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 31 毫秒
1.
Initial rate and biotransformation studies were applied to refine and validate a mathematical model for enzymatic (R)-phenylacetylcarbinol (PAC) production from pyruvate and benzaldehyde using Candida utilis pyruvate decarboxylase (PDC). The rate of PAC formation was directly proportional to the enzyme activity level up to 5.0 U ml-1 carboligase. Michaelis-Menten kinetics were determined for the effect of pyruvate concentration on the reaction rate. The effect of benzaldehyde followed the sigmoidal shape of the Monod-Wyman-Changeux (MWC) model. The biotransformation model, which also included a term for PDC inactivation by benzaldehyde, was used to determine the overall rate constants for the formation of PAC, acetaldehyde, and acetoin. These values were determined from data for three batch biotransformations performed over a range of initial concentrations (viz. 50-150 mM benzaldehyde, 60-180 mM pyruvate, 1.1-3.4 U ml-1 enzyme activity). The finalized model was then used to predict a batch biotransformation profile at 120/100 mM initial pyruvate/benzaldehyde (initial enzyme activity 3.0 U ml-1). The simulated kinetics gave acceptable fitting (R2 = 0.9963) to the time courses of these latter experimental data for substrates pyruvate and benzaldehyde, product PAC, by-products acetaldehyde and acetoin, as well as enzyme activity level.  相似文献   

2.
Aqueous/organic two-phase systems have been evaluated for enhanced production of (R)-phenylacetylcarbinol (PAC) from pyruvate and benzaldehyde using partially purified pyruvate decarboxylase (PDC) from Candida utilis. In a solvent screen, octanol was identified as the most suitable solvent for PAC production in the two-phase system in comparison to butanol, pentanol, nonanol, hexane, heptane, octane, nonane, dodecane, methylcyclohexane, methyl tert butyl ether, and toluene. The high partitioning coefficient of the toxic substrate benzaldehyde in octanol allowed delivery of large amounts of benzaldehyde into the aqueous phase at a concentration less than 50 mM. PDC catalyzed the biotransformation of benzaldehyde and pyruvate to PAC in the aqueous phase, and continuous extraction of PAC and byproducts acetoin and acetaldehyde into the octanol phase further minimized enzyme inactivation, and inhibition due to acetaldehyde. For the rapidly stirred two-phase system with a 1:1 phase ratio and 8.5 U/mL carboligase activity, 937 mM (141 g/L) PAC was produced in the octanol phase in 49 h with an additional 127 mM (19 g/L) in the aqueous phase. Similar concentrations of PAC could be produced in the slowly stirred phase separated system at this enzyme level, although at a much slower rate. However at lower enzyme concentration very high specific PAC production (128 mg PAC/U carboligase at 0.9 U/mL) was achieved in the phase separated system, while still reaching final PAC levels of 102 g/L in octanol and 13 g/L in the aqueous phase. By comparison with previously published data by our group for a benzaldehyde emulsion system without octanol (50 g/L PAC, 6 mg PAC/U carboligase), significantly higher PAC concentrations and specific PAC production can be achieved in an octanol/aqueous two-phase system.  相似文献   

3.
Summary Production of phenylacetylcarbinol (PAC) was measured in various yeast species. The yeast strains tested were cultivated under submerged conditions in a medium containing corn steep and sucrose as the main components; sucrose, acetaldehyde and benzaldehyde were added to the grown cultures. In a first series of experiments the initial rate of PAC production, i.e. the PAC production determined 30 min after the addition of benzaldehyde was determined in 38 yeast strains, mostly of the generaSaccharomyces andCandida. The amount of PAC produced varied from zero (12 strains) to 1.24 mg ml–1. In a second series of experiments, 15 strains, which in the first series had shown a higher PAC production, were further tested. Sucrose, acetaldehyde and benzaldehyde were added to the cultures until the PAC production ceased. The highest PAC production (6.3 mg ml–1) was reached in the strainSaccharomyces carlsbergensis Budvar; the production was slightly lower in 4 strains of the generaSaccharomyces, Candida andHansenula.  相似文献   

4.
Loss of substrate, pyruvate, a limitation for enzymatic batch production of (R)-phenylacetylcarbinol (PAC), resulted from two phenomena: temperature dependent non-enzymatic concentration decrease due to the cofactor Mg2+ and formation of by-products, acetaldehyde and acetoin, by pyruvate decarboxylase (PDC). In the absence of enzyme, pyruvate stabilization was achieved by lowering the Mg2+ concentration from 20 to 0.5 mM. With 0.5 mM Mg2+ Rhizopus javanicus and Candida utilis PDC produced similar levels of PAC (49 and 51 g l–1, respectively) in 21 h at 6 °C; however C. utilis PDC formed less by-product from pyruvate and was more stable during biotransformation. The process enhancements regarding Mg2+ concentration and source of PDC resulted in an increase of molar yield (PAC/consumed pyruvate) from 59% (R. javanicus PDC, 20 mM Mg2+) to 74% (R. javanicus PDC, 0.5 mM Mg2+) to 89% (C. utilis PDC, 0.5 mM Mg2+).  相似文献   

5.
(R)-Phenylacetylcarbinol (PAC), a pharmaceutical precursor, was produced from benzaldehyde and pyruvate by pyruvate decarboxylase (PDC) of Candida utilis in an aqueous/organic two-phase emulsion reactor. When the partially purified enzyme in this previously established in vitro process was replaced with C. utilis cells and the temperature was increased from 4 to 21 °C, a screen of several 1-alcohols (C4–C9) confirmed the suitability of 1-octanol as the organic phase. Benzyl alcohol, the major by-product in the commercial in vivo conversion of benzaldehyde and sugar to PAC by Saccharomyces cerevisiae, was not formed. With a phase volume ratio of 1:1 and 5.6 g C. utilis l−1 (PDC activity 2.5 U ml−1), PAC levels of 103 g l−1 in the octanol phase and 12.8 g l−1 in the aqueous phase were produced in 15 h at 21 °C. In comparison to our previously published process with partially purified PDC in an aqueous/octanol emulsion at 4 °C, PAC was produced at a 4-times increased specific rate (1.54 versus 0.39 mg U−1 h−1) with simplified catalyst production and reduced cooling cost. Compared to traditional in vivo whole cell PAC production, the yield on benzaldehyde was 26% higher, the product concentration increased 3.9-fold (or 6.9-fold based on the organic phase), the productivity improved 3.1-fold (3.9 g l−1 h−1) and the catalyst was 6.9-fold more efficient (PAC/dry cell mass 10.3 g g−1).*Dedicated with gratitude to Prof. Dr. Franz Lingens – “Theo”.  相似文献   

6.
In addition to the decarboxylation of 2-oxo acids, thiamin diphosphate (ThDP)-dependent decarboxylases/dehydrogenases can also carry out so-called carboligation reactions, where the central ThDP-bound enamine intermediate reacts with electrophilic substrates. For example, the enzyme yeast pyruvate decarboxylase (YPDC, from Saccharomyces cerevisiae) or the E1 subunit of the Escherichia coli pyruvate dehydrogenase complex (PDHc-E1) can produce acetoin and acetolactate, resulting from the reaction of the central thiamin diphosphate-bound enamine with acetaldehyde and pyruvate, respectively. Earlier, we had shown that some active center variants indeed prefer such a carboligase pathway to the usual one [Sergienko, Jordan, Biochemistry 40 (2001) 7369–7381; Nemeria et al., J. Biol. Chem. 280 (2005) 21,473–21,482]. Herein is reported detailed analysis of the stereoselectivity for forming the carboligase products acetoin, acetolactate, and phenylacetylcarbinol by the E477Q and D28A YPDC, and the E636A and E636Q PDHc-E1 active-center variants. Both pyruvate and β-hydroxypyruvate were used as substrates and the enantiomeric excess was analyzed by a combination of NMR, circular dichroism and chiral-column gas chromatographic methods. Remarkably, the two enzymes produced a high enantiomeric excess of the opposite enantiomer of both acetoin-derived and acetolactate-derived products, strongly suggesting that the facial selectivity for the electrophile in the carboligation is different in the two enzymes. The different stereoselectivities exhibited by the two enzymes could be utilized in the chiral synthesis of important intermediates.  相似文献   

7.
Considerable evidence indicates that acetaldehyde is released from the leaves of a variety of plants. The conventional explanation for this is that ethanol formed in the roots is transported to the leaves where it is converted to acetaldehyde by the alcohol dehydrogenase (ADH) found in the leaves. It is possible that acetaldehyde could also be formed in leaves by action of pyruvate decarboxylase (PDC), an enzyme with an uncertain metabolic role, which has been detected, but not characterized, in cottonwood leaves. We have found that leaf PDC is present in leaf veins and petioles, as well as in non-vein tissues. Veins and petioles contained measurable pyruvate concentrations in the range of 2 mM. The leaf vein form of the enzyme was purified approximately 143-fold, and, at the optimum pH of 5.6, the Km value for pyruvate was 42 μM. This Km is lower than the typical millimolar range seen for PDCs from other sources. The purified leaf PDC also decarboxylates 2-ketobutyric acid (Km = 2.2 mM). We conclude that there are several possible sources of acetaldehyde production in cottonwood leaves: the well-characterized root-derived ethanol oxidation by ADH in leaves, and the decarboxylation of pyruvate by PDC in leaf veins, petioles, and other leaf tissues. Significantly, the leaf vein form of PDC with its high affinity for pyruvate, could function to shunt pyruvate carbon to the pyruvate dehydrogenase by-pass and thus protect the metabolically active vascular bundle cells from the effects of oxygen deprivation.  相似文献   

8.
Summary The rate of production ofl-phenylacetyl carbinol bySaccharomyces cerevisiae in reaction mixtures containing benzaldehyde with sucrose or pyruvate as cosubstrate was investigated in short 1 h incubations. The effect of yeast dose rate, sucrose and benzaldehyde concentration and pH on the rate of reaction was determined. Maximum biotransformation rates were obtained with concentrations of benzaldehyde, sucrose and yeast of 6 g, 40 g and 60 g/l, respectively. Negligible biotransformation rates were observed at a concentration of 8 g/l benzaldehyde. The reaction had a pH optimum of 4.0–4.5. Rates of bioconversion of benzaldehyde and selected substituted aromatic aldehydes using both sucrose and sodium pyruvate as cosubstrate were compared. The rate of aromatic alcohol production was much higher when sucrose was used rather than pyruvate.o-Tolualdehyde and 1-chlorobenzaldehyde were poor substrates for aromatic carbinol formation although the latter produced significant aromatic alcohol in sucrose-containing media. Yields of 2.74 and 3.80 g/l phenylacetyl carbinol were produced from sucrose and pyruvate, respectively, in a 1 h reaction period.  相似文献   

9.
Biotransformation of benzaldehyde and pyruvate into (R)-phenylacetylcarbinol (PAC) catalysed by Candida utilis pyruvate decarboxylase (PDC) at low buffer concentration (20 mM MOPS) was enhanced by maintenance of neutral pH through acetic acid addition. PDC was very stable in this buffer (half-life 138 h at 6 degrees C), however a benzaldehyde emulsion (400 mM) caused rapid deactivation. The inclusion of 2M glycerol did not protect PDC from inactivation by benzaldehyde but initial rates were increased by 50% and the final PAC level was enhanced from 40 to 51 g l(-1). Low levels of by-products acetaldehyde (0.1-0.15 g l(-1)) and acetoin (1.1-1.3 g l(-1)) were formed in both the presence and absence of 2 M glycerol. Interestingly PDC was more stable towards benzaldehyde when pyruvate was present: no activity was lost during the first hour of biotransformation (2 M glycerol, benzaldehyde concentration decreased from 400 to 345 mM, pyruvate from 480 to 420 mM) but PDC was completely inactivated in less than 30 min when exposed to the same concentrations of benzaldehyde in the absence of pyruvate. Thus the enzyme in catalytic action was more stable than the resting enzyme.  相似文献   

10.
Several 2-keto acid decarboxylases catalyse an acyloin condensation-like carboligase reaction beside their physiological decarboxylase activity. Although many data concerning stability and catalytic potential of these enzymes are available, a standard evaluation under similar reaction conditions is lacking. In this comprehensive survey we assemble already published data combined with new studies of three bacterial pyruvate decarboxylases, yeast pyruvate decarboxylase, benzoylformate decarboxylase from Pseudomonas putida (BFD) and the branched-chain 2-keto acid decarboxylase from Lactococcus lactis (KdcA). The obtained results proof that the optima for activity and stability are rather similar if comparable reaction conditions are used. Although the substrate ranges of the decarboxylase reaction of the various pyruvate decarboxylases are similar as well, they differ remarkably from those of BFD and KdcA. We further show that the range of acceptable donor aldehydes for the carboligase reaction of a respective enzyme can be reliably predicted from the substrate range of decarboxylase reaction.  相似文献   

11.
Summary Two yeast strains, Saccharomyces fermentati and S. delbrueckii, have been found that promote acyloin condensation with yields higher than previously reported with other microorganisms. Benzaldehyde is the best substrate for the reaction in which a C2 unit from pyruvate is condensed. The ability to condense other aldehydes and other oxoacids were significant only with S. cerevisiae. Purified pyruvate decarboxylase from yeast was effective in condensing benzaldehyde with the C2 unit coming from pyruvate, but surprisingly less efficient than S. cerevisiae in reaction with other substrates in water. Offprint requests to: S. Servi  相似文献   

12.
Recent progress in enzymatic (R)-phenylacetylcarbinol (PAC) production has established the need for low cost and efficient biocatalyst preparation. Pyruvate decarboxylase (PDC) added in the form of Candida utilis cells showed higher stability towards benzaldehyde and temperature in comparison with partially purified preparations. In the presence of 50 mM benzaldehyde and at 4°C, a half-life of 228 h was estimated for PDC added as C. utilis cells, in comparison with 24 h for the partially purified preparation. Increasing the temperature from 4 to 21°C for PAC production with C. utilis cells resulted in similar final PAC levels of 39 and 43 g l−1 (258 and 289 mM), respectively, from initial 300 mM benzaldehyde and 364 mM pyruvate. The overall volumetric productivity was enhanced 2.8-fold, which reflected the 60% shorter reaction time at the higher temperature. Enantiomeric excess values of 98 and 94% for R-PAC were obtained at 4 and 21°C, respectively, and benzyl alcohol (a potential by-product from benzaldehyde) was not formed.  相似文献   

13.
Summary In Mexico there are different alcoholic beverages produced from agave juices from different agave plants, which are cooked, fermented and distilled. For tequila production only Agave tequilana is allowed. In this study we compared yeast strains of different species from different origin (agave and grape juice) for parameters of technological interest, such as SO2 and copper resistance, ethanol tolerance and enzymatic activities. All agave strains were found to be more resistant to SO2 and agave non-Saccharomyces yeasts were more tolerant to ethanol, whereas grape strains exhibited positive results for β-glucosidase and β-xylosidase activities. As regards fermentations of Agave tequilana juice with ethanol added at different concentrations, only agave Saccharomyces strains were more tolerant to ethanol than grape strains.  相似文献   

14.
Fourteen strains of the yeastSaccharomyces cerevisiae were isolated from three wineries in the Salnés wine region (N.W. Spain) at the three different periods of the natural fermentation. Each wild yeast was screened for production of acetaldehyde, ethyl acetate, isobutanol,n-propanol, amylic alcohol and other important enological compounds during laboratory scale fermentations of grape juice. After 25 days at 20°C, the analytical results evidenced variations in the production of acetaldehyde (from 13.1 to 24.3 mg/l), isobutanol (from 27.7 to 51.1 mg/l), amyl alcohols (from 111 to 183 mg/l) and ethyl acetate (from 19.3 to 43.7 mg/l). Although isolated from the same wine region, differences in the wine composition were observed depending on the particular yeast strain used for the vinification experiments.  相似文献   

15.
The effect of decreasing the organic (octanol) to aqueous phase volume ratio was evaluated in a two-phase enzymatic process for (R)-phenylacetylcarbinol (PAC) production. Decreasing the ratio from 1:1 to 0.43:1 at 4°C increased PAC in the organic phase from 112 g/l to 183 g/l with a 10% improvement in overall productivity. Interestingly, the rate of enzyme (pyruvate decarboxylase) activity loss was unaffected by the reduced phase ratio over the reaction period (48 h). At 20°C and 0.43:1 phase ratio the organic phase PAC concentration increased to 212 g/l and the overall productivity increased by 30% although the PAC yield (based on pyruvate) declined by about 10% due to greater byproduct acetoin formation at the higher temperature. Product recovery in such a system is facilitated both by the higher PAC concentration and the reduced organic phase volume.  相似文献   

16.
Esters are an important group of volatile compounds that can contribute to wine flavour. Wine lactic acid bacteria (LAB) have been shown to produce esterases capable of hydrolysing ester substrates. This study aims to characterise the esterase activities of nine LAB strains under important wine conditions, namely, acidic conditions, low temperature (to 10°C) and in the presence of ethanol (2–18% v/v). Esterase substrate specificity was also examined using seven different ester substrates. The bacteria were generally found to have a broad pH activity range, with the majority of strains showing maximum activity close to pH 6.0. Exceptions included an Oenococcus oeni strain that retained most activity even down to a pH of 4.0. Most strains exhibited highest activity across the range 30–40°C. Increasing ethanol concentration stimulated activity in some of the strains. In particular, O. oeni showed an increase in activity up to a maximum ethanol concentration of around 16%. Generally, strains were found to have greater activity towards short-chained esters (C2–C8) compared to long-chained esters (C10–C18). Even though the optimal physicochemical conditions for enzyme activity differed from those found in wine, these findings are of potential importance to oenology because significant activities remained under wine-like conditions.  相似文献   

17.
Glycerol formation is vital for reoxidation of nicotinamide adenine dinucleotide (reduced form; NADH) under anaerobic conditions and for the hyperosmotic stress response in the yeast Saccharomyces cerevisiae. However, relatively few studies have been made on hyperosmotic stress under anaerobic conditions. To study the combined effect of salt stress and anaerobic conditions, industrial and laboratory strains of S. cerevisiae were grown anaerobically on glucose in batch-cultures containing 40 g/l NaCl. The time needed for complete glucose conversion increased considerably, and the specific growth rates decreased by 80–90% when the cells were subjected to the hyperosmotic conditions. This was accompanied by an increased yield of glycerol and other by-products and reduced biomass yield in all strains. The slowest fermenting strain doubled its glycerol yield (from 0.072 to 0.148 g/g glucose) and a nearly fivefold increase in acetate formation was seen. In more tolerant strains, a lower increase was seen in the glycerol and in the acetate, succinate and pyruvate yields. Additionally, the NADH-producing pathway from acetaldehyde to acetate was analysed by overexpressing the stress-induced gene ALD3. However, this had no or very marginal effect on the acetate and glycerol yields. In the control experiments, the production of NADH from known sources well matched the glycerol formation. This was not the case for the salt stress experiments in which the production of NADH from known sources was insufficient to explain the formed glycerol.  相似文献   

18.
Cell suspensions of Zymomonas mobilis and Saccharomyces carlsbergensis and the pyruvate decarboxylases from the two organisms were compared with respect to their efficiencies of acyloin formation. Although Z. mobilis contained five times more pyruvate decarboxylase activity than yeast, sugar-fermenting suspensions of Z. mobilis produced, in the presence of benzaldehyde, 4-5 times less phenylacetylcarbinol than the yeast. The pyruvate decarboxylases of both organisms catalyzed acetoin and phenylacetylcarbinol synthesis from pyruvate and acetaldehyde or benzaldehyde, but the affinity of the Z. mobilis pyruvate decarboxylase towards the aldehyde reactants was lower than that of the yeast enzyme. Because of the limited solubility of benzaldehyde, neither enzyme could be saturated with this substrate for phenyl-acetylcarbinol synthesis. Studies with 2-toluidinonaphthalene-6-sulfonate and substrate analogues showed that the catalytic sites of pyruvate decarboxylase from Z. mobilis were less lipophilic than those of the enzyme from yeast. This difference could explain the lower affinity for benzaldehyde of the Z. mobilis enzyme.  相似文献   

19.
Utilization of Candida glabrata overproducing pyruvate is a promising strategy for high-level acetoin production. Based on the known regulatory and metabolic information, acetaldehyde and thiamine were fed to identify the key nodes of carboligase activity reaction (CAR) pathway and provide a direction for engineering C. glabrata. Accordingly, alcohol dehydrogenase, acetaldehyde dehydrogenase, pyruvate decarboxylase, and butanediol dehydrogenase were selected to be manipulated for strengthening the CAR pathway. Following the rational metabolic engineering, the engineered strain exhibited increased acetoin biosynthesis (2.24 g/L). In addition, through in silico simulation and redox balance analysis, NADH was identified as the key factor restricting higher acetoin production. Correspondingly, after introduction of NADH oxidase, the final acetoin production was further increased to 7.33 g/L. By combining the rational metabolic engineering and cofactor engineering, the acetoin-producing C. glabrata was improved stepwise, opening a novel pathway for rational development of microorganisms for bioproduction.  相似文献   

20.
An octanol/aqueous two-phase process for the enzymatic production of (R)-phenylacetylcarbinol (PAC) has been investigated further with regard to optimal pH control and replacement of 2.5?M MOPS buffer by a low cost solute. The specific rate of PAC production in the 2.5?M MOPS system controlled at pH?7 was 0.60?mg?U?1?h?1 (reaction completed at 34?h), a 1.6 times improvement over the same 2.5?M MOPS system without pH control (0.39?mg?U?1?h?1 at 49?h). An improved stability of PDC was evident at the end of biotransformation for the pH-controlled system with 84% residual carboligase activity, while 23% of enzyme activity remained in the absence of pH control. Lowering the MOPS concentration to 20?mM resulted in a lower benzaldehyde concentration in the aqueous phase with a major increase in the formation of by-product acetoin and three times decreased PAC production (0.21?mg?U?1?h?1). Biotransformation with 20?mM MOPS and 2.5?M DPG as inexpensive replacement of high MOPS concentrations provided similar aqueous phase benzaldehyde concentrations compared to 2.5?M MOPS and resulted in a comparable PAC concentration (92.1?g?L?1 in the total reaction volume in 47?h) with modest formation of acetoin.  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