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 765 毫秒
1.
A new sesquiterpenoid quinone, Acyl hibiscone B (1), together with five known compounds, (R)-lasiodiplodin (2), (R)-de-O-methyllasiodiplodin, (3) dibutyl phthalate (4), (R)-9-phenylnonan-2-ol (5) and hibiscone B (6), was obtained from the stem tuber of Abelmoschus sagittifolius. The structure of compound 1 was elucidated by analysing its 1H and 13C NMR, 1H–1H COSY, HSQC, HMBC, NOESY and HR-ESI-MS values. Compound 1 showed significant cytotoxicity against Hela and HepG-2 human cancer cell lines.  相似文献   

2.
The high specific activity tritiation of (+)-SKF-10,047 (1) and N,N-dimethyltryptamine (4) is described. [N-allyl-3H] (+)-SKF-10,047 (3) was prepared by Lindlar catalyst tritiation of (+)-N-propargylnormetazocine (2) and [N-methyl-3H] N,N-dimethyltryptamine (6) was synthesized by the alkylation of N-methyltryptamine (5) with [3H] methyl iodide. Both sigma-1 synthetic agonist 3 and endogenous agonist 6 have been useful in studying this receptor.  相似文献   

3.
The vanadium(III) complexes, V(S2CNMe2)3 (1) and V(S2CN i Pr2)3 (2) were prepared and characterized by analysis, IR, electronic and 1H NMR spectra. The complexes show reversible thermochromic behaviour. MM2 calculations were used to simulate the molecular structure of 1. For 2, variable temperature 1H NMR revealed hindered rotation about C–N bonds. The rotational energy barrier (38?kJ?mol?1) was obtained by molecular mechanics force-field calculations.  相似文献   

4.
The reactions of Ru3(CO)12with 4-phenylbut-3-an-2-ine (1a), 3-phenyl-1-p-tolylprop-2-an-1-ine (1b), and 1,3-diferrocenylprop-2-an-1-ine (1c) afforded the Ru2(CO)6(-H)(O=C(R1)C(H)=C(R2)) (2) and Ru3(CO)8(O=C(R1)C(H)=C(R2))2(3) complexes. Dissolution of these complexes in CHCl3or CH2Cl2gave rise to the Ru2(CO)4(-Cl)2(O=C(R1)C(H)=C(R2)) complexes (4). The thermal transformations of complexes 2and 3in the presence of an excess of the ligand yielded the Ru2O2(CO)4(3-OC(R1)C(H)(CH2R2)C(R2)C(H)C(R1))2(5) and Ru(CO)2(O=C(R1)C(H)=C(R2))2(6) complexes. Analogous complexes were obtained upon more prolonged heating of the starting reaction mixtures. The structures of complexes 4a, 5a, and 6cwere established by X-ray diffraction analysis and confirmed by spectroscopic data.  相似文献   

5.
Summary 5-Fluoro-uracil (1) reacts with chloro-diphenylphosphane to 5-fluoro-N(1), N(3)-bis(diphenylphosphanyl)uracil (3) which was characterized by X-ray crystallography, IR, and mass spectra.19F,31P{1H},13C{1H}, and1H NMR spectra indicate that3 rearranges inTHF solution to some extent to the tautomeric 5-fluoro-O(2), O(4)-bis(diphenylphosphanyl)uracil (4). If the solvent contains traces of water, Ph2P(O)-PPh2 (6) and uracil derivatives are formed by hydrolysis.
Herrn Professor Dr.Walter Siebert zum 60. Geburtstag gewidmet.  相似文献   

6.
Iridium pincer complexes [C6H3-2,6-(OPBut 2)2]Ir(H)Cl (10) and [4-EtOOCC6H2-2,6-(OPBut 2)2]Ir(H)Cl (11) react with protic acids undergoing metallation of one of the tert-butyl groups to form double cyclometallated products [4-R-C6H2-2-(OPBut 2)-6-(OP(But)CMe2CH2)]IrCl (12, R = H; 13, R = COOEt), which are stable in air. Complex 12 reacts with CO and ButNC giving the corresponding 18-electron complexes [C6H3-2-(OP-But 2)-6-(OP(But)CMe2CH2)]Ir(L)Cl (14, L = CO; 15, L = CNBut). The structure of compound 14 was established by X-ray diffraction analysis.  相似文献   

7.
The reaction of triallylborane with 3-phenylprop-1-yne at 135–140 °C followed by treatment of the reaction mixture with MeOH afforded 7-benzyl-3-methoxy-3-borabicyclo[3.3.1]non-6-ene (1) in 81% yield. Hydroboration of compound1 with a solution of BH3 in THF yielded the tetrahydrofuran complex of 2-phenyl-1-boraadamantane (2). The reactions of trimethylamine or pyridine with compound2 afforded the trimethylamine (3) or pyridine (4) complexes of 2-phenyl-1-boraadamantane. respectively. Hindered rotation about the C(2)−Ph bond in adduct3 was observed by1H and13C NMR spectroscopy. The activation energy of this process is 58.6 kJ mol−1 (determined by 2D1H−1H EXSY spectroscopy).  相似文献   

8.
The mechanism of “self-spin-probing” reactions between carbanion of nitroalkanes (1) and ω-hydroperfluorodiacyl peroxides (2) initiated by electron transfer has been verified by EPR detection of new bis(ω-hydroperfluoro-dialkyl) nitroxides (11) and ω-hydroperfluoroalkyl nitroalkyl nitroxides (12) — spin adducts of radical intermediates H(CF2CF2)n · (n=1,2,3) and R1R2C·NO2 to the nitroso trap H(CFCF)nNO (9) generated in the reactions.  相似文献   

9.
Chemical investigation of the cultures of a sponge-derived fungus Simplicillium sp. YZ-11 led to the isolation of a new minor diketopiperazine alkaloid cyclo-(2-hydroxy-Pro-Gly) (1) and a natural lactone (S)-dihydro-5-[(S)- hydroxyphenylmethyl]-2(3H)-furanone (2), together with five known ergostane-type sterols (37). Their structures were established based on extensive spectroscopic methods (1H and 13C NMR, 1H-1H COSY, HSQC and HMBC) and optical rotation analysis.  相似文献   

10.
Equimolar reactions of BuSn(OPri)3 with diethanolamines, RN(CH2CH2 OH) 2 (abbreviated as RdeaH2, where R = H or Me), afford dimeric isopropoxo-bridged six-coordinate butyltin(IV) complexes [{Bu(η3-Rdea)Sn(μ-OPri)}2] (R = H ( 1 ), Me ( 2 )). Interactions between BuSn(OPri)3 and diethanolamines (RdeaH2) in a 1:2 molar ratio yield monomeric derivatives of the type [BuSn(Rdea)(RdeaH)] (R = H ( 3 ), R = Me ( 4 )). These homometallic complexes on 1:1 reactions with an appropriate metal alkoxide form monomeric heterobimetallic complexes of the type [BuSn (Rdea)2 {M(OR′)n}] (R = H, M = Al, R′ = Pri, n = 2 ( 5 ); R = H, M = Ti, R = Pri, n = 3 ( 6 ); R = H, M = Zr, R′ = Pri, n = 3 ( 7 ); R = Me, M = Al, R′ = Pri, n = 2 ( 8 ); R = Me, M = Ti, R′ = Pri, n = 3 ( 9 ); R = Me, M = Ge, R′ = Et, n = 3 ( 10 )). The driving force behind this work was (i) to explore the utility of homometal complexes ( 1 ) ( 4 ) in assembling a metal alkoxide fragment via a condensation reaction and (ii) to gain insights into the structures of new compounds by NMR spectral data. All of these derivatives have been characterized by elemental analysis, spectroscopic (IR, NMR; 1H, 27Al, and 119Sn) studies, and molecular weight measurements. 119Sn NMR spectral studies indicate that both the homometallic ( 3 ) and ( 4 ) and heterobimetallic ( 5 ) ( 9 ) complexes exist in a solution in an equilibrium of six- and five-coordinated tin(IV) species.  相似文献   

11.
Abstract

α,β-(1→4)-Glucans were devised as models for heparan sulfate with the simplifying assumptions that carboxyl-reduction and sulfation of heparan sulfate does not decrease the SMC antiproliferative activity and that N-sulfates in glucosamines can be replaced by O-sulfates. The target oligo-saccharides were synthesized using maltosyl building blocks. Glycosylation of methyl 2,3,6,2′,3′,6′-hexa-O-benzyl-β-maltoside (1) with hepta-O-acetyl-α-maltosyl bromide (2) furnished tetrasaccharide 3 which was deprotected to α-D-Glc-(1→4)-β-D-Glc-(1→4)-α-D-Glc-(1→4)-β-D-Glc-(1→OCH3) (5) or, alternatively, converted to the tetrasaccharide glycosyl acceptor (8) with one free hydroxyl function (4?′-OH). Further glycosylation with glucosyl or maltosyl bromide followed by deblocking gave the pentasaccharide [β-D-Glc-(1→4)-α-D-Glc-(1→4)]2-β-D-Glc-(1→OCH3) (11) and hexasaccharide [α-D-Glc-(1→4)-β-D-Glc-(1→4)2-α-D-Glc-(1→4)-β-D-Glc-(1→OCH3) (14). The protected tetrasaccharide 3 and hexasaccharide 12 were fully characterized by 1H and 13C NMR spectroscopy. Assignments were possible using 1D TOCSY, T-ROESY, 1H,1H 2D COSY supplemented by 1H-detected one-bond and multiple-bond 1H,13C 2D COSY experiments.  相似文献   

12.
以3,4-二溴环戊砜(1)为原料, 在无水吡啶作用下发生消除反应, 得到反应中间体4-溴-2-环戊烯砜(2), 再分别与一系列取代苯甲酸盐3a3c以及茜素黄GG (3d)发生酯化反应, 合成出4种新环戊烯砜衍生物4a4d, 并用IR, 1H NMR, MS, 元素分析等表征了它们的结构.  相似文献   

13.
A new six-coordinate organotin(IV)-phosphoric triamide complex of OP[NC5H10]3 = OP was synthesized ([Cl2Sn(CH3)2(OP)2], 1) and characterized by X-ray crystallography and spectroscopic methods (FT-IR, UV–Vis, and 1H/13C/31P-NMR). The crystal structures of 1 and the analogous previously reported five-coordinate complex [Cl2Sn(CH3)2(OP)] (IZOVIE) were compared on a structural level and by computational means using Hirshfeld surface analysis, density functional theory calculations and the atom in molecule method. The investigation of intermolecular interactions in the crystal structures of the two complexes by the Hirshfeld surface method indicates that in the absence of normal hydrogen bonds, the chlorine-based interactions H?Cl/Cl?H (for 1 and IZOVIE) and Cl?Sn/Sn?Cl (for IZOVIE) play a determinant role in the molecular assemblies. However, the prominent contacts are of H?H type. From calculated electronic parameters such as bond order, Mulliken charge and electron delocalization energy, it was found that the Sn-OP contact has a lower strength in IZOVIE than in 1, suggesting more ionic character of the metal-oxygen contact in five-coordinate complex IZOVIE. Furthermore, we discuss the similarities and differences of the two complexes 1 and IZOVIE derived from the same ligand OP by density functional theory calculations to present an insight into the organotin(IV)-phosphoric triamide coordination chemistry affected by different geometries and coordination numbers.  相似文献   

14.
Template reactions of salicylaldehyde or pentanedione with 3-aminopropanethiol (Hapt) in the presence of Ni(II) ions are described. When salicylaldehyde was used, a dinuclear Ni(II) complex [Ni(bit′)]2 (2) (H2bit′?=?2-(3′-mercaptopropyliminomethyl)phenol) was obtained instead of the reported trinuclear one [Ni(bit)]3 (1) (H2bit?=?2-(2′-mercaptoethyliminomethyl)phenol) containing 2-aminoethanethiol (Haet). Starting from pentanedione, the expected dinuclear complex [Ni(pit′)]2 (H2pit′?=?2-(3′-mercaptopropylimino)pentanol) was not obtained, nor was [Ni(pit)]2 (3) (H2pit?=?2-(2′-mercaptoethylimino)pentanol). The complex was found to be a trinuclear Ni(II) complex [Ni{Ni(apt)2}2]2+ (4), as confirmed by elemental analysis, electronic and NMR spectra. Complexes 1 and 3 were also synthesized and their 13C, 1H–1H and 13C–1H?NMR spectra are discussed in detail. The X-ray crystal structure of 2 shows that two Ni(II) ions are connected by the thiolate donor atom from each ligand, resulting in a four-membered ring. Differences in reactivity and properties is due to the presence of an additional methylene group in the aminoalkane arm of the ligand.  相似文献   

15.
Four Ru(II) complexes with tridentate ligands viz. (4-hydroxy-N′-(pyridin-2-yl-ethylene) benzohydrazide [Ru(L1)(PPh3)2(Cl)] (1), N′-(pyridin-2-yl-methylene) nicotinohydrazide [Ru(L2)(PPh3)2(Cl)] (2), N′-(1H-imidazol-2-yl-methylene)-4-hydroxybenzohydrazide [Ru(L3)(PPh3)2(Cl)] (3), and N′-(1H-imidazol-2-yl-methylene) nicotinohydrazide [Ru(L4)(PPh3)2(Cl)] (4) have been synthesized and characterized. The methoxy-derivative of L3H (abbreviated as L3H*) exists in E configuration with torsional angle of 179.4° around C7-N8-N9-C10 linkage. Single crystal structures of acetonitrile coordinated ruthenium complexes of 1 and 3 having compositins as [Ru(L1)(PPh3)2(CH3CN)]Cl (1a) and [Ru(L3)(PPh3)2(CH3CN)]Cl (3a) revealed coordination of tridentate ligands with significantly distorted octahedral geometry constructed by imine nitrogen, heterocyclic nitrogen, and enolate amide oxygen, forming a cis-planar ring with trans-placement of two PPh3 groups and a coordinated acetonitrile. Ligands (L1H-L4H) and their ruthenium complexes (1–4) are characterized by 1H, 13C, 31P NMR, and IR spectral analysis. Ru(II) complexes have reversible to quasi-reversible redox behavior having Ru(II)/Ru(III) oxidation potentials in the range of 0.40–0.71 V. The DNA binding constants determined by absorption spectral titrations with Herring Sperm DNA (HS-DNA) reveal that L4H and 1 interact more strongly than other ligands and Ru(II) complexes. Complexes 1–3 exhibit DNA cleaving activity possibly due to strong electrostatic interactions while 4 displays intercalation.  相似文献   

16.
Polycrystalline (CH3)4NOH·5 H2O (I) and (CH3)4NOD·5D2O (II) have been studied by1H NMR lineshapes, second moments and spin-lattice relaxation times and by2H NMR lineshapes as a function of temperature. From low temperatures the first motion to occur is reorientation of the internally rigid (CH3)4N+ ion about a uniqueC 3 axis (E ta = 8.37 kJ/mol forI,E a = 9.00 kJ/mole forII), followed closely by pseudo isotropic reorientation of the whole ion (E a = 18.10 kJ/mol). Motion of the cage molecules (water and hydroxide ion) occurs at higher temperatures with an apparentE a = 11.30 kJ/mol. There is some evidence of a phase transition inII but notI in the 220–230 K region.2H NMR lineshapes ofII below 220 K indicate static cage molecules. Some of the2H quadrupole coupling constants derived from these spectra correspond to O·O hydrogen-bond distances which are incompatible with the known room temperature structure ofI. Above the possible transition inII the anisotropic2H lineshapes indicate rapid motion of2H among all possible hydrogen-bond sites via transfer along the bonds and molecular reorientation. This motion persists in the high temperature phase but the lineshape becomes isotropic due to the cubic symmetry of this phase. It is possible that1H or2H tunnelling plays an important part in the motion of the cage molecules and the different phase behaviour ofI andII.Dedicated to Dr D. W. Davidson in honor of his great contributions to the sciences of inclusion phenomena.  相似文献   

17.
Four monomeric [n-Bu2SnL2 (1), Et2SnL2 (2), Me2SnL2 (3), and n-Oct2SnL2 (7)] and three polymeric {[n-Bu3SnL]n (4), [Me3SnL]n (5), and [Ph3SnL]n (6)} organotin(IV) carboxylates, where L?=?4-chlorophenylethanoate, were synthesized and characterized by elemental analysis, FT-IR, and multinuclear NMR (1H, 13C, and 119Sn). Compounds 2 and 5 were also analyzed by X-ray single-crystal analysis showing monomeric and zigzag structures, respectively. Two types of O…H (2.641?Å) and Cl…H (2.943?Å) non-covalent interactions generate a 2-D supramolecular structure for 2. Layer-by-layer supramolecular structure was observed for 5 in which polymeric chains are connected via non-covalent interactions {Cl…H (2.869?Å), H…π (2.899?Å)}, and unconventional dihydrogen {H…H (2.381?Å)} interactions.  相似文献   

18.
A density functional theory (DFT) study reveals that dehydrogenation of ethanol catalyzed by aliphatic PNP pincer cobalt complexes, [(PNPEt)Co(H)(OMe)] (1a) and [(PNPEt)Co(H)(CH2SiMe3)] (1b) (PNPEt = bis(2-(diethylphosphino)ethyl)amine), undergoes a self-promoted mechanism, in which an ethanol assists the formation of H2 as a bridge for proton transfer. The calculated total free energy barriers of ethanol dehydrogenation catalyzed by 1a and 1b are 23.9 and 22.2 kcal mol?1, respectively, which indicate that 1b is a promising catalyst for the dehydrogenation of ethanol under mild conditions.  相似文献   

19.
Reaction of [Pd(dppe)Cl2/Br2] with AgOTf in a dichloromethane medium followed by ligand addition led to [Pd(dppe)(OSO2CF3)2] and then [Pd(dppe)(RaaiR)](OSO2CF3)2 [RaaiR′ = p-R-C6H4-N=N-C3H2-NN-1-R′, (1–3), abbreviated as a N,N′-chelator, where N(imidazole) and N(azo) are represented by N and N′, respectively; R = H (a), Me (b), Cl (c) and R′ = Me (1), CH2CH3 (2), CH2Ph (3), OSO2CF3 is the triflate anion, dppe = 1,2-bis-(diphenylphosphinoethane)]. 31P “1H” NMR confirmed that due to the two phosphorus atom interaction in the azoimine symmetrical environment one sharp peak was formed. The 1H NMR spectral measurements suggest that azo-imine link with lot of phenyl protons in the aromatic region. 13C (1H) NMR spectrum, 1H, 1H COSY and 1H, 13C HMQC spectrum assign the solution structure and stereo-retentive conformation in each complex.  相似文献   

20.
Oxidative addition of epoxides to [PtCl4]2– in aqueous acids affords the platinum(iv) derivatives [PtCl5(CH2CR(OH)CH2Cl)]2– (R = H (1) and Me (2)). Complex 1 was isolated as a cesium salt and characterized by IR spectroscopy. The complex is stable in acidic media; under basic conditions, the original epoxide and platinum(ii) are recovered. The formation of complex 2 was detected by 1H NMR spectroscopy.  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