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 469 毫秒
1.
Summary The effects of pH, temperature and concentration of citrate were investigated to achieve an optimal production of diacetyl, acetoin and C2 compounds such as acetaldehyde, acetate and ethanol for free and immobilized cells. The critical conditions of culture, 22°C, pH 4.8, increased the production of C4 compounds (diacetyl, acetoin, 2, 3 butylene glycol), C2 compounds (acetaldehyde, ethanol, acetate) and formate. A higher yield of C2 and C4 compounds was observed for the immobilized cells than for the free cells in continuous culture. At 75 mMol/l of citrate, the citrate bioconversion yield was 42.8% and 80% for free and immobilized cells, respectively. This paper discusses citrate and lactose utilization and NADH2 part on diacetyl reduction.  相似文献   

2.
Citrate utilization and acetoin, diacetyl, acetaldehyde, and lactic acid production in milk at 21 C by five different mixed-strain starters, containing Streptococcus diacetilactis (D type), Leuconostoc (B type), and S. diacetilactis and Leuconostoc (BD type), were measured. BD and D cultures utilized citrate more rapidly and produced more diacetyl, acetoin, and acetaldehyde than B types. All cultures produced much more acetoin than diacetyl, with the BD and D cultures producing four to five times larger amounts of acetoin than the B cultures. Reduction of diacetyl and acetoin toward the end of the normal incubation period was characteristic of BD and D cultures, whereas a similar reduction of acetaldehyde was characteristic of BD and especially of B cultures. Continued incubation of B cultures beyond 17 h also resulted in reduction of diacetyl and acetoin. Addition of citrate to the milk retarded diacetyl and acetoin reduction. Mn2+ had no effect on diacetyl production by a BD culture but increased citrate utilization and, as a consequence, caused greater diacetyl destruction in one of the B cultures.  相似文献   

3.
The production of aroma compounds (acetoin and diacetyl) in fresh unripened cheese by Lactococcus lactis subsp. lactis biovar diacetylactis CNRZ 483 was studied at 30°C at different initial oxygen concentrations (0, 21, 50, and 100% of the medium saturation by oxygen). Regardless of the initial O2 concentration, maximal production of these compounds was reached only after all the citrate was consumed. Diacetyl and acetoin production was 0.01 and 2.4 mM, respectively, at 0% oxygen. Maximum acetoin concentration reached 5.4 mM at 100% oxygen. Diacetyl production was increased by factors of 2, 6, and 18 at initial oxygen concentrations of 21, 50, and 100%, respectively. The diacetyl/acetoin concentration ratio increased linearly with initial oxygen concentration: it was eight times higher at 100% (3.3%) than at 0% oxygen (0.4%). The effect of oxygen on diacetyl and acetoin production was also shown with other lactococci. At 0% oxygen, specific activity of α-acetolactate synthetase (0.15 U/mg) and NADH oxidase (0.04 U/mg) was 3.6 and 5.4 times lower, respectively, than at 100% oxygen. The increasing α-acetolactate synthetase activity in the presence of oxygen would explain the higher production of diacetyl and acetoin. The NADH oxidase activity would replace the role of the lactate dehydrogenase, diacetyl reductase, and acetoin reductase in the reoxidation of NADH, allowing accumulation of these two aroma compounds.  相似文献   

4.
Summary The effects of citrate on diacetyl, acetoin and 2,3-butylene glycol (2,3-BG) production by Leuconostoc mesenteroides subsp. cremoris grown in continuous culture at pH 5.2 were studied. In glucose alone end-product production agreed with the theoretical stoichiometry. In the presence of citrate, lactate and acetate production was higher than the theoretical stoichiometry from glucose. Lactate production was constant when the initial citrate concentration was increased whereas ethanol production strongly decreased. In the absence of citrate, citrate lyase (CL) exhibited weak activity. Diacetyl reductase (DR) and acetoin reductase (AR) exhibited basal activity. When varying citrate concentrations ranging from 10 to 75 mm were added to glucose broth, DR, AR, lactate dehydrogenase, NADH oxidase and alcohol dehydrogenase decreased as the initial citrate concentration increased suggesting that they were partly repressed by citrate. In contrast, CL increased and the specific citrate utilization rate also increased in the same way, indicating no saturation of the first step of citrate metabolism. Acetate kinase (AK) was slightly higher in the presence of citrate and increased when the initial citrate concentration increased. This result was correlated with an increase of acetate from the acetyl phosphate pathway. More ATP was produced in the presence of citrate, which could explain the increase in biomass formation. Citrate bioconversion into diacetyl, acetoin and 2,3-BG increased as the initial citrate increased. Correspondence to: C. Diviès  相似文献   

5.
The influence of pH on the type and concentration of metabolites produced from pyruvate by Lactobacillus plantarum ATCC 8014 was examined in pH-controlled fermentors at pH values of 4.5 to 6.5. Specific growth rates, cell dry weights, and diacetyl concentrations were highest at pH 5.5, with values of 0.78 h−1, 190 mg/liter, and 1.2 mM, respectively. While the conversion efficiency (millimoles of acetoin formed per millimoles of pyruvate utilized) was highest (94.6%) at pH 4.5, acetoin levels were similar (20 mM) between pH 4.5 and 5.5. Feeding stationary-phase cells exogenous pyruvate increased acetoin levels to 78 mM.  相似文献   

6.
Glycerol and glucose fermentation redox routes by Escherichia coli and their regulation by oxidizing and reducing reagents were investigated at different pHs. Cell growth was followed by decrease of pH and redox potential (E h ). During glycerol utilization at pH 7.5 ?pH, the difference between initial and end pH, was lower compared with glucose fermentation. After 8 h growth, during glycerol utilization E h dropped down to negative values (?150 mV) but during glucose fermentation it was positive (+50 mV). In case of glycerol H2 was evolved at the middle log phase while during glucose fermentation H2 was produced during early log phase. Furthermore, upon glycerol utilization, oxidizer potassium ferricyanide (1 mM) inhibited both cell growth and H2 formation. Reducing reagents dl-dithiothreitol (3 mM) and dithionite (1 mM) inhibited growth but stimulated H2 production. The findings point out the importance of reductive conditions for glycerol fermentation and H2 production by E. coli.  相似文献   

7.
A decrease in citric acid and increases in acetic acid, acetoin and diacetyl were found in the test red wine after inoculation of intact cells of Leuconostoc mesenteroides subsp. lactosum ATCC 27307. a malo-lactic bacterium, grown on the malate plus citrate-medium. Citric acid in the buffer solution was transformed to acetic acid, acetoin and diacetyl in the pH range of 2 to 6 after inoculation with intact cells of this bacterial species. It was concluded that citric acid in wine making involving malolactic fermentation, at first, was converted by citrate lyase to acetic and oxaloacetic acids, and the latter was successively transformed by decarboxylation to pyruvic acid which was subsequently converted to acetoin, diacetyl and acetic acid.

Both the activities of citrate lyase and acetoin formation from pyruvic acid in the dialyzed cell-free extract were optimal at pH 6.0. Divalent cations such as Mn2+, Mg2+, Co2+ and Zn2+ activated the citrate lyase. The citrate lyase was completely inhibited by EDTA, Hg2+ and Ag2+ . The acetoin formation from pyruvic acid was significantly stimulated by thiamine pyrophosphate and CoCl2, and inhibited by oxaloacetic acid. Specific activities of the citrate lyase and acetoin formation were considerably variable among the six strains of malo-lactic bacteria examined. Some activities of irreversible reduction of diacetyl to acetoin were found in the cell-free extracts of four of the malo-lactic bacteria strains and the optimal pH was 6.0 for this activity of Leu. mesenteroides.  相似文献   

8.
The effects of citrate on diacetyl and acetoin level by fully grown cells ofStreptococcus lactis subsp.diacetylactis CNRZ 124 were studied. In the absence of citrate, diacetyl synthase as well as acetolactate synthase and acetoin and diacetyl reductases exhibited a basal activity confirming their constitutive nature. However, when initial citrate concentration ranged from 8.8 to 59 mM, the enzyme levels increased in the same way, indicating no saturation rate of citrate metabolism. These results were reflected by a similar enhancement in acetoin and diacetyl production. When citrate was added in fed-batch conditions, its utilization by the fully grown cells led to a twofold increase in diacetyl yield over batch conditions.  相似文献   

9.
Citrate stimulated growth rate, increased the specific lactose consumption rate and enhanced the molar growth yield of Leuconostoc mesenteroides subsp. cremoris growing on lactose at pH 5.2 or 6.2 and at 22 or 30°C. As soon as citrate utilization began, diacetyl and acetoin were produced: 2,3-butylene glycol appeared later while acetoin decreased.  相似文献   

10.
Acidaminococcus fermentans utilized citrate or the citrate analog aconitate as an energy source for growth, and these tricarboxylates were used simultaneously. Citrate utilization and uptake showed biphasic kinetics. High-affinity citrate uptake had a Kt of 40 μM, but the Vmax was only 25 nmol/mg of protein per min. Low-affinity citrate utilization had a 10-fold higher Vmax, but the Ks was greater than 1.0 mM. Aconitate was a competitive inhibitor (Ki = 34μM) of high-affinity citrate uptake, but low-affinity aconitate utilization had a 10-fold-lower requirement for sodium than did low-affinity citrate utilization. On the basis of this large difference in sodium requirements, it appeared that A. fermentans probably has two systems of tricarboxylate uptake: (i) a citrate/aconitate carrier with a low affinity for sodium and (ii) an aconitate carrier with a high affinity for sodium. Citrate was catabolized by a pathway involving a biotin-requiring, avidin-sensitive, sodium-dependent, membrane-bound oxaloacetate decarboxylase. The cells also had aconitase, but this enzyme was unable to convert citrate to isocitrate. Since cell-free extracts converted either aconitate or glutamate to 2-oxoglutarate, it appeared that aconitate was being catabolized by the glutaconyl-CoA decarboxylase pathway. Exponentially growing cultures on citrate or citrate plus aconitate were inhibited by the sodium/proton antiporter, monensin. Because monensin had no effect on cultures growing with aconitate alone, it appeared that citrate metabolism was acting as an inducer of monensin sensitivity. A. fermentans cells always had a low proton motive force (<50 mV), and cells treated with the protonophore TCS (3,3′,4′,5-tetrachlorosalicylanide) grew even though the proton motive force was less than 20 mV. On the basis of these results, it appeared that A. fermentans was depending almost exclusively on a sodium motive force for its membrane energetics.  相似文献   

11.
Two 2,3-butanediol dehydrogenases (enzymes 1 and 2; molecular weight of each, 170,000) have been partially purified from Lactococcus lactis subsp. lactis (Streptococcus diacetylactis) D10 and shown to have reductase activity with either diacetyl or acetoin as the substrate. However, the reductase activity with 10 mM diacetyl was far greater for both enzymes (7.0- and 4.7-fold for enzymes 1 and 2, respectively) than with 10 mM acetoin as the substrate. In contrast, when acetoin and diacetyl were present together, acetoin was the preferred substrate for both enzymes, with enzyme 1 showing the more marked preference for acetoin. meso-2,3-Butanediol was the only isomeric product, with enzyme 1 independent of the substrate combinations. For enzyme 2, both the meso and optical isomers of 2,3-butanediol were formed with acetoin as the substrate, but only the optical isomers were produced with diacetyl as the substrate. With batch cultures of strain D10 at or near the point of citrate exhaustion, the main isomers of 2,3-butanediol present were the optical forms. If the pH was sufficiently high (>pH 5), acetoin reduction occurred over time and was followed by diacetyl reduction, and meso-2,3-butanediol became the predominant isomer. Interconversion of the optical isomers into the meso isomer did occur. The properties of 2,3-butanediol dehydrogenases are consistent with diacetyl and acetoin removal and the appearance of the isomers of 2,3-butanediol.  相似文献   

12.
Summary While the ability of lactobacilli to catabolize pyruvate to a variety of industrially important catabolites is well known, the mechanisms which regulate pyruvate distribution among alternative catabolic pathways is unclear. This paper demonstrates that environmental acidity regulates the catabolic activities ofLactobacillus plantarum cells in chemostat cultures.L. plantarum cells grown in medium containing 100 mM exogenous pyruvate, diverted pyruvate away from lactate to acetoin. Pyruvate uptake and acetoin generation increased under acidic conditions; on a molar basis, pyruvate utilization increased twice as fast as acetoin production, reflecting the 21 stoichiometry of pyruvate incorporation into acetoin. Lactate production increased under alkaline conditions when glucose was fermented to provide endogenous pyruvate. Acetate was formed only at pH 7.5 and 8.0, although acetoin production decreased at elevated pH values. These data indicate thatL. plantarum adjusts to changes in environmental pH by altering its distribution of pyruvate among various catabolites.  相似文献   

13.
14.
Citrate (si)-synthase (citrate oxaloacetate-lyase, EC 4.1.3.7) was purified as an electrophoretically homogeneous protein from a nitrite-oxidizing chemoautotrophic bacterium, Nitrobacter agilis ATCC 14123. The molecular mass (Mr) of the native enzyme was estimated to be about 250,000 by gel filtration, whereas SDS-PAGE gave two bands with Mr values of 45,000 and 80,000, respectively, suggesting that the enzyme is a tetramer consisting of two different subunits (α: 45,000, β: 80,000). The isoelectric point of the enzyme was 5.4. The pH and temperature optima on the citrate synthase activity were about 7.5–8.0 and 30–35°C, respectively. The citrate synthase was stable in the pH range of 6.0–9.0 and up to 55°C. The apparent Km values for oxaloacetate and acetyl-CoA were about 27 μM and 410 μM, respectively. The activity of citrate synthase was not inhibited by ATP (1 mM), NADH (1 mM) or 2-oxoglutarate (10 mM), but was strongly inhibited by SDS (1 mM). Activation by metal ions was not observed.  相似文献   

15.
The extracellular protease of Pseudomonas fluorescens NC 3 was optimally active at 40°C in a reaction mixture containing: 50 mM HEPES (N-2-hydroxyethylpiperazine-N′-2-ethanesulfonic acid) buffer (pH 6.6), 0.5 mM CaCl2, and 25 mg hide powder azure in 5 ml total volume. Divalent cation chelators, i.e., EDTA, o-phenanthroline, citrate or phosphate, inhibited the enzyme. Protease production by P. fluorescens NC 3 was initiated during late-logarithmic-growth phase in a sodium caseinate medium and reached its maximum at the onset of the stationary phase.  相似文献   

16.
Potassium citrate (10 mM, pH 6) inhibits the growth of cultured (Glycine max L.) cells when urea is the sole nitrogen source. Ureadependent citrate toxicity is overcome by three separate additions to the growth medium: (a) NH4Cl (20 mM); (b) high levels of MgCl2 (10 mM) or CaCl2 (5-10 mM); (c) low levels of NiSO4 (10−2 mM). Additions of 10−2 mM NiSO4 not only overcome citrate growth inhibition but the resultant growth is usually better than urea-supported growth in basal medium (neither added citrate nor added nickel). In the absence of added citrate, exceedingly low levels of NiSO4 (10−4 mM) strongly stimulate urea-supported growth in suspension cultures.  相似文献   

17.
Citrate(si)-synthase (citrate oxaloacetate-lyasem EC 4.1.3.7) was purified as an electrophoretically homogeneous protein from an ammonia-oxidizing chemoautotrophic bacterium, Nitrosomonas sp. TK794. The molecular mass of the native enzyme was estimated to be about 287 kDa by gel filtration, whereas SDS-PAGE produced one band with Mr values of 44.7 kDa, suggesting that the enzyme is a hexamer consisting of identical subunits. The isoelectric point of the enzyme was 5.0. The pH and temperature optima for citrate synthase (CS) activity was about 7.5–8.0 and 40°C, respectively. The citrate synthase was stable over a pH range of 6.0–8.5 and up to 40°C. The apparent Km values for oxaloacetate and acetyl-CoA were about 11 μM and 247 μM, respectively. The activity of the citrate synthase was not inhibited by ATP, NADH or 2-oxoglutarate at 5mM, and was activated by potassium chloride at 0.1–100 mM. The N-terminal amino acid sequence of the enzyme protein was PPQDVATLSPGENKKTIELPILG.  相似文献   

18.
Citrate and pyruvate metabolism by nongrowing cells of Leuconostoc oenos was investigated. (sup13)C nuclear magnetic resonance (NMR) spectroscopy was used to elucidate the pathway of citrate breakdown and to probe citrate or pyruvate utilization, noninvasively, in living cell suspensions. The utilization of isotopically enriched substrates allowed us to account for the end products derived from the metabolism of endogenous reserves. The effect of environmental parameters, e.g., pH, gas atmosphere, and presence of malate, on the end products of citrate utilization was studied. Approximately 10% of the citrate supplied was converted to aspartate which remained inside the cells. A metabolic shift with pH was observed, with acetoin production being favored at pH 4, whereas lactate and acetate production increased significantly at higher pH values. The information obtained with NMR was complemented with studies on the relevant enzyme activities in the metabolic pathway of citrate breakdown. The intracellular pH of the cells was strongly dependent on the external pH; this result, together with the determination of the pH profile of the enzymic activities, allowed us to establish the basis for pH regulation; lactate dehydrogenase activity was optimal at pH 7, whereas the acetoin-forming enzymes displayed maximal activities below pH 5. Citrate utilization was also monitored in dilute cell suspensions for comparison with NMR experiments performed with dense suspensions.  相似文献   

19.
Citrate metabolism by Enterococcus faecalis FAIR-E 229 was studied in various growth media containing citrate either in the presence of glucose or lactose or as the sole carbon source. In skim milk (130 mM lactose, 8 mM citrate), cometabolism of citrate and lactose was observed from the first stages of the growth phase. Lactose was stoichiometrically converted into lactate, while citrate was converted into acetate, formate, and ethanol. When de Man-Rogosa-Sharpe (MRS) broth containing lactose (28 mM) instead of glucose was used, E. faecalis FAIR-E 229 catabolized only the carbohydrate. Lactate was the major end product, and small amounts of ethanol were also detected. Increasing concentrations of citrate (10, 40, 70, and 100 mM) added to MRS broth enhanced both the maximum growth rate of E. faecalis FAIR-E 229 and glucose catabolism, although citrate itself was not catabolized. Glucose was converted stoichiometrically into lactate, while small amounts of ethanol were produced as well. Finally, when increasing initial concentrations of citrate (10, 40, 70, and 100 mM) were used as the sole carbon sources in MRS broth without glucose, the main end products were acetate and formate. Small amounts of lactate, ethanol, and acetoin were also detected. This work strongly supports the suggestion that enterococcal strains have the metabolic potential to metabolize citrate and therefore to actively contribute to the flavor development of fermented dairy products.  相似文献   

20.
Having a sufficient supply of energy, usually in the form of ATP, is essential for all living organisms. In this study, however, we demonstrate that it can be beneficial to reduce ATP availability when the objective is microbial production. By introducing the ATP hydrolyzing F1-ATPase into a Lactococcus lactis strain engineered into producing acetoin, we show that production titer and yield both can be increased. At high F1-ATPase expression level, the acetoin production yield could be increased by 10 %; however, because of the negative effect that the F1-ATPase had on biomass yield and growth, this increase was at the cost of volumetric productivity. By lowering the expression level of the F1-ATPase, both the volumetric productivity and the final yield could be increased by 5 % compared to the reference strain not overexpressing the F1-ATPase, and in batch fermentation, it was possible to convert 176 mM (32 g/L) of glucose into 146.5 mM (12.9 g/L) acetoin with a yield of 83 % of the theoretical maximum. To further demonstrate the potential of the cell factory developed, we complemented it with the lactose plasmid pLP712, which allowed for growth and acetoin production from a dairy waste stream, deproteinized whey. Using this cheap and renewable feedstock, efficient acetoin production with a titer of 157 mM (14 g/L) acetoin was accomplished.  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