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 181 毫秒
1.
Although the medicinal plant and food Nigella glandulifera Freyn has been researched for decades, isobenzofuranones have never been isolated before. Two isobenzofuranone derivatives and two saponins were successfully separated and purified from seeds of N. glandulifera Freyn by high-speed counter-current chromatography (HSCCC) with the optimized two-phase solvent system, n-hexane-ethyl acetate–methanol–water (7:3:5:5, v/v). Salfredin B11 (22.1 mg, HPLC purity 95.3%), 5, 7-dihydroxy-6-(3-methybut-2-enyl) isobenzofuran-1(3H)-one (18.9 mg, HPLC purity 97.3%) and crude sample 2 (555 mg) were separated from 600 mg of ethyl acetate extract of N. glandulifera Freyn. Following a cleaning-up step by chromatography on Sephadex LH-20, hederagenin (12 mg) and 3-O-[β-d-xylopyranosyl-(1 → 3)-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl]-hederagenin (45 mg) were separated from sample 2. All of the fractions before peak II were collected and subjected to a Sephadex LH-20 column and eluted by methanol, two of triterpene saponins (12 mg of hederagenin and 45 mg of 3-O-[β-d-xylopyranosyl-(1 → 3)-α-l-rhamnopyranosyl-(1 → 2)-α-l-arabinopyranosyl]-hederagenin) were isolated. The structures of peak fractions were identified by IR, electron ionization MS, 1H NMR and 13C NMR. 5, 7-Dihydroxy-6-(3-methybut-2-enyl) isobenzofuran-1(3H)-one was isolated for the first time from higher plant and salfredin B11 was isolated for the first time in this plant.  相似文献   

2.
The two major steroidal saponins from the roots of Asparagus racemosus were isolated by RP-HPLC and their structure determined by extensive NMR studies. Their structures did not match those reported previously for shatavarins I and IV and were found to be 3-O-{[β-d-glucopyranosyl(1→2)][α-l-rhamnopyranosyl(1→4)]-β-d-glucopyranosyl}-26-O-(β-d-glucopyranosyl)-(25S)-5β-furostan-3β,22α,26-triol and 3-O-{[β-d-glucopyranosyl(1→2)][α-l-rhamnopyranosyl(1→4)]-β-d-glucopyranosyl}-(25S)-5β-spirostan-3β-ol.  相似文献   

3.
The proanthocyanidin extract from tea (Camellia sinensis) leaves was purified for the further study of the biological role of proanthocyanidins in blister blight leaf disease of tea, which is caused by the fungus Exobasidium vexans. An aqueous acetone extract of proanthocyanidins prepared from healthy tea leaves was partially purified using Sephadex LH-20 chromatography. The crude proanthocyanidin extract obtained was fractionated with high-speed counter-current chromatography (HSCCC) using the solvent system n-hexane–EtOAc–MeOH–water (1:5:1:5). The purity of the each isolated fraction after a single HSCCC run was evaluated by high-performance liquid chromatography (HPLC). Seven fractions of high purity were isolated. The identity of the compound present in each fraction isolated was established using electrospray ionization mass spectrometry (ESI-MS) and nuclear magnetic resonance (NMR) spectroscopy. Five proanthocyanidins and two flavanol digallates, (−)-epigallocatechin digallate (EGCDG) and (−)-epicatechin digallate (ECDG) were isolated. Comparison of spectral data of the proanthocyanidins isolated with those previously reported indicated that all five were known B-type proanthocyanidins with 2,3-cis stereochemistry in both the upper (u-unit) and the terminal (t-unit) units, and 4R configuration of the C-ring in the u-unit. The proanthocyanidins were established to be dimers composed of (−)-epigallocatechin gallate (EGCG), (−)-epicatechin gallate (ECG) and (−)-epiafzelechin gallate (EAG) units with the following structures: EGCG-(4β → 6)-EGCG, ECG-(4β → 6)-EGCG, EGCG-(4β → 6)-ECG, EAG-(4β → 6)-EGCG, ECG-(4β → 6)-ECG by analysis of spectral data. Therefore HSCCC offers a powerful method for the separation of a group of closely related naturally occurring compounds.  相似文献   

4.
A new steroidal saponin, shatavarin V, (3-O-{[α-l-rhamnopyranosyl(1→2)][β-d-glucopyranosyl(1→4)]-β-d-glucopyranosyl}-(25S)-5β-spirostan-3β-ol), was isolated from the roots of Asparagus racemosus by RP-HPLC, and its structure determined by 1D and 2D NMR studies. This data permits clarification of the structures reported for several known saponins: asparinins A and B; asparosides A and B; curillin H; curillosides G and H and shavatarins I and IV.  相似文献   

5.
A general method has been developed for the synthesis of oligosaccharides consisting of (1→2)- and (1→3)-linked rhamnans with GlcNAc side chains. As examples, highly effective and convergent syntheses of two decasaccharides in the O polysaccharide moiety of the lipopolysaccharide of the phytopathogenic bacterium Pseudomonas syringae pv. ribicola NCPPB 1010 were achieved. The two decasaccharides consist of O polysaccharide repeating units I+II and II+I, respectively. Allyl 3-O-acetyl-4-O-benzoyl-α-l-rhamnopyranoside, allyl 2-O-benzoyl-3-O-chloroacetyl-α-l-rhamnopyranoside, 2,4-di-O-benzoyl-3-O-chloroacetyl-α-l-rhamnopyranosyl trichloroacetimidate, and 3-O-acetyl-2,4-di-O-benzoyl-α-l-rhamnopyranosyl trichloroacetimidate, which were obtained by highly regioselective 3-O-acylations, were used as the key synthons to obtain the required α-(1→2)- and α-(1→3)-linked rhamnoocta saccharide acceptors with 33- and 37-free hydroxyl groups. Therefore, several disaccharides were synthesized, from which tetrasaccharides and hexasaccharides were then synthesized. Coupling of the hexasaccharide donors with the disaccharide acceptors gave the octasaccharide acceptors. Finally, the coupling of 3,4,6-tri-O-acetyl-2-deoxy-2-phthalimido-β-d-glucopyranosyl trichloroacetimidate with the octasaccharide acceptors, followed by deprotection, afforded the two target decasaccharides. A repeating hexasaccharide unit of the cell wall polysaccharide of β-hemolytic Streptococci Group A was also synthesized in a similar way.  相似文献   

6.
A detailed study of iron (III)–citrate speciation in aqueous solution (θ = 25 °C, Ic = 0.7 mol L−1) was carried out by voltammetric and UV–vis spectrophotometric measurements and the obtained data were used for reconciled characterization of iron (III)–citrate complexes. Four different redox processes were registered in the voltammograms: at 0.1 V (pH = 5.5) which corresponded to the reduction of iron(III)–monocitrate species (Fe:cit = 1:1), at about −0.1 V (pH = 5.5) that was related to the reduction of FeL25−, FeL2H4− and FeL2H23− complexes, at −0.28 V (pH = 5.5) which corresponded to the reduction of polynuclear iron(III)–citrate complex(es), and at −0.4 V (pH = 7.5) which was probably a consequence of Fe(cit)2(OH)x species reduction. Reversible redox process at −0.1 V allowed for the determination of iron(III)–citrate species and their stability constants by analyzing Ep vs. pH and Ep vs. [L4−] dependence. The UV–vis spectra recorded at varied pH revealed four different spectrally active species: FeLH (log β = 25.69), FeL2H23− (log β = 48.06), FeL2H4− (log β = 44.60), and FeL25− (log β = 38.85). The stability constants obtained by spectrophotometry were in agreement with those determined electrochemically. The UV–vis spectra recorded at various citrate concentrations (pH = 2.0) supported the results of spectrophotometric–potentiometric titration.  相似文献   

7.
Eight chromium(III) complexes of tetradentate Schiff bases have been prepared in situ by condensing of a substituted salicylaldehyde compound with ethylenediamine. These were characterized by elemental analysis, m.p., IR, molar conductivity, magnetic moment measurements, and electronic spectra. The free ligands were also characterized by 1H and 13C NMR spectra. The 13C NMR spectra are discussed in terms of possible substituent effects. The IR and electronic spectra of the free ligand and the complexes are compared and discussed. The electrospray ionization (ESI) mass spectra of four free ligands and their complexes were measured. The deconvolution of the visible spectra of the complexes, C2v symmetry, in DMSO yields three peaks at ca. 15 600–17 600, 18 400–20 400 and 20 000–23 100, and are assigned to the three d–d transitions, 4B1g → 4Eg(4T2g); 4B1g → 4B2g(4T2g); 4B1g → 4Eg(4T1g), respectively. The complexes showed magnetic moment in the range of 3.5–4.2 BM which corresponds to three unpaired electrons.  相似文献   

8.
The upconversion luminescent properties of YF3:Yb3+(20%)/Tm3+(1%) nanobundles with different sizes (240-500 nm in length) were studied under 980-nm excitation. Ultraviolet (1I6 → 3F4/3H6 and 1D2 → 3H6), blue (1D2 → 3F4 and 1G4 → 3H6), red (1D2 → 3H4, 1G4 → 3F4, and 3F3 → 3H6), and near infrared (3H4 → 3H6) emissions were observed. The results indicated that the relative intensity of the ultraviolet to the blue as well as the blue to the near infrared increased with decreasing the size of nanobundles. Especially, the position of the dominant red emission peak varied with the size of nanobundles. As the length of nanobundles increased to 500 nm, unusual 3F3 → 3H6 transition was observed, which was theoretically explained considering the decrease of the nonradiative transition rate of 3F3 → 3H4.  相似文献   

9.
The nucleophilic substitution reaction of S2O32− with [Ru(HaaiR′)2(OH2)2](ClO4)2 (1) [HaaiR′ = 1-alkyl-2-(phenylazo)imidazole] and [Ru(ClaaiR′)2(OH2)2](ClO4)2 (2) [ClaaiR′ = 1-alkyl-2-(chlorophenylazo)imidazole] [where R′ = Me(a), Et(b) or Bz(c)] in acetonitrile–water (50% v/v) medium to yield Na2[Ru(HaaiR′)2(S2O3)2] (3a, 3b or 3c) and Na2[Ru(ClaaiR′)2(S2O3)2] (4a, 4b or 4c) has been studied. The products were characterized by microanalytical data and spectroscopic techniques (UV–Vis, NMR and mass spectroscopy). The reaction proceeds in two consecutive steps (A → B → C); each step follows first order kinetics with respect to each complex and S2O32−, and the first step second order rate constant (k2) is greater than the second step one (k2). An increase in the π-acidity of the ligand increases the rate. Thermodynamic parameters, the standard enthalpy of activation (ΔH0) and the standard entropy of activation (ΔS0), have been calculated for both steps using the Eyring equation from variable temperature kinetic studies. The low ΔH0 and large negative ΔS0 values indicate an associative mode of activation for both aqua ligand substitution processes.  相似文献   

10.
New bimetallic complex [Cp2ZrH2 · ClAlEt2]2 (1) was synthesized, and its reactivity in hydrometallation reaction with the following alkenes was studied: hept-1-ene, okt-1-ene, α-methylstyrene, (1S)-β-pinene, (+)-camphene. Complex 1 shows the highest reactivity among the other known Al,Zr-bimetallic complexes: [Cp2ZrH2 · ClAlBui2]2 (2), [Cp2ZrH2 · AlEt3]2 (3), [Cp2ZrH2 · AlBui3]2 (4) and [Cp2ZrH2 · HAlBui2] (5) as well as organoaluminium compounds (OAC): iBu2AlH, iBu3Al and iBu2AlCl in presence of Zr catalysts. Chlorine containing complexes 1 and 2 appear to be more effective in alkene hydrometallation, and relative hydrometallation rates are (1S)-β-pinene ? (+)-camphene < α-methylstyrene < oct-1-ene < hept-1-ene. Hydrometallation of (1S)-β-pinene and its subsequent oxidation with I2 run with high diastereoselectivity and yield trans-myrtanol. However, the diastereoselectivity of (+)-camphene hydrometallation is less than that for (1S)-β-pinene, and the reaction gives predominately endo-camphanol.  相似文献   

11.
A method of using high-speed counter-current chromatography (HSCCC) was established for preparative isolation and purification of antimycin A components from antimycin fermentation broth. Six antimycin A components were successfully purified for the first time by HSCCC with a two-phase solvent system composed of n-hexane–ethyl acetate–methanol–water (5:2:4:1, by volume). Total of 20 mg antimycin A4(a or b), 25 mg antimycin A3(a or b), 21 mg antimycin A8(a or b), 34 mg antimycin A2(a or b), 26 mg antimycin A1(a or b) and 34 mg antimycin A1(a or b) with the purities of 93.2, 98.6, 96.2, 94.1, 94.9 and 96.7%, respectively, determined by high-performance liquid chromatography (HPLC), were yielded from 200 mg crude sample only in one HSCCC run.  相似文献   

12.
The reaction pathways and energetics for the reaction of methane with CaO are discussed on the singlet spin state potential energy surface at the B3LYP/6-311+G(2df,2p) and QCISD/6-311++G(3df,3pd)//B3LYP/6-311+G(2df,2p) levels of theory. The reaction of methane with CaO is proposed to proceed in the following reaction pathways: CaO + CH4 → CaOCH4 → [TS] → CaOH + CH3, CaO + CH4 → OCaCH4 → [TS] → HOCaCH3 → CaOH + CH3 or [TS] → CaCH3OH → Ca + CH3OH, and OCaCH4 → [TS] → HCaOCH3 → CaOCH3 + H or [TS] → CaCH3OH → Ca + CH3OH. The gas-phase methane–methanol conversion by CaO is suggested to proceed via two kinds of important reaction intermediates, HOCaCH3 and HCaOCH3, and the reaction pathway via the hydroxy intermediate (HOCaCH3) is energetically more favorable than the other one via the methoxy intermediate (HCaOCH3). The hydroxy intermediate HOCaCH3 is predicted to be the energetically most preferred configuration in the reaction of CaO + CH4. Meanwhile, these three product channels (CaOH + CH3, CaOCH3 + H and Ca + CH3OH) are expected to compete with each other, and the formation of methyl radical is the most preferable pathway energetically. On the other hand, the intermediates HCaOCH3 and HOCaCH3 are predicted to be the energetically preferred configuration in the reaction of Ca + CH3OH, which is precisely the reverse reaction of methane hydroxylation.  相似文献   

13.
Tm3+/Yb3+ codoped rod-like YF3 nanocrystals were synthesized through a facile hydrothermal method. After annealing in an argon atmosphere, the nanocrystals emitted bright blue and intense ultraviolet (UV) light under a 980-nm continuous wave diode laser excitation. Up-conversion emissions centered at ∼291 nm (1I6 → 3H6), ∼347 nm (1I6 → 3F4), ∼362 nm (1D2 → 3H6), ∼452 nm (1D2 → 3F4), ∼476 nm (1G4 → 3H6), ∼642 nm (1G4 → 3F4), and ∼805 nm (3H4 → 3H6) were recorded using a fluorescence spectrophotometer. Especially, enhanced UV emissions were studied by changing Yb3+/Tm3+ doping concentrations, the annealing temperatures, and the excitation power densities. A possible mechanism, energy transfer-cross relaxation-energy transfer (ET-CR-ET), was proposed based on a simple rate-equation model to elucidate the process of the enhanced UV emissions.  相似文献   

14.
The complexation of native β-cyclodextrin (CD) and seven aromatic compounds, namely, phenetole, toluene, m-xylene, naphthalene, biphenyl, fluorene and phenanthrene, has been studied for first time utilizing a solid-phase microextraction (SPME)–high-performance liquid chromatography (HPLC) method. The stoichiometries of the analyte:β-CD complexes were found to be either 1:1 or 1:2. The formation of 1:2 complexes was confirmed for naphthalene, biphenyl, fluorene, and phenanthrene only when utilizing relatively high concentrations of β-CD (up to 6.6 mM). The 1:2 stoichiometries were confirmed using the classical modified Benesi–Hildebrand (BH) method. The calculated binding constants for 1:1 stoichiometries (K1) using the SPME method varied from 115.3 M−1 for toluene to 3510 M−1 for phenanthrene, whereas the corresponding values to the 1:2 stoichiometries (K3) varied from 7.30 × 105 M−2 for biphenyl to 9.03 × 106 M−2 for naphthalene.  相似文献   

15.
8-Quinolinol (HQ) reacts with [Pd(α-/β-NaiR)Cl2] [α-/β-NaiR = 1-alkyl-2-(naphthyl-α-/β-azo)imidazole] in acetonitrile (MeCN) solution to give [Pd(α-/β-NaiR)(Q)](ClO4). The products are characterized by spectroscopic techniques (FT-IR, UV–Vis, NMR). The reaction kinetics show a first order dependence of rate on each of the concentration of the metal complex and HQ. Addition of LiCl to the reaction retarded the rate of reaction and has proved the cleavage of the Pd–Cl bond as the rate-determining step. Thermodynamic parameters (ΔH° and ΔS°) are determined from variable temperature kinetic studies. The magnitude of the second order rate constant, k2, increases as in the order: Pd(NaiEt)Cl2 < Pd(NaiMe)Cl2 < Pd(NaiBz)Cl2 as well as Pd(β-NaiR)Cl2 < Pd(α-NaiR)Cl2.  相似文献   

16.
The reaction of PhHgOAc with N-NHCO-2-C4H3S-Htpp (5) and N-p-HNSO2C6H4tBu-Htpp (4) gave a mercury (II) complex of (phenylato) (N-2-thiophenecarboxamido-meso-tetra phenylporphyrinato)mercury(II) 1.5 methylene chloride solvate [HgPh(N-NHCO-2-C4H3S-tpp) · CH2Cl2 · 0.5C6H14;  6 · CH2Cl2 · 0.5C6H14] and a bismercury complex of bisphenylmercury(II) complex of 21-(4-tert-butyl-benzenesulfonamido)-5,10,15,20-tetraphenylporphyrin, [(HgPh)2(N-p-NSO2C6H4tBu-tpp); 7], respectively. The crystal structures of 6 · CH2Cl2 · 0.5C6H14 and 7 were determined. The coordination sphere around Hg(1) in 6 · CH2Cl2 · 0.5C6H14 and Hg(2) in 7 is a sitting-atop derivative with a seesaw geometry, whereas for the Hg(1) in 7, it is a linear coordination geometry. Both Hg(1) in 6 · CH2Cl2 · 0.5C6H14 and Hg(2) in 7 acquire 4-coordination with four strong bonds [Hg(1)–N(1) = 2.586(3) Å, Hg(1)–N(2) = 2.118(3) Å, Hg(1)–N(3) = 2.625(3) Å, and Hg(1)–C(50) = 2.049(4) Å for 6 · CH2Cl2 · 0.5C6H14; Hg(2)–N(1) = 2.566(6) Å, Hg(2)–N(2) = 2.155(6) Å, Hg(2)–N() = 2.583(6) Å, and Hg(2)–C(61) = 2.064(7) Å for 7]. The plane of the three pyrrole nitrogen atoms [i.e., N(1)–N(3)] strongly bonded to Hg(1) in 6 · CH2Cl2 · 0.5C6H14 and to Hg(2) in 7 is adopted as a reference plane 3N. For the Hg2+ complex in 6 · CH2Cl2 · 0.5C6H14, the pyrrole nitrogen bonded to the 2-thiophenecarboxamido ligand lies in a plane with a dihedral angle of 33.4° with respect to the 3N plane, but for the bismercury(II) complex in 7, the corresponding dihedral angle for the pyrrole nitrogen bonded to the NSO2C6H4tBu group is found to be 42.9°. In the former complex, Hg(1)2+ and N(5) are located on different sides at 1.47 and −1.29 Å from its 3N plane, and in the latter one, Hg(2)2+ and N(5) are also located on different sides at −1.49 and 1.36 Å form its 3N plane. The Hg(1)?Hg(2) distance in 7 is 3.622(6) Å. Hence, no metallophilic Hg(II)?Hg(II) interaction may be anticipated. NOE difference spectroscopy, HMQC and HMBC were employed to unambiguous assignment for the 1H and 13C NMR resonances of 6 · CH2Cl2 ·  0.5C6H14 in CD2Cl2 and 7 in CDCl3 at 20 °C. The 199Hg chemical shift δ for a 0.05 M solution of 7 in CDCl3 solution is observed at −1074 ppm for Hg(2) nucleus with a coordination number of four and at −1191 ppm for Hg(1) nucleus with a coordination number of two. The former resonance is consistent with that chemical shift for a 0.01 M solution of 6 in CD2Cl2 having observed at −1108 ppm for Hg(1) nucleus with a coordination number of four.  相似文献   

17.
Studies of heavy lanthanide chlorides may provide important information on the degree of Ln3+–ligand bond covalency. Monocrystals of LnCl3·6H2O, where Ln = Dy, Ho and Er, were grown and spectroscopic investigations were performed at room temperature and at low temperatures down to 4.2 K in order to understand the nature of the Ln3+–L bonds. The intensities of the electronic lines and the Judd–Ofelt parameters were calculated and compared with those obtained for chlorides of light lanthanides (i.e. Ce(III), Pr(III) and Nd(III)). Room temperature Raman and IR studies of the compounds under investigation were also performed. The relationship between hypersensitivity and covalency is discussed. The change of vibronic coupling strength along the lanthanide ion series does not modify monotonically. The ion-pair interactions are especially visible for the 5I8 → 5F2 and 5I8 → 5F3 transitions in the HoCl3·6H2O low temperature spectra.  相似文献   

18.
Three complexes of composition [CrL(X)3], where L = 4′-(2-pyridyl)-2,2′:6′,2″-terpyridine and X = Cl, N3, NCS are synthesized. They are characterized by IR, UV–Vis, fluorescence, EPR spectroscopic, and X-ray crystallographic studies. Structural studies reveal that the Cr(III) ion is coordinated by three N atoms of L in a meridional fashion. The three anions occupy the other three coordination sites completing the mer-N3Cl3 (1) and mer-N3N3 (2 and 3), distorted octahedral geometry. The Cr–N2 has a shorter length than the Cr–N1 and Cr–N3 distances and the order Cr–N(NCS) < Cr–N(N3) < Cr–Cl is observed. They exhibit some of the d–d transitions in the visible and intra-ligand transitions in the UV regions. The lowest energy d–d transition follows the trend [CrLCl3] < [CrL(N3)3] < [CrL(NCS)3] consistent with the spectrochemical series. In DMF, they exhibit fluorescence having π → π character. All the complexes show a rhombic splitting as well as zero-field splitting (zfs) in X-band EPR spectra at 77 K.  相似文献   

19.
The condensed phase, acid-base reaction enthalpy for Ln(tfa)·3H2O (s) + 3aza (s) → Ln(tfa)3·3aza (s) + 3H2O; (kJ mol−1) = −33.90 ± 1.54, −2.10 ± 1.25, −9.40 ± 2.10, 0.05 ± 2.27 and 2.46 ± 1.45 for the Pr, Nd, Sm, Eu and Tb compounds, respectively, where tfa, the trifluoroacetate, and aza, the 2-azacyclononanone, were measured by calorimetry.  相似文献   

20.
In this paper, we describe the structural and sensing properties of high-k PrYxOy sensing films deposited on Si substrates through reactive co-sputtering. Secondary ion mass spectrometry and atomic force microscopy were employed to analyze the compositional and morphological features of these films after annealing at various temperatures. The electrolyte-insulator-semiconductor (EIS) device incorporating a PrYxOy sensing membrane that had been annealed at 800 °C exhibited good sensing characteristics, including a high sensitivity (59.07 mV pH−1 in solutions from pH 2 to 12), a low hysteresis voltage (2.4 mV in the pH loop 7 → 4 → 7 → 10 → 7), and a small drift rate (0.62 mV h−1 in the buffer solution at pH 7). The PrYxOy EIS device also showed a high selective response towards H+. This improvement can be attributed to the small number of crystal defects and the large surface roughness. In addition, the enzymatic EIS-based urea biosensor incorporating a high-k PrYxOy sensing film annealed at 800 °C allowed the potentiometric analysis of urea, at concentrations ranging from 1 to 16 mM, with a sensitivity of 9.59 mV mM−1.  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