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.
Leuconostoc mesenteroides NCDO 518, provided with oxygen and pyruvate, preferentially used oxygen as accessory electron acceptor and converted pyruvate to acetoin. With glucose, 5.6 mM, as sole energy source only small amounts of acetoin were formed (0.08–0.21 mM). With glucose, 5.6 mM, and pyruvate, 20 mM, substantial amounts of acetoin were produced in growing, aerated cultures at pH 5 (2.8 mM, equivalent to 0.5 mol [mol glucose fermented]–1). On exhaustion of glucose, growth ceased but metabolism of pyruvate continued with the formation of acetate and a little acetoin. In aerated cultures at pH 6 the general pattern was similar to that at pH 5 but less acetoin (0.6 mM) was formed during the growth phase and, after the exhaustion of glucose, pyruvate was converted very slowly to acetate only. Leuc. mesenteroides did not grow with pyruvate as sole energy source.  相似文献   

2.
Using the pyruvate production strain Escherichia coli YYC202 ldhA::Kan different process alternatives are studied with the aim of preventing potential product inhibition by appropriate product separation. This strain is completely blocked in its ability to convert pyruvate into acetyl-CoA or acetate, resulting in acetate auxotrophy during growth in glucose minimal medium. Continuous experiments with cell retention, repetitive fed-batch, and an in situ product recovery (ISPR) process with fully integrated electrodialysis were tested. Although the continuous approach achieved a high volumetric productivity (QP) of 110 g L(-1) d(-1), this approach was not pursued because of long-term production strain instabilities. The highest pyruvate/glucose molar yield of up to 1.78 mol mol(-1) together with high QP 145 g L(-1) d(-1) and high pyruvate titers was achieved by the repetitive fed-batch approach. To separate pyruvate from fermentation broth a fully integrated continuous process was developed. In this process electrodialysis was used as a separation unit. Under optimum conditions a (calculated) final pyruvate titer of >900 mmol L(-1) (79 g L(-1)) was achieved.  相似文献   

3.
Corynebacterium glutamicum was engineered for the production of L-valine from glucose by deletion of the aceE gene encoding the E1p enzyme of the pyruvate dehydrogenase complex and additional overexpression of the ilvBNCE genes encoding the L-valine biosynthetic enzymes acetohydroxyacid synthase, isomeroreductase, and transaminase B. In the absence of cellular growth, C. glutamicum DeltaaceE showed a relatively high intracellular concentration of pyruvate (25.9 mM) and produced significant amounts of pyruvate, L-alanine, and L-valine from glucose as the sole carbon source. Lactate or acetate was not formed. Plasmid-bound overexpression of ilvBNCE in C. glutamicum DeltaaceE resulted in an approximately 10-fold-lower intracellular pyruvate concentration (2.3 mM) and a shift of the extracellular product pattern from pyruvate and L-alanine towards L-valine. In fed-batch fermentations at high cell densities and an excess of glucose, C. glutamicum DeltaaceE(pJC4ilvBNCE) produced up to 210 mM L-valine with a volumetric productivity of 10.0 mM h(-1) (1.17 g l(-1) h(-1)) and a maximum yield of about 0.6 mol per mol (0.4 g per g) of glucose.  相似文献   

4.
Malic acid is a potential biomass-derivable "building block" for chemical synthesis. Since wild-type Saccharomyces cerevisiae strains produce only low levels of malate, metabolic engineering is required to achieve efficient malate production with this yeast. A promising pathway for malate production from glucose proceeds via carboxylation of pyruvate, followed by reduction of oxaloacetate to malate. This redox- and ATP-neutral, CO(2)-fixing pathway has a theoretical maximum yield of 2 mol malate (mol glucose)(-1). A previously engineered glucose-tolerant, C(2)-independent pyruvate decarboxylase-negative S. cerevisiae strain was used as the platform to evaluate the impact of individual and combined introduction of three genetic modifications: (i) overexpression of the native pyruvate carboxylase encoded by PYC2, (ii) high-level expression of an allele of the MDH3 gene, of which the encoded malate dehydrogenase was retargeted to the cytosol by deletion of the C-terminal peroxisomal targeting sequence, and (iii) functional expression of the Schizosaccharomyces pombe malate transporter gene SpMAE1. While single or double modifications improved malate production, the highest malate yields and titers were obtained with the simultaneous introduction of all three modifications. In glucose-grown batch cultures, the resulting engineered strain produced malate at titers of up to 59 g liter(-1) at a malate yield of 0.42 mol (mol glucose)(-1). Metabolic flux analysis showed that metabolite labeling patterns observed upon nuclear magnetic resonance analyses of cultures grown on (13)C-labeled glucose were consistent with the envisaged nonoxidative, fermentative pathway for malate production. The engineered strains still produced substantial amounts of pyruvate, indicating that the pathway efficiency can be further improved.  相似文献   

5.
A mutant of Methanosarcina barkeri (Fusaro) is able to grow on pyruvate as the sole carbon and energy source. During growth, pyruvate is converted to CH4 and CO2, and about 1.5 mol of ATP per mol of CH4 is formed (A.-K. Bock, A. Prieger-Kraft, and P. Schönheit, Arch. Microbiol. 161:33-46, 1994). The pyruvate-utilizing mutant of M. barkeri could also grow on pyruvate when methanogenesis was completely inhibited by bromoethanesulfonate (BES). The mutant grew on pyruvate (80 mM) in the presence of 2 mM BES with a doubling time of about 30 h up to cell densities of about 400 mg (dry weight) of cells per liter. During growth on pyruvate, the major fermentation products were acetate and CO2 (about 0.9 mol each per mol of pyruvate). Small amounts of acetoin, acetolactate, alanine, leucine, isoleucine, and valine were also detected. CH4 was not formed. The molar growth yield (Yacetate) was about 9 g of cells (dry weight) per mol of acetate, indicating an ATP yield of about 1 mol/mol of acetate formed. Growth on pyruvate in the presence of BES was limited; after six to eight generations, the doubling times increased and the final cell densities decreased. After 9 to 11 generations, growth stopped completely. In the presence of BES, suspensions of pyruvate-grown cells fermented pyruvate to acetate, CO2, and H2. CH4 was not formed. Conversion of pyruvate to acetate, in the complete absence of methanogenesis, was coupled to ATP synthesis. Dicyclohexylcarbodiimide, an inhibitor of H(+)-translocating ATP synthase, did not inhibit ATP formation. In the presence of dicyclohexylcarbodiimide, stoichiometries of up to 0.9 mol of ATP per mol of acetate were observed. The uncoupler arsenate completely inhibited ATP synthesis, while the rates of acetate, CO2, and H2 formation were stimulated up to fourfold. Cell extracts of M. barkeri grown on pyruvate under nonmethenogenic conditions contained pyruvate: ferredoxin oxidoreductase (0.5 U/mg), phosphate acetyltransferase (12 U/mg), and acetate kinase (12 U/mg). From these data it is concluded that ATP was synthesized by substrate level phosphorylation during growth of the M. barkeri mutant on pyruvate in the absence of methanogenesis. This is the first report of growth of a methanogen under nonmethanogenic conditions at the expense of a fermentative energy metabolism.  相似文献   

6.
Spirochaeta thermophila RI 19.B1 (DSM 6192) fermented glucose to lactate, acetate, CO2, and H2 with concomitant formation of cell material. The cell dry mass yield was 20.0 g/mol of glucose. From the fermentation balance data and knowledge of the fermentation pathway, a YATP of 9.22 g of dry mass per mol of ATP was calculated for pH-uncontrolled batch-culture growth on glucose in a mineral medium. Measurement of enzyme activities in glucose-grown cells revealed that glucose was taken up by a permease and then subjected to ATP-dependent phosphorylation by a hexokinase. Glucose-6-phosphate was further metabolized to pyruvate through the Embden-Meyerhof-Parnas pathway. The phosphoryl donor for phosphofructokinase activity was PPi rather than ATP. This was also found for the type strain of S. thermophila, Z-1203 (DSM 6578). PPi was probably formed by pyrophosphoroclastic cleavage of ATP, with recovery of the resultant AMP by the activity of adenylate kinase. All other measured kinase activities utilized ATP as the phosphoryl donor. Pyruvate was further metabolized to acetyl coenzyme A with concomitant production of H2 and CO2 by pyruvate synthase. Lactate was also produced from pyruvate by a fructose-1,6-diphosphate-insensitive lactate dehydrogenase. Evidence was obtained for the transfer of reducing equivalents from the glycolytic pathway to hydrogenase to produce H2. No formate dehydrogenase or significant ethanol-producing enzyme activities were detected.  相似文献   

7.
The recombinant Bacillus megaterium strain WH323 was employed for the inducible production and secretion of recombinant Thermobifida fusca hydrolase (TFH). Continuous cultivations were carried out in a chemostat using either glucose or pyruvate as sole carbon source. A remarkable increase of produced TFH was detected for the pyruvate-dependent cultivation compared to glucose-dependent growth. Estimation of intracellular carbon fluxes through the central metabolism for both growth conditions using (13)C-labelled substrates revealed noticeable changes of the fluxes through the tricarboxylic acid cycle, the pentose phosphate pathway and around the pyruvate node when protein production was induced. With pyruvate as sole carbon source the observed alterations of the fluxes yielded an increased production of ATP and NADPH both required for the anabolism. Additionally, the analysis of the corresponding secretome revealed significantly reduced amounts of extracellular proteases in the medium compared to glucose-grown cultivations. Thus, pyruvate-dependent chemostat cultivation was identified as a favourable condition for production and secretion of recombinant TFH using B. megaterium as production host.  相似文献   

8.
Acetoin production in Leuconostoc mesenteroides NCDO 518   总被引:1,自引:0,他引:1  
Abstract Cell suspensions of Leuconostoc mesenteroides NCDO 518 converted pyruvate to acetoin and a small amount of lactate and acetate. Acetoin was not produced from mixtures of pyruvate and glucose unless the ratio of pyruvate to glucose was greater than 2:1. In the presence of glucose, external pyruvate was first used as an electron acceptor, being reduced to lactate, and was converted to acetoin only after the exhaustion of glucose. Use of added pyruvate as an electron acceptor suppressed ethanol formation and the products of glucose fermentation were then lactate and acetate; 2 mol of pyruvate per mol of glucose were required to completely suppress ethanol formation. It is suggested that acetoin is produced by heterofermentative organisms when available pyruvate is in excess of that required for reoxidation of all NADH produced during glucose fermentation.  相似文献   

9.
The effect of glucose concentration as a carbon source in the range of 5-20 g/L on the fermentative productions of intra-and extra-cellular ethanol, acetate, formate, oxalate, lactate, and pyruvate, as well as pyruvate decarboxylase in A. orientalis were investigated, depending on the incubation period. Intra-and extra-cellular pyruvate levels increased with rising glucose concentrations up to 15 and 20 g/L of glucose, respectively. In addition, intra-cellular pyruvate levels reached their maximum on the 48th hour in the range of 12.5-20 g/L of glucose, except for 5 and 10 g/L while extra-cellular pyruvate were at the 48th and 60th hours. As a fermentative end product, intra-and extra-cellular ethanol levels increased with increasing glucose concentrations of the growth medium and with incubation period. Activity of pyruvate decarboxylase, one of the key enzymes of the alcoholic fermentation, increased significantly with increasing glucose concentrations up to the 48th hour. Intra-and extra-cellular acetate levels increased significantly with increasing glucose concentrations of the growth medium and reached their maximums on the 48th hour, as was the case also for pyruvate. Intra-cellular formate levels increased up to 15 g/L, while extra-cellular levels increased with increasing glucose concentration. The maximum intra-and extra-cellular lactate levels were determined at 12.5 g/L and 20 g/L of glucose on the 48th hour, respectively. The results suggest that elevated ethanol production suppressed lactate and formate production, supported via possibly formed CO(2). In addition, pyruvate, as well as acetate, were used as carbon sources due to the depletion of glucose contents in the growth medium.  相似文献   

10.
Growth, substrate consumption, metabolite formation, biomass composition and respiratory parameters of Kluyveromyces marxianus ATCC 26548 were determined during aerobic batch and chemostat cultivations, using mineral medium with glucose as the sole carbon source, at 30 degrees C and pH 5.0. Carbon balances closed within 95-101% in all experiments. A maximum specific growth rate of 0.56 h(-1), a biomass yield on glucose of 0.51 g g(-1), and a maximum specific consumption of oxygen of 11.1 mmol g(-1) h(-1) were obtained during batch cultures. The concentration of excreted metabolites was very low at the culture conditions applied, representing 6% of the consumed carbon at most. Acetate and pyruvate were excreted to a larger extent than ethanol under the batch conditions, and the protein content accounted for 54.6% of the biomass dry weight. Steady states were obtained during chemostats at dilution rates of 0.1, 0.25 and 0.5 h(-1). At the two former dilution rates, cells grew at carbon limitation and the biomass yield on glucose was similar to that obtained under the batch conditions. Metabolite formation was rather low, accounting for a total of 0.005 C-mol C-mol(-1) substrate. At 0.5 h(-1), although the biomass yield on glucose was similar to the value obtained under the above-mentioned conditions, the cultivation was not under carbon limitation. Under this condition, 2-oxoglutarate, acetate, pyruvate and ethanol were the prevalent metabolites excreted. Total metabolite formation only accounted to 0.056 C-mol C-mol(-1) of substrate. A very high protein and a low carbohydrate content (71.9% and 9.6% of biomass dry weight, respectively) were measured in cells under this condition. It is concluded that K. marxianus aligns with the so-called aerobic-respiring or Crabtree-negative yeasts. Furthermore, it has one of the highest growth rates among yeasts, and a high capacity of converting sugar into biomass, even when carbon is not the limiting nutrient. These results provide useful data regarding the future application of K. marxianus in processes aimed at the production of biomass-linked compounds, with high yields and productivities.  相似文献   

11.
Derivatives of Escherichia coli C were previously described for succinate production by combining the deletion of genes that disrupt fermentation pathways for alternative products (ldhA::FRT, adhE::FRT, ackA::FRT, focA-pflB::FRT, mgsA, poxB) with growth-based selection for increased ATP production. The resulting strain, KJ073, produced 1.2 mol of succinate per mol glucose in mineral salts medium with acetate, malate, and pyruvate as significant co-products. KJ073 has been further improved by removing residual recombinase sites (FRT sites) from the chromosomal regions of gene deletion to create a strain devoid of foreign DNA, strain KJ091(DeltaldhA DeltaadhE DeltaackA DeltafocA-pflB DeltamgsA DeltapoxB). KJ091 was further engineered for improvements in succinate production. Deletion of the threonine decarboxylase (tdcD; acetate kinase homologue) and 2-ketobutyrate formate-lyase (tdcE; pyruvate formate-lyase homologue) reduced the acetate level by 50% and increased succinate yield (1.3 mol mol(-1) glucose) by almost 10% as compared to KJ091 and KJ073. Deletion of two genes involved in oxaloacetate metabolism, aspartate aminotransferase (aspC) and the NAD(+)-linked malic enzyme (sfcA) (KJ122) significantly increased succinate yield (1.5 mol mol(-1) glucose), succinate titer (700 mM), and average volumetric productivity (0.9 g L(-1) h(-1)). Residual pyruvate and acetate were substantially reduced by further deletion of pta encoding phosphotransacetylase to produce KJ134 (DeltaldhA DeltaadhE DeltafocA-pflB DeltamgsA DeltapoxB DeltatdcDE DeltacitF DeltaaspC DeltasfcA Deltapta-ackA). Strains KJ122 and KJ134 produced near theoretical yields of succinate during simple, anaerobic, batch fermentations using mineral salts medium. Both may be useful as biocatalysts for the commercial production of succinate.  相似文献   

12.
The metabolism of Desulfovibrio vulgaris Hildenborough grown on medium containing lactate or pyruvate plus a high concentration of sulfate (36 mM) was studied. Molecular growth yields were 6.7 +/- 1.3 and 10.1 +/- 1.7 g/mol for lactate and pyruvate, respectively. Under conditions in which the energy source was the sole growth-limiting factor, we observed the formation of 0.5 mol of hydrogen per mol of lactate and 0.1 mol of hydrogen per mol of pyruvate. The determination of metabolic end products revealed that D. vulgaris produced, in addition to normal end products (acetic acid, carbon dioxide, hydrogen sulfide) and molecular hydrogen, 2 and 5% of ethanol per mol of lactate and pyruvate, respectively. Power-time curves of growth of D. vulgaris on lactate and pyruvate were obtained, by the microcalorimetric Tian-Calvet apparatus. The enthalpies (delta Hmet) associated with the oxidation of these substrates and calculated from growth thermograms were -36.36 +/- 5 and -70.22 +/- 3 kJ/mol of lactate and pyruvate, respectively. These experimental values were in agreement with the homologous values assessed from the theoretical equations of D. vulgaris metabolism of both lactate and pyruvate. The hydrogen production by this sulfate reducer constitutes an efficient regulatory system of electrons, from energy source through the pathway of sulfate reduction. This hydrogen value may thus facilitate interactions between this strain and other environmental microflora, especially metagenic bacteria.  相似文献   

13.
Pyruvate decarboxylase is a key enzyme in the production of low-molecular-weight byproducts (ethanol, acetate) in biomass-directed applications of Saccharomyces cerevisiae. To investigate whether decreased expression levels of pyruvate decarboxylase can reduce byproduct formation, the PDC2 gene, which encodes a positive regulator of pyruvate-decarboxylase synthesis, was inactivated in the prototrophic strain S. cerevisiae CEN. PK113-7D. This caused a 3-4-fold reduction of pyruvate-decarboxylase activity in glucose-limited, aerobic chemostat cultures grown at a dilution rate of 0.10 h(-1). Upon exposure of such cultures to a 50 mM glucose pulse, ethanol and acetate were the major byproducts formed by the wild type. In the pdc2Delta strain, formation of ethanol and acetate was reduced by 60-70%. In contrast to the wild type, the pdc2Delta strain produced substantial amounts of pyruvate after a glucose pulse. Nevertheless, its overall byproduct formation was ca. 50% lower. The specific rate of glucose consumption after a glucose pulse to pdc2Delta cultures was about 40% lower than in wild-type cultures. This suggests that, at reduced pyruvate-decarboxylase activities, glycolytic flux is controlled by NADH reoxidation. In aerobic, glucose-limited chemostat cultures, the wild type exhibited a mixed respiro-fermentative metabolism at dilution rates above 0.30 h(-1). Below this dilution rate, sugar metabolism was respiratory. At dilution rates up to 0.20 h(-1), growth of the pdc2Delta strain was respiratory and biomass yields were similar to those of wild-type cultures. Above this dilution rate, washout occurred. The low micro(max) of the pdc2Delta strain in glucose-limited chemostat cultures indicates that occurrence of respiro-fermentative metabolism in wild-type cultures is not solely caused by competition of respiration and fermentation for pyruvate. Furthermore, it implies that inactivation of PDC2 is not a viable option for reducing byproduct formation in industrial fermentations.  相似文献   

14.
Escherichia coli strains (KJ060 and KJ073) that were previously developed for succinate production have now been modified for malate production. Many unexpected changes were observed during this investigation. The initial strategy of deleting fumarase isoenzymes was ineffective, and succinate continued to accumulate. Surprisingly, a mutation in fumarate reductase alone was sufficient to redirect carbon flow into malate even in the presence of fumarase. Further deletions were needed to inactivate malic enzymes (typically gluconeogenic) and prevent conversion to pyruvate. However, deletion of these genes (sfcA and maeB) resulted in the unexpected accumulation of D-lactate despite the prior deletion of mgsA and ldhA and the absence of apparent lactate dehydrogenase activity. Although the metabolic source of this D-lactate was not identified, lactate accumulation was increased by supplementation with pyruvate and decreased by the deletion of either pyruvate kinase gene (pykA or pykF) to reduce the supply of pyruvate. Many of the gene deletions adversely affected growth and cell yield in minimal medium under anaerobic conditions, and volumetric rates of malate production remained low. The final strain (XZ658) produced 163 mM malate, with a yield of 1.0 mol (mol glucose(-1)), half of the theoretical maximum. Using a two-stage process (aerobic cell growth and anaerobic malate production), this engineered strain produced 253 mM malate (34 g liter(-1)) within 72 h, with a higher yield (1.42 mol mol(-1)) and productivity (0.47 g liter(-1) h(-1)). This malate yield and productivity are equal to or better than those of other known biocatalysts.  相似文献   

15.
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 CO2). 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   

16.
This study was conducted to examine the effect of energy substrates in a serum-free culture medium on in vitro development of porcine embryos. Presumptive zygotes derived from in vitro fertilization were cultured in glucose-free North Carolina State University (NCSU)-23 medium with glucose, pyruvate, fructose and lactate added to the culture medium singly or in various combinations. In experiment 1, a higher percentage of embryos cleaved (53-63% vs 10-13%) and developed to the blastocyst stage (18-27% vs 0) after the single addition of glucose (5.6 mM), pyruvate (0.5 mM) or lactate (10 mM) than with no energy substrate addition or the addition only of fructose (5.6 mM). In experiment 2, the addition of pyruvate and lactate resulted in higher blastocyst formation (25%) than other combinations (6-22%), while the addition of glucose and pyruvate significantly inhibited blastocyst formation. Increasing lactate concentration, as a single energy supplement, from 5 to 20 mM significantly improved blastocyst formation (7% vs 14-18%), while no benefit was achieved from increasing pyruvate concentration up to 2 mM (experiment 3). Glucose-free NCSU-23 medium supplemented with 0.5 mM pyruvate and 5 mM lactate significantly improved blastocyst formation (28% vs 17%) compared with NCSU-23 medium supplemented with 5.6 mM glucose (experiment 4). In conclusion, pyruvate and lactate are preferable energy substrates to support in vitro development of porcine embryos cultured in a serum-free NCSU-23 medium.  相似文献   

17.
An approach to broaden the product range of the ethanologenic, gram-negative bacterium Zymomonas mobilis by means of genetic engineering is presented. Gene alaD for L-alanine dehydrogenase (EC 1.4.1.1.) from Bacillus sphaericus was cloned and introduced into Z. mobilis. Under the control of the strong promoter of the pyruvate decarboxylase (pdc) gene, the enzyme was expressed up to a specific activity of nearly 1 mu mol . min -1 . mg of protein -1 in recombinant cells. As a results of this high L-alanine dehydrogenase activity, growing cells excreted up to 10 mmol of alanine per 280 mmol of glucose utilized into a mineral salts medium. By the addition of 85 mM NH4+ to the medium, growth of the recombinant cells stopped, and up to 41 mmol alanine was secreted. As alanine dehydrogenase competed with pyruvate decarboxylase (PDC) (EC 4.1.1.1.) for the same substrate (pyruvate), PDC activity was reduced by starvation for the essential PDC cofactor thiamine PPi. A thiamine auxotrophy mutant of Z. mobilis which carried the alaD gene was starved for 40 h in glucose-supplemented mineral salts medium and then shifted to mineral salts medium with 85 mM NH4+ and 280 mmol of glucose. The recombinants excreted up to 84 mmol of alanine (7.5 g/liter) over 25 h. Alanine excretion proceeded at an initial velocity of 238 nmol . min-1 . mg [dry weight]-1. Despite this high activity, the excretion rate seemed to be a limiting factor, as the intracellular concentration of alanine was as high as 260 mM at the beginning of the excretion phase and decreased to 80 to 90 mM over 24 h.  相似文献   

18.
An approach to broaden the product range of the ethanologenic, gram-negative bacterium Zymomonas mobilis by means of genetic engineering is presented. Gene alaD for L-alanine dehydrogenase (EC 1.4.1.1.) from Bacillus sphaericus was cloned and introduced into Z. mobilis. Under the control of the strong promoter of the pyruvate decarboxylase (pdc) gene, the enzyme was expressed up to a specific activity of nearly 1 mu mol . min -1 . mg of protein -1 in recombinant cells. As a results of this high L-alanine dehydrogenase activity, growing cells excreted up to 10 mmol of alanine per 280 mmol of glucose utilized into a mineral salts medium. By the addition of 85 mM NH4+ to the medium, growth of the recombinant cells stopped, and up to 41 mmol alanine was secreted. As alanine dehydrogenase competed with pyruvate decarboxylase (PDC) (EC 4.1.1.1.) for the same substrate (pyruvate), PDC activity was reduced by starvation for the essential PDC cofactor thiamine PPi. A thiamine auxotrophy mutant of Z. mobilis which carried the alaD gene was starved for 40 h in glucose-supplemented mineral salts medium and then shifted to mineral salts medium with 85 mM NH4+ and 280 mmol of glucose. The recombinants excreted up to 84 mmol of alanine (7.5 g/liter) over 25 h. Alanine excretion proceeded at an initial velocity of 238 nmol . min-1 . mg [dry weight]-1. Despite this high activity, the excretion rate seemed to be a limiting factor, as the intracellular concentration of alanine was as high as 260 mM at the beginning of the excretion phase and decreased to 80 to 90 mM over 24 h.  相似文献   

19.
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   

20.
Substrate and energy costs of the production of exocellular enzymes from glucose and citrate by B. Iicheniformis S1684 as well as molar growth yields corrected for these costs of product formation were calculated using data from chemostat experiments. The calculations showed that 1.46-1.73 mol glucose and 2.31-2.77 mol citrate are needed for formation and excretion of 1 mol protein. Consequently, the values of the maximal product yield from substrate (Y(psm') g/mol) are 80 < Y(psm) < 95 when product is formed from glucose and 50 < Y(psm) < 60 when product is formed from citrate. The higher substrate costs for product formation from citrate are due to a higher level of CO(2) production during protein formation and a higher substrate requirement for the energy supply of product formation and excretion than when product is formed from glucose. The theoretical ATP requirement for protein synthesis could be determined reasonably well, but the energy costs of protein excretion could not be determined exactly. The energy costs of protein formation are higher than those of biomass formation or protein excretion. Molar growth yields corrected for the substrate costs of product formation were high, indicating a high efficiency of growth.Growth and production parameters were determined as well from experimental data of recycling fermentor experiments using a parameter optimization procedure based on a mathematical model describing biomass growth as a linear function of the substrate consumption rate and the rate of product formation as a linear function of biomass growth rate. The fitting procedure yielded two growth and production domains during glucose limitation. In the first domain the values for the maximal growth yield and maintenance coefficient were in agreement with those found in chemostat experiments at corresponding values of Y(spm). Domain 2 could be described best with linear growth and product formation. In domain 2 the rate of product formation decreased and more substrate became available for biomass formation. As a consequence the specific growth rate increased in the shift from domain 1 to 2. Domain 2 behavior most probably is caused by the rel-status of B. Iicheniformis S1684.  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