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 15 毫秒
1.
Hexabenzo[a,cd,f,j,lm,o]perylene (HBP), an undecacyclic condensed polycyclic aromatic hydrocarbon with two severely crowded fjord regions, was synthesized by Clar's method; the starting material 8H-benzo[fg]-naphthacene-8-one was newly prepared by the glycerol condensation of 5,12-dihydro-naphthacene-5,12-dione. 1H and 13C NMR spectra of HBP were measured and the chemical shifts were completely assigned. Absorption and fluorescence spectra of HBP were also measured and the effects of nonplanarity on its spectroscopic properties was discussed.  相似文献   

2.
The assignment of nonexchanging protons of a small microcrystalline protein, the alpha-spectrin SH3 domain (7.2 kDa, 62 residues), was achieved by means of three-dimensional (3D) heteronuclear (1H-13C-13C) magic-angle spinning (MAS) NMR dipolar correlation spectroscopy. With the favorable combination of a high B(0)-field, a moderately high spinning frequency, and frequency-switched Lee-Goldburg irradiation applied during 1H evolution, a proton linewidth < or =0.5 ppm at 17.6 Tesla was achieved for the particular protein preparation used. A comparison of the solid-state 1H chemical shifts with the shifts found in solution shows a remarkable similarity, which reflects the identical protein structures in solution and in the solid. Significant differences between the MAS solid- and liquid-state 1H chemical shifts are only observed for residues that are located at the surface of the protein and that exhibit contacts between different SH3 molecules. In two cases, aromatic residues of neighboring SH3 molecules induce pronounced upfield ring-current shifts for protons in the contact area.  相似文献   

3.
1-Bromonaphthanthrone (1-bromo-6H-benzo[cd]pyren-6-one), a useful starting material for the synthesis of undecacyclic aromatic hydrocarbons, was synthesized by means of glycerol condensation of 1-bromopyrene. The structure of the synthetic product was confirmed by NMR measurements and the 1 H and 13 C chemical shifts were completely assigned. Although the synthetic method provided three possible isomers, 1-, 3-, and 5-bromonaphthanthrone, only 1-bromonaphthanthorone was obtained as a final product. An explanation for the regioselectivity was given in terms of the charge distribution obtained by molecular orbital calculations.  相似文献   

4.
A detailed characterization of two humic fractions was performed: One with low relative molecular mass (LMr < 3,500 Da) and one with high relative molecular mass (HMr > 3,500 Da). Distinct 1H NMR spectroscopic patterns were observed for the two fractions. HMr showed an aromatic proton region, an intense and broad region (3.0–5.0 ppm) attributed to sugar-like and polyether components, and an intense doublet at 1.33 ppm (identified as protons of the β-CH3 in lactate). In contrast, LMr did not show resonances due to aromatic protons and was characterized by a broad unresolved region, assigned to sugar-like components. The 13C NMR spectra showed that the LMr humic fraction was richer in carboxylic and aliphatic C groups compared to HMr fraction. These substances were fluorescein-labeled [fluorescein isothiocyanate (FITC)], and their interaction with carrot cells in culture was monitored for 10 d, and compared to FITC–indole-3-acetic acid (IAA) to clarify their mechanisms of biological activity. After different incubation times, fluorescein staining of carrot cells and decrease of fluorescein concentration in the culture medium were evaluated. Fluorescent membrane staining was only present in IAA and the LMr humic fraction treated cell cultures. A consequential decrease of fluorescein concentration in the culture media was also observed. Pretreatment of carrot cells with unconjugated IAA or LMr humic fraction markedly reduced fluorescein staining of both FITC–IAA and FITC–LMr humic fraction. Blocking tests gave indirect evidence of possible binding of the LMr humic fraction to IAA cell membrane receptors. These results indicate that the two humic fractions behave differently. Only LMr humic fraction, like IAA, interacts with cellular membranes in carrot cell cultures.  相似文献   

5.
We have synthesized undecacyclic aromatic hydrocarbons using condensation of dibenzoanthrones. For example, dibenzo[ a,o ]dinaphtho[3,2,1- cd ;1,2,3- lm ]perylene, diphenanthra[1,2,3- cd ; 3,2,1- lm ]perylene and diphenanthra[2,3,4- cd ;4,3,2- lm ]perylene were produced from 5,6-, 8,9- and 10,11-dibenzoanthrones, respectively. In the present study, we have extended our work and aimed to synthesize new undecacyclic aromatic hydrocarbons by the use of 9,10-dibenzoanthrone (7 H -benzo[ de ]naphthacene-7-one) as a starting material. 9,10-Dibenzoanthrone was prepared by the cyclization of 1,2'-dinaphthylketone in aluminum chloride anhydride and sodium chloride, and the structure was confirmed by NMR measurements. It dissolves into ethanol and crystallizes in pale yellow needles; its melting point is 198.2-199.3°. The condensation of 9,10-dibenzoanthrone should give undecacyclic aromatic hydrocarbons such as benzo[ vwx ]naphtha[12,1,2- cde ] hexaphene.  相似文献   

6.
Abstract

Substituent effects on the chemical shifts of aromatic carbons in lignin model compounds have been elucidated from 13C NMR spectra of guaiacyl and syringyl type monomerio and β-O-4 model compounds and guaiacyl type β-5 model compounds. Evaluation of the observed values of substituent chemical shift (SCS) for the aromatic carbons leads to elucidation of a generalized SCS additivity rule, for estimation of the chemical shifts of aromatic carbons in ring A of β-O-4 and β-5 type substructures in model compounds and in ring B of β-O-4 substructures in lignin preparations, with errors of less than 1 ppm. The rule is applicable to substructures of both guaiacyl and syringyl types, using an appropriate parent compound as reference instead of benzene. Signals in the aromatic region of the 13C NMR spectra of β-O-4 and β-5 type model compounds are reassigned on the basis of the observed SCS's as well as APT spectra of the compounds.  相似文献   

7.
The polycyclic aromatic ketone, 13H-dibenzo[a,de]anthracene-13-one (5,6-BBz) is useful as a starting material for synthesis of undecacyclic aromatic compounds by means of condensation. To synthesize 5,6-BBz, glycerol condensation of benzo[a]anthraquinone was conducted. The condensation, however, gave two structural isomers besides 5,6-BBz and the isolation was very difficult; separation of the crude products by column chromatography or high-vacuum sublimation was unsuccessful because of their similarity in structure and vapor pressure. Only a little 5,6-BBz was obtained by repeated recrystallization, but the amount was insufficient for condensation. Thus, we developed a new synthetic method which affords 5,6-BBz selectively. 9-(o-Chlorobenzoyl)anthracene was synthesized by the Friedel-Crafts' reaction of anthracene with o-chlorobenzoyl chloride and aluminum chloride anhydride. In order to suppress the production of 9,10-di(o-chlorobenzoyl) anthracene, the reaction was performed at low temperature. The crude products obtained were purified by column chromatography on activated alumina and then recrystallized with benzene, which yielded yellow crystals of 9-(o-chlorobenzoyl)anthracene. The structure of 9-(o-chlorobenzoyl)anthracene was determined by X-ray diffraction analysis for the first time. Cyclo-dehydrohalogenation of 9-(o-chlorobenzoyl) anthracene gave 5,6-BBz selectively, the amount of which was sufficient for synthesizing undecacyclic aromatic hydrocarbons.  相似文献   

8.
This study observed the activities of bromelain in the presence of various cationic surfactants at different temperatures and the conformational changes in bromelain by 1H NMR spectroscopy. We found that the bromelain activity was enhanced by tens to hundreds of micromoles per liter of the surfactant. In the presence of the surfactants, bromelain exhibited good tolerance to a range of substrate temperatures and its thermal stability was also increased. The 1H NMR experiments indicated that when the temperature was increased from 25.0 to 45.0 °C, the protons of bromelain having chemical shifts (δ) between 3.7 and 5.2 ppm moved upfield, while those having δ values between 3.2 and 3.7 ppm moved downfield. In the bromelain/cationic surfactant mixture, the values of δ for the protons in both bromelain and the cationic surfactants decreased, accompanied by the broadening of the half-peak width of the surfactant protons. These results indicated that both increasing temperature and adding a cationic surfactant made the bromelain chain more flexible and hence, increased the bromelain activity. To the best of our knowledge, this was the first time that the relationship between the protein activity and the 1H NMR data was expounded.  相似文献   

9.
Humic-like substances (HULIS) constitute a class of organic compounds identified in atmospheric samples that influence many properties of aerosols in the atmosphere. In this study, 6 HULIS samples were isolated from atmospheric total suspended particle (TSP) samples collected at 3 locations in Guangzhou of China, 1 each in summer and winter. On the basis of analyses by elemental analyzer, more detailed chemical species of elements were explored by proton nuclear magnetic resonance (1H NMR) spectroscopy for hydrogen and X-ray photoelectron spectroscopy (XPS) for carbon, oxygen, nitrogen, and sulfur. The sources of HULIS were identified by carbon isotopic techniques. The results show that HULIS made up an important component of water-soluble organic carbon (WSOC). Carbon and oxygen were the predominant components in the HULIS. The 1H NMR spectra indicated that H functional groups consisted of aliphatic C?H, aromatic C?H, H?C?C?, and H?C?O groups. They were all characterized by the highest contents of aliphatic protons and the lowest contents of aromatic protons. The XPS analysis of the HULIS fraction isolated from winter samples at the urban site Wushan (WS) showed that C, O, N, and S have various chemical states. The oxidized carbon groups accounted for 30.5% of total carbon. The nitrogen species contained amide, quaternary, and nitrate groups, and the sulfur species consisted of thiophenes, thioesters, mercaptanes, sulfones, and sulfates. The δ13C values of HULIS showed no obvious differences among the 6 HULIS samples. The δ15N values of summer HULIS samples in Maofengshan Mountain Forest Park (MFS) and Universities Town (UT) were significantly higher than those of winter samples. It is noteworthy that 14C data clearly indicate that even in urban areas biomass-derived carbon was the major component in the atmospheric HULIS samples.

Copyright 2012 American Association for Aerosol Research  相似文献   

10.
The degrees of swelling of rigid aromatic polyamide networks in various solvents, nonsolvents, and solvent mixtures were used to determine their solubility parameters. This was made possible because of the amorphous nature of the rigid networks and the low levels of intersegmental H-bonding. The solubility parameters are: δ = 23.0 (MPa)1/2, δd = 18.0 (MPa)1/2, δp = 11.9 (MPa)1/2, and δH = 7.9 (MPa)1/2. The determined values are believed to be good approximations of the solubility parameters of stiff linear aromatic polyamides, because of the essential identity of their structure and the stiff segments in the networks.  相似文献   

11.
A series of main chain phosphorus-containing copolyesters were synthesized by polycondensation of terephthalic acid (TPA), ethylene glycol (EG) and phenyl phosphonic acid (PPA). Chemical structures of these main chain phosphorus-containing copolyesters were characterized by 1H NMR. Experimental results show that the resonance intensity of PPA aromatic protons increases with increase of the phosphorus content. The chemical shifts of the ethylene protons in the ethylene glycol units vary with different sequences. The resonance chemical shift of the ethylene protons of the T-E-T unit is higher than those of P-E-T (T-E-P) and P-E-P units. The monomer molar fraction, sequential distribution and degree of randomness of the phosphorus-containing copolyesters were determined through analyses of the ethylene protons in the ethylene glycol units. The molar fractions of the PPA comonomer determined by 1H NMR analyses are close to the values determined by a UV method. The degree of randomness for the copolyesters was found to be in the range 0·66–0·83. © 1998 SCI.  相似文献   

12.
Ozonation reactions are very important in vegetable oil chemistry since their ozonation products are involved in antimicrobial effect in therapeutical uses for several microbiological etiology diseases. Information on the spectroscopic characterization of the products generated by ozonolysis of sunflower oil is limited. In the present study ozonized sunflower oil with 650 mmol-equiv/kg of peroxide index is chemically characterized. Ozonation of sunflower oil produced ozonides, aldehydes and hydroperoxides which were identified by 1H, 13C and two-dimensional 1H Nuclear Magnetic Resonance (NMR). The virgin sunflower oil and ozonized sunflower oil show very similar 1H NMR spectra except for the resonances at δ = 9.74 and δ = 9.63 ppm that correspond to both triplet from aldehydic protons, δ = 5.6 ppm (olefinic signal from hydroperoxides), and δ = 5.15 ppm (multiplet from ozonides methylic protons). Other resonance assignments are based on the connectivities provided by the proton scalar coupling constants. These are the following: δ = 3.15 ppm (doublet from methylenic group in α position respect to olefinic proton), δ = 2.45 ppm (multiplet from methylenic group allylic to ozonides methynic protons) and δ = 1.62 ppm (multiplet methylenic protons in β position respect to ozonides methynic protons). From the 13C NMR and 1H-13C two- dimensional spectrum of the ozonized sunflower oil, the presence of ozonides was confirmed by the signals δ = 103.43 and δ = 103.49 ppm, respectively. The others new signals found in δ = 42.5 and δ = 42.76 ppm confirm the presence of methylenic carbons from hydroperoxides and ozonides. These results indicate that NMR Spectroscopy can provide valuable information about the amount of reaction compounds of ozonized vegetable oil. From the chemical structural elucidation of ozonated sunflower oils, relevant biochemical and chemical information can be achieved.  相似文献   

13.
Computational chemistry is an important tool for signal assignment of 27Al nuclear magnetic resonance spectra in order to elucidate the species of aluminum(III) in aqueous solutions. The accuracy of the popular theoretical models for computing the 27Al chemical shifts was evaluated by comparing the calculated and experimental chemical shifts in more than one hundred aluminum(III) complexes. In order to differentiate the error due to the chemical shielding tensor calculation from that due to the inadequacy of the molecular geometry prediction, single-crystal X-ray diffraction determined structures were used to build the isolated molecule models for calculating the chemical shifts. The results were compared with those obtained using the calculated geometries at the B3LYP/6-31G(d) level. The isotropic chemical shielding constants computed at different levels have strong linear correlations even though the absolute values differ in tens of ppm. The root-mean-square difference between the experimental chemical shifts and the calculated values is approximately 5 ppm for the calculations based on the X-ray structures, but more than 10 ppm for the calculations based on the computed geometries. The result indicates that the popular theoretical models are adequate in calculating the chemical shifts while an accurate molecular geometry is more critical.  相似文献   

14.
Oxidation reactions of DL -selenomethionine (DL -SeMet) with hydrogen peroxide and ozone were investigated by 1H and 13C NMR spectroscopy. Chemical shifts of the reaction products were measured at different pD values in D2O solutions. In the moderate (4–8) pD range, a pair of singlets of equal intensity appeared at 2·71 and 2·80 ppm, with the disappearance of the DL -SeMet methyl peak at 2·04 ppm. As the pD of the solution decreased, the chemical shifts of both peaks increased relative to 3-trimethylsilylpropionic acid-d4 sodium salt. In strongly acidic solutions (pD<3), the pair collapsed to one singlet at 3·18 ppm. As pD values increased the chemical shifts decreased in numerical value with only a singlet at 2·69 ppm observed in strongly basic solutions (pD>10). These data were assigned to Met selenoxide in basic solution, stabilized by intermolecular hydrogen bonding, and to the hydration of Met selenoxide to Met dihydroxyselenide in acidic solutions. In moderate pD solutions (pD = 4–8), both forms of the selenoxide can exist. For the reaction of DL -SeMet with CNBr, two parallel reactions occur, i.e. oxidation and bond cleavage with cyanization: the bond cleavage giving the expected products, i.e. 2-amino-4-butyrolactone and methyl selenocyanate, with an additional singlet observed at 2·54 ppm. © 1988 SCI  相似文献   

15.
Summary Copolymers of carbazole-containing monomers such as 4- and 3-(9-carbazolylmethyl)styrene (1 and 2), 2(9-carbazolyl)ethyl methacrylate (3) and 2-(9-carbazolylacetyloxy)ethyl methacrylate (4) with optically active menthyl acrylate (5) and methacrylate (6) were prepared by free radical polymerization.In all cases optically active copolymers were obtained and for copolymer samples of 1 and 2 with 5 and 6, CD spectra indicate an appreciable, even if small, dissymmetric perturbation of heteroaromatic moiety.NMR and fluorescence emission spectra of the above copolymers are consistent with a monomer-type behaviour which can be associated with a high dishomogeneity of the conformational environment in which the aromatic chromophores are located.Profiles of the UV band in the 230–240nm region are analysed in terms of the different structural features of the copolymers. 1H-NMR spectra recorded on partially deuterated po1y(1) eliminate any contribution by the carbazolyl protons to the upfield signal (6.5–5.5ppm) of the complex aromatic protons resonance.Dedicated to Prof. C.I. Simionescu on the occasion of the 60th anniversary of his birthday  相似文献   

16.
Conformational Investigations by Means of 13C-N.M.R.-Spectroscopy. II. Conformational Preferences; Substituent Effects and Barriers to Ring Inversion in the Ketals of Methyl-substituted Cyclohexanones The comparison of the experimental determined 13C-NMR-chemical shifts of 8 ketal s with those, which were calculated by the well-known additivity rules allows some conclusions about the conformation of the compounds respectively about γ-gauche-effects and δ-syn-axial interactions in question. 13C-NMR-measurements at low temperatures make it possible to determine the barriers to ring inversion of some ketals. The obtained values are discussed in dependence of the substitution pattern.  相似文献   

17.
Positional isomers (1-butyryl-2X-3Y-rac-glycerol and 2-butyryl-1X-3Y-rac-glycerol;X,Y=long-chain acyls) of saturated triacylglycerols (TAG) with 34 and 40 acyl carbons were shown to separate in two chromatographic peaks on immobilized phenyl(65%) methylsilicone column by gas-liquid chromatography, and on reversed-phase ODS-1 column by high-performance liquid chromatography. The analysis of 500-MHz1H nuclear magnetic resonance (NMR) spectra showed distinct differences between 2-butyryl-1X-3Y-rac-glycerol and 1-butyryl-2X-3Y-rac-glycerol isomers in the resonance signals of methylene and methine protons of glycerol backbone, and carbon-2 methylene of acyl groups, and methyl protons of butyryl group. The1H NMR spectra of three interesterified mixtures of three monoacid TAG containing saturated butyrate and caproate TAG and unsaturated butyrate TAG showed that triplets of methyl protons of butyryl groups atsn-1(3)- andsn-2-positions in saturated and unsaturated TAG had similar chemical shifts and that the chemical shift of caproyl methyl protons was different from those of butyryl methyl protons. The positional distribution of butyryl groups in isolated positional isomers of butyrate TAG, interesterified TAG mixtures, and natural and interesterified butteroil can be determined by integration of these signals.  相似文献   

18.
Copolymers of acrylonitrile with furan, 2-methylfuran and with 2,5-dimethyl furan have been prepared by free radical copolymerizations, and their 13C spectra examined for evidence of the three types of structure that formed, according to previous studies. We have characterized the Diels-Alder adducts of acrylonitrile with furan and with 2-methylfuran, and have found that they may also copolymerize with acrylonitrile by addition at their double bonds. The enchained adducts were most readily recognized in the 1H spectrum by shifts at 4.8 and 4.6 ppm respectively from protons at the oxygen bridgehead. The 13C spectra of these various polymers have been assigned with the aid of model compounds. Below δ=60 ppm the nitrile shift was dispersed over a larger range of shifts than is found in PAN itself, and was joined by the shifts of unsaturated carbon atoms and of carbon atoms linked to the oxygen atoms within these structures. Above 60 ppm the shifts of carbons in acrylonitrile units were found to be perturbed systematically from their normal places by the presence of neighbouring structures formed from the different furan-containing molecules. In both the 1H and 13C spectra shifts characteristic of the end groups from the AIBN initiator were noted. Fine structure in the spectra was attributed to monomer sequence effects and to the tactic structure of the polymer. Chiral centres appeared to be created from the furan rings at random, at least when protonated carbons were the site of initial attachment, but the mode of addition to the ring—whether cis, trans or both, was not certain. Assignments were made of bands in the i.r. spectra to modes of vibration of the units formed from the furan monomers.  相似文献   

19.
Knothe G 《Lipids》2006,41(4):393-396
Cyclopropane FA occur in nature in the phosphoplipids of bacterial membranes, in oils containing cyclopropene FA, and in Litchi sinensis oil. Dihydrosterculic acid (2-octyl cyclopropaneoctanoic acid) and its methyl ester were selected for 1H and 13C NMR analysis as compounds representative of cyclopropane FA. The 500 MHz 1H NMR spectra acquired with CDCl3 as solvent show two individual peaks at −0.30 and 0.60 ppm for the methylene protons of the cyclopropane ring. Assignments were made with the aid of 2D correlations. In accordance with previous literature, the upfield signal is assigned to the cis proton and the downfield signal to the trans proton. This signal of the trans proton is resolved from the peak of the two methine protons of the cyclopropane ring, which is located at 0.68 ppm. The four protons attached to the two methylene carbons α to the cyclopropane ring also show a split signal. Two of these protons, one from each methylene moiety, display a distinct shift at 1.17 ppm, and the signal of the other two protons is observed at 1.40 ppm, within the broad methylene peak. The characteristic peaks in the 13C spectra are also assigned.  相似文献   

20.
Mono‐ or di(meth)acrylate‐terminated derivatives of diglycidyl hexahydrophthalate (ER) were prepared by reacting 1 : 1 or 1 : 2M ratio of ER and methacrylic acid or acrylic acid. These vinyl ester (VE) resins were characterized by determining epoxy equivalent weight, acid number, and molecular weight by gel permeation chromatography. Structural characterization was done by FTIR and 1H NMR spectroscopy. In the 1H NMR spectra of acrylate‐terminated VE resins, three proton resonance signals were observed in the region 5.8–6.4 ppm due to vinyl group while in methacrylate‐terminated VE resins only two proton resonance signals due to vinylidene protons were observed at 5.6–6.1 ppm. The Brookfield viscosity (room temperature (25 ± 2)°C) of these resins diluted with varying amounts of MMA was determined at 20 rpm. Curing behavior was monitored by determination of gel time and differential scanning calorimetry. An exothermic transition was observed in the DSC scans in the temperature range of (81–150)°C. Isothermal curing of MMA‐diluted VE resins containing AIBN as an initiator was done at 60°C for 2 h in N2 atmosphere, and then heating for another 2 h in static air atmosphere. Thermal stability of isothermally cured resins in N2 atmosphere was evaluated by thermogravimetric analysis. All cured resins decomposed above 310°C in single step. Thermal stability of the cured resins having acrylate end caps was marginally higher than the resins having methacrylate end groups. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci, 2006  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