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 62 毫秒
1.
The fatty acids (FAs) composition of lipids extracted from Rhodobacter sphaeroides 2.4.1 was investigated by gas chromatography–mass spectrometry (GC–MS) analysis of the corresponding FA methyl esters (FAMEs), obtained through trans-esterification of the original lipid species. A GC stationary phase based on a highly polar ionic liquid (IL) was selected, aimed to enhance the separation of isomeric FAMEs with particular emphasis on positional and geometrical isomers of monounsaturated 16:1 and 18:1 fatty acyl chains. The occurrence of 18:1 cis9 (oleic) acid, a positional isomer of the well-known and most predominant 18:1 cis11 (cis-vaccenic) acid, has been demonstrated here for the first time. Furthermore a methyl branched 18:1 FA was also identified and its structure tentatively assigned as 11-methyl-Δ12-octadecenoic acid (most likely as trans isomer). The unprecedented observation about 18:1 cis9 FA occurrence in R. sphaeroides 2.4.1 is, even indirectly, supported by a biosynthetic pathway postulated with the aid of the Kyoto Encyclopedia of Genes and Genomes (KEGG) database. The concurrent presence of 16:1 cis7 and 18:1 cis9 FAs suggested the existence of parallel and/or complementary processes to those invoked for the formation of most common 16:1 cis9 and 18:1 cis11 FAs. A further route was hypothesized for the trans FAs biosynthesis in wild-type cells of R. sphaeroides.  相似文献   

2.
The equilibrium between fluoral in dichloromethane solution and live condensed liquid polyfluoral has been investigated between 22 and 43°C. Equilibrium monomer concentrations gave: ΔHac°(298 K) = -50-8 ± 2·3 kJ mol?1 and ΔSsc° (298 K) = -142·7 ± 7·4 J K-1 mol-1. With the aid of calibration and monomer vaporization data, thermodynamic values for the polymerization of liquid monomer to liquid polymer were also calculated: ΔHtc° (298 K) = -47 ± 3 kJ mol-1 and ΔS1e° (298 K) = -97 ± 10 J K-1 mol-1.  相似文献   

3.
The availability of tetrahydrocannabinols (Δ9-THC), tetrahydrocannabivarins (Δ9-THCV), and their metabolites in both their undeuterated and deuterated forms is critical for the analysis of biological and toxicological samples. We report here a concise methodology for the syntheses of (−)-Δ9-THC and (−)-Δ9-THCV metabolites in significantly improved overall yields using commercially available starting materials. Our approach allowed us to obtain the key intermediates (6aR,10aR)-9-nor-9-oxo-hexahydrocannabinols in four steps from (+)-(1R)-nopinone. This was followed by an optimized Shapiro reaction to give the (−)-11-nor-9-carboxy-metabolites, which were converted to their respective (−)-11-hydroxy analogs. The synthetic sequence involves a minimum number of steps, avoids undesirable oxidative conditions, and incorporates the costly deuterated resorcinols near the end of the synthetic sequence. This methodology enabled us to synthesize eight regiospecifically deuterated (−)-Δ9-THC and (−)-Δ9-THCV metabolites in a preparative scale and high optical purity without deuterium scrambling or loss.  相似文献   

4.
The synthesis of endo-3,6-epoxy-Δ4-tetrahydrophthalic anhydride from the endo-adduct of furan and maleic acid is described. Reduction of endo- and exo-3,6-epoxy-Δ4-tetrahydrophthalic anhydride with sodium borohydride gave the corresponding lactones, while catalytic hydrogenation over 10% Pd/C gave anhydride and/or hemi-acylals, depending on the solvent.  相似文献   

5.
Asymmetric hydrogenations of Ac-ΔTyr(Ac)-(S)-Ala-Gly-OMe (6), Ac-ΔTyr(Ac)-(R)-AlaGly-(S)-Phe-OMe (7), Ac-ΔPhe-NH-CH(R) CH2-OCH2Ph (10), HCO-ΔPhe-(S)-Leu-OMe (16), X-AA-ΔPhe-AA'-OMe (5: X tBOC, CBZ, CF3CO; AA, AA' = α-amino acid), and tBOC-AA-ΔPhe-AA'-NH-Y (21: Y=H, NH-AA'-ΔPhe-AA-tBOC, NHPh), catalyzed by cationic Rh complexes, [L*Rh(NBD)]+ClO4 (L* = chiral diphosphine), were performed to give the corresponding chiral oligopeptides with high stereoselectivities. It was found that the nature of the N-protecting group of dehydrotripeptides (5) exerted a significant influence on the asymmetric induction as well as catalyst efficiency. The chiral centers in AA' and AA' amino acid residues in 5 were also found to have some influence on the catalytic asymmetric induction. Possible mechanism which can accommodate these effects are discussed. Asymmetric reduction of RCOCO-AA-OMe (13) via hydrosilylation was carried out to give chiral depsipeptide units. The asymmetric hydrogenation of dehydropeptides was successfully applied to the synthesis of enkephalin analogs, Ac-(R)Tyr-(R) -Ala-Gly-(S)-Phe-(S)-Leu-OMe (23) and Ac-(S)Tyr-(R) -Ala-Gly-(S)-Phe-(S)- Leu-OMe (29)  相似文献   

6.
A geometrical interpretation of intermittency in fully developed turbulence is realized through an hierarchy of fractal structures Ωp of dimensions Δp linked each other by the relations Ωp + 1 − Ωp (i.e. Δp + 1 < Δp) and γ = (Δp + 1Δ)/(ΔpΔ) with γ = ((1 + 3/√8)1/3 + (1 − 3/√8)1/3)3 and Δ = 1 and Δ = 1. This is obtained by the introduction of an entropy jump, defined at the scale r, ΔSp(r) = (Δp + 1Δp) ln (r/r0) characterizing the order level of each sub-structure Ωp and verifying a linear relation ΔSp(r) = γ ΔSp − 1(r).  相似文献   

7.
Reaction of Δ1,4-3-keto steroids with zinc and acid produces a characteristic fluorescence. Estrogenic, 3-keto and Δ4-3-keto steroids did not yield this fluorescence. Stabilization of the dienone fluorescence by extraction into butyl ether allowed quantitation of triamcinolone acetonide in liquid formulations in which this dienone was present at the 0.01% level. A mechanism for the reaction is proposed.  相似文献   

8.
α-Substituted (CH3, C6H5, OCH3, OC6H5, Cl, and F) hydantoin-Δ5,α -acetates were obtained from diethyl oxaloacetates and urea, and their 3-methyl derivative s were obtained from N-methylurea or by methylation of the nitrogen-unsubstituted hydantoins with diazomethane; theα-nitro derivative was obtained by nitration of unsubstituted ethyl hydantoin-Δ5,α -acetate with nitric acid in acetic acid, and ethyl hydantoin-Δ5,α -glycolate was obtained from hydantoin and diethyl oxalate. All of the synthesized hydantoin-Δ5,α -acetates, except for theα-nitro andα-hydroxy derivatives, are converted to the corresponding 3- and 5-substituted orotic acids.  相似文献   

9.
Gas chromatography in combination with electron ionization mass spectrometry (GC/EI-MS) was used to determine the fatty acids of a membrane lipid from Bacillus megaterium. Special attention was put on the structure and absolute configuration of a monoenoic fatty acid previously described in this sample. GC/EI-MS operated in the selected ion monitoring mode was used to determine twelve fatty acids in the bacterium. Methyl esters were prepared to verify the presence of a 14-methylhexadecenoic acid (a17:1) isomer. The position of the double bond of the a17:1 isomer and four further monoenoic fatty acids was elucidated by means of their picolinyl esters produced by the transesterification of the phospholipid. For the a17:1 isomer, the double bond was located between C-5 and C-6. Silver ion liquid chromatography was used to verify that the double bond was in cis-configuration. The bacterial 14-methylhexadec-5-enoic acid (a17:1Δ5) is chiral due to the stereogenic C-14 carbon. Initial enantioselective measurements were carried out with isomers of a17:1Δ5 which were available in form of racemic and (S)-enantiopure cis- and trans-isomers of a17:1Δ12 previously synthesized. The cis-a17:1Δ12 enantiomers were partly resolved on a chiral stationary phase coated with 50% heptakis(6-O-tert.-butyldimethylsilyl-2,3-di-O-methyl)-β-cyclodextrin in OV-1701 (β-TBDM). However, resolution of the enantiomers of the trans-isomer of a17:1Δ12 failed. Only one peak was also observed for the a17:1Δ5 isomer from B. megaterium. Thus, it remained unclear whether the compound a17:1Δ5 was racemic or enantiopure in the sample. To clarify this point, we separated the cis-monoenoic fraction from the saturated fatty acids. Then, the monoenoic fraction was hydrogenated in order to transform a17:1Δ5 into 14-methylhexadecanoic acid (a17:0). This chiral fatty acid was known to be sufficiently enantioseparated on the β-TBDM column and was found to be (S)-enantiopure in the sample. Hence, these measurements verified that the B. megaterium sample contained enantiopure (S)-a17:1Δ5.  相似文献   

10.
Reversed-phase high-performance liquid chromatographic methods were developed for the separation of enantiomers of eleven unnatural β2-amino acids on a new chiral stationary phase, using the 11-methylene-unit spacer of aminoundecylsilica gel for the bonding of (+)-(18-crown-6)-2,3,11,12-tetracarboxylic acid as selector. The nature and concentration of the acidic and organic modifiers, the pH, the mobile phase composition, and the structures of the analytes substantially influenced the retention and resolution. Separations were carried out at constant mobile phase compositions in the temperature range 7–40 °C and the changes in enthalpy, Δ(ΔH°), entropy, Δ(ΔS°), and free energy, Δ(ΔG°) were calculated. The elution sequence was determined in some cases: the S enantiomers eluted before the R enantiomers.  相似文献   

11.
Proton-ligand dissociation constants of five biologically important pyrazole derivatives, viz. [5-(2-hydroxyphenyl)-3-(pyridin-3-yl)-4-benzoyl]-pyrazol (HPPBP), [5-(2-hydroxyphenyl)-3-(3-nitrophenyl)-4-(3-pyridinoyl)]-pyrazol (HPNPPP), [5-(2-hydroxyphenyl)-3-(3-nitrophenyl)-4-benzoyl]-pyrazol (HPNPBP), [5-(2-hydroxyphenyl)-3-phenyl-4-(3-pyridinoyl)]-pyrazol (HPPPP), and [5-(2-hydroxyphenyl)-3-(3-nitrophenyl)-4-(2-furoyl) pyrazol (HPNPFP) and metal ligand stability constants of their Ni(II) complexes in 70% (v/v) dioxane-water and 0.1 M KNO3 were determined at 298.15, 303.15, and 308.15 K by potentiometric method. Thermodynamic functions, such as, free energy change (ΔG ), enthalpy change (ΔH ) and entropy (ΔS ) change for dissociation and complex formation have been estimated form temperature dependence of proton-ligand and metal-ligand stability constants and interpreted in terms of feasibility of these processes.  相似文献   

12.
The modified statistical theory developed previously for potentials appropriate to interactions in neutral-neutral collisions, is now extended to more strongly attractive potentials involved in ion-neutral collisions. The model system is the collisional deactivation of C5H9+ by a variety of both polar and non-polar neutral molecules. A 12 - 6 - 4 potential is used for ion interaction with non-polar neutrals, and a 12 - 6 - 4 - 2 potential, as modified by Su and Bowers to take into account the rotational energy of the neutral, for interaction with polar neutrals. Calculated is (ΔE), the average energy lost by the ion in a collision, and compared with experiment. For C5H9+-CH4 collisions, the calculated (ΔE) agrees with experiment within 5%. Predictions of the theory, namely that (ΔE) should increase with excitation energy and should decrease with the size of the excited reactant, are found to be in fair agreement with the somewhat ambiguous experimental evidence.  相似文献   

13.
The effects of oxygen reduction treatments on the magnetic properties of La-deficient manganites, La1−ΔMnO3+δ and Sr- and Ca-doped manganites, La1−xMxMnO3+δ (M: Sr, Ca) have been investigated to confirm the contrasting oxygen reduction effects on the magnetization properties. It is found that oxygen reduction treatments in reduced oxygen pressures of 103- for La1−ΔMnO3+δ result in a continuous change in the magnetization but the reduction treatments for La1−xMxMnO3+δ result in a negligible change under the same reduction conditions. To interpret the contrasting behavior of the La-deficient manganites, several possible models have been discussed. Among the models, the most probable model is that vacancies generated by the La deficiency Δ are partially replaced by Δ2(=ΔΔ1?Δ1) Mn ions to give both La and Mn site vacancies according to the formula La1−ΔVΔMnO3+δ→{La1−ΔMnΔ2VΔ1}{Mn1−Δ2VΔ2}O3+δ. Details of thermodynamic basis of this model have been presented.  相似文献   

14.
  1. By means of differential scanning calorimetry the phase transition temperatures (T u) and phase transition enthalpies (ΔH u) ofα- andβ-branched lecithins were determined. The lecithin samples contained 50% water per weight.
  2. The occurrence of the pre-transition was dependent on the position of the methyl group in the acid residue as well as on the length of the carbon chains. TheΔH u values of the pre-transition were about 2.5 kJ·mol?1.
  3. The hydration number (HZ) (number of water molecules per lecithin headgroup) was independent on the position of the methyl group and the length of the acyl group. The average HZ-values for all the lecithins studied is about HZ=14+-1.
  4. The influence of the chain length of theα- andβ- branched acyl groups affects the thermodynamical parameters of the main transitions (gel→L α). TheΔH m andT m-values increased as the chain length increases.
  相似文献   

15.
Heats of fusion of polyethylene-adipate, pimelate, suberate and azelaate have been determined by two methods, viz. DTA and the pressure dependence of the melting point up to 6000 bar. Degrees of crystallization were measured dilatometrically. Entropies of fusion were divided into volume and conformational entropy terms, only the latter alternates. Alternation of melting points depends on enthalpy of fusion ΔHm, entropy of fusion ΔSm and volume change Δυ on melting; influence of the functions increases thus ΔVm < ΔHm < ΔSm and ΔSm is dominant. Entropy and heat of fusion alternation is explained by the conformational change on melting governed by the driving force of maximum H-bridge formation.  相似文献   

16.
This paper focuses on the thermal properties, the microstructure, and the molecular dynamics of water in the hydrogels (1.5, 2, 3, 4, and 5% [g mL−1]) formed by sugar-based low molecular-weight gelator methyl-4,6-O-(p-nitrobenzylidene)-α-d-glucopyranoside. The energy needed to break the non-covalent interactions such as the hydrophobic, dipole-dipole, and π-π stacking interactions responsible for the gel formation was calculated to be 43 kJ mol−1. The microstructure of the 4% [g mL−1] hydrogel shows a characteristic fibril structure of the gel network with individual gel fibers, the junction points of thicker fibers, and pores occupied by water. The single mode diffusion of water molecules inside the gel network was detected irrespective of the diffusion time Δ (8-75 ms) and hydrogel concentration. For Δ of 10 ms the water diffusion is almost free and characterized by the diffusion coefficient in the range from 2.17×10−9 to 1.84×10−9 m2 s−1 for studied hydrogels. For larger Δ values, so-called restricted diffusions are observed and manifested in the linear decreases of the diffusion coefficient with diffusing time Δ, as shown for 5% gel. Only the one average proton spin-lattice relaxation time T1 of water was determined for the studied hydrogels, irrespective of gelator concentration.  相似文献   

17.
《Chemical physics letters》1987,140(5):531-536
Trifluoromethoxy radical formation (by O-atom addition to trifluoromethyl) and dissociation (by F-atom elimination) are studied by ab initio molecular-orbital theory. The activation enthalpy (298 K) for F-atom elimination is 35.3 kcal mol−1 at the UMP4SDQ/6-31 G1//UHF/6-31 G1+ΔZPE+Δ(H-E0 level. The implication of calculated RRKM dissociation rate constants is discussed.  相似文献   

18.
The molar enthalpies of dissolution for 3,6-bis(1H-1,2,3,4-tetrazol-5-yl-amino)-1,2,4,5-tetrazine (BTATz) were measured in N-methyl-2-pyrrolidone (NMP) and dimethyl sulfoxide (DMSO) using an RD496-2000 Calvet microcalorimeter at 298.15 K under atmospheric pressure. Empirical formulae for the calculation of the molar enthalpies of dissolution (Δdiss H), relative partial molar enthalpies (Δdiss H partial), and relative apparent molar enthalpies (Δdiss H apparent) were obtained from the experimental results of the dissolution processes of BTATz in NMP or DMSO. Furthermore, the corresponding kinetic equations describing the two dissolution processes are /dt = 10?3.55(1 ? α)0.57 for the dissolution of BTATz in NMP, and /dt = 10?3.74(1 ? α)0.63 for the dissolution of BTATz in DMSO, respectively.  相似文献   

19.
A. Bodor  A. Barabás 《Tetrahedron》1979,35(2):233-240
O-Benzyl- and/or O-iso-propyl-oximes of a number of 3-, 7- or 20-oxo-steroids and of some 3,20-di-oxosteroids have been prepared. The products were characterized by their C,H,N-content and physical properties and studied by UV, 1H- and 13C-NMR spectroscopy. The ketoximes of 3-oxo-Δ4-steroids were invariably found to consist of a mixture of two geometrical isomers, denominated syn and anti. In some cases both isomers could be obtained in a pure state by column-chromatography. Oximes of the 7-oxo-Δ5-steroids were found to consist only of the syn isomer, whereas the 20-oximino compounds showed no isomerism.  相似文献   

20.
Synthetic Na-magadiite sample was used for organofunctionalization process with N-propyldiethylenetrimethoxysilane and bis[3-(triethoxysilyl)propyl]tetrasulfide, after expanding the interlayer distance with polar organic solvents such as dimethylsulfoxide (DMSO). The resulted materials were submitted to process of adsorption with arsenic solution at pH 2.0 and 298±1 K. The adsorption isotherms were adjusted using a modified Langmuir equation with regression nonlinear; the net thermal effects obtained from calorimetric titration measurements were adjusted to a modified Langmuir equation. The adsorption process was exothermic (ΔintH=−4.15-5.98 kJ mol−1) accompanied by increase in entropy (ΔintS=41.32-62.20 J k−1 mol−1) and Gibbs energy (ΔintG=−22.44−24.56 kJ mol−1). The favorable values corroborate with the arsenic (III)/basic reactive centers interaction at the solid-liquid interface in the spontaneous process.  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