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.
5-Methylfurfural diethyl acetal, 5-methylfuran acid ester, and 4-hydroxypentanoic acid ethyl ester were first identified in high-sacchariferous corn media subjected to alcoholic fermentation by Mucor X-I culture and in aged sweet wines by means of GLC, GLC-MS, and TLC, as well as by UV and IR spectroscopic examinations. These substances are products of enzymatic esterification of sugars followed by their dehydration by heat treatment or long seasoning of grape wines.  相似文献   

2.
The esterification rates (ratio of the concentration of an acid in the neutral ethyl ester form to total concentration of the acid) of main organic acids in wines were determined to study the extent of ethyl ester formation of organic acids. The esterification rates ranged from zero to 24.6%. The averaged values of table wines were from 6 to 16% for acetic and lactic acid, from 0.3 to 3.6% for succinic and malic acid, and from zero to 0.1% for tartaric acid. Sherries had higher esterification rates, about 1.6 to 6 times larger, than table wines. It was found that storage time and temperature influence the formation of ethyl esters, and it was suggested that the aging period required for the ester equilibrium is about one year for acetic and lactic acid, and more than two years for succinic, malic and tartaric acid. The possibility and the procedure to control wine quality during the aging process were discussed.  相似文献   

3.
4-Chloroindole-3-acetic acid methyl ester was identified unequivocally in Lathyrus latifolius L., Vicia faba L. and Pisum sativum L. by thin layer chromatography, gas chromatography and mass spectrometry. The gas chromatographic system was able to separate underivatized chloroindole-3-acetic acid methyl ester isomers. The quantitative determination of 4-chloroindole-3-acetic acid methyl ester in immature seeds of these three species was performed by gas chromatography – mass spectrometry using deuterium labelled 4-chloro-indole-3-acetic acid methyl ester as an internal standard. P. sativum contained approximately 25 mg kg-1, V. faba 1–2 mg kg-1 and L. latifolius 2 mg kg-1 dry weight.  相似文献   

4.
Abstract

The reaction of 5-iodouridine with esters of acrylic acid in the palladium-catalysed Heck reaction was used to generate a series of esters of the acid (E) -5- (2-carboxyvinyl)uridine in poor to moderate yield. An alternative method starts from the acid by protection of the sugar moiety followed by in situ generation of the acid chloride then ester formation followed by simultaneous sugar deprotection. Treatment of the methyl ester with ammonia and methylamine gave the corresponding amides, while treatment with dimethylamine gave 5-(1-dimethylamino-2-carboxy)ethyluridine as the major product.  相似文献   

5.
E. Vieitez    E. Seoame    D. V. Gesto    C. Mato    A. Vazquez  A. Carnicer 《Physiologia plantarum》1966,19(2):294-307
Woody cuttings of Ribes rubrum, an easy plant to root, were extracted with methanol. One fraction of this extract was both water and ether soluble; a part of it was only water soluble. From the water and ether soluble fraction a crystalline compound was isolated which was fully identified with p-hydroxybenzoic acid (PHB) by m.p., mixed m.p., infrared spectrum, ultraviolet absorption in neutral and alkaline methanol and chemical analyses. For Avena coleoptiles straight growth test, p-hydroxyhenzoic acid was found to have a significant growth promoting activity in the range 20–100 μg/ml. and it showed strong growth inhibition at concentrations higher than 200 μg/ml. From the same extract fraction a phenolic ester was isolated by paper chromatography which showed a weak growth inhibiting activity. This compound was identified as an ester of p-hydroxybenzoic acid by micro-scale reactions. From the extract soluble only in water further amounts of crystalline p-hydroxybenzoic acid were obtained by alkaline hydrolysis. This PHB may have arisen from the breakdown of pelargonidin glucosides.  相似文献   

6.
Volatile aroma compounds are synthesized by wine yeast during wine fermentation. In this study the volatile aroma composition of two varieties of mango wine were determined to differentiate and characterize the wines. The wine was produced from the fruits of two varieties of mango cultivars namely Banginapalli and Alphonso. The volatile compounds formed in mango wine were analyzed by gas chromatography coupled with mass spectrometry (GC-MS). Thirty-two volatile compounds in wines were determined of which four were new and unidentified present in lower concentration. Apart from the ethanol (8.5 ± 0.28 and 7.2 ± 0.28% v/v), 1-propanol (54.11 ± 0.33 and 42.32 ± 0.57 mg/l), isobutyl alcohol (102 ± 1.57 and 115.14 ± 2.88 mg/l) and isoamyl alcohol (123 ± 2.88 and 108.40 ± 0.23 mg/1) were found to be the major flavouring higher alcohols in the mango wines produced from the fruits of Banginapalli and Alphonso respectively. Ethyl acetate (35 ± 0.57 and 30.42 ±1.15 mg/l) was the major ester component in both wines produced. Besides, other esters like ethyl octonoate, ethyl hexanoate and ethyl decanoate were also present in the wines. Cyclohexane methanol (1.45 ± 0.11 mg/l) was present only in wine made from Banginapalli and β-phenylethyl butanoate (0.62 ± 0.01 mg/1) was found only in Alphonso wine. The results demonstrate that the wine prepared from Banginapalli variety had better aroma composition and good taste than that from the Alphonso variety.  相似文献   

7.
Candida rugosa lipase was encapsulated within a sol–gel procedure and improved considerably by fluoride-catalyzed hydrolysis of mixtures of octyltriethoxysilane and tetraethoxysilane in the presence of magnetic sporopollenin. The catalytic properties of the immobilized lipases were evaluated into model reactions, i.e., the hydrolysis of p-nitrophenylpalmitate (p-NPP), and the enantioselective hydrolysis of racemic naproxen methyl ester, mandelic acid methyl ester or 2-phenoxypropionic acid methyl ester that were studied in aqueous buffer solution/isooctane reaction system. The encapsulated magnetic sporopollenin (Spo-M-E) was found to give 319 U/g of support with 342% activity yield. It has been observed that the percent activity yields and enantioselectivity of the magnetic sporopollenin encapsulated lipase were higher than that of the encapsulated lipase without support. The substrate specificity of the encapsulated lipase revealed more efficient hydrolysis of the racemic naproxen methyl ester and 2-phenoxypropionic acid methyl ester than racemic mandelic acid methyl ester. It was observed that excellent enantioselectivity (E > 400) was obtained for encapsulated lipase with magnetic sporopollenin with an ee value of S-Naproxen and R-2 phenoxypropionic acid about 98%.  相似文献   

8.
As a potential source of biomass supplies, cassava (Manihot esculenta Crantz) has been studied for bioethanol production, but not for the production of biodiesel. In this study, we used cassava hydrolysate as an alternative carbon source for the growth of microalgae (Chlorella protothecoides) which accumulated oil in vivo, with high oil content up to 53% by dry mass under a 5-L scale fermentation condition. The oils were extracted and converted into biodiesel by transesterification. The biodiesel obtained consisted of mainly unsaturated fatty acids methyl ester (over 82%), cetane acid methyl ester, linoleic acid methyl ester, and oleic acid methyl ester. This work suggests the feasibility of an alternative choice for producing biodiesel from cassava by microalgae fermentation. We report herewith the optimized condition for the fermentation and for the hydrolysis of cassava as the carbon source.  相似文献   

9.
A preliminary investigation on 20 Aglianico del Vulture commercial wines from the Basilicata region proved the existence of a significant variability in total antioxidant capacity which can exert a potential impact on wine quality. Nineteen Saccharomyces cerevisiae strains were tested in Aglianico del Vulture on pilot scale fermentation and the experimental wines obtained were evaluated for the antioxidant capacity, ethanol and total polyphenols. At the ninth day of fermentation the experimental wines had an antioxidant capacity, measured by photochemiluminescence, between 2.88 and 6.25 mM of ascorbic acid equivalent, ethanol concentration, measured by GC, between 5.49% and 10.99% and total polyphenols, determined by Folin Ciocalteau reagent, from 1153 to 1867 mg catechins/l. After 12 days the total antioxidant capacity was increased in most wines but decreased in some wines. These results, statistically analysed by principal component analysis, revealed a significant influence of S. cerevisiae strain on total antioxidant capacity and total polyphenols content of wine.  相似文献   

10.
A stable ESR signal, centred at g = 2.0037 ± 0.0002, characterised by a single resonance and assignable to a free radical, was found in all the bottled red wines, both commercial and experimental, that we have examined. The radical concentration was calculated to be in the range of 5–82 nM. After exposure of the wines to air for a few minutes a two fold increase of the ESR signal, followed by a slow decrease with time, was observed. The intensity of ESR signal in experimental red wines, was found to increase with the ageing of the wines and was strictly correlated to the total content of polyphenols. The formation of semiquinone radicals of polyphenols is suggested as one possible mechanism leading to the presence of stable free radicals in red wines.  相似文献   

11.
A large number of strains of Oenococcus oeni (formerly Leuconostoc oenos) that had been isolated from wines were checked for lysogeny with mitomycin C as inducer. As a result of this test, 45% of the strains proved to be lysogenic, suggesting that lysogeny is widespread among bacteria isolated from wines during malolactic fermentation. The sensitivity of bacteria to phages was very different, depending on the strain. All the lysogenic strains were resistant to infection by the temperate phage they released. Some phages infected none of the strains. Phages of Oenoc. oeni had a classical morphology, an isometric head, and a long striated tail. With the broadest host strain as an indicator, phages were detected in wines after malolactic fermentation. Received: 28 November 1997 / Accepted: 5 January 1998  相似文献   

12.
Herein, we report the influence of different combinations of initial concentration of acetic acid and ethanol on the removal of acetic acid from acidic wines by two commercial Saccharomyces cerevisiae strains S26 and S29. Both strains reduced the volatile acidity of an acidic wine (1.0 g l−1 acetic acid and 11% (v/v) ethanol) by 78% and 48%, respectively. Acetic acid removal by strains S26 and S29 was associated with a decrease in ethanol concentration of 0.7 and 1.2% (v/v), respectively. Strain S26 revealed better removal efficiency due to its higher tolerance to stress factors imposed by acidic wines. Sulfur dioxide (SO2) in the concentration range 95–170 mg l−1 inhibits the ability of both strains to reduce the volatile acidity of the acidic wine used under our experimental conditions. Therefore, deacidification should be carried out either in wines stabilized by filtration or in wines with SO2 concentrations up to 70 mg l−1. Deacidification of wines with the better performing strain S26 was associated with changes in the concentration of volatile compounds. The most pronounced increase was observed for isoamyl acetate (banana) and ethyl hexanoate (apple, pineapple), with an 18- and 25-fold increment, respectively, to values above the detection threshold. The acetaldehyde concentration of the deacidified wine was 2.3 times higher, and may have a detrimental effect on the wine aroma. Moreover, deacidification led to increased fatty acids concentration, but still within the range of values described for spontaneous fermentations, and with apparently no negative impact on the organoleptical properties.  相似文献   

13.
Synthesis of (±)-trans-chrysanthemic acid from (±)-1′-hydroxydihydro-trans-chrysanthemic acid by the dehydration with p-toluene-sulfonic acid was attempted. However, the attempt was found to be unsuccessful giving a compound believed to be methyl methyl 2,6 dimethylhepta-3.6-diene-5-carboxylate upon dehydration.

A cleavage upon cyclopropane ring was confirmed by deriving the acid obtained by the hydrolysis of the above ester to already known 2,6-dimethyl-heptane-5-carboxylic acid.

Analogous mode of dehydration and cleavage upon the ester of (±)-2,2-dimethyl-3-trans-hydroxylbenzyl-cyclopropane-l-carboxylic acid was also observed to give 1-phenyl-4-methyl-penta-1,3-diene-3-carboxylic acid. On the other hand, (±)-trans-caronic acid being derived to (±)-1′-oxo-2′-hydroxy-dihydro-trans-chrysanthemic acid, the synthesis of (±)-trans-chrysanthemic acid from (±)-trans-caronic acid became possible using (±)-1′-oxo-2′-hydroxy-dihydro-trans-chrysanthemic acid as a relay substance.  相似文献   

14.
The lipase‐catalyzed irreversible transesterification procedure using vinyl esters was applied to the resolution of racemic 2‐phenoxypropanoic acids. Aspergillus niger lipase showed high enantioselectivities and reasonable reaction rates. The enantioselectivity was found to be affected profoundly by several variables, e.g., the alcohol as nucleophile, the organic solvent used, and the reaction temperature. A gram‐scale resolution of (RS)‐2‐phenoxypropanoic acid was achieved after optimization of the reaction conditions. Then this irreversible transesterification procedure was applied to the resolution of some related 2‐substituted carboxylic acids. Thus, racemic 2‐methoxy‐2‐phenylacetic acid was resolved via the A. niger lipase‐catalyzed transesterification of the corresponding vinyl ester. 2‐Phenylpropanoic acid and 2‐phenylbutanoic acid were resolved using Pseudomonas sp. lipase. A gram‐scale resolution of 2‐phenylbutanoic acid was achieved by this procedure coupled with the porcine liver esterase‐catalyzed hydrolysis of the resulting methyl ester. Chirality 11:554–560, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

15.
Aims: To evaluate the ability of grapevine ecosystem fungi to degrade histamine, tyramine and putrescine in synthetic medium and in wines. Methods and Results: Grapevine and vineyard soil fungi were isolated from four locations of Spain and were subsequently identified by PCR. A total of 44 fungi were evaluated for in vitro amine degradation in a microfermentation system. Amine degradation by fungi was assayed by reversed‐phase (RP)‐HPLC. All fungi were able to degrade at least two different primary amines. Species of Pencillium citrinum, Alternaria sp., Phoma sp., Ulocladium chartarum and Epicoccum nigrum were found to exhibit the highest capacity for amine degradation. In a second experiment, cell‐free supernatants of P. citrinum CIAL‐274,760 (CECT 20782) grown in yeast carbon base with histamine, tyramine or putrescine, were tested for their ability to degrade amines in three different wines (red, white and synthetic). The highest levels of biogenic amine degradation were obtained with histamine‐induced enzymatic extract. Conclusion: The study highlighted the ability of grapevine ecosystem fungi to degrade biogenic amines and their potential application for biogenic amines removal in wine. Significance and Impact of Study: The fungi extracts described in this study may be useful in winemaking to reduce the biogenic amines content of wines, thereby preventing the possible adverse effects on health in sensitive individuals and the trade and export of wine.  相似文献   

16.
Two new flavonoid glucuronate esters, named scuregeliosides A and B ( 1 and 2 ), as well as three known ones, chrysin‐7‐O‐β‐d ‐glucuronic acid methyl ester ( 3 ), 5,7,4′‐trihydroxyflavone‐8‐O‐β‐d ‐glucuronic acid methyl ester ( 4 ) and apigenin‐7‐O‐β‐d ‐glucuronic acid ethyl ester ( 5 ), were isolated from the ethanolic extract of the whole plant of Scutellaria regeliana. Their chemical structures were elucidated on the basis of comprehensive spectroscopic analyses. Five compounds were screened for anti‐inflammatory activity in vitro. As the results, the inhibition rates of release of β‐glucuronidase from rat polymorphonuclear leukocytes were in the range of 42.2 – 47.1% at a concentration of 10 μm .  相似文献   

17.
We have expanded on the suitability ofp-aminobenzoic acid ethyl ester as an ultraviolet-absorbing reagent [Wanget al., (1984) Anal Biochem 141:366–81] for the analysis of asparagine-linked oligosaccharides derived from glycoproteins. The oligosaccharides released from glycoproteins by hydrazinolysis/N-reacetylation were derivatized withp-aminobenzoic acid ethyl ester and the derivatives were purified and separated into neutral and acidic oligosaccharides on a PRE-SEP C18 cartridge. The acidic oligosaccharides could be further separated into a few species by high-voltage paper electrophoresis. p-Aminobenzoic acid ethyl ester derivatives of neutral oligosaccharides were analyzed by gel permeation chromatography on Bio-Gel P-4 and HPLC on a silica-based amide column. The elution profile and the proportion of the oligosaccharides were in agreement with literature values. The overall yield of oligosaccharides from glycoproteins was approximately 70%. Fifty pmol of oligosaccharide were detectable on Bio-Gel P-4 and 4–5 pmol on HPLC.Abbreviations HPLC high performance liquid chromatography - NABEE p-aminobenzoic acid ethyl ester - FAB-MS fast-atom bombardment mass spectrometry - (GlcNAc)2, (GlcNAc)3, (GlcNAc)4, (GlcNAc)5 and (GlcNAc)6 chito-oligosaccharides containing 2,3,4,5 and 6 residues ofN-acetylglucosamine  相似文献   

18.
Resveratrol is a polyphenolic compound with diverse beneficial effects on human health. Red wine is the major dietary source of resveratrol but the amount that people can obtain from wines is limited. To increase the resveratrol production in wines, two expression vectors carrying 4‐coumarate: coenzyme A ligase gene (4CL) from Arabidopsis thaliana and resveratrol synthase gene (RS) from Vitis vinifera were transformed into industrial wine strain Saccharomyces cerevisiae EC1118. When cultured with 1 mM p‐coumaric acid, the engineered strains grown with and without the addition of antibiotics produced 8.249 and 3.317 mg/L of trans‐resveratrol in the culture broth, respectively. Resveratrol content of the wine fermented with engineered strains was twice higher than that of the control, indicating that our engineered strains could increase the production of resveratrol during wine fermentation. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:650–655, 2015  相似文献   

19.
Poddar‐Sarkar, M., Raha, P., Bhar, R., Chakraborty, A. and Brahmachary, R.L. 2011. Ultrastructure and lipid chemistry of specialized epidermal structure of Indian porcupines and hedgehog. —Acta Zoologica (Stockholm) 92 : 134–140. In the present study, we investigated the ultrastructural variations of specialized epidermal structure of Indian porcupines (Hystrix indica and Atherurus macrourus) and hedgehog (Hemiechinus collaris) as well as the variation in the fatty acid composition of total lipid fraction. Scanning electron microscope images reveal the usual scaly structure in surface view and network of channels in cross‐section but with different orientation of partition walls. The lipid profile reveals the presence of free sterol, long‐chain alcohol, free fatty acids, wax ester and sterol ester in all the three cases and trace amount of triglyceride, diglyceride and monoglyceride. Gas chromatography–mass spectrometry analysis of fatty acid methyl ester of total lipid fraction indicates the presence of C8‐C22 fatty acids in Hystrix indica, C8‐C18 in Atherurus macrourus and C8‐C20 fatty acids in Hemiechinus collaris. It is interesting to note that the total lipid fraction of hedgehog shows no branched‐chain, unsaturated and odd‐carbon fatty acids. Odd‐carbon fatty acid and branched‐chain fatty acids detected in the adult H. indica but were absent in juvenile H. indica as well as in A. macrourus. With the exception of C18:1, the other unsaturated fatty acids were also absent in both juvenile H. indica and A. macrourus.  相似文献   

20.
Yang L  Wei DZ 《Biotechnology letters》2003,25(14):1195-1198
In the enzymatic synthesis of cefaclor, 3-chloro-7-d-(2-phenylglycinamide)-3-cephem-4-carboxylic acid, from phenylglycine methyl ester and 7-aminodesacetoxymethyl-3-chlorocephalosporanic acid, the in situ product could influence both the overall conversion and hydrolysis of the ester. Optimization of the parameters, such as pH 6.2, 5 °C and substrate molar ratio of 2:1, made in situ product removal improve the overall conversion from 64% to 85% (mol/mol).  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