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 718 毫秒
1.
The heterofermentative lactic acid bacteria Oenococcus oeni and Leuconostoc mesenteroides are able to grow by fermentation of pyruvate as the carbon source (2 pyruvate --> 1 lactate + 1 acetate + 1 CO(2)). The growth yields amount to 4.0 and 5.3 g (dry weight)/mol of pyruvate, respectively, suggesting formation of 0.5 mol ATP/mol pyruvate. Pyruvate is oxidatively decarboxylated by pyruvate dehydrogenase to acetyl coenzyme A, which is then converted to acetate, yielding 1 mol of ATP. For NADH reoxidation, one further pyruvate molecule is reduced to lactate. The enzymes of the pathway were present after growth on pyruvate, and genome analysis showed the presence of the corresponding structural genes. The bacteria contain, in addition, pyruvate oxidase activity which is induced under microoxic conditions. Other homo- or heterofermentative lactic acid bacteria showed only low pyruvate fermentation activity.  相似文献   

2.
A high yield of lactic acid per gram of glucose consumed and the absence of additional metabolites in the fermentation broth are two important goals of lactic acid production by microrganisms. Both purposes have been previously approached by using a Kluyveromyces lactis yeast strain lacking the single pyruvate decarboxylase gene (KlPDC1) and transformed with the heterologous lactate dehydrogenase gene (LDH). The LDH gene was placed under the control the KlPDC1 promoter, which has allowed very high levels of lactate dehydrogenase (LDH) activity, due to the absence of autoregulation by KlPdc1p. The maximal yield obtained was 0.58 g g−1, suggesting that a large fraction of the glucose consumed was not converted into pyruvate. In a different attempt to redirect pyruvate flux toward homolactic fermentation, we used K. lactis LDH transformant strains deleted of the pyruvate dehydrogenase (PDH) E1α subunit gene. A great process improvement was obtained by the use of producing strains lacking both PDH and pyruvate decarboxylase activities, which showed yield levels of as high as 0.85 g g−1 (maximum theoretical yield, 1 g g−1), and with high LDH activity.  相似文献   

3.
The effect of the concentration of a mixture (1/1 [wt/wt]) of yeast extract and bioTrypcase (YE+bT) on the growth and physiology of a new species, Bacillus thermoamylovorans, a moderately thermophilic, non-spore-forming, lactic acid-producing bacterium isolated from palm wine, was studied. At an initial glucose concentration of 100 mM, B. thermoamylovorans growth was limited when the concentration of YE+bT was lower than 5.0 g liter−1; under these conditions, cellular yield reached a maximum value of 0.4 g of cells per g of YE+bT. Growth limitation due to deficiency in growth factors led to a significant shift in glucose metabolism towards lactate production. Lactate constituted 27.5 and 76% of the end products of glucose fermentation in media containing YE+bT at 20.0 and 1.0 g liter−1, respectively. This result markedly differed from published data for lactic bacteria, which indicated that fermentative metabolism remained homolactic regardless of the concentration of YE. Our results showed that the ratio between cellular synthesis and energy production increased with the concentration of YE+bT in the culture medium. They indicate that the industrial production of lactic acid through glucose fermentation by B. thermoamylovorans can be optimized by using a medium where glucose is present in excess and the organic additives are limiting.  相似文献   

4.
During anaerobic growth of Escherichia coli, pyruvate formate-lyase (PFL) and lactate dehydrogenase (LDH) channel pyruvate toward a mixture of fermentation products. We have introduced a third branch at the pyruvate node in a mutant of E. coli with a mutation in pyruvate dehydrogenase (PDH*) that renders the enzyme less sensitive to inhibition by NADH. The key starting enzymes of the three branches at the pyruvate node in such a mutant, PDH*, PFL, and LDH, have different metabolic potentials and kinetic properties. In such a mutant (strain QZ2), pyruvate flux through LDH was about 30%, with the remainder of the flux occurring through PFL, indicating that LDH is a preferred route of pyruvate conversion over PDH*. In a pfl mutant (strain YK167) with both PDH* and LDH activities, flux through PDH* was about 33% of the total, confirming the ability of LDH to outcompete the PDH pathway for pyruvate in vivo. Only in the absence of LDH (strain QZ3) was pyruvate carbon equally distributed between the PDH* and PFL pathways. A pfl mutant with LDH and PDH* activities, as well as a pfl ldh double mutant with PDH* activity, had a surprisingly low cell yield per mole of ATP (YATP) (about 7.0 g of cells per mol of ATP) compared to 10.9 g of cells per mol of ATP for the wild type. The lower YATP suggests the operation of a futile energy cycle in the absence of PFL in this strain. An understanding of the controls at the pyruvate node during anaerobic growth is expected to provide unique insights into rational metabolic engineering of E. coli and related bacteria for the production of various biobased products at high rates and yields.In Escherichia coli as well as in other aerobic organisms, sugars such as glucose are metabolized in two separate steps: glycolysis, which converts glucose to pyruvate, and tricarboxylic acid (TCA) cycle enzymes, which oxidize acetyl coenzyme A (acetyl-CoA) to CO2 (5, 9). The pyruvate dehydrogenase complex (PDH) connects the glycolytic reactions to TCA cycle enzymes by catalyzing the production of acetyl-CoA from pyruvate. Because of its unique central role in metabolism, PDH is regulated at both the genetic and the biochemical level (7, 12, 27, 33, 34). The NADH generated during the complete oxidation of sugar is reoxidized to NAD+ by O2 through the respiratory electron transport pathway with accompanying energy production (11). Optimum coupling of these enzyme reactions helps to maintain the internal ratios of [NADH] to [NAD+] (redox balance) and of [ATP] to [ADP] plus [AMP] in order to support growth at the highest rate.The absence of O2 or another external electron acceptor during the growth of E. coli (anaerobic conditions) forces the bacterium to minimize the contribution of the TCA cycle enzymes to biosynthesis from catabolism (4, 14). Under these conditions, pyruvate or acetyl-CoA derived from pyruvate serves as the electron acceptor (reduced to lactate or ethanol, respectively) to maintain the redox balance. The enzymes responsible for redox balance in anaerobic E. coli are pyruvate formate-lyase (PFL), lactate dehydrogenase (LDH), and alcohol/aldehyde dehydrogenase (adhE; ADH-E). The main products of the fermentation of E. coli are a mixture of organic acids, such as acetate, lactate, and formate, in addition to ethanol (2, 4). Succinate, derived from phosphoenolpyruvate (PEP), is a minor product of fermentation and normally accounts for less than 5% of the total products produced from glucose by the culture.Anaerobic growth of E. coli, compared to aerobic growth, is also limited by energy, leading to an increase in glycolytic flux (19). The conversion of pyruvate to acetate and ethanol yields an additional ATP per glucose, suggesting that this would be the preferred route for pyruvate oxidation during anaerobic growth. This is accomplished by the PFL-dependent production of acetyl-CoA and further conversion to acetate (Fig. (Fig.1).1). This preference for PFL has been demonstrated `with several bacteria under carbon limitation conditions imposed either in a chemostat or in the presence of a poor carbon source (10, 20, 23). This additional ATP also elevates the ATP yield per glucose to 3, with an increase in the growth rate, and has been shown to be essential for the anaerobic growth of E. coli in xylose-mineral salts medium (13). The absence of this third ATP in a pfl mutant has been reported to increase glycolytic flux to lactate to compensate for this decrease in ATP yield per glucose (39). However, the flow of pyruvate carbon to acetate is tempered by the need to maintain redox balance, and this is achieved by the conversion of a second acetyl-CoA to ethanol by ADH-E. Under conditions of energy excess due to a declining growth rate, lactate production is expected to support redox balance maintenance without the additional ATP from the PFL-ADH-E pathway (Fig. (Fig.1).1). The production of this mixture of products in an appropriate ratio helps to maintain the redox balance under anaerobic conditions while also maximizing the ATP yield per glucose to support high growth rates and cell yields.Open in a separate windowFIG. 1.Anaerobic metabolic pathways of E. coli carrying the lpd101 mutation (PDH*).No PDH-based fermentation reaction to ethanol that can also help maintain cellular redox balance in an anaerobic cell has evolved in E. coli or other closely related bacteria. PDH activity is inhibited by NADH, normally found at higher levels in anaerobically growing cultures than in aerobic cultures (12, 18, 34, 35). Based on genome sequences available in GenBank, the genes encoding the components of PDH are not found in strictly anaerobic bacteria.We have recently described a mutation (lpd101) in the dihydrolipoamide dehydrogenase (LPD) of the PDH that allowed the enzyme to function in anaerobic cells (designated PDH* here) (17, 18). With this altered PDH*, an anaerobic cell can have three different pathways for pyruvate metabolism (Fig. (Fig.1).1). The three main enzymes that utilize pyruvate as a substrate, PDH*, PFL, and LDH, have different apparent Km values for pyruvate (0.4, 2.0, and 7.2 mM, respectively) (1, 18, 37, 41). PDH requires NAD+ for activity (apparent Km, 0.07 mM), while LDH is dependent on NADH (apparent Km, 0.2 mM) as the second substrate (18, 37).The PDH* serves as the first enzyme in a pathway that oxidatively decarboxylates pyruvate to acetyl-CoA and NADH, followed by reduction of the acetyl-CoA by alcohol dehydrogenase to ethanol in a two-step process using 2 NADHs (Fig. (Fig.1).1). The NADH produced during the conversion of glucose to acetyl-CoA dictates that the acetyl-CoA generated by PDH be used for redox balance (ethanol) and not for ATP generation (acetate), unless some of the NADH is used for biosynthesis by the growing cell (17). PDH* and LDH serve essentially the same physiological role in the cell, oxidizing NADH to support continued operation of glycolysis, although it is not readily apparent with PDH*. Although PDH* contributes to an increase in NADH pool, the redox balance is still maintained by coupling PDH* to NADH-dependent reduction of acetyl-CoA to ethanol by ADH-E (Fig. (Fig.1).1). This potential competition between LDH and PDH has been eliminated in the wild type through inhibition of the activity of PDH by NADH (12, 18, 32). However, the in vivo role of PDH* in a mutant that has all three pathways has not been investigated, since the flow of pyruvate through any of the three reactions during growth and postgrowth fermentation of sugars to products is expected to be dependent on the redox state, the ATP requirement, and other physiological conditions of the anaerobic cell. Using a combination of metabolic flux analysis and mutations in one or more of the genes encoding these enzymes, we have evaluated the flow of pyruvate carbon among the three potential pathways. The results are presented in this communication.  相似文献   

5.
Fermentation patterns of Escherichia coli with and without the phosphoenolpyruvate carboxylase (PPC) and pyruvate carboxylase (PYC) enzymes were compared under anaerobic conditions with glucose as a carbon source. Time profiles of glucose and fermentation product concentrations were determined and used to calculate metabolic fluxes through central carbon pathways during exponential cell growth. The presence of the Rhizobium etli pyc gene in E. coli (JCL1242/pTrc99A-pyc) restored the succinate producing ability of E. coli ppc null mutants (JCL1242), with PYC competing favorably with both pyruvate formate lyase and lactate dehydrogenase. Succinate formation was slightly greater by JCL1242/pTrc99A-pyc than by cells which overproduced PPC (JCL1242/pPC201, ppc+), even though PPC activity in cell extracts of JCL1242/pPC201 (ppc+) was 40-fold greater than PYC activity in extracts of JCL1242/pTrc99a-pyc. Flux calculations indicate that during anaerobic metabolism the pyc+ strain had a 34% greater specific glucose consumption rate, a 37% greater specific rate of ATP formation, and a 6% greater specific growth rate compared to the ppc+ strain. In light of the important position of pyruvate at the juncture of NADH-generating pathways and NADH-dissimilating branches, the results show that when PPC or PYC is expressed, the metabolic network adapts by altering the flux to lactate and the molar ratio of ethanol to acetate formation.  相似文献   

6.
In vivo 13C nuclear magnetic resonance spectroscopy was used to elucidate the pathways and the regulation of pyruvate metabolism and pyruvate-lactate cometabolism noninvasively in living-cell suspensions of Propionibacterium freudenreichii subsp. shermanii. The most important result of this work concerns the modification of fluxes of pyruvate metabolism induced by the presence of lactate. Pyruvate was temporarily converted to lactate and alanine; the flux to acetate synthesis was maintained, but the flux to propionate synthesis was increased; and the reverse flux of the first part of the Wood-Werkman cycle, up to acetate synthesis, was decreased. Pyruvate was consumed at apparent initial rates of 148 and 90 μmol · min−1 · g−1 (cell dry weight) when it was the sole substrate or cometabolized with lactate, respectively. Lactate was consumed at an apparent initial rate of 157 μmol · min−1 · g−1 when it was cometabolized with pyruvate. P. shermanii used several pathways, namely, the Wood-Werkman cycle, synthesis of acetate and CO2, succinate synthesis, gluconeogenesis, the tricarboxylic acid cycle, and alanine synthesis, to manage its pyruvate pool sharply. In both types of experiments, acetate synthesis and the Wood-Werkman cycle were the metabolic pathways used most.  相似文献   

7.
Clostridium formicoaceticum homofermentatively converted lactate to acetate at mesophilic temperatures (30 to 42°C) and at pHs between 6.6 and 9.6. The production of acetate was found to be growth associated. Approximately 0.96 g of acetic acid and 0.066 g of cells were formed from each gram of lactic acid consumed at 37°C. The concentration of the substrate (lactate) had little or no effect on the growth rate; however, the fermentation was inhibited by acetic acid. The bacterium grew at an optimal pH of 7.6 and an optimal temperature of 37°C. Small amounts of bicarbonate were stimulatory to bacterial growth. Bacterial growth was enhanced, however, by the use of higher concentrations of bicarbonate in the media, only because higher buffer capacities were obtained and proper medium pH could be maintained for growth. Based on its ability to convert lactate to acetate, this homoacetic bacterium may be important in the anaerobic methanogenic process when lactate is a major intermediary metabolite.  相似文献   

8.
The hyperthermophilic anaerobe Pyrococcus furiosus was found to grow on pyruvate as energy and carbon source. Growth was dependent on yeast extract (0.1%). The organism grew with doublings times of about 1 h up to cell densities of 1–2×108 cells/ml. During growth 0.6–0.8 mol acetate and 1.2–1.5 mol CO2 and 0.8 mol H2 were formed per mol of pyruvate consumed. The molar growth yield was 10–11 g cells(dry weight)/mol pyruvate. Cell suspensions catalyzed the conversion of 1 mol of pyruvate to 0.6–0.8 mol acetate, 1.2–1.5 mol CO2, 1.2 mol H2 and 0.03 mol acetoin. After fermentation of [3-14C]pyruvate the specific radioactivities of pyruvate, CO2 and acetate were equal to 1:0.01:1. Cellfree extracts contained the following enzymatic activities: pyruvate: ferredoxin (methyl viologen) oxidoreductase (0.2 U mg-1, T=60°C, with Clostridium pasteurianum ferredoxin as electron acceptor; 1.4 U mg-1 at 90°C, with methyl viologen as electron acceptor); acetyl-CoA synthetase (ADP forming) [acetyl-CoA+ADP+Piacetate+ATP+CoA] (0.34 U mg-1, T=90°C), and hydrogen: methyl viologen oxidoreductase (1.75 U mg-1). Phosphate acetyl-transferase activity, acetate kinase activity, and carbon monoxide:methyl viologen oxidoreductase activity could not be detected. These findings indicate that the archaebacterium P. furiosus ferments pyruvate to acetate, CO2 and H2 involving only three enzymes, a pyruvate:ferredoxin oxidoreductase, a hydrogenase and an acetyl-CoA synthetase (ADP forming).Non-standard abbreviations DTE dithioerythritol - MV methyl viologen - MOPS morpholinopropane sulfonic acid - Tricine N-tris(hydroxymethyl)-methylglycine Part of the work was performed at the Laboratorium für Mikrobiologie, Fachbereich Biologie, Philipps-Universität, Karlvon-Frisch-Strasse, W-3550 Marburg/Lahn, Federal Republic of Germany  相似文献   

9.
The amount of energy that can be conserved via halorespiration by Desulfitobacterium dehalogenans JW/IU-DC1 was determined by comparison of the growth yields of cells grown with 3-chloro-4-hydroxyphenyl acetate (Cl-OHPA) and different electron donors. Cultures that were grown with lactate, pyruvate, formate, or hydrogen as an electron donor and Cl-OHPA as an electron acceptor yielded 3.1, 6.6, 1.6, and 1.6 g (dry weight) per mol of reduction equivalents, respectively. Fermentative growth on pyruvate yielded 14 g (dry weight) per mol of pyruvate oxidized. Pyruvate was not fermented stoichiometrically to acetate and lactate, but an excess of acetate was produced. Experiments with 13C-labeled bicarbonate showed that during pyruvate fermentation, approximately 9% of the acetate was formed from the reduction of CO2. Comparison of the growth yields suggests that 1 mol of ATP is produced per mol of acetate produced by substrate-level phosphorylation and that there is no contribution of electron transport phosphorylation when D. dehalogenans grows on lactate plus Cl-OHPA or pyruvate plus Cl-OHPA. Furthermore, the growth yields indicate that approximately 1/3 mol of ATP is conserved per mol of Cl-OHPA reduced in cultures grown in formate plus Cl-OHPA and hydrogen plus Cl-OHPA. Because neither formate nor hydrogen nor Cl-OHPA supports substrate-level phosphorylation, energy must be conserved through the establishment of a proton motive force. Pyruvate ferredoxin oxidoreductase, lactate dehydrogenase, formate dehydrogenase, and hydrogenase were localized by in vitro assays with membrane-impermeable electron acceptors and donors. The orientation of chlorophenol-reductive dehalogenase in the cytoplasmic membrane, however, could not be determined. A model is proposed, which may explain the topology analyses as well as the results obtained in the yield study.  相似文献   

10.
Acetic acid plays a crucial role in the organoleptic balance of many fermented products. We have investigated the factors controlling the production of acetate by Saccharomyces cerevisiae during alcoholic fermentation by metabolic engineering of the enzymatic steps involved in its formation and its utilization. The impact of reduced pyruvate decarboxylase (PDC), limited acetaldehyde dehydrogenase (ACDH), or increased acetoacetyl coenzyme A synthetase (ACS) levels in a strain derived from a wine yeast strain was studied during alcoholic fermentation. In the strain with the PDC1 gene deleted exhibiting 25% of the PDC activity of the wild type, no significant differences were observed in the acetate yield or in the amounts of secondary metabolites formed. A strain overexpressing ACS2 and displaying a four- to sevenfold increase in ACS activity did not produce reduced acetate levels. In contrast, strains with one or two disrupted copies of ALD6, encoding the cytosolic Mg2+-activated NADP-dependent ACDH and exhibiting 60 and 30% of wild-type ACDH activity, showed a substantial decrease in acetate yield (the acetate production was 75 and 40% of wild-type production, respectively). This decrease was associated with a rerouting of carbon flux towards the formation of glycerol, succinate, and butanediol. The deletion of ALD4, encoding the mitochondrial K+-activated NAD(P)-linked ACDH, had no effect on the amount of acetate formed. In contrast, a strain lacking both Ald6p and Ald4p exhibited a long delay in growth and acetate production, suggesting that Ald4p can partially replace the Ald6p isoform. Moreover, the ald6 ald4 double mutant was still able to ferment large amounts of sugar and to produce acetate, suggesting the contribution of another member(s) of the ALD family.  相似文献   

11.
We report the homofermentative production of lactate in Escherichia coli strains containing mutations in the aceEF, pfl, poxB, and pps genes, which encode the pyruvate dehydrogenase complex, pyruvate formate lyase, pyruvate oxidase, and phosphoenolpyruvate synthase, respectively. The process uses a defined medium and two distinct fermentation phases: aerobic growth to an optical density of about 30, followed by nongrowth, anaerobic production. Strain YYC202 (aceEF pfl poxB pps) generated 90 g/liter lactate in 16 h during the anaerobic phase (with a yield of 0.95 g/g and a productivity of 5.6 g/liter · h). Ca(OH)2 was found to be superior to NaOH for pH control, and interestingly, significant succinate also accumulated (over 7 g/liter) despite the use of N2 for maintaining anaerobic conditions. Strain ALS961 (YYC202 ppc) prevented succinate accumulation, but growth was very poor. Strain ALS974 (YYC202 frdABCD) reduced succinate formation by 70% to less than 3 g/liter. 13C nuclear magnetic resonance analysis using uniformly labeled acetate demonstrated that succinate formation by ALS974 was biochemically derived from acetate in the medium. The absence of uniformly labeled succinate, however, demonstrated that glyoxylate did not reenter the tricarboxylic acid cycle via oxaloacetate. By minimizing the residual acetate at the time that the production phase commenced, the process with ALS974 achieved 138 g/liter lactate (1.55 M, 97% of the carbon products), with a yield of 0.99 g/g and a productivity of 6.3 g/liter · h during the anaerobic phase.  相似文献   

12.
We report pyruvate formation in Escherichia coli strain ALS929 containing mutations in the aceEF, pfl, poxB, pps, and ldhA genes which encode, respectively, the pyruvate dehydrogenase complex, pyruvate formate lyase, pyruvate oxidase, phosphoenolpyruvate synthase, and lactate dehydrogenase. The glycolytic rate and pyruvate productivity were compared using glucose-, acetate-, nitrogen-, or phosphorus-limited chemostats at a growth rate of 0.15 h−1. Of these four nutrient limitation conditions, growth under acetate limitation resulted in the highest glycolytic flux (1.60 g/g · h), pyruvate formation rate (1.11 g/g · h), and pyruvate yield (0.70 g/g). Additional mutations in atpFH and arcA (strain ALS1059) further elevated the steady-state glycolytic flux to 2.38 g/g · h in an acetate-limited chemostat, with heterologous NADH oxidase expression causing only modest additional improvement. A fed-batch process with strain ALS1059 using defined medium with 5 mM betaine as osmoprotectant and an exponential feeding rate of 0.15 h−1 achieved 90 g/liter pyruvate, with an overall productivity of 2.1 g/liter · h and yield of 0.68 g/g.  相似文献   

13.
Due to a growing market for the biodegradable and renewable polymer polylactic acid, the world demand for lactic acid is rapidly increasing. The tolerance of yeasts to low pH can benefit the process economy of lactic acid production by minimizing the need for neutralizing agents. Saccharomyces cerevisiae (CEN.PK background) was engineered to a homofermentative lactate-producing yeast via deletion of the three genes encoding pyruvate decarboxylase and the introduction of a heterologous lactate dehydrogenase (EC 1.1.1.27). Like all pyruvate decarboxylase-negative S. cerevisiae strains, the engineered strain required small amounts of acetate for the synthesis of cytosolic acetyl-coenzyme A. Exposure of aerobic glucose-limited chemostat cultures to excess glucose resulted in the immediate appearance of lactate as the major fermentation product. Ethanol formation was absent. However, the engineered strain could not grow anaerobically, and lactate production was strongly stimulated by oxygen. In addition, under all conditions examined, lactate production by the engineered strain was slower than alcoholic fermentation by the wild type. Despite the equivalence of alcoholic fermentation and lactate fermentation with respect to redox balance and ATP generation, studies on oxygen-limited chemostat cultures showed that lactate production does not contribute to the ATP economy of the engineered yeast. This absence of net ATP production is probably due to a metabolic energy requirement (directly or indirectly in the form of ATP) for lactate export.  相似文献   

14.
Heterofermentative lactic acid bacteria (LAB) such as Leuconostoc, Oenococcus, and Lactobacillus strains ferment pentoses by the phosphoketolase pathway. The extra NAD(P)H, which is produced during growth on hexoses, is transferred to acetyl-CoA, yielding ethanol. Ethanol fermentation represents the limiting step in hexose fermentation, therefore, part of the extra NAD(P)H is used to produce erythritol and glycerol. Fructose, pyruvate, citrate, and O2 can be used in addition as external electron acceptors for NAD(P)H reoxidation. Use of the external acceptors increases the growth rate of the bacteria. The bacteria are also able to ferment organic acids like malate, pyruvate, and citrate. Malolactic fermentation generates a proton potential by substrate transport. Pyruvate fermentation sustains growth by pyruvate disproportionation involving pyruvate dehydrogenase. Citrate is fermented in the presence of an additional electron donor to acetate and lactate. Thus, heterofermentative LAB are able to use a variety of unusual fermentation reactions in addition to classical heterofermentation. Most of the reactions are significant for food biotechnology/microbiology.  相似文献   

15.
The pyruvate dehydrogenase complex was partially purified and characterized from etiolated maize (Zea mays L.) shoot mitochondria. Analysis by sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed proteins of 40, 43, 52 to 53, and 62 to 63 kD. Immunoblot analyses identified these proteins as the E1β-, E1α-, E2-, and E3-subunits, respectively. The molecular mass of maize E2 is considerably smaller than that of other plant E2 subunits (76 kD). The activity of the maize mitochondrial complex has a pH optimum of 7.5 and a divalent cation requirement best satisfied by Mg2+. Michaelis constants for the substrates were 47, 3, 77, and 1 μm for pyruvate, coenzyme A (CoA), NAD+, and thiamine pyrophosphate, respectively. The products NADH and acetyl-CoA were competitive inhibitors with respect to NAD+ and CoA, and the inhibition constants were 15 and 47 μm, respectively. The complex was inactivated by phosphorylation and was reactivated after the removal of ATP and the addition of Mg2+.  相似文献   

16.
Based on the presence and absence of enzyme activities, the biochemical pathways for the fermentation of inulin by Clostridium thermosuccinogenes DSM 5809 are proposed. Activities of nine enzymes (lactate dehydrogenase, phosphoenolpyruvate carboxylase, malate dehydrogenase, fumarase, fumarate reductase, phosphotransacetylase, acetate kinase, pyruvate kinase, and alcohol dehydrogenase) were measured at four temperatures (37, 47, 58, and 70°C). Each of the enzymes increased 1.5 to 2.0-fold in activity between 37 and 58°C, but only lactate dehydrogenase, fumarate reductase, malate dehydrogenase, and fumarase increased at a similar rate between 58 and 70°C. No acetate kinase activity was observed at 70°C. Arrhenius energies were calculated for each of these nine enzymes and were in the range of 9.8 to 25.6 kcal/mol. To determine if a relationship existed between product formation and enzyme activity, serum bottle fermentations were completed at the four temperatures. Maximum yields (in moles per mole hexose unit) for succinate (0.23) and acetate (0.79) and for biomass (29.5 g/mol hexose unit) occurred at 58°C, whereas the maximum yields for lactate (0.19) and hydrogen (0.25) and the lowest yields for acetate (0.03) and biomass (19.2 g/mol hexose unit) were observed at 70°C. The ratio of oxidized products to reduced products changed significantly, from 0.52 to 0.65, with an increase in temperature from 58 to 70°C, and there was an unexplained detection of increased reduced products (ethanol, lactate, and hydrogen) with a concomitant decrease in oxidized-product formation at the higher temperature.  相似文献   

17.
Lactobacillus spp. from an inoculant and Weissella and Leuconostoc spp. from forage crops were characterized, and their influence on silage fermentation was studied. Forty-two lactic acid-producing cocci were obtained from forage crops and grasses. All isolates were gram-positive, catalase-negative cocci that produced gas from glucose, and produced more than 90% of their lactate in the d-isomer form. These isolates were divided into groups A and B by sugar fermentation patterns. Two representative strains from the two groups, FG 5 and FG 13, were assigned to the species Weissella paramesenteroides and Leuconostoc pseudomesenteroides, respectively, on the basis of DNA-DNA relatedness. Strains FG 5, FG 13, and SL 1 (Lactobacillus casei), isolated from a commercial inoculant, were used as additives to alfalfa and Italian ryegrass silage preparations. Lactic acid bacterium counts were higher in all additive-treated silages than in the control silage at an early stage of ensiling. During silage fermentation, inoculation with SL 1 more effectively inhibited the growth of aerobic bacteria and clostridia than inoculation with strain FG 5 or FG 13. SL 1-treated silages stored well. However, the control and FG 5- and FG 13-treated silages had a significantly (P < 0.05) higher pH and butyric acid and ammonia nitrogen contents and significantly (P < 0.05) lower lactate content than SL 1-treated silage. Compared with the control silage, SL 1 treatments reduced the proportion of d-(−)-lactic acid, gas production, and dry matter loss in two kinds of silage, but the FG 5 and FG 13 treatments gave similar values in alfalfa silages and higher values (P < 0.05) in Italian ryegrass silage. The results confirmed that heterofermentative strains of W. paramesenteroides FG 5 and L. pseudomesenteroides FG 13 did not improve silage quality and may cause some fermentation loss.Silage is now the most common preserved cattle feed in many countries, including Japan. It is well established that lactic acid bacteria (LAB) play an important role in silage fermentation. Epiphytic microflora, the microorganisms naturally present on forage crops, are responsible for silage fermentation and also influence silage quality (3, 11, 15). Lactobacilli and lactic acid-producing cocci, e.g., leuconostocs, lactococci, streptococci, pediococci, and Weissella species, are major components of the microbial flora in various types of forage crops (3). Stirling and Whittenbury (21) reported that leuconostocs were the most numerous and widely distributed on forages and that lactobacilli occurred mostly on grasses. Cai et al. (3) examined a large number of forage crops and grasses and also found that the predominant LAB were lactic acid-producing cocci and that lactobacilli were the least numerous and mostly homofermentative. Ruser (17) found that although all LAB groups were present in chopped-maize samples, homofermentative lactobacilli and heterofermentative leuconostocs were present in the highest numbers.In order to improve silage quality, many LAB-containing biological additives have been developed and are currently available (13, 20, 25). These inoculants may inhibit the growth of harmful bacteria and enhance lactic acid fermentation during ensiling periods. The epiphytic LAB influence the effectiveness of silage inoculants because the introduced bacteria must compete with these LAB (12). Therefore, the LAB species and their characteristics in the silage environment require further study. However, while an increasing number of studies have reported positive benefits from using some bacterial inoculants as silage additives, relatively few have reported the effect of epiphytic LAB, especially Leuconostoc and Weissella species, on silage fermentation. In the present study, the characterization of Leuconostoc and Weissella species isolated from forage crops and their influence on silage fermentation were examined.  相似文献   

18.
In Escherichia coli K12, succinate was not the dominant fermentation product from xylose. To reduce by-product formation and increase succinate accumulation, pyruvate formate lyase and lactate dehydrogenase, encoded by pflB and ldhA genes, were inactivated. However, these mutations eliminated cell growth and xylose utilization. During anaerobic growth of bacteria, organic intermediates, such as pyruvate, serve as electron acceptors to maintain the overall redox balance. Under these conditions, the ATP needed for cell growth is derived from substrate level phosphorylation. In E. coli K12, conversion of xylose to pyruvate only yielded 0.67 net ATP per xylose during anaerobic fermentation. However, E. coli produces equimolar amounts of acetate and ethanol from two pyruvates, and these reactions generate one additional ATP. Conversion of xylose to acetate and ethanol increases the net ATP yield from 0.67 to 1.5 per xylose, which could meet the ATP needed for xylose metabolism. A pflB deletion strain cannot convert pyruvate to acetyl coenzyme A, the precursor for acetate and ethanol production, and could not produce the additional ATP. Thus, the double mutations eliminated cell growth and xylose utilization. To supply the sufficient ATPs, overexpression of ATP-forming phosphoenolpyruvate-carboxykinase from Bacillus subtilis 168 in an ldhA, pflB, and ppc deletion strain resulted in a significant increase in cell mass and succinate production. In addition, fermentation of corn stalk hydrolysate containing a high percentage of xylose and glucose produced a final succinate concentration of 11.13 g l−1 with a yield of 1.02 g g−1 total sugars during anaerobic fermentation.  相似文献   

19.
During anaerobic growth of Klebsiella pneumoniae on citrate, 9.4 mmol of H2/mol of citrate (4-kPa partial pressure) was formed at the end of growth besides acetate, formate, and CO2. Upon addition of NiCl2 (36 μM) to the growth medium, hydrogen formation increased about 36% to 14.8 mmol/mol of citrate (6 kPa), and the cell yield increased about 15%. Cells that had been harvested and washed under anoxic conditions exhibited an H2-dependent formation of NAD(P)H in vivo. The reduction of internal NAD(P)+ was also achieved by the addition of formate. In crude extracts, the H2:NAD+ oxidoreductase activity was 0.13 μmol min−1 mg−1, and 76% of this activity was found in the washed membrane fraction. The highest specific activities of the membrane fraction were observed in 50 mM potassium phosphate, with 1.6 μmol of NADPH formed min−1 mg−1 at pH 7.0 and 1.7 μmol of NADH formed min−1 mg−1 at pH 9.5. In the presence of the protonophore carbonyl cyanide m-chlorophenylhydrazone and the Na+/H+ antiporter monensin, the H2-dependent reduction of NAD+ by membrane vesicles decreased only slightly (about 16%). The NADP+- or NAD+-reducing hydrogenases were solubilized from the membranes with the detergent lauryldimethylamine-N-oxide or Triton X-100. NAD(P)H formation with H2 as electron donor, therefore, does not depend on an energized state of the membrane. It is proposed that hydrogen which is formed by K. pneumoniae during citrate fermentation is recaptured by a novel membrane-bound, oxygen-sensitive H2:NAD(P)+ oxidoreductase that provides reducing equivalents for the synthesis of cell material.  相似文献   

20.
Hexamita sp. is an amitochondriate free-living diplomonad which inhabits O2-limited environments, such as the deep waters and sediments of lakes and marine basins. 13C nuclear magnetic resonance spectroscopy reveals ethanol, lactate, acetate, and alanine as products of glucose fermentation under microaerobic conditions (23 to 34 μM O2). Propionic acid and butyric acid were also detected and are believed to be the result of fermentation of alternative substrates. Production of organic acids was greatest under microaerobic conditions (15 μM O2) and decreased under anaerobic (<0.25 μM O2) and aerobic (200 to 250 μM O2) conditions. Microaerobic incubation resulted in the production of high levels of oxidized end products (70% acetate) compared to that produced under anoxic conditions (20% acetate). In addition, data suggest that Hexamita cells contain the arginine dihydrolase pathway, generating energy from the catabolism of arginine to citrulline, ornithine, NH4+, and CO2. The rate of arginine catabolism was higher under anoxic conditions than under microaerobic conditions. Hexamita cells were able to grow in the absence of a carbohydrate source, albeit with a lower growth rate and yield.  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