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 87 毫秒
1.
One of the two bridging protons of the aza‐nido‐decaboranes RNB9H10X can be removed by certain bases to give nido‐anions [RNB9H9X] [R/X = H/H ( 1 a ), Ph/H ( 1 b ), p‐MeC6H4/H ( 1 c ), Bzl/H ( 1 d ), H/N3 ( 1 ′ a )]; the stericly demanding base 1,8‐bis(dimethylamino)naphthalene (“proton sponge”, ps) is ideal. In the case of tBu anion, the deprotonation (→ C4H10) may be accompanied by a hydridation (→ C4H8), yielding the arachno‐anions [RNB9H11X] ( 2 a , b , d , 2 ′ a ); these are the main products, when stericly non‐demanding bases like H are applied. The Lewis acid BH3 is added to 1 a and 1 ′ a to give the aza‐arachno‐undecaborates HNB10H12X [X = H ( 3 a ), N3 (in position 2) ( 3 ′ a )]. Thia‐ and selenaaza‐arachno‐undecaborates, [S(RN)B9H10] ( 4 b , c ) and [Se(RN)B9H10] ( 4 ′ b , c ), are obtained from 1 b , c by the addition of sulfur or selenium, respectively. The methylation of the anions 4 c and 4 ′ c gives the thia‐ and selenaazaarachno‐undecaboranes (MeS)(RN)B9H10 ( 5 c ) and (MeSe)(RN)B9H10 ( 5 ′ c ), respectively. The action of HBF4 on the arachno‐borates [HNB10H12X] ( 3 a , 3 ′ a ) leads to a mixture of nido‐HNB9H10X and nido‐HNB10H11X by the elimination of BH3 or H2, respectively; the aza‐nido‐decaborane predominates in the case of 3 ′ a and the aza‐nido‐undecaborane in the case of 3 a . The action of HBF4 on the anion 4 c yields the hypho‐undecaborate [S(RN)B9H10F2] ( 6 c ). The structures of the products are elucidated on the basis of 1H and 11B NMR spectra, supported by 2D COSY and HMQC techniques. Two types of 11‐vertex‐arachno structures and an 11‐vertex‐hypho structure are found for the products. The crystal structures of 5 c and [Hps] 6 c · CH2Cl2 are reported.  相似文献   

2.
Jing Cao  Yang Li  Junxiang Feng 《中国化学》2012,30(7):1571-1574
Two kinds of novel styryl chemosensory 2‐FMNC and 3‐FMNC, were designed and synthesized by an apporiate introduction of 9‐benzylidene‐9H‐fluorene group as fluorophore with the aim at avoiding photoisomerisation. These 9‐benzylidene‐9H‐fluorene derivatives showed the similar selectivity and sensitivity upon addition of metal ions. The sensitivity of FMNC to alkaline earth metal ions was Ba2+>Sr2+>Ca2+≈Mg2+.  相似文献   

3.
In Cannabis sativa, Δ9‐Tetrahydrocannabinolic acid‐A (Δ9‐THCA‐A) is the non‐psychoactive precursor of Δ9‐tetrahydrocannabinol (Δ9‐THC). In fresh plant material, about 90% of the total Δ9‐THC is available as Δ9‐THCA‐A. When heated (smoked or baked), Δ9‐THCA‐A is only partially converted to Δ9‐THC and therefore, Δ9‐THCA‐A can be detected in serum and urine of cannabis consumers. The aim of the presented study was to identify the metabolites of Δ9‐THCA‐A and to examine particularly whether oral intake of Δ9‐THCA‐A leads to in vivo formation of Δ9‐THC in a rat model. After oral application of pure Δ9‐THCA‐A to rats (15 mg/kg body mass), urine samples were collected and metabolites were isolated and identified by liquid chromatography‐mass spectrometry (LC‐MS), liquid chromatography‐tandem mass spectrometry (LC‐MS/MS) and high resolution LC‐MS using time of flight‐mass spectrometry (TOF‐MS) for accurate mass measurement. For detection of Δ9‐THC and its metabolites, urine extracts were analyzed by gas chromatography‐mass spectrometry (GC‐MS). The identified metabolites show that Δ9‐THCA‐A undergoes a hydroxylation in position 11 to 11‐hydroxy‐Δ9‐tetrahydrocannabinolic acid‐A (11‐OH‐Δ9‐THCA‐A), which is further oxidized via the intermediate aldehyde 11‐oxo‐Δ9‐THCA‐A to 11‐nor‐9‐carboxy‐Δ9‐tetrahydrocannabinolic acid‐A (Δ9‐THCA‐A‐COOH). Glucuronides of the parent compound and both main metabolites were identified in the rat urine as well. Furthermore, Δ9‐THCA‐A undergoes hydroxylation in position 8 to 8‐alpha‐ and 8‐beta‐hydroxy‐Δ9‐tetrahydrocannabinolic acid‐A, respectively, (8α‐Hydroxy‐Δ9‐THCA‐A and 8β‐Hydroxy‐Δ9‐THCA‐A, respectively) followed by dehydration. Both monohydroxylated metabolites were further oxidized to their bishydroxylated forms. Several glucuronidation conjugates of these metabolites were identified. In vivo conversion of Δ9‐THCA‐A to Δ9‐THC was not observed. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
1,3‐Dichloro‐tetra‐n‐butyl‐distannoxane was firstly used to catalyze the direct substitution of 9H‐xanthen‐9‐ol with indoles at room temperature to afford a class of 3‐(9H‐xanthen‐9‐yl)‐1H‐indole derivatives in good to excellent isolating yield. Moreover, other nucleophiles (such as diketone and pyrrole) could also proceed smoothly in this methodology. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

5.
The solvolysis of eight 9‐aryl‐9‐bromofluorenes ( 6b~6i ) in a variety of solvents were studied. Correlation analysis using single‐parameter Grunwald‐Winstein equation (Eqn. 1) with different Y scales showed good linearity (R ≥ 0.98) for most cases if YxBnBr was employed. Linear relationships were observed from Hammett‐type analysis of logarithm of rate constants using Brown‐Okamoto σ+ constants (Eqn. 3), although inverse order of k(p‐CF3)/k(m‐ CF3) was realized in a number of cases. The ρ values were found to vary slightly with different solvent systems. Calculated atomic charge reveals the similarity between 9‐phenyl‐9‐fluorenyl cation ( 7 ) and triphenylmethyl cation ( 8 ). An extended charge delocalization throughout the fluorenyl ring led to the conclusion of the insignificance of antiaromaticity, which was in harmony with that obtained in previous studies. The variation of relative kBr/kCl rate ratios was attributed to the electrophilic pull by solvents in solvolysis.  相似文献   

6.
For GC-MS analysis of delta-9-tetrahydrocannabinol (delta-9-THC), perfluoroacid anhydrides in combination with perfluoroalcohols are commonly used for derivatization. This reagent mixture is preferred because it allows simultaneous derivatization of delta-9-THC and its acid metabolite, 11-nor-delta-9-THC-9-carboxylic acid present in biological samples. When delta-9-THC was derivatized by trifluoroacetic anhydride/hexafluoroisopropanol (TFAA/HFIPOH) and analyzed by GC-MS using full scan mode (50-550 amu), two peaks (P1 and P2) with an identical molecular mass of 410 amu were observed. On the basis of the total ion chromatogram (TIC), P1 with a shorter retention time (RT) was the major peak (TIC 84%). To identify the peaks, delta-8-THC was also tested under the same conditions. The RT and spectra of the major peak (TIC 95%) were identical with that of P1 for delta-9-THC. A minor peak (5%) present also correlated well with the latter peak (P2) for the delta-9-THC derivative. The fragmentation pathway of P1 was primarily demethylation followed by retro Diels-Alder fragmentation (M - 15-68, base peak 100%) indicating P1 as a delta-8-THC-trifluoroacetyl compound. This indicated that delta-9-THC isomerized to delta-8-THC during derivatization with TFAA/HFIPOH. Similar results were also observed when delta-9-THC was derivatized with pentafluoropropionic anhydride/pentafluoropropanol or heptafluorobutyric anhydride/heptafluorobutanol. No isomerization was observed when chloroform was used in derivatization with TFAA. In this reaction, the peaks of delta-8-THC-TFA and delta-9-THC-TFA had retention times and mass spectra matching with P1 and P2, respectively. Because of isomerization, perfluoroacid anhydrides/perfluoroalcohols are not suitable derivatizing agents for analysis of delta-9-THC; whereas the TFAA in chloroform is suitable for the analysis.  相似文献   

7.
The1H and 13C NMR resonances for acridine derivatives 9‐substituted with chloro, allylamino and propargylamino groups were completely assigned using a concerted application of gs‐COSY, gs‐HMQC and gs‐HMBC experiments. 9‐(N‐Allyl)‐ and 9‐(N‐propargyl)acridinamine derivatives present amino–imino tautomerism including a large broadening of 1H and 13C NMR signals at room temperature. To obtain suitable resolution, therefore, these latter compounds were studied at 370 K in DMSO‐d6 solutions and showed a complete shift towards the imino tautomers. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

8.
Delta9-tetrahydrocannabinolic acid A (Delta9-THCA-A) is the precursor of Delta9-tetrahydrocannabinol (Delta9-THC) in hemp plants. During smoking, the non-psychoactive Delta9-THCA-A is converted to Delta9-THC, the main psychoactive component of marihuana and hashish. Although the decarboxylation of Delta9-THCA-A to Delta9-THC was assumed to be complete--which means that no Delta9-THCA-A should be detectable in urine and blood serum of cannabis consumers--we found Delta9-THCA-A in the urine and blood serum samples collected from police controls of drivers suspected for driving under the influence of drugs (DUID). For LC-MS/MS analysis, urine and blood serum samples were prepared by solid-phase extraction. Analysis was performed with a phenylhexyl column using gradient elution with acetonitrile. For detection of Delta9-THCA-A, the mass spectrometer (MS) (SCIEX API 365 triple-quadrupole MS with TurboIonSpray source) was operated in the multiple reaction monitoring (MRM) mode using the following transitions: m/z357 --> 313, m/z357 --> 245 and m/z357 --> 191. Delta9-THCA-A could be detected in the urine and blood serum samples of several cannabis consumers in concentrations of up to 10.8 ng/ml in urine and 14.8 ng/ml in serum. The concentration of Delta9-THCA-A was below the Delta9-THC concentration in most serum samples, resulting in molar ratios of Delta9-THCA-A/Delta9-THC of approximately 5.0-18.6%. Only in one case, where a short elapsed time between the last intake and blood sampling is assumed, the molar ratio was 18.6% in the serum. This indicates differences in elimination kinetics, which need to be investigated in detail.  相似文献   

9.
The crystal structure of the title bifunctional silicon‐bridged compound, C35H31NSi, (I), has been determined. The compound crystallizes in the centrosymmetric space group P21/c. In the crystal structure, the pairs of aryl rings in the two different chromophores, i.e. 9‐phenyl‐9H‐carbazole and 9,9‐dimethyl‐9H‐fluorene, are positioned orthogonally. In the crystal packing, no classical hydrogen bonding is observed. UV–Vis absorption and fluorescence emission spectra show that the central Si atom successfully breaks the electronic conjugation between the two different chromophores, and this was further analysed by density functional theory (DFT) calculations.  相似文献   

10.
Bi9Rh2Br3, Bi9Rh2I3, and Bi9Ir2I3 – A New Structure Family of Quasi One‐dimensional Metals Bi9Rh2Br3, Bi9Rh2I3, and Bi9Ir2I3 were synthesized from the elements using niobium bromides or iodides as auxiliaries to modify the partial pressures in the course of the reaction. X‐ray diffraction on single crystals showed that the compounds are not isomorphous. However they have a common structural principle: strands of condensed [MBi8] polyhedra, which are separated by halide anions. The spatial arrangement of the [MBi1/1Bi7/2] strands differs with the combination of elements: In Bi9Rh2I3 (monoclinic, P21/m (no. 11), a = 775.6(1), b = 1374.9(2), c = 901.1(2) pm, β = 109.29(2)°) all strands are oriented parallel to each other. Bi9Rh2Br3 (monoclinic, P21/m (no. 11), a = 927.98(8), b = 1372.1(1), c = 1992.7(2) pm, β = 100.77(1)°) and Bi9Ir2I3 (orthorhombic, Pnma (no. 62), a = 2677.5(5), b = 1394.2(2), c = 967.6(1) pm) are ordered polytypes with two orientations changing in alternating layers of characteristic widths. The experimental proof of metallic conductivity in Bi9Ir2I3 supports the assumption of delocalised electrons inside the  [MBi1/1Bi7/2] strands. The magnetic susceptibility of Bi9Rh2Br3 increases slowly with decreasing temperature and shows a local maximum at about 14 K.  相似文献   

11.
The first preparation of acridin‐9(10H)‐ones carrying a tertiary thiocarbamoyl group at C(10), i.e., N,N‐dialkyl‐9‐oxoacridine‐10(9H)‐carbothioamides 9 , is described. The method is based on the reaction of (2‐halophenyl)(2‐isothiocyanatophenyl)methanones 7 , prepared from (2‐aminophenyl)(2‐halophenyl)methanones 5 by a convenient three‐step sequence, with secondary amines in DMF at room temperature to generate the corresponding thiourea derivatives 8 in situ, which are treated with NaH at 100–120° to provide the desired products in one‐pot reactions in generally good yields.  相似文献   

12.
The synthesis of 2‐ethynyl‐9‐substituted carbazole and 3‐ethynyl‐9‐substituted carbazole monomers containing first‐generation chiral and achiral dendritic (i.e., minidendritic) substituents, 2‐ethynyl‐9‐[3,4,5‐tris(dodecan‐1‐yloxy)benzyl]carbazole (2ECz), 3‐ethynyl‐9‐[3,4,5‐tris(dodecan‐1‐yloxy)benzyl]carbazole (3ECz), 2‐ethynyl‐9‐{3,4,5‐tris[(S)‐2‐methylbutan‐1‐yloxy]benzyl}carbazole (2ECz*), and 3‐ethynyl‐9‐{3,4,5‐tris[(S)‐2‐methylbutan‐1‐yloxy]benzyl}carbazole (3ECz*), is presented. All monomers were polymerized and copolymerized by stereospecific polymerization to produce cis‐transoidal soluble stereoisomers. A structural analysis of poly(2ECz), poly(2ECz*), poly(3ECz), poly(3ECz*), poly(2ECz*‐co‐2ECz), and poly(3ECz*‐co‐3ECz) by a combination of techniques, including 1H NMR, ultraviolet–visible, and circular dichroism spectroscopy, thermal optical polarized microscopy, and X‐ray diffraction experiments, demonstrated that these polymers had a helical conformation that produced cylindrical macromolecules exhibiting chiral and achiral nematic phases. Individual chains of these cylindrical macromolecules were visualized by atomic force microscopy. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 3509–3533, 2002  相似文献   

13.
In(DS)3 catalyzes formation of 9H‐xanthen‐9‐ol with indoles at room temperature in water to afford a class of 3‐(9H‐xanthen‐9‐yl)‐1H‐indole derivatives in high yields.  相似文献   

14.
An investigation into the preparation of poly(9‐alkyl‐9H‐carbazole‐3,6‐diyl)s with palladium catalyzed cross‐coupling reactions of 3‐halo‐6‐halomagnesio‐9‐alkyl‐9H‐carbazoles, generated in situ from their corresponding 3,6‐diiodo‐ and 3,6‐dibromo‐derivatives was undertaken. Monomers with a range of alkyl group substituents with different steric requirements were investigated and their effects on the polymerization were studied. The effects of the nature of halogen substituents on the polymerization reaction were also investigated. Structural analysis of the polymers revealed exclusive 3,6‐linkage between consecutive carbazole repeat units on the polymer chains. The physical properties of these polymers were investigated with spectroscopic, thermal gravimetric analysis, and electrochemical studies. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 6041–6051, 2004  相似文献   

15.
The antitumor activity of a colon‐specific N‐(2‐hydroxypropyl)methacrylamide (HPMA) copolymer – 9‐aminocamptothecin (9‐AC) conjugate (P‐9‐AC) was assessed in orthotopic and subcutaneous animal (HT29 xenograft) tumor models. P‐9‐AC treatment of mice bearing orthotopic colon tumors, with a dose of 3 mg/kg of 9‐AC equivalent every other day for 6 weeks, resulted in regression of tumors in 9 of 10 mice. A lower dose of P‐9‐AC (1.25 mg/kg of 9‐AC equivalent) every other day for 8 weeks inhibited subcutaneous tumor growth in all mice. No liver metastases were observed. Colon‐specific release of 9‐AC from polymer conjugates enhanced antitumor activity and minimized the systemic toxicity.

  相似文献   


16.
A novel acrylic monomer‐bearing carbazole chromophore, 3‐methacrylamide‐9‐ethyl‐carbazole and its model compound 3‐isobutyramide‐9‐ethylcarbazole were synthesized by reaction of 3‐amino‐9‐ethyl carbazole and the corresponding acyl chloride in the presence of triethylamine. It can be polymerized easily by using azo‐bisisobutyronitrile as an initiator or photopolymerized without any sensitizer. The photochemical behavior of 3‐methacrylamide‐9‐ethyl‐carbazole, its polymer and 3‐isobutyramide‐9‐ethylcarbazole were investigated by recording the fluorescence spectra in N,N‐dimethylformamide. It was found that the fluorescence intensity of the monomer is dramatically lower than those of its polymer and the model compound in the same chromophore concentration. This phenomenon, termed as the ‘structural self‐quenching effect’, was commonly observed for acrylic monomers bearing chromophore moieties and ascribed to the coexistence of the electron‐donating chromophore and the electron‐accepting double bond within one molecule. The strong fluorescence of the polymer can be quenched by adding electron‐deficient monomers having no chromophore moieties such as methyl methacrylate and acrylonitrile, and the Stern–Volmer constants were determined. It is observed that the higher the electron deficiencies of the quenchers, the higher the Stern–Volmer constants, implying a stronger quenching effect.Copyright © 2000 John Wiley & Sons, Ltd.  相似文献   

17.
Poly[N‐(9‐fluorenylmethoxycarbonyl)‐L‐phenylalanine] (PN9FPA) films with good fluorescence properties and chirality were prepared electrochemically by direct anodic oxidation of N‐(9‐fluorenylmethoxycarbonyl)‐L‐phenylalanine (N9FPA) in boron trifluoride diethyletherate (BFEE). Fourier transform infrared spectroscopy measurement showed that the polymerization of N9FPA occurred mainly at the C(2) and C(7) positions. The fluorescence spectra indicated that PN9FPA films were blue‐light emitters. In addition, the structures and properties of the monomer and the polymers were characterized and evaluated with CV, UV, TGA and SEM.  相似文献   

18.
Boiadjiev SE  Lightner DA 《Tetrahedron》2007,63(36):8962-8976
Yellow 9-methyldipyrrinones can be converted readily and in high yields to symmetric linear tetrapyrroles, blue biliverdinoids, which are cleaved in half, smoothly at room temperature to afford yellow 9-H dipyrrinones, and 9-CHO dipyrrinones as their violet to orange colored adducts with the carbon acid used for the scission: thiobarbituric acid (TBA), N,N′-diethylthiobarbituric acid, barbituric acid, N,N′-dimethylbarbituric acid, and Meldrum's acid. The adducts, usually only of passing interest, are formally Knövenagel condensation products of a 9-CHO dipyrrinone with TBA and other carbon acids of this work, and a reverse Knövenagel reaction of such adducts leads to 9-CHO dipyrrinones. Under a set of improved reaction conditions the sequence thus efficiently converts 9-CH3 dipyrrinones to 9-H and 9-CHO dipyrrinones.  相似文献   

19.
The twelve isotypic intermetallic compounds R2Ru3Ga9 with R = Y, La–Nd, Sm, Gd–Tm were prepared by arc‐melting of the elemental components. Their crystal structure was determined from single‐crystal X‐ray data of Dy2Ru3Ga9: Cmcm, a = 1279.3(2) pm, b = 755.6(1) pm, c = 964.7(1) pm, Z = 4, R = 0.020 for 671 structure factors and 42 variable parameters. All atomic positions have within two standard deviations ideal occupancies (occupancy values vary between 98.8(5) and 101.2(6)%). The structure is briefly discussed, emphasizing its relation to other structures with a high content of gallium or aluminum.  相似文献   

20.
The syntheses and crystal structures of two new ternary hafnium compounds, Cs[Hf2Br9] ( 1 ) and Rb[Hf2Br9] ( 2 ), are described. Both compounds are obtained in high yield from the chemical reaction of HfBr4 and CsBr or RbBr, respectively, in the presence of a small amount of elemental Al at 450 °C in sealed silica tubes. They crystallize isostructurally in the monoclinic space group P2/n. The lattice parameters are a = 9.946(1) ( 1 ) and 9.9388(4) Å ( 2 ), b = 6.6580(9) ( 1 ) and 6.6695(3) Å ( 2 ), c = 12.930(2) ( 1 ) and 12.8435(6) Å ( 2 ), and β = 112.479(6)° ( 1 ) and 112.726(2)° ( 2 ). The crystals of the two compounds contain dinuclear tri‐μ‐bromido‐hexabromido‐dihafnate(–) complex anions, [Hf2Br9], besides the alkali metal cations. The complex anions can be described as face‐sharing bioctahedral units.  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