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.
The 13C NMR spectra of some derivatives of bryonolic acid (1) (D:C-friedoolean-8-en-3β-ol-29-oic acid) were assigned by means of 13C-enrichment, lanthanide-induced shifts (LIS) and comparison of chemical shift data between derivatives. The 13C-enriched species of 1, i.e., 1a, 1b and 1c were biosynthesized by Luffa cylindrica (Cucurbitaceae) callus fed with [1-13C]-, [2-13C]- or [1,2-13C2]-acetate, respectively. Methyl acetylbryonolates 2, 2a, 2b and 2c, methyl bryonolates 3, 3a, 3b and 3c, methyl bryononates 4 and 4a, diacetyl-3β,29-diols (3,29-diacetyl-D:C-friedoolean-8-en-β,29-diol) 5, 5a, 5b and 5c, and 3-acetyl-3β,29-diols 6, 6a and 6b were prepared from 1, 1a, 1b and 1c, and their 13C NMR spectra were recorded. The 13C concentration of the 13C-enriched species was high enough to exhibit the satellite peaks clearly, and the analysed data were very useful for this study. Thus, total assignments for 2, 3, 4, 5 and 6 were established. It was found that conversion of the methoxycarbonyl group at C-29 into an acetoxymethyl group caused complex changes in the chemical shifts of the C, D- and E-ring carbons and those of the methyl carbons linked to these rings.  相似文献   

2.
250 MHz 1H NMR of two monosubstituted [2.2]paracyclophanes shows that whether the substituent is an electron releasing group, OMe, or a withdrawing group, CO2Me, the transannular effect is deshielding. 62.86 MHz 13C NMR shows that among the six transannular effects, only one has a sign which changes with the nature of the substituent (–4.6 ppm for OMe but +3.3 ppm for CO2Me).  相似文献   

3.
The synthesis of [13C6]3,4-diaminobenzoic acid from commercially available [13C6]aniline is described in six steps. Salient features of this route include the preparation of a differentially protected 3,4-diaminobenzonitrile, hydrogen gas free aromatic nitro group reduction with ammonium formate and facile benzimidazole ring closure of the ortho-arylenediamine with triethylorthoformate. This stable-labeled 3,4-diaminobenzoic acid is an ideal [M+6]isotopomer to synthesize complex benzimidazole fragments for mass spectrometry internal assays.  相似文献   

4.
Carbon-13 chemical shift assignments are reported for benzo[b]thiophene and 1-(X-benzo[b]thienyl)ethyl acetate derivatives, where X=? CH(OAc)CH3 substituted at positions 2-7. Substituent chemical shift (SCS) effects for the ethyl acetate group are additive at all positions. A substantial upfield shift was observed at C-3, arising from the peri interaction of H-3 and the 4-ethyl acetate substituent. Carbon-13 relaxation times (T1) and nuclear Overhauser enhancements (η) have been measured for benzo[b]thiophene and its derivatives, and the contributions of dipolar, TDD1, and spin rotation, TSR1, relaxation have been determined. Intramolecular dipole–dipole interactions are found to provide by far the most important spin-lattice relaxation mechanism whenever protons are bound directly to the carbons under investigation. Nonprotonated ring carbons are relaxed by both DD and SR mechanisms. Anisotropic motion has an easily observable effect on the DD contribution to T1, and can form the basis for spectral assignments, as in 1-phenylethyl acetate. Long-range 13C? 1H coupling constants were observed both between ring carbons and between ring carbons with ring side-chain hydrogens. These results have been used for the structure determination of the title compounds.  相似文献   

5.
Dynamic nuclear polarization (DNP) of (13)C-labeled metabolic substrates in vitro and their subsequent intravenous administration allow both the location of the hyperpolarized substrate and the dynamics of its subsequent conversion into other metabolic products to be detected in vivo. We report here the hyperpolarization of [1-(13)C]-ascorbic acid (AA) and [1-(13)C]-dehydroascorbic acid (DHA), the reduced and oxidized forms of vitamin C, respectively, and evaluate their performance as probes of tumor redox state. Solution-state polarization of 10.5 ± 1.3% was achieved for both forms at pH 3.2, whereas at pH 7.0, [1-(13)C]-AA retained polarization of 5.1 ± 0.6% and [1-(13)C]-DHA retained 8.2 ± 1.1%. The spin-lattice relaxation times (T(1)'s) for these labeled nuclei are long at 9.4 T: 15.9 ± 0.7 s for AA and 20.5 ± 0.9 s for DHA. Extracellular oxidation of [1-(13)C]-AA and intracellular reduction of [1-(13)C]-DHA were observed in suspensions of murine lymphoma cells. The spontaneous reaction of DHA with the cellular antioxidant glutathione was monitored in vitro and was approximately 100-fold lower than the rate observed in cell suspensions, indicating enzymatic involvement in the intracellular reduction. [1-(13)C]-DHA reduction was also detected in lymphoma tumors in vivo. In contrast, no detectable oxidation of [1-(13)C]-AA was measured in the same tumors, consistent with the notion that tumors maintain a reduced microenvironment. This study demonstrates that hyperpolarized (13)C-labeled vitamin C could be used as a noninvasive biomarker of redox status in vivo, which has the potential to translate to the clinic.  相似文献   

6.
Abstract

Solid-state NMR data are presented to clearly support the formation of a 1:1 supramolecular complex between C60 and the tetra-tert-butylated analogue of calix[4]azulene by a simple mechanochemical hand-grinding of host and guest in a mortar and pestle. The experimental results are supported by a DFT study.  相似文献   

7.
8.
The chemical shifts and coupling constants of [1,2-15N2]pyrazole, 2-(1-[1,2- 15N2]pyrazolyl)-2-[l,3-2H6]propanol, 1-nitro[1,215N2] and 3-nitro[1,2-15N2]pyrazole are reported.  相似文献   

9.
The solution conformation of L-6-methylperhydroimidazo[1,5-c]thiazole-5,7-dione (γ-thiaprolinehydantoin) has been determined from an extensive 1H and 13C NMR study, allowing the extraction of vicinal inter-proton and carbon-hydrogen coupling constants. The major conformation of the thiazolidine ring is an envelope with C-δ as the flap exo?). In solution the preferred solid state (twist) conformer with C-α exo and C-β endo (αβT) is only a minor contributor. 13C spin–lattice relaxation data reveal the flexibility of the thiazolidine ring.  相似文献   

10.
A comparison of HSQC and HMQC pulse schemes for recording (1)H[bond](13)C correlation maps of protonated methyl groups in highly deuterated proteins is presented. It is shown that HMQC correlation maps can be as much as a factor of 3 more sensitive than their HSQC counterparts and that the sensitivity gains result from a TROSY effect that involves cancellation of intra-methyl dipolar relaxation interactions. (1)H[bond](13)C correlation spectra are recorded on U-[(15)N,(2)H], Ile delta 1-[(13)C,(1)H] samples of (i) malate synthase G, a 723 residue protein, at 37 and 5 degrees C, and of (ii) the protease ClpP, comprising 14 identical subunits, each with 193 residues (305 kDa), at 5 degrees C. The high quality of HMQC spectra obtained in short measuring times strongly suggests that methyl groups will be useful probes of structure and dynamics in supramolecular complexes.  相似文献   

11.
The 1H and 13C nmr spectra of 11,12-dimethoxy[1]benzothieno[3,2-a]-4,7-phenanthroline and its 8-chloro precursor were totally assigned using a combination of two-dimensional nmr techniques. After correlation of the majority of the proton signals by a COSY spectrum and one-bond heteronuclear correlation, the full assignment of the 1H and 13C nmr spectra of the novel heterocyclic compounds required the application of long-range CH coupling information particularly for quaternary carbon resonance assignments and the orientations of individual spin systems relative to one another. Key long-range heteronuclear couplings in both compounds served to confirm the one-bond heteronuclear correlations. Unequivocal interpretation of the spectral data leads to the complete assignments of the resonances.  相似文献   

12.
High-resolution (13)C NMR spectra (150 MHz) have been obtained on the complete series of D-aldohexoses (D-allose 1, D-altrose 2, D-galactose 3, D-glucose 4, D-gulose 5, D-idose 6, D-mannose 7, D-talose 8) selectively labeled with (13)C at C1 in order to detect and quantify the percentages of acyclic forms, and to measure and/or confirm percentages of furanoses and pyranoses, in aqueous solution. Aldehyde and hydrate signals were detected for all aldohexoses, and percentages of these forms at 30 degrees C ranged from 0.006 to 0.7% (hydrate) and 0.0032 to 0.09% (aldehyde). Aldehyde percentages are largest for the altro, ido, and talo configurations, ranging from 0.01 to 0.09%; the ido configuration yielded the most hydrate (0.74%). Hydrate/aldehyde ratios vary with aldohexose configuration, ranging from 1.5 to 13, with gluco exhibiting the smallest ratio and gulo the largest. (2)H Equilibrium isotope effects (EIEs) on aldohexose anomerization were measured in D-galactose 3 and D-talose 8 selectively (13)C- and (2)H-labeled at C1 and H1. The (2)H isotope effect on (13)C chemical shift, and broadband (1)H- and (2)H-decoupling, were exploited to permit simultaneous observation and quantitation of the protonated and deuterated molecules in NMR samples containing equimolar mixtures of D-[1-(13)C]aldose and D-[1-(13)C; 1-(2)H]aldose. Small (2)H EIEs were observed for 8, but were undetectable for 3. These results suggest that configuration at C2 influences the magnitude of the (2)H isotope effect at H1 and/or that the observed effect cannot be reliably interpreted due to complications arising from the involvement of acyclic aldehyde forms as intermediates in the interconversion of cyclic forms. The observed (2)H isotope effects on aldohexose tautomeric equilibria provide new insights into the important question of whether (2)H substitutions can alter aldofuranose ring conformation, and lead to the identification of an optimal (2)H- and (13)C-substituted 2-deoxyribofuranose isotopomer on which to investigate this potential effect.  相似文献   

13.
The Rh1(diolefin)complexes [Rh(nbd)( 2 )][PF6] [Rh(1,5-cod)( 2 )][PF6], and [Rh((Z)-α -acetamidocinnamic acid)( 2 )][PF6] ( 2 = the chiral P,N-ligand (S)-1-[bis(p-methylphenyl)phospino]-2-[p-methoxybenzyl)amino]-3-methylbutane have been prepared and characterized. These complexes exit as a mixture of isomers arising from different five-membered-ring conformations and diastereoisomers due to both the prochiral nitrogen and olefin ligands. The three-dimensional solutions structures of these complexes have been studied with the specific aim of understanding how the chiral pocket is built. Aspects of the exchange dynamics and their possible relevance to homogeneous hydrogenation are discussed The solid-state structure for the nbd complex, [Rh(nbd)( 2 )][PF6], as well as detailed one- and two-dimensional 31P-, 13C-, and 1H-NMR results are presented.  相似文献   

14.
1H, 13C and 15N chemical shifts of some pyridines and mesoionic oxatriazole aminides were recorded in the absence and presence of the complex dirhodium tetrakis(trifluoroacetate). The adduct formation shifts prove that the nitrogen atom in the pyridine derivatives and the N‐6 atom of the aminides are the binding sites in the adducts. At low temperature, adduct species can be identified separately by their individual signals. The 15N chemical shift responds very sensitively even to small concentration changes in the adduct. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

15.
Non linear modelling of data in photomineralization kinetics of organic micropollutants, by photocatalytic membranes immobilizing semiconductors (TiO2 particularly) has been previously applied to methane, phenol and to 2,4-dichlorophenol as model molecules, by using a four parameters kinetic modelling based on substrate disappearance and total organic carbon (TOC) in laboratory scale experiments, as a function of initial concentration of substrate and of irradiance.In the present paper, the photocatalytic degradation of diclofenac as model molecule was investigated in a pilot plant module, fitted with 2–3 concentric membranes. Maximum allowable quantum efficiencies corresponded to equal distances between 3 membranes immobilizing photocatalyst. This arrangement fully behaves as if a photocatalyst nanopowder would be homogeneously suspended in the reactor, but obviating all drawbacks of a nanopowder suspension.  相似文献   

16.
The 1H and 13C nmr resonances of pyrazolo[1,5-b][1,2,4]triazine ( 1 ) were assigned and the structure of its 2,3-dimethyl derivative 2 was determined using X-ray crystallography.  相似文献   

17.
1H and 13C NMR spectra of cis- and trans-3-(2-[2-(4-methylphenyl)ethenyl]phenyl])sydnones, the first stilbene-substituted mezoionic oxadiazolium rings, were fully assigned combining the information in various solvents, such as deuterated benzene, acetone and chloroform, using 2D NMR techniques.  相似文献   

18.
Membrane protein orientation has traditionally been determined by NMR using mechanically or magnetically aligned samples. Here we show a new NMR approach that abolishes the need for preparing macroscopically aligned membranes. When the protein undergoes fast uniaxial rotation around the bilayer normal, the 0 degrees -frequency of the motionally averaged powder spectrum is identical to the frequency of the aligned protein whose alignment axis is along the magnetic field. Thus, one can use unoriented membranes to determine the orientation of the protein relative to the bilayer normal. We demonstrate this approach on the M2 transmembrane peptide (M2TMP) of influenza A virus, which is known to assemble into a proton-conducting tetrameric helical bundle. The fast uniaxial rotational diffusion of the M2TMP helical bundle around the membrane normal is characterized via 2H quadrupolar couplings, C-H and N-H dipolar couplings, 13C chemical shift anisotropies, and 1H T1rho relaxation times. We then show that 15N chemical shift anisotropy and N-H dipolar coupling measured on these powder samples can be analyzed to yield precise tilt angles and rotation angles of the helices. The data show that the tilt angle of the M2TMP helices depends on the membrane thickness to reduce the hydrophobic mismatch. Moreover, the orientation of a longer M2 peptide containing both the transmembrane domain and cytoplasmic residues is similar to the orientation of the transmembrane domain alone, suggesting that the transmembrane domain regulates the orientation of this protein and that structural information obtained from M2TMP may be extrapolated to the longer peptide. This powder-NMR approach for orientation determination is generally applicable and can be extended to larger membrane proteins.  相似文献   

19.
Preparation is described of stereoisomeric (p-nitrobenzoyl)-5-aminomethylbicyclo[2.2.1]hept-2-enes and their epoxidation by peracetic acid. By an example of one of amides was demonstrated a possibility of selective reduction of separated fragments in the polyfunctional compound using sulfur in alkaline medium, hydrazine hydrate in the presence of a nickel catalyst, and lithium aluminum hydride. By reaction with electrophilic reagents from the amines synthesized, (p-aminobenzoyl)-endo-5-aminomethylbicyclo[2.2.1]hept-2-ene and (p-aminobenzoyl)-endo-2-aminomethylbicyclo[2.2.1]hept-2-ene, were obtained new bicyclic compounds containing alongside the amide group also sulfonamide, carboxamide, urea, and thiourea moieties. The structure of compounds obtained was confirmed by 1Hand 13C NMR spectroscopy, by two-dimensional spectra measured along COSY and NOESY procedures.  相似文献   

20.
1H, 13C and 15N NMR studies demonstrated that 9-hydrazono-6,7,8,9-tetrahydro-4-oxo-4H-pyrido-[1,2-a] pyrimidlnes exist as an equilibrium mixture of Z-E isomers in the hydrazono–imino tautomeric form having an exocyclic double bond. Proton-catalysed Z-E interconversion is fast. Substituent and solvent effects revealed that the decisive factors controlling the Z:E ratio are internal hydrogen bonding in the Z-isomer, stabilization by solvation and steric interaction.  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