首页 | 官方网站   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.
Recombinant Escherichia coli cells expressing linoleate 13‐hydratase from Lactobacillus acidophilus were permeabilized by treating with 0.2 M NaCl. The optimal conditions for the production of 13‐hydroxy‐9,15(Z,Z)‐octadecadienoic acid (13‐HODE) from α‐linolenic acid by permeabilized recombinant cells were pH 6.0, 40 °C, 7.5 % (v/v) methanol, 60 g/l permeabilized cells, and 15 g/l α‐linolenic acid. Under these conditions, permeabilized cells produced 7.5 g/l 13‐HODE after 6 h, with a conversion yield of 50 % (w/w) and a volumetric productivity of 1.25 g/l/h. These values were 161 and 160 % of those obtained by nonpermeabilized cells, respectively. To the best of our knowledge, this is the first report on the process optimization for the biotechnological production of 13‐HODE.  相似文献   

2.
Whole cells of recombinant Escherichia coli expressing diol synthase from Aspergillus nidulans produced 5,8‐dihydroxy‐9,12,15(Z,Z,Z)‐octadecatrienoic acid from α‐linolenic acid via 8‐hydroperoxy‐9,12,15(Z,Z,Z)‐octadecatrienoic acid as an intermediate. The optimal conditions for 5,8‐dihydroxy‐9,12,15(Z,Z,Z)‐octadecatrienoic acid production using whole recombinant cells were exhibited at pH 7.0, 40 °C, and 250 rpm with 40 g/L cells, 12 g/L, α‐linolenic acid, and 5 % (v/v) dimethyl sulfoxide in a 250‐mL baffled flask containing 50 mL reaction solution. Under these conditions, whole recombinant cells produced 9.1 g/L 5,8‐dihydroxy‐9,12,15(Z,Z,Z)‐octadecatrienoic acid for 100 min, with a conversion yield of 75 % (w/w), a volumetric productivity of 5.5 g/L/h, and specific productivity of 137 mg/g‐cells/h. As an intermediate, 8‐hydroperoxy‐9,12,15(Z,Z,Z)‐octadecatrienoic acid was observed at approximately 1.4 g/L after 100 min. With regard to dihydroxy fatty acid production, this is the highest reported volumetric and specific productivities thus far. This is the first report on the biotechnological production of 5,8‐dihydroxy‐9,12,15(Z,Z,Z)‐octadecatrienoic acid.  相似文献   

3.
Macrophage apoptosis, a key process in atherogenesis, is regulated by oxidation products, including hydroxyoctadecadienoic acids (HODEs). These stable oxidation products of linoleic acid (LA) are abundant in atherosclerotic plaque and activate PPARγ and GPR132. We investigated the mechanisms through which HODEs regulate apoptosis. The effect of HODEs on THP‐1 monocytes and adherent THP‐1 cells were compared with other C18 fatty acids, LA and α‐linolenic acid (ALA). The number of cells was reduced within 24 hours following treatment with 9‐HODE (p < 0.01, 30 μM) and 13 HODE (p < 0.01, 30 μM), and the equivalent cell viability was also decreased (p < 0.001). Both 9‐HODE and 13‐HODE (but not LA or ALA) markedly increased caspase‐3/7 activity (p < 0.001) in both monocytes and adherent THP‐1 cells, with 9‐HODE the more potent. In addition, 9‐HODE and 13‐HODE both increased Annexin‐V labelling of cells (p < 0.001). There was no effect of LA, ALA, or the PPARγ agonist rosiglitazone (1μM), but the effect of HODEs was replicated with apoptosis‐inducer camptothecin (10μM). Only 9‐HODE increased DNA fragmentation. The pro‐apoptotic effect of HODEs was blocked by the caspase inhibitor DEVD‐CHO. The PPARγ antagonist T0070907 further increased apoptosis, suggestive of the PPARγ‐regulated apoptotic effects induced by 9‐HODE. The use of siRNA for GPR132 showed no evidence that the effect of HODEs was mediated through this receptor. 9‐HODE and 13‐HODE are potent—and specific—regulators of apoptosis in THP‐1 cells. Their action is PPARγ‐dependent and independent of GPR132. Further studies to identify the signalling pathways through which HODEs increase apoptosis in macrophages may reveal novel therapeutic targets for atherosclerosis.  相似文献   

4.
An enzyme from the alga Chlorella pyrenoidosa, previously identified as a hydroperoxide lyase (HPLS), cleaves the 13‐hydroperoxide derivatives of linoleic and linolenic acids into a volatile C5 fragment and a C13 oxo‐product, 13‐oxo‐9(Z),11(E)tridecadienoic acid (13‐OTA). Gas chromatography/mass spectrometry (GC/MS) headspace analysis of the volatile products indicated the formation of pentane when the substrate was the 13‐hydroperoxide derivative of linoleic acid, whereas a more complex mixture of hydrocarbons was formed when the 13‐hydroperoxide derivative of linolenic acid was the substrate. Analysis of the nonvolatile products by GC/MS and liquid chromatography/mass spectrometry (LC/MS) indicated the formation of 13‐OTA along with the 13‐ketone derivative. This enzymatic activity was inhibited by oxygen but was restored with nitrogen. The enzymatic cleavage activity was coincidental in purified fractions with lipoxygenase activity that produced the 13‐ and 9‐hydroperoxide derivatives of linolenic acid. The results suggest that the enzymatic cleavage activity in Chlorella pyrenoidosa was not a consequence of hydroperoxide lyase activity as previously thought, but was due to anaerobic lipoxygenase activity. This enzyme fraction was purified by (NH4)2 SO4 precipitation, gel filtration, and hydrophobic interaction chromatography. The purified enzyme has an approximate MW of 120 KDa and maximum activity at pH 8.0.  相似文献   

5.
Dietary trans monoenes have been associated with an increased risk of heart disease in some studies and this has caused much concern. Trans polyenes are also present in the diet, for example, trans α‐linolenic acid is formed during the deodorisation of α‐linolenic acid‐rich oils such as rapeseed oil. One would expect the intake of trans α‐linolenic acid to be on the increase since the consumption of rapeseed oil in the western diet is increasing. There are no data on trans α‐linolenic acid consumption and its effects. We therefore carried out a comprehensive study to examine whether trans isomers of this polyunsaturated fatty acid increased the risk of coronary heart disease. Since inhibition of Δ6‐desaturase had also been linked to heart disease, the effect of trans α‐linolenic acid on the conversion of [U‐13C]‐labelled linoleic acid to dihomo‐γ‐linolenic and arachidonic acid was studied in 7 healthy men recruited from the staff and students of the University of Edinburgh. Thirty percent of the habitual fat was replaced using a trans ‘free’‐ or ‘high’ trans α‐linolenic acid fat. After at least 6 weeks on the experimental diets, the men received 3‐oleyl, 1,2‐[U‐13C]‐linoleyl glycerol (15 mg twice daily for ten days). The fatty acid composition of plasma phospholipids and the incorporation of 13C‐label into n‐6 fatty acids were determined at day 8, 9 and 10 and after a 6‐week washout period by gas chromatography‐combustion‐isotope ratio mass spectrometry. Trans α‐linolenic acid of plasma phospholipids increased from 0.04 ? 0.01 to 0.17 ? 0.02 and cis ? ‐linolenic acid decreased from 0.42 ? 0.07 to 0.29 ? 0.08 g/100 g of fatty acids on the high trans diet. The composition of the other plasma phospholipid fatty acids did not change. The enrichment of phosphatidyl 13C‐linoleic acid reached a plateau at day 10 and the average of the last 3 days did not differ between the low and high trans period. Both dihomo‐γ‐linolenic and arachidonic acid in phospholipids were enriched in 13C, both in absolute and relative terms (with respect to 13C‐linoleic acid). The enrichment was slightly and significantly higher during the high trans period (P<0.05). Our data suggest that a diet rich in trans α‐linolenic acid (0.6% of energy) does not inhibit the conversion of linoleic acid to dihomo‐γ‐linolenic and arachidonic acid in healthy middle‐aged men consuming a diet rich in linoleic acid.  相似文献   

6.
The asymmetric Sharpless epoxidation of methyl 13S‐hydroxy‐9Z, 11E‐octadeca‐dienoate (13S‐HODE, 1 ) with tert‐butyl hydroperoxide (TBHP) catalysed by titanium tetraisopropoxide {Ti(iOPr)4} in the presence of L(+)‐diisopropyl tartrate (L‐DIPT) gave methyl 13S‐hydroxy‐11S, 12S‐epoxy‐9Z‐octadecenoate 2 (erythro isomer) in 84% diastereomeric excess (de). The epoxidation of 1 with TBHP catalysed by Ti(iOPr)4 in the presence of D(‐)‐DIPT yielded methyl 13S‐hydroxy‐11RR12R‐epoxy‐9Z‐octadecenoate (threo isomer) 3 in 76% de.  相似文献   

7.
BACKGROUND: Biopolymers produced by microbes are in demand as their biodegradable and biocompatible properties make them suitable for disposable products and for potential use as biomaterials for medical applications. The effective microbial production of copolyesters of 3‐hydroxybutyrate (3HB) and 4‐hydroxybutyrate(4HB) with high molar fractions of 4HB unit by a wild‐type Wautersia eutropha H16 was investigated in culture media containing 4‐hydroxybutyric acid (4HBA) and different carbon substrates in the presence of various α‐amino acids. RESULTS: The addition of carbon sources such as glucose, fructose and acetic acid to the culture medium containing 4HBA in the presence of α‐amino acids resulted in the production of random poly(3HB‐co‐4HB) with compositions of up to 77 mol% 4HB unit, but the yields of copolyesters with 60–77 mol% 4HB units were less than 15 wt% of dried cell weights. In contrast, when carbon sources such as propionic acid and butyric acid were used as the co‐substrates of 4HBA in the presence of α‐amino acids, poly(3HB‐co‐4HB) copolyesters with compositions of 72–86 mol% 4HB were produced at maximally 47.2 wt% of dried cell weight (11.3 g L?1) and the molar conversion yield of 4HBA to 4HB fraction in copolyesters was as high as 31.4 mol%. Further, poly(3HB‐co‐4HB) copolyesters with compositions of 93–96 mol% 4HB were isolated at up to 35.2 wt% of dried cell weights by fractionation of the above copolymers with chloroform/n‐hexane. CONCLUSION: The productivity of copolyesters with over 80 mol% 4HB fractions was as high as 0.146 g L?1 h?1 (3.51 g L?1 for 24 h) by flask batch cultivation. Copyright © 2007 Society of Chemical Industry  相似文献   

8.
BACKGROUND: Arthrobacter simplex cells immobilised in sodium cellulose sulfate/poly‐dimethyl‐diallyl‐ammonium chloride microcapsules were used for the microbial dehydrogenation of 11α‐hydroxy‐16α,17‐epoxyprogesterone to 11α‐hydroxy‐16α,17α‐epoxypregn‐1,4‐diene‐3,20‐dione in an aqueous/organic solvent two‐liquid‐phase system, which is a key reaction in the production of glucocorticoid pharmaceuticals. The aim of the study was to establish a suitable aqueous/organic solvent two‐liquid‐phase system for performing semi‐continuous production in an airlift loop reactor by encapsulated A. simplex cells with the addition of suitable surfactants to achieve a higher yield of the product. RESULTS: n‐Hexane was selected as the most suitable organic solvent. In optimised Tween‐80 emulsion feed mode the conversion in the airlift loop reactor was as high as 97.54% when the time of reaction was 2 h, and the reaction time was greatly shortened. In semi‐continuous production the cultivation with immobilised cells was carried out for five batches in total. The conversion in each batch was above 95% and the enzymatic activity still remained quite high after five batches of biotransformation. CONCLUSION: The results showed that performing the conversion by this method shortened the reaction time and increased the productivity, thus demonstrating the great potential of the method for the dehydrogenation of 11α‐hydroxy‐16α,17‐epoxyprogesterone. Copyright © 2008 Society of Chemical Industry  相似文献   

9.
A new immobilized biocatalyst based on Rhizopus oryzae fungal cells entrapped in poly(vinyl alcohol)‐cryogel was evaluated in both the batch and semi‐batch processes of L (+)‐lactic acid (LA) production, when glucose, acid hydrolysates of starch or gelatinized potato starch were used as the main substrates. Under the batch conditions, the immobilized biocatalyst developed produced LA with yields of 94% and 78% from glucose and acid starch hydrolysates, respectively. Semi‐batch conditions enabled product yields of 52% and 45% to be obtained with the corresponding substrates. The highest process productivity (up to 173 g L?1) was reached under semi‐batch conditions. Potato starch (5–70 g L?1) was also transformed into lactic acid by immobilized R. oryzae. It was shown that long‐term operation of the immobilized biocatalyst (for at least 480 h) produced a low decrease in metabolic activity. Copyright © 2006 Society of Chemical Industry  相似文献   

10.
The seed oil of Arum maculatum has been found to contain 13‐phenyltridec‐9‐enoic (0.4%) and 15‐phenyl‐pentadec‐9‐enoic (1%) acids, detected by gas chromatographymass spectrometry of the picolinyl ester and related derivatives.  相似文献   

11.
Highly purified cellulose preparations were obtained by pretreatment of dewaxed barley straw, oil palm frond fiber, poplar wood, maize stems, wheat straw, rice straw, and rye straw with 2.0% H2O2 at 45°C and pH 11.6 for 16 h, and sequential purification with 80% acetic acid–70% nitric acid (10/1, v/v) at 120°C for 15 min. The purified cellulose obtained was relatively free of bound hemicelluloses (2.3–3.2%) and lignin (0.4–0.6%) and had a yield of 35.5% from barley straw, 39.6% from oil palm frond fiber, 40.8% from poplar wood, 36.0% from maize stems, 34.1% from wheat straw, 23.4% from rice straw, and 35.8% from rye straw. The weight‐average molecular weights of the purified cellulose ranged from 39,030 to 48,380 g/mol. The thermal stability of the purified cellulose was higher than that of the corresponding crude cellulose. In comparison, the isolated crude and purified cellulose samples were also studied by Fourier transform IR and cross‐polarization/magic‐angle spinning 13C‐NMR spectroscopy. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 322–335, 2005  相似文献   

12.
Linoleate 9R-lipoxygenase (9R-LOX) from Nostoc sp. SAG 25.82 was identified as arachidonate (ARA) 11R-LOX by the determination of the product obtained from the conversion of ARA. The specific activity and catalytic efficiency (kcat/Km) of the enzyme for C20 and C22 polyunsaturated fatty acids followed the order ARA > eicosapentaenoic acid > docosahexaenoic acid. The production of the lipid mediator 11R-hydroxyeicosatetraenoic acid (11R-HETE) was performed using Escherichia coli cells expressing ARA 11R-LOX from Nostoc sp. The reaction conditions, such as pH, temperature, solvent and its concentration, and substrate and cell concentrations, were optimized using the recombinant cells, and the optimal conditions for the production of 11R-HETE from ARA were pH 7.0, 25°C, 10 g L−1 cells, 5.0 mM ARA, 4% (v/v) ethanol, and 10 mM cysteine as a reducing agent. Under these optimized conditions, E. coli cells expressing 11R-LOX converted 5.0 mM ARA into 4.74 mM 11R-HETE in 60 min, with a molar conversion yield of 95% a volumetric productivity of 79 μM min−1 and a specific productivity of 7.9 μM min−1 g−1. To the best of our knowledge, this is the first report on the quantitative biotechnological production of 11R-HETE.  相似文献   

13.
Δ6‐desaturase is located in a pivotal position in the metabolism of essential fatty acids (EFA). Various methods have been used to estimate Δ6‐desaturase activity, including the assessment of: (i) tissue fatty acid compositions (and associated product/precursor ratios), (ii) Δ6‐desaturase activities ex vivo, and (iii) isotopically labelled linoleic acid metabolism in vivo. This review critically examines these methods and considers their appropriateness and reliability in assessing linoleic acid metabolism in diabetes and cardiovascular disease. In general, there was a good agreement between the three methods and the effect of experimental diabetes on linoleic acid metabolism. In humans, however, the effect of diabetes on tissue fatty acid composition was inconsistent, and there was a paucity of data on linoleic acid metabolism ex vivo and in vivo. The inconsistency in human fatty acid compositional data may relate to variable and uncontrolled intakes of linoleic acid and its n‐6 metabolites, but also to a less extreme insulin deficiency as studied in animals. Risk markers for cardiovascular disease generally reduced rat liver Δ6‐desaturase activity ex vivo. This was not, however, reflected in tissue fatty acid compositions in these controlled studies. Linoleic acid metabolism, as determined by tissue fatty acid composition in humans, is reduced in cardiovascular disease; however, the omnivorous dietary patterns and decreased linoleic acid intakes make this conclusion potentially unreliable. Few stable‐isotope studies have been conducted on the effect of cardiovascular risk markers on linoleic acid metabolism, and there is a requirement for this type of work to be standardised to facilitate inter‐study comparisons. These studies may eventually help optimise EFA intake in health and disease conditions.  相似文献   

14.
A novel enzymatic production system of optically pure β‐hydroxy α‐amino acids was developed. Two enzymes were used for the system: an N‐succinyl L ‐amino acid β‐hydroxylase (SadA) belonging to the iron(II)/α‐ketoglutarate‐dependent dioxygenase superfamily and an N‐succinyl L ‐amino acid desuccinylase (LasA). The genes encoding the two enzymes are part of a gene set responsible for the biosynthesis of peptidyl compounds found in the Burkholderia ambifaria AMMD genome. SadA stereoselectively hydroxylated several N‐succinyl aliphatic L ‐amino acids and produced N‐succinyl β‐hydroxy L ‐amino acids, such as N‐succinyl‐L ‐β‐hydroxyvaline, N‐succinyl‐L ‐threonine, (2S,3R)‐N‐succinyl‐L ‐β‐hydroxyisoleucine, and N‐succinyl‐L ‐threo‐β‐hydroxyleucine. LasA catalyzed the desuccinylation of various N‐succinyl‐L ‐amino acids. Surprisingly, LasA is the first amide bond‐forming enzyme belonging to the amidohydrolase superfamily, and has succinylation activity towards the amino group of L ‐leucine. By combining SadA and LasA in a preparative scale production using N‐succinyl‐L ‐leucine as substrate, 2.3 mmol of L ‐threo‐β‐hydroxyleucine were successfully produced with 93% conversion and over 99% of diastereomeric excess. Consequently, the new production system described in this study has advantages in optical purity and reaction efficiency for application in the mass production of several β‐hydroxy α‐amino acids.

  相似文献   


15.
Linoleic acid isomerases (LAI) are enzymes responsible for conversion of linoleic acid (LA) to conjugated linoleic acids (CLA) in different isomeric forms. CLAs are well known for numerous beneficial effects as functional foods. Despite identification of several LAI producing‐bacteria and release of their LAI nucleotide sequences to Bio‐Banks, no related bioinformatics study on these important enzymes is addressed yet. To gain insights into the structural/functional and phylogentic relations of LAIs as well as recombinant production of the desired enzyme, we employed several bioinformatical tools to analyze their primary structure and physicochemical properties. The results indicated that LAIs produced in different bacterial strains might be divided in two distinct families (Propionibacterium acnes and myosin cross‐reactive antigen (MCRA)‐like LAIs) with specific isomerase activities. In another part of the study, physicochemical properties, solubility upon over expression in E. coli, disulfide bond formation, and potential glycosylation sites in LAI sequences were predicted using bioinformatics tools and the most appropriate strategy for recombinant production of each LAI was determined. Our predicted data may be useful for further experimental studies toward production of the desired LAI.  相似文献   

16.
N‐Acetyl‐D ‐neuraminic acid (Neu5Ac) was efficiently synthesized from lactate and a mixture of N‐acetyl‐D ‐glucosamine (GlcNAc) and N‐acetyl‐D ‐mannosamine (ManNAc) by whole cells. The biotransformation utilized Escherichia coli cells (Neu5Ac aldolase), Pseudomonas stutzeri cells (lactate oxidase components), GlcNAc/ManNAc and lactate. By this process, 18.32±0.56 g/liter Neu5Ac were obtained from 65.61±2.70 g/liter lactate as an initial substrate input. Neu5Ac (98.4±0.4 % purity, 80.87±0.79 % recovery yield) was purified by anionic exchange chromatography. Our results demonstrate that the reported Neu5Ac biosynthetic process can compare favorably with natural product extraction or chemical synthesis processes.  相似文献   

17.
A new monomer, exo‐3,6‐epoxy‐1,2,3,6‐tetrahydrophthalimidocaproic acid (ETCA), was prepared by reaction of maleimidocaproic acid and furan. The homopolymer of ETCA and its copolymers with acrylic acid (AA) or with vinyl acetate (VAc) were obtained by photopolymerizations using 2,2‐dimethoxy‐2‐phenylacetophenone as an initiator at 25 °C. The synthesized ETCA and its polymers were identified by FTIR, 1H NMR and 13C NMR spectroscopies. The apparent average molecular weights and polydispersity indices determined by gel permeation chromatography (GPC) were as follows: Mn = 9600 g mol?1, Mw = 9800 g mol?1, Mw/Mn = 1.1 for poly(ETCA); Mn = 14 300 g mol?1, Mw = 16 200 g mol?1, Mw/Mn = 1.2 for poly(ETCA‐co‐AA); Mn = 17 900 g mol?1, Mw = 18 300 g mol?1, Mw/Mn = 1.1 for poly(ETCA‐co‐VAc). The in vitro cytotoxicity of the synthesized compounds against mouse mammary carcinoma and human histiocytic lymphoma cancer cell lines decreased in the following order: 5‐fluorouracil (5‐FU) ≥ ETCA > polymers. The in vivo antitumour activity of the polymers against Balb/C mice bearing sarcoma 180 tumour cells was greater than that of 5‐FU at all doses tested. © 2001 Society of Chemical Industry  相似文献   

18.
The effect of pretreatment of molasses, nitrogen sources, natural oils, fatty acids, antioxidant, precursors, and mixtures of the above substances on β‐carotene production by Blakeslea trispora in shake flask culture was investigated. Also, a central composite design was employed to determine the maximum β‐carotene concentration at optimum values for the process variables (linoleic acid, kerosene, antioxidant). The highest concentration of the carotenoid pigment was obtained in molasses solution treated with invertase. Corn steep liquor and yeast extract at concentrations of 5.0% and 0.5% (w/v), respectively, increased slightly the concentration of β‐carotene, while the natural oils, fatty acids, and precursors (except kerosene) did not improve the production of pigment when they were added separately to the medium. On the other hand, the mixture of linoleic acid, kerosene and antioxidant increased significantly the concentration of β‐carotene. The fit of the model was found to be good. Linoleic acid, kerosene and antioxidant had a strong linear effect on β‐carotene concentration. The concentration of β‐carotene was significantly affected by linoleic acid–antioxidant and kerosene–antioxidant interactions as well as by the negative quadratic effects of these variables. The interaction between linoleic acid–kerosene had no significant linear effect. Maximum β‐carotene concentration (790.0 mg dm?3) was obtained in culture grown in molasses solution supplemented with linoleic acid (30.74 g dm?3), kerosene (27.79 g dm?3) and antioxidant (10.22 g dm?3). © 2002 Society of Chemical Industry  相似文献   

19.
Polyunsaturated fatty acids (PUFA) are important ingredients of human diet because of their prominent role in the function of human brain, eye and kidney. α‐Linolenic acid (ALA), a C18, n‐3 PUFA is a precursor of long chain PUFA in humans. Commercial lipases of Candida rugosa, Pseudomonas cepacea, Pseudomonas fluorescens, and Rhizomucor miehei were used for hydrolysis of flax seed oil. Reversed phase high performance liquid chromatography followed by gas chromatography showed that the purified oil contained 12 triacylglycerols (TAGs) with differences in fatty acid compositions. Flax seed oil TAGs contained α‐linolenic acid (50%) as a major fatty acid while palmitic, oleic, linoleic made up rest of the portion. Among the four commercial lipases C. rugosa has preference for ALA, and that ALA was enriched in free fatty acids. C. rugosa lipase mediated hydrolysis of the TAGs resulted in a fatty acid mixture that was enriched in α‐linolenic to about 72% yield that could be further enriched to 80% yield by selective removal of saturated fatty acids by urea complexation. Such purified ALA can be used for preparation of ALA‐enriched glycerides. Practical applications : This methodology allows purifying ALA from fatty acid mixture obtained from flax seed oil by urea complexation.  相似文献   

20.
4‐Hydroxy‐2‐trans‐nonenal (HNE) is a toxic aldehyde produced mostly in oils containing polyunsaturated fatty acid due to heat‐induced lipid peroxidation. The present study examined the effects of the heating time, the degree of unsaturation, and the antioxidant potential on the formation of HNE in two light olive oils (LOO) and two sunflower oils (one high oleic and one regular) at frying temperature. HNE concentrations in these oil samples heated for 0, 1, 3, and 5 hours at 185 °C were measured using high‐performance liquid chromatography. The fatty‐acid distribution and the antioxidant capacity of these four oils were also analyzed. The results showed that all oils had very low HNE concentrations (<0.5 μg g?1 oil) before heating. After 5 hours of heating at 185 °C, HNE concentrations were increased to 17.98, 25.00, 12.51, and 40.00 μg g?1 in the two LOO, high‐oleic sunflower oil (HOSO), and regular sunflower oil (RSO), respectively. Extending the heating time increased HNE formation in all oils tested. It is related to their fatty‐acid distributions and antioxidant capacities. RSO, which contained high levels of linoleic acid (59.60%), a precursor for HNE, was more susceptible to degradation and HNE formation than HOSO and LOO, which contained only 6–8% linoleic acid.  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-23

京公网安备 11010802026262号