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 46 毫秒
1.
So far, “proton sponges” have been defined as bis(dialkylamino)arenes whose dialkylamino groups are in close spatial proximity.[1] The unusual basicity of these compounds is ascribed to the destabilizing overlap of the lone electron pairs on the nitrogen atoms, to the formation of especially strong hydrogen bonds in the monoprotonated diamines, and to the hydrophobic shielding of these hydrogen bonds. In order to differentiate and assess the relative importance of these factors, we were interested in quino[7,8-h]quinoline 1 , whose nitrogen atoms exhibit a mutual orientation similar to that in 1,8-bis(dimethylamino)naphthalene 2 (“proton sponge”). In contrast to 2 , however, 1 lacks the hydrophobic shielding of the hydrogen bonds of its monoprotonated derivative. This shielding is considered to be responsible for the low rates of proton transfer, which make the “proton sponges” reported so far unsuitable as auxiliary bases in chemical reactions.  相似文献   

2.
Because of their extreme instability, it is generally difficult to synthesize and fully characterize open chain peroxides, also known as peroxols. In our attempt to investigate the mechanism of the Skraup–Doebner–Von Miller quinoline synthesis, we were able to obtain an unusual open chain peroxy‐quinoline, namely, 4‐(8‐ethoxy‐2,3‐dihydro‐1H‐cyclopenta[c]quinolin‐4‐yl)butane‐1‐peroxol (1), and its alcohol counterpart, namely 4‐(8‐ethoxy‐2,3‐dihydro‐1H‐cyclopenta[c]quinolin‐4‐yl)butan‐1‐ol (2) obtained as a side product during the same reaction. Although structurally similar, these two compounds appeared to display some very distinct physical and spectroscopic characteristics. This work reports detailed NMR studies and full 1H and 13 C NMR assignments for these two compounds. These assignments are based upon the analysis of the NMR spectra of these compounds including 1H, 13 C, COSY, gHSQC and gHMBC. The effect of the peroxide functional group on the chemical shift of neighboring carbons and protons was also investigated by comparing the NMR data of these two compounds. Furthermore, the effects of potential hydrogen bondings in 1, 2, and possible 1–1 dimer, 2–2 dimer and in prototypical model systems, as well as the stability of these compounds, were investigated computationally. The computed dissociation energies and NMR data support the interpretation of the experimental data. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

3.
A new 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS)-radical scavenging and antiproliferative agents of pyrrolo[1,2-a]quinoline derivatives have been synthesized. An efficient method for the synthesis of 14 novel diversified pyrrolo[1,2-a]quinoline derivatives has been described using 4-(1,3-dioxolan-2-yl)quinoline and different phenacyl bromides in acetone and followed by reacting with different acetylenes in dimethylformamide/K2CO3. The structure of the newly synthesized compounds was determined by infrared, 1H NMR, 13C NMR, mass spectrometry, and elemental analysis. The in vitro antioxidant activity revealed that among all the tested compounds 5n exhibited maximum scavenging activity with ABTS. Compound 5b has showed good antiproliferative activity as an inhibitor of epidermal growth factor receptor (EGFR) tyrosine kinase.  相似文献   

4.
The classical (Johnson-Bovey) and quantum mechanical (Haigh-Mallion) approach accounting for the shielding contribution to protons in the proximity of aromatic nuclei are compared with respect to their applicability in conformational analysis. This is accomplished by means of the1H-NMR spectra of various cyclophanes their geometries being fairly well known from X-ray data or other independent methods. The chemical shifts of the monitor protons (all of them being situated above or below the plane of the benzene ring considered) are consistent with shielding effects predicted by the classical theory notwithstanding the fact that the benzene hexagons are not strictly planar in any of these systems. In contrast, the quantum mechanical treatment considerably underestimates the shielding of protons in those regions.

Mit 2 Abbildungen  相似文献   

5.
The synthesis of the two remaining isomeric monothieno-analogues of phenanthridine N-oxide, thieno-[3,2-c]quinoline N-oxide and thieno[3,2-c]isoquinoline N-oxide, is described. The 1H and 13C nmr spectra of all six isomeric thieno-fused quinoline and isoquinoline N-oxides are discussed.  相似文献   

6.
Excess volumes of V E of binary liquid mixtures of quinoline with each of the nitroalkanes have been measured at 30°C as a function of composition using a continuous dilution dilatometer. The excess volumes are negative over the whole mole-fraction range for all the mixtures except quinoline + nitromethane, which exhibits positive V E over the whole range. V E results have been analyzed in the light of the Prigogine-Flory-Patterson theory, which divides V E into three contributions. The agreement between experimental and theoretical values is reasonably good.  相似文献   

7.
The 1H nmr spectra of phenanthro[9′,10′:4,5]thieno[2,3-c]quinoline, benzo[f]phenanthro-[9′,10′:4,5]thieno[2,3-c]quinoline and benzo[h]phenanthro[9′,10′:4,5]thieno[2,3-c]quinoline are highly congested. For each compound, all protons abide in an aromatic environment complicated by pseudo-symmetric regions which result in multiple overlap of the different spin systems these molecules contain. We illustrate here the utility of the HMQC-TOCSY experiment to identify spin systems when the proton spectrum is highly congested. To complete the assignment of the 1H and 13C nmr spectra of each compound the HMBC experiment is used to assign the quaternary carbons.  相似文献   

8.
In the two title optically active tetra­hydro­iso­quinoline derivatives, namely 3‐hydroxy­methyl‐4‐phenyl‐1,2,3,4‐tetra­hydro­isoquinolin‐2‐ium bromide methanol hemisolvate, C16H18NO+·Br?·0.5CH3OH, (IIb), and 2‐formyl‐3‐hydroxy­methyl‐4‐phenyl‐1,2,3,4‐tetra­hydro­iso­quinoline, C17H17NO2, (III), the absolute configurations have been confirmed as 3R,4R by structure refinement using Bijvoet‐pair reflections. The hydroxy­methyl and phenyl groups in (IIb) are oriented in equatorial and pseudo‐equatorial positions, respectively, whereas in (III), the corresponding groups are in axial and pseudo‐axial positions, respectively; the hydroxy­methyl and phenyl groups are trans with respect to one another in both structures. The heterocyclic rings in (IIb) and (III) adopt envelope conformations inverted with respect to each other. In both structures, the mol­ecules are linked through hydrogen bonds.  相似文献   

9.
Hydrogenation of 10H-indeno[1,2-b]quinoline in benzene in the presence of R2S7 gave 5,5a,10a,11-tetrahydro-10H-indeno[1,2-b]quinoline, the structure of which was established by mass-, IR, UV,13C, and1H NMR spectra. Thecis fusion of the indan and tetrahydroquinoline fragments, the axial orientation of the proton at C(5a), and the equatorial orientation of the proton at C(10a) were confirmed by molecular mechanics calculations using the PC MODEL program.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1098–1101, June, 1994.  相似文献   

10.
A series of novel and synthetically important quinoline selenium compounds have been successfully synthesized using an efficient and simple strategy. The method employed leads to the synthesis of both cyclic as well as open chain quinoline selenium compounds. The prepared selenium compounds have been characterized with the help of various spectroscopic techniques viz., NMR (1H, 13C), FT-IR, mass spectrometry. The structure of 6-methoxy-3H-[1,2]diselenolo[3,4-b]quinoline has been determined by X-ray crystallography.  相似文献   

11.
The syntheses of the K-imine derivatives of benzo[h]quinoline ( 1 ), benzo[f]quinoline ( 2 ) and 1,10-phenanthroline ( 3 ) are described. The parent nitrogen heterocycles were oxidized with sodium hypochlorite to the corresponding K-oxides, 4, 6 and 8 , which in turn were reacted with sodium azide. The resulting azido alcohols were then cyclized with triethyl phosphite to the title compounds 5, 7 and 9 . The oxirane ring cleavage in benzo[h]quinoline 5,6-oxide ( 4 ) and in benzo[f]quinoline 5,6-oxide ( 6 ) by sodium azide proceeded by the predicted regioselectivity: 4 gave trans-5-azido-5,6-dihydro-6-benzo[h]quinolinol ( 11 ) and trans-6-azido-5,6-dihydro-5-benzo[h]quinolinol ( 10 ) as the major and minor products respectively, and 6 yielded solely trans-6-azido-5,6-dihydro-5-benzo[f]quinolinol ( 12 ). The latter compound proved by X-ray analysis to crystallize as a hydrogen bonded dimer.  相似文献   

12.
Functionalization of quinoline aldehydes, derived from nimesulide framework was carried out using Morita–Baylis–Hillman (MBH) chemistry. A number of novel quinoline‐based diverse MBH adducts was prepared via the reaction of derivatives of 2‐chloroquinoline‐3‐carbaldehyde and various activated alkenes in good yields. Many of these compounds were found to be potent when tested against human prostate cancer (Pc‐3) cell line in vitro. Among all the compounds tested N‐(2‐chloro‐3‐(2‐cyano‐1‐hydroxyallyl)‐7‐phenoxyquinolin‐6‐yl)formamide (IC50 = 1.2 μg mL?1) was identified as the most potent compound in this series. J. Heterocyclic Chem., (2012).  相似文献   

13.
Excess volumes V E of binary liquid mixtures of quinoline with alkanols have been determined from densities at 30°C as a function of composition. The excess volumes are negative over the whole mole fraction range for all the mixtures and decrease with increasing length of alkanol (C1–C10). The VE data have been analyzed in terms of an approach which uses graph theoretical connectivity parameters of the third degrees for two components. The analysis gives information regarding associated species in the pure state and in the mixture. It is suggested that, in the mixture state, no change occurs in the association of alkanols.  相似文献   

14.
A novel Schiff base ligand (H‐DPPMHQ) derived from 2‐hydrazineylquinoline and 1,3‐diphenyl‐1H‐pyrazole‐5‐carbaldehyde and its dimeric complexes with compositions [Cr(DPPMHQ)Cl]2?2Cl and [M(DPPMHQ)Cl]2 (where M = Cu(II), Co(II), Ni(II) and Zn(II)) have been synthesized and characterized using physicochemical methods like elemental analysis, magnetic susceptibility and molar conductivity measurements, multispectral techniques and electrochemical studies. The molar conductance data reveal that all metal chelates are non‐electrolytes, except the Cr(III) complex which shows a ΛM value of 146.82 Ω?1 cm2 mol?1, indicating that it is a 1:2 electrolyte. Infrared spectral results show that the metal is organized through four nitrogen atoms (azomethine and deprotonated imine groups, pyrazole and quinoline rings) besides chlorine atoms. The NH proton is also displaced during complexation, as indicated by 1H NMR spectral data. Based on the electron spin resonance and ligand field parameter data, the bonding parameters of these complexes have been calculated. Using Coats–Redfern and Horowitz–Metzger equations, thermodynamic parameters were determined. The spectral data indicate that the dimeric complexes have octahedral geometry around the central metal ions. The cytotoxic activities of all compounds were evaluated towards human breast cancer (MCF‐7) and lung cancer (A549) cell lines.  相似文献   

15.
The 1H- and 13C-nmr spectra of [1]benzothieno[2,3-c]naphtho[1,2-h]quinoline and [1]benzothieno[2,3-c]-naphtho[1,2-h][1,2,4]triazolo[4,3-a]quinoline were totally assigned using a combination of two-dimensional nmr techniques. After correlation of the proton signals by a COSY spectrum and one-bond heteronuclear correlation, complete assignment of the 1H- and 13C-nmr spectra of the novel heterocyclic compounds required the application of long-range CH coupling information particularly for quarternary resonance assignments and the orientations of individual spin systems relative to one another.  相似文献   

16.
Novel coated graphite electrode (CGE) for the determination of quinoline yellow (QY), an artificial food colorant, was evaluated based on the use of tetraphenylphosponium-QY ion-pair complex as the electroactive substance. The electrode revealed linear emf-pQY response over wide concentration ranges (1.0 × 10?1 ~ 5.0 × 10?5 M) with a slope ?30.1 ± 0.2 mV decade?1 and a limit of detection of 4.0 × 10?5 M. The electrode showed good stability, reproducibility and fast response (<5 s). It can be used in wide pH range (4.7–10.7) and was successfully applied to determination of quinoline yellow in artificial mixtures and commercial soft drinks. The results obtained with the electrode were in good agreement with the value obtained by using HPLC measurements, as an official method.  相似文献   

17.
In the closely related quinoline compounds 8‐nitro‐2‐(trichloromethyl)quinoline, (I), 6‐nitro‐2‐(trichloromethyl)quinoline, (II), and 5‐nitro‐2‐(trichloromethyl)quinoline, (III), all C10H5Cl3N2O2, which are of both reactivity and pharmacological interest, and for which the biological activity and cytotoxicity appear to be based on the positions of the CCl3 and nitro substituents, the nitro group is only coplanar with its aromatic substrate in (II). The deviation of the nitro group from coplanarity is concluded to be a function of both its position with respect to the trichloromethyl group and the intermolecular contacts in which it participates. The discrepancies between the crystal structures and the molecular shapes predicted by ab initio calculations are also explained in these terms. The quinoline ring is not rigorously planar in any of the structures, which may be explained by stress produced by the CCl3 substituent.  相似文献   

18.
Several new poly(pyridinium salt)s with quinoline diamine moieties in their backbones with tosylate and triflimide counterion were prepared by either a ring‐transmutation polymerization reaction with bis(pyrylium tosylate) with a series of isomeric quinoline diamines in dimethyl sulfoxide (DMSO) for 48 h at 130–140 °C or a metathesis reaction of the tosylate polymers with lithium triflimide in DMSO at about 60 °C. Their chemical structures were confirmed by FTIR, 1H and 13C NMR spectroscopy, and elemental analysis. Their number‐average molecular weights (Mn) were in the range of 18,000–58,000, and their polydispersities were in the range of 1.12–1.53 as determined by gel permeation chromatography. They had thermal stability in the temperature range of 353–455 °C and glass‐transition temperatures >240 °C. They had good solubility in common organic solvents and were characterized with polarizing optical microscopy (POM) studies for their lyotropic liquid‐crystalline properties. Their light‐emission properties were examined by spectrofluorometry in both the solution and film states. Their quantum yields were also determined. Additionally, their morphologies in the thin‐film states and melt‐drawn fibers were examined with POM, scanning electron microscopy, and transmission electron microscopy techniques. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

19.
Under electron impact, the molecular ions of quinoline N-oxide, carbostyril and 8-hydroxyquinoline lose carbon monoxide giving a fragment ion C8H7N (m/z 117), which was shown by collision-activated dissociation in each case to have the structure of the molecular ion of indole. Its formation from 8-hydroxyquinoline requires an unusual rearrangement. Isoquinoline N-oxide loses HCN rather than CO and gives a fragment which has the structure of the molecular ion of benzofuran. When the first three compounds were subjected to flash vacuum pyrolysis, quinoline N-oxide at 500–700°C gave carbostyril and indole was detected by gas chromatography/mass Spectrometry. At 900°C carbostyril and 8-hydroxyquinoline both gave indole in small amounts, detected by gas chromatography/mass Spectrometry.  相似文献   

20.
The UV absorption spectra of imidazo[4, 5-f]quinoline and its quaternary salts were examined. The causes of the development of luminescence in solutions of the quaternary salts of imidazo[4, 5-f]quinoline were elucidated.For part I, see [6].  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