinaphthol) was prepared from 1-naphthol. Then diester was prepared from the reaction of dinaphthol and methylchloroacetate. Aza crowns were synthesized from the reaction of diester and diamines (ethylene diamine, 1,3-diaminopropane, 1,4-diaminobutane, and 1,6-diaminohexane) in methanol.

Supplemental materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements to view the free supplemental file.

GRAPHICAL ABSTRACT   相似文献   

4.
Some dinaphthosulfide aza macrocycles (3–9) were synthesized based on the conventional route from the reaction of corresponding dinaphthosulfide diester and aliphathic diamines in refluxing methanol in good yields. Dinaphthosulfide diester were synthesized from the reaction of 1,1′-thiobis (2-hydroxy naphthalene) and methyl chloroacetate. The structures of these compounds were confirmed using IR, 1H NMR, 13C NMR, MASS spectroscopy and elemental analysis. Conductometric studies of the complexation of some metal ions with aza derivative 8 (TDN) in methanol as solvent implied the formation of 1:1 complexes. The stability of the 1:1 complexes of TDN decreases in the order Hg2+ >> Pb2+ > Cu2+ > Zn2+ > Ca2+ > Mg2+ > Cd2+ > Sr2+ > Ag+ > Ba2+.  相似文献   

5.
6.
The lipopolysaccharide (LPS) of enterobacteria frequently contains various numbers of charged non‐stoichiometric substituents such as phosphate (P) and ethanolamine (EtN) groups and a third residue of 3‐deoxy‐D ‐manno‐2‐octulosonic acid (KDO) on the R‐core polysaccharide backbone. These substituents can modify the biological activities of LPS including varying the stability of the outer membrane, tolerance to cationic antibiotics, pathogenicity, and sensitivity to enterobacteria bacteriophages. These diverse substituents can be clearly detected in degraded samples of LPS from E. coli C using electrospray ionization single quadrupole mass spectrometry (ESI‐Q‐MS) from a 0.1 mg/mL solution in a 50:50 mixture of methanol and 10 mM ammonium acetate (pH 6.8). The O‐deacylated derivative showed multiple peaks of [M–3H]3? ions which corresponded to species having up to eight phosphates, two ethanolamines, and an additional KDO on the backbone of Hex5 Hep3 KDO2 GlcN2 C14:0(3‐OH)2. The major components of the O,N‐deacylated derivative were the species associated with four and five phosphates on Hex5 Hep3 KDO2 GlcN2. The polysaccharide portion of LPS also revealed species which corresponded to Hex5 Hep3 KDO associated with two to four phosphates and an ethanolamine. The present method was proved to be useful to investigate the structural diversity of enterobacterial LPS. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

7.
Epoxides of fatty acids are hydrolyzed by epoxide hydrolases (EHs) into dihydroxy fatty acids which are of particular interest in the mammalian leukotriene pathway. In the present report, the analysis of the configuration of dihydroxy fatty acids via their respective hydroxylactones is described. In addition, the biotransformation of (±)‐erythro‐7,8‐ and ‐3,4‐dihydroxy fatty acids in the yeast Saccharomyces cerevisiae was characterized by GC/EI‐MS analysis. Biotransformation of chemically synthesized (±)‐erythro‐7,8‐dihydroxy(7,8‐2H2)tetradecanoic acid ((±)‐erythro‐ 1 ) in the yeast S. cerevisiae resulted in the formation of 5,6‐dihydroxy(5,6‐2H2)dodecanoic acid ( 6 ), which was lactonized into (5S,6R)‐6‐hydroxy(5,6‐2H2)dodecano‐5‐lactone ((5S,6R)‐ 4 ) with 86% ee and into erythro‐5‐hydroxy(5,6‐2H2)dodecano‐6‐lactone (erythro‐ 8 ). Additionally, the α‐ketols 7‐hydroxy‐8‐oxo(7‐2H1)tetradecanoic acid ( 9a ) and 8‐hydroxy‐7‐oxo(8‐2H1)tetradecanoic acid ( 9b ) were detected as intermediates. Further metabolism of 6 led to 3,4‐dihydroxy(3,4‐2H2)decanoic acid ( 2 ) which was lactonized into 3‐hydroxy(3,4‐2H2)decano‐4‐lactone ( 5 ) with (3R,4S)‐ 5 =88% ee. Chemical synthesis and incubation of (±)‐erythro‐3,4‐dihydroxy(3,4‐2H2)decanoic acid ((±)‐erythro‐ 2 ) in yeast led to (3S,4R)‐ 5 with 10% ee. No decano‐4‐lactone was formed from the precursors 1 or 2 by yeast. The enantiomers (3S,4R)‐ and (3R,4S)‐3,4‐dihydroxy(3‐2H1)nonanoic acid ((3S,4R)‐ and (3R,4S)‐ 3 ) were chemically synthesized and comparably degraded by yeast without formation of nonano‐4‐lactone. The major products of the transformation of (3S,4R)‐ and (3R,4S)‐ 3 were (3S,4R)‐ and (3R,4S)‐3‐hydroxy(3‐2H1)nonano‐4‐lactones ((3S,4R)‐ and (3R,4S)‐ 7 ), respectively. The enantiomers of the hydroxylactones 4, 5 , and 7 were chemically synthesized and their GC‐elution sequence on Lipodex® E chiral phase was determined.  相似文献   

8.
Abstract

A simple and eco-compatible synthesis of podands as 4a-i is performed using amines (3a-c ), CS2, and dichlordiamides (DCDs) (2a, c ) in the absence of a catalyst in water. Three reacting DCDs (2a-c ) were obtained in the reaction of aromatic diamines (1a, c ) with chloroacetyl chloride. Dynamic NMR spectroscopic data of three series of podands (4a-c , 4d-f , and 4g-i ) are discussed and their free energies of activations are calculated (ΔG c s) at coalescence temperatures. The ΔG c of these podands were attributed to conformational isomerization in the range of 14.7–17 kcal/mol due to rotation and resonance effects about the thioamide C?N bond.

[Supplementary materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfer, and Silicon and the Related Elements for the following free supplemental files: Additional figures.]  相似文献   

9.
The synthesis of unsymmetrical diphosphine ligands ( 3 a – g ) with an o‐tolyl backbone and tert‐butyl, adamantyl, cyclohexyl and isopropyl substituents on the phosphorus moiety is described (1,2‐(CH2PR2)(PR′2)C6H4; 3 a : R=tBu, R′=tBu, 3 b : R=tBu, R′=Cy, 3 c : R=tBu, R′=iPr, 3 d : R=Ad, R′=tBu, 3 e : R=Ad, R′=Cy, 3 f : R=Cy, R′=Cy, 3 g : R=Ad, R′=Ad). The corresponding diphosphine–PdII ditriflate complexes [(P^P)Pd(OTf)2] ( 5 a – g ) were prepared and structurally characterised by X‐ray crystallography. These new complexes were studied as catalyst precursors in the isomerising methoxycarbonylation of methyl oleate, and were found to convert methyl oleate into the corresponding linear α,ω‐diester ( L ) with 70–80 % selectivity. The products of this catalytic reaction with the known [{1,2‐(tBu2PCH2)2C6H4}Pd(OTf)2] complex ( 5 h ) were fully analysed, and revealed the formation of the linear α,ω‐diester ( L , 89.0 %), the methyl‐branched diester B1 (4.3 %), the ethyl‐branched diester B2 (1.0 %), the propyl‐branched diester B3 (0.6 %) and all diesters from butyl‐ to hexadecyl‐branched diesters B4 – B16 (overall 4.8 %) at 90 °C and 20 bar CO. The productivity of the catalytic conversion of methyl oleate with complexes 5 a – g varied with the steric bulk of the alkyl substituent on the phosphorus. Ligands with more bulky groups, like tert ‐ butyl or adamantyl (e.g., 5 a , 5 d , 5 g ), were more productive systems. The formation of the catalytically active hydride species [(P^P)Pd(H)(MeOH)]+ ( 6‐MeOH ) was investigated and observed directly for complexes 5 a – e and 5 g , respectively. These hydride species were isolated as the corresponding triphenylphosphine complexes ( 6‐PPh3 ) and fully characterised, including by X‐ray crystallography. The catalytic productivity of 6 a‐PPh3 was virtually identical to that of 5 a , thereby confirming the efficient hydride formation of 5 a under catalytic conditions.  相似文献   

10.
A convenient method of synthesis of phosphoryl-substituted podands with dipehnylphosphorylmethyl end groups of the general formula Ph2P(O)CH2O(CH2CH2O)nCH2P(O)Ph2 (Ln, n = 0–6) is described. The stability constants of the complexes of the podands with alkali metal 2,4-dinitrophenolates were determined by conductometry. The ion-selective properties of the podands with respect to alkali and alkaline-earth metal cations were assessed by ionometry. The crystal structure of 1,3-bis(diphenylphosphoryl)-2-oxapropane was established by X-ray diffraction analysis.  相似文献   

11.
The synthesis of tetramethoxyresorcinarene podands bearing p-toluene arms connected by -SO3- ( 1 ) and -CH2O- ( 2 ) linkers is presented herein. In the solid state, the resorcinarene podand 1 forms an intramolecular self-inclusion complex with the pendant p-toluene group of a podand arm, whereas the resorcinarene podand 2 does not show self-inclusion. The conformations of the flexible resorcinarene podands in solution were investigated by variable-temperature experiments using 1D and 2D NMR spectroscopic techniques as well as by computational methods, including a conformational search and subsequent DFT optimisation of representative structures. The 1H NMR spectra of 1 and 2 at room temperature show a single set of proton signals that are in agreement with C4v symmetry. At low temperatures, the molecules exist as a mixture of boat conformations featuring slow exchange on the NMR timescale. Energy barriers (ΔG298) of 55.5 and 52.0 kJ mol−1 were calculated for the boat-to-boat exchange of 1 and 2 , respectively. The results of the ROESY experiments performed at 193 K and computational modelling suggest that in solution the resorcinarene podand 1 adopts a similar conformation to that present in its crystal structure, whereas podand 2 populates a more versatile range of conformations in solution.  相似文献   

12.
Some novel bis-(substituted-phenoxy) ended glycols were synthesised usinghydroxy aromatics of vanillin, o-vanillin, iso-vanillin and 4-hydroxy coumarin which reacted with bis-dihalides of polyglycols in the presence ofDMSO/alkali carbonate. The novel podands, Ar-(CH2CH2O)m-Ar,(m = 1–4), were identified with IR, 1H-NMR, 13C-NMR and mass spectrometry. The various (formyl-methoxy)phenyl and 4-oxycoumarin derivatives of glycols were studied to estimate the cation binding selectivity of SCN- salts ofLi+, Na+, K+ and Zn2+ cations in acetonitrile using steady statefluorescence spectroscopy. The relevant structures of podands have shown goodselectivity depending on the cation and the glycollength, although the chromophoreend groups have no specific contribution on binding.  相似文献   

13.
Photochromic 6‐bromomethyl‐6′‐methyl‐[2,2′‐bi‐1H‐indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 2 ), 6,6′‐ bis(bromomethyl)‐[2,2′‐bi‐1H‐indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 3 ) and 6,6′‐bis(dibromomethyl)‐[2,2′‐ bi‐1H‐indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 4 ) have been synthesized from 6,6′‐dimethyl‐[2,2′‐bi‐1H‐ indene]‐3,3′‐diethyl‐3,3′‐dihydroxy‐1,1′‐dione ( 1 ). The single crystal of 4 was obtained and its crystal structure was analyzed. The results indicate that in crystal 4 , molecular arrangement is defective tightness compared with its precursor 1 . Besides, UV‐Vis absorption spectra in CH2Cl2 solution, photochromic and photomagnetic properties in solid state of 2 , 3 and 4 were also investigated. The results demonstrate that when the hydrogen atoms in the methyl group on the benzene rings of biindenylidenedione were substituted by bromines, its properties could be affected considerably.  相似文献   

14.
Rate coefficients have been determined for the gas‐phase reaction of the hydroxyl (OH) radical with the aromatic dihydroxy compounds 1,2‐dihydroxybenzene, 1,2‐dihydroxy‐3‐methylbenzene and 1,2‐dihydroxy‐4‐methylbenzene as well as the two benzoquinone derivatives 1,4‐benzoquinone and methyl‐1,4‐benzoquinone. The measurements were performed in a large‐volume photoreactor at (300 ± 5) K in 760 Torr of synthetic air using the relative kinetic technique. The rate coefficients obtained using isoprene, 1,3‐butadiene, and E‐2‐butene as reference hydrocarbons are kOH(1,2‐dihydroxybenzene) = (1.04 ± 0.21) × 10−10 cm3 s−1, kOH(1,2‐dihydroxy‐3‐methylbenzene) = (2.05 ± 0.43) × 10−10 cm3 s−1, kOH(1,2‐dihydroxy‐4‐methylbenzene) = (1.56 ± 0.33) × 10−10 cm3 s−1, kOH(1,4‐benzoquinone) = (4.6 ± 0.9) × 10−12 cm3 s−1, kOH(methyl‐1,4‐benzoquinone) = (2.35 ± 0.47) × 10−11 cm3 s−1. This study represents the first determination of OH radical reaction‐rate coefficients for these compounds. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 696–702, 2000  相似文献   

15.
The new phytosphingosine‐type ceramide 1 , named paxillamide (=2,3‐dihydroxy‐N‐[(1S,2S,3R)‐2,3‐dihydroxy‐1‐(hydroxymethyl)heptadecyl]tetracosanamide), was isolated from the CHCl3/MeOH extract of the fruiting bodies of the Basidiomycete Paxillus panuoides, and its structure was elucidated by spectroscopic and chemical methods.  相似文献   

16.
The structures of three related keto diester and diester ylides, namely diethyl 3‐oxo‐2‐(triphenylphosphoranylidene)glutarate, C27H27O5P, (I), diethyl 3‐oxo‐2‐(triphenylphosphoranylidene)glutarate acetic acid monosolvate, C27H27O5P·C2H4O2, (II), and diethyl 2‐(triphenylphosphoranylidene)succinate, C26H27O4P, (III), are presented. The syn‐keto anti‐ester conformations in the crystalline keto diesters are governed by electronic delocalization between the P—C and ylidic bonds and an acyl group, and by intra‐ and intermolecular interactions. There are also intramolecular attractive and repulsive interactions of different types (C—H...O and C—H...π) controlling the molecular conformations. The mono‐ylidic diester (III) has an anti‐ester conformation, while those for (I) and (II) are related to pyrolytic formation of acetylene derivatives. The terminal nonylidic ester group in (I) was disordered over two sets of almost equally populated positions.  相似文献   

17.
The reaction of dithiocarbamate salts ( IV a − c ) with bis(naphthalene chloroacetates) ( II a,b ) and bis(naphthalene ethoxybromide) ( II c ) in dimethylformamide (DMF) furnished corresponding podands as V a − i in high to excellent yields. Three reacting ligands, ( II a,b ) and ( II c ), were obtained in the reaction of bis(naphthalene) ( I a,b ) with chloroacetylchloride and 1,2‐dibromoethane. Dynamic NMR spectroscopic data of three series of podands ( V a − c , V d−f , and V g − i ) are discussed, and their free energy of activation (Δ GC) at coalescence temperatures are figured out. The Δ GC s of these podands were attributed to conformational isomerization in the range of 14.5–18.3 kcal mol−1 due to rotation and resonance effects about thioamide C N bond. © 2011 Wiley Periodicals, Inc. Heteroatom Chem 22:659–668, 2011; View this article online at wileyonlinelibrary.com . DOI 10.1002/hc.20730  相似文献   

18.
Stilbenoids are the main components of leaves and stems of Pholidota chinensis. In the present investigation, high‐speed counter‐current chromatography was used for the separation and purification of two classes of stilbenoids, namely, bibenzyls and 9,10‐dihydrophenanthrenes, on a preparative scale from whole plants of P. chinensis with different solvent systems after silica gel column chromatography fractionation. n‐Hexane/ethyl acetate/methanol/water (1.2:1:1:0.8, v/v/v/v) was selected as the optimum solvent system to purify 1‐(3,4,5‐trimethoxyphenyl)‐1′,2′‐ethanediol ( 1 ), coelonin ( 2 ), 3,4′‐dihydroxy‐5,5′‐dimethoxybibenzyl ( 3 ), and 2,?7‐?dihydroxy‐?3,?4,?6‐?trimethoxy‐?9,?10‐?dihydrophenanthrene ( 4 ). While 2,7‐dihydroxy‐3,4,6‐trimethoxy‐?9,?10‐?dihydrophenanthrene ( 5 ), batatasin III ( 6 ), orchinol ( 7 ), and 3′‐O‐methylbatatasin III ( 8 ) were purified by n‐hexane/ethyl acetate/methanol/water (1.6:0.8:1.2:0.4, v/v/v/v). After the high‐speed counter‐current chromatography isolation procedure, the purity of all compounds was over 94% assayed by ultra high performance liquid chromatography. The chemical structure identification of all compounds was carried out by mass spectrometry and 1H and 13C NMR spectroscopy. To the best of our knowledge, the current investigation is the first study for the separation and purification of bibenzyls and 9,10‐dihydrophenanthrenes by high‐speed counter‐current chromatography from natural resources.  相似文献   

19.
Three types of silicon podands, Me2Si(OR)2, EtSi(OR)3 and PhSi(OR)3, where R is a polyoxaethylene chain with different numbers of oxygen atoms (two, three or four), were obtained and studied by 1H, 13C and 29Si NMR methods. NMR spectra of 1H, 7Li, 13C, 23Na and 29Si nuclei were also used for the study of lithium, sodium and rubidium complexes with the silicon podands. Theoretical calculations were performed using the PM3 hamiltonian. The heats of reactions between the compounds obtained and SbCl5 were determined.  相似文献   

20.
In our present study, two groups of xanthones isomers (1‐hydroxy‐3,5,8‐trimethoxyxanthone and 1‐hydroxy‐3,7,8‐trimethoxyxanthone; 1,8‐dihydroxy‐3,7‐dimethoxyxanthone and 1,8‐dihydroxy‐3,5‐dimethanolxanthone) and other two xanthones (3‐methoxy‐1,5,8‐trihydroxyxanthone and 3,5‐dimethoxy‐1‐hydroxyxanthone) were separated from Swertia franchetiana . First, a solvent system composed of petroleum ether/methanol/water (2:1:0.6, v/v) was developed for the liquid–liquid extraction of these xanthones from the crude extract. Then, an efficient method was established for the one‐step separation of these six xanthones by high‐speed countercurrent chromatography using n‐hexane/ethyl acetate/methanol/ethanol/water (HEMEW; 6:4:4:2:4, v/v) as the solvent system. The results showed that liquid–liquid extraction could be well developed for efficient enrichment of target compounds. Additionally, high‐speed countercurrent chromatography could be a powerful technology for separation xanthones isomers. It was found ethanol could be a good methanol substitute when the HEMEW system could not provide good separation factors.  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