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.
A Corynebacterium glutamicum strain with inactivated pyruvate dehydrogenase complex and a deletion of the gene encoding the pyruvate:quinone oxidoreductase produces about 19 mM l-valine, 28 mM l-alanine and about 55 mM pyruvate from 150 mM glucose. Based on this double mutant C. glutamicumaceEpqo, we engineered C. glutamicum for efficient production of pyruvate from glucose by additional deletion of the ldhA gene encoding NAD+-dependent l-lactate dehydrogenase (LdhA) and introduction of a attenuated variant of the acetohydroxyacid synthase (△C–T IlvN). The latter modification abolished overflow metabolism towards l-valine and shifted the product spectrum to pyruvate production. In shake flasks, the resulting strain C. glutamicumaceEpqoldhA △C–T ilvN produced about 190 mM pyruvate with a Y P/S of 1.36 mol per mol of glucose; however, it still secreted significant amounts of l-alanine. Additional deletion of genes encoding the transaminases AlaT and AvtA reduced l-alanine formation by about 50%. In fed-batch fermentations at high cell densities with adjusted oxygen supply during growth and production (0–5% dissolved oxygen), the newly constructed strain C. glutamicumaceEpqoldhA △C–T ilvNalaTavtA produced more than 500 mM pyruvate with a maximum yield of 0.97 mol per mole of glucose and a productivity of 0.92 mmol g(CDW)−1 h−1 (i.e., 0.08 g g(CDW) −1 h−1) in the production phase.  相似文献   

2.
Exchange of the native Corynebacterium glutamicum promoter of the aceE gene, encoding the E1p subunit of the pyruvate dehydrogenase complex (PDHC), with mutated dapA promoter variants led to a series of C. glutamicum strains with gradually reduced growth rates and PDHC activities. Upon overexpression of the l-valine biosynthetic genes ilvBNCE, all strains produced l-valine. Among these strains, C. glutamicum aceE A16 (pJC4 ilvBNCE) showed the highest biomass and product yields, and thus it was further improved by additional deletion of the pqo and ppc genes, encoding pyruvate:quinone oxidoreductase and phosphoenolpyruvate carboxylase, respectively. In fed-batch fermentations at high cell densities, C. glutamicum aceE A16 Δpqo Δppc (pJC4 ilvBNCE) produced up to 738 mM (i.e., 86.5 g/liter) l-valine with an overall yield (YP/S) of 0.36 mol per mol of glucose and a volumetric productivity (QP) of 13.6 mM per h [1.6 g/(liter × h)]. Additional inactivation of the transaminase B gene (ilvE) and overexpression of ilvBNCD instead of ilvBNCE transformed the l-valine-producing strain into a 2-ketoisovalerate producer, excreting up to 303 mM (35 g/liter) 2-ketoisovalerate with a YP/S of 0.24 mol per mol of glucose and a QP of 6.9 mM per h [0.8 g/(liter × h)]. The replacement of the aceE promoter by the dapA-A16 promoter in the two C. glutamicum l-lysine producers DM1800 and DM1933 improved the production by 100% and 44%, respectively. These results demonstrate that C. glutamicum strains with reduced PDHC activity are an excellent platform for the production of pyruvate-derived products.  相似文献   

3.
4.
Mitochondria isolated from the Crassulacean acid metabolism plant Sedum praealtum were demonstrated to decarboxylate added malate at basal rates of 30–50 μmol mg?1 original chlorophyll h?1. The basal rate could be stimulated markedly by the addition of ADP, oxaloacetic acid, an uncoupler of oxidative phosphorylation, or NAD, with maximum rates of 70–100 μmol mg?1 original chlorophyll h?1 observed. These observed rates were high enough to account for a large proportion of the estimated rate of malate decarboxylation in vivo. The major products of malate oxidation by the mitochondria in most cases were found to be pyruvate and CO2, indicating that malate oxidation in these mitochondria proceeds mainly through NAD malic enzyme rather than NAD malate dehydrogenase. Under conditions employed little of the pyruvate formed was further oxidized, suggesting a fate other than oxidation (conversion to starch) for this pyruvate. Malate decarboxylation by mitochondria and by partially purified NAD malic enzyme was markedly inhibited by NaHCO3. A possible physiological role is suggested for this inhibition as a feedback control on the enzyme.  相似文献   

5.
Malic acid, a petroleum-derived C4-dicarboxylic acid that is used in the food and beverage industries, is also produced by a number of microorganisms that follow a variety of metabolic routes. Several members of the genus Aspergillus utilize a two-step cytosolic pathway from pyruvate to malate known as the reductive tricarboxylic acid (rTCA) pathway. This simple and efficient pathway has a maximum theoretical yield of 2 mol malate/mol glucose when the starting pyruvate originates from glycolysis. Production of malic acid by Aspergillus oryzae NRRL 3488 was first improved by overexpression of a native C4-dicarboxylate transporter, leading to a greater than twofold increase in the rate of malate production. Overexpression of the native cytosolic alleles of pyruvate carboxylase and malate dehydrogenase, comprising the rTCA pathway, in conjunction with the transporter resulted in an additional 27 % increase in malate production rate. A strain overexpressing all three genes achieved a malate titer of 154 g/L in 164 h, corresponding to a production rate of 0.94 g/L/h, with an associated yield on glucose of 1.38 mol/mol (69 % of the theoretical maximum). This rate of malate production is the highest reported for any microbial system.  相似文献   

6.
A sensitive isotope exchange method was developed to assess the requirements for and compartmentation of pyruvate and oxalacetate production from malate in proliferating and nonproliferating human fibroblasts. Malatedependent pyruvate production (malic enzyme activity) in the particulate fraction containing the mitochondria was dependent on either NAD+ or NADP+. The production of pyruvate from malate in the soluble, cytosolic fraction was strictly dependent on NADP+. Oxalacetate production from malate (malate dehydrogenase, EC 1.1.1.37) in both the particulate and soluble fraction was strictly dependent on NAD+. Relative to nonproliferating cells, NAD+-linked malic enzyme activity was slightly reduced and the NADP+-linked activity was unchanged in the particulate fraction of serum-stimulated, exponentially proliferating cells. However, a reduced activity of particulate malate dehydrogenase resulted in a two-fold increase in the ratio of NAD(P)+-linked malic enzyme to NAD+-linked malate dehydrogenase activity in the particulate fraction of proliferating fibroblasts. An increase in soluble NADP+-dependent malic enzyme activity and a decrease in NAD+-linked malate dehydrogenase indictated an increase in the ratio of pyruvate-producing to oxalacetate-producing malate oxidase activity in the cytosol of proliterating cells. These coordinate changes may affect the relative amount of malate that is oxidized to oxalacetate and pyruvate in proliferating cells and, therefore, the efficient utilization of glutamine as a respiratory fuel during cell proliferation.  相似文献   

7.
The effect of anaplerotic pathways activation on CO2-dependent anaerobic glucose utilization by Escherichia coli strains deficient in the main fermentation pathways and possessing a modified system of glucose transport and phosphorylation was studied. Intracellular CO2 generation in the strains was ensured resulting from oxidative decarboxylation of pyruvic acid by pyruvate dehydrogenase. Sodium bicarbonate dissolved in the medium was used as an external source of CO2. The genes of heterologous pyruvate carboxylase and native NADH-dependent malic enzyme were overexpressed in the strains to allow anaplerotic carboxylation of pyruvic acid to oxaloacetic or malic acid. The ability of the strains to reoxidize NADH utilizing carboxylation products was additionally increased due to enhanced expression of malate dehydrogenase gene. In the case of endogenous CO2 formation, the activation of anaplerotic pathways did not cause a notable increase in the anaerobic glucose consumption by the constructed strains. At the same time, the expression of pyruvate carboxylase led to a pronounced decrease in the secretion of pyruvic acid with the concomitant increase in the yield of four-carbon metabolites. Further enhancement of NADH-dependent malic enzyme expression provoked activation of a pyruvate–oxaloacetate–malate–pyruvate futile cycle in the strains. The availability in the medium of the external CO2 source sharply increased the anaerobic utilization of glucose by strains expressing pyruvate carboxylase. The activity of the futile cycle has raised with the increased malic enzyme expression and dropped upon enhancement of malate dehydrogenase expression. As a result, the efficiency of CO2-dependent anaerobic glucose utilization coupled to the formation of four-carbon carboxylation products increased in the studied strains resulting from the primary anaplerotic conversion of pyruvic acid into oxaloacetic acid followed by the involvement of the precursor formed in NADH-consuming biosynthetic reactions dominating over the reactions of the revealed futile cycle.  相似文献   

8.
We engineered P. putida for the production of isobutanol from glucose by preventing product and precursor degradation, inactivation of the soluble transhydrogenase SthA, overexpression of the native ilvC and ilvD genes, and implementation of the feedback‐resistant acetolactate synthase AlsS from Bacillus subtilis, ketoacid decarboxylase KivD from Lactococcus lactis, and aldehyde dehydrogenase YqhD from Escherichia coli. The resulting strain P. putida Iso2 produced isobutanol with a substrate specific product yield (YIso/S) of 22 ± 2 mg per gram of glucose under aerobic conditions. Furthermore, we identified the ketoacid decarboxylase from Carnobacterium maltaromaticum to be a suitable alternative for isobutanol production, since replacement of kivD from L. lactis in P. putida Iso2 by the variant from C. maltaromaticum yielded an identical YIso/S. Although P. putida is regarded as obligate aerobic, we show that under oxygen deprivation conditions this bacterium does not grow, remains metabolically active, and that engineered producer strains secreted isobutanol also under the non‐growing conditions.  相似文献   

9.
On the basis of our previous studies of microbial L-valine production under oxygen deprivation, we developed isobutanol-producing Corynebacterium glutamicum strains. The artificial isobutanol synthesis pathway was composed of the first three steps of the L-valine synthesis pathway; and the subsequent Ehrlich Pathway: pyruvate was converted to 2-ketoisovalerate in the former reactions; and the 2-keto acid was decarboxylated into isobutyraldehyde, and subsequently reduced into isobutanol in the latter reactions. Although there exists redox cofactor imbalance in the overall reactions, i.e., NADH is generated via glycolysis whereas NADPH is required to synthesize isobutanol, it was resolved by taking advantage of the NAD-preferring mutant acetohydroxy acid isomeroreductase encoded by ilvCTM and the NAD-specific alcohol dehydrogenase encoded by adhA. Each enzyme activity to synthesize isobutanol was finely tuned by using two kinds of lac promoter derivatives. Efficient suppression of succinate by-production and improvement of isobutanol yield resulted from inactivation of pckA, which encodes phosphoenolpyruvate carboxykinase, whereas glucose consumption and isobutanol production rates decreased because of the elevated intracellular NADH/NAD+ ratio. On the other hand, introduction of the exogenous Entner–Doudoroff pathway effectively enhanced glucose consumption and productivity. Overexpression of phosphoenolpyruvate:carbohydrate phosphotransferase system specific to glucose and deletion of ilvE, which encodes branched-chain amino acid transaminase, further suppressed by-products and improved isobutanol productivity. Finally, the produced isobutanol concentration reached 280 mM at a yield of 84% (mol/mol glucose) in 24 h.  相似文献   

10.
A recent effort to improve malic acid production by Saccharomyces cerevisiae by means of metabolic engineering resulted in a strain that produced up to 59 g liter−1 of malate at a yield of 0.42 mol (mol glucose)−1 in calcium carbonate-buffered shake flask cultures. With shake flasks, process parameters that are important for scaling up this process cannot be controlled independently. In this study, growth and product formation by the engineered strain were studied in bioreactors in order to separately analyze the effects of pH, calcium, and carbon dioxide and oxygen availability. A near-neutral pH, which in shake flasks was achieved by adding CaCO3, was required for efficient C4 dicarboxylic acid production. Increased calcium concentrations, a side effect of CaCO3 dissolution, had a small positive effect on malate formation. Carbon dioxide enrichment of the sparging gas (up to 15% [vol/vol]) improved production of both malate and succinate. At higher concentrations, succinate titers further increased, reaching 0.29 mol (mol glucose)−1, whereas malate formation strongly decreased. Although fully aerobic conditions could be achieved, it was found that moderate oxygen limitation benefitted malate production. In conclusion, malic acid production with the engineered S. cerevisiae strain could be successfully transferred from shake flasks to 1-liter batch bioreactors by simultaneous optimization of four process parameters (pH and concentrations of CO2, calcium, and O2). Under optimized conditions, a malate yield of 0.48 ± 0.01 mol (mol glucose)−1 was obtained in bioreactors, a 19% increase over yields in shake flask experiments.In recent years, biologically produced 1,4-dicarboxylic acids (succinate, malate, and fumarate) have attracted great interest as more sustainable replacements for oil-derived commodity chemicals, such as maleic anhydride (50). Malate is currently mainly produced via petrochemical routes for use in food and beverages (18). Development of a biotechnological production process started in the early 1960s with the investigation of the natural malate producer Aspergillus flavus (2). Although process improvements eventually resulted in high product yields and productivities (6), the potential production of aflatoxins (20) prevented the use of this filamentous fungus in industry. Other investigated natural malate-producing fungi (listed in reference 51) produced insufficient malate for industrial use. With the rational design options of metabolic engineering, microorganisms that do not naturally produce large amounts of malic acid may also be considered as production platforms. Wild-type Saccharomyces cerevisiae strains produce little if any malate but would be an interesting starting point for the construction of an efficient malate producer. This yeast has a relatively high tolerance to organic acids and low pH, and due to its role as a model organism in research, a well-developed metabolic engineering toolbox is available. In addition, wild-type S. cerevisiae strains have GRAS (Generally Regarded As Safe) status, so that modified strains are more likely to be allowed in the production of food-grade malic acid.One of the main challenges in the development of an organic acid-producing strain of S. cerevisiae has been the elimination of ethanol formation, which in wild-type strains occurs even under aerobic conditions when glucose concentrations are high (45). Deletion of the pyruvate decarboxylase-encoding genes was found to prevent ethanolic fermentation (17). After evolutionary engineering to remove the growth defects usually associated with pyruvate decarboxylase-negative S. cerevisiae strains, a strain was obtained that produced large amounts of pyruvate, a direct precursor to malate, when grown on glucose (42). Subsequent overexpression of the anaplerotic enzyme pyruvate carboxylase, a cytosolically relocalized malate dehydrogenase and a heterologous malate transporter from Schizosaccharomyces pombe led to a strain that produced significant amounts of malate (51). Cultivation in calcium carbonate (CaCO3)-buffered shake flasks resulted in malate titers of up to 59 g liter−1 at a yield of 0.42 mol (mol glucose)−1.There are many differences between cultivation in shake flasks and cultivation in (laboratory or industrial) bioreactors. As shake flask cultures lack online pH monitoring and control, there is often significant pH variation over time. The pH is of particular importance. If the yeast can be persuaded to produce organic acids at lower pH values, this reduces the need for active neutralization and thereby reduces by-product formation such as gypsum. However, thermodynamic constraints on acid export, as well as increased stress levels from (undissociated) acid and the low pH, often limit the ability of the microorganisms to produce acids at low pH (32, 43). For this reason, the poorly soluble compound CaCO3 has traditionally been used to maintain a near-neutral pH in malic acid-producing microbial cultures (6, 29, 51). Adding CaCO3 also gives increased concentrations of bicarbonate (and thereby CO2), a substrate for pyruvate carboxylase in the carboxylation of pyruvate (a C3 carbon molecule) to oxaloacetate (C4 carbon), as well as calcium. Calcium is known to be involved in cellular signaling pathways (22, 26, 33, 46) and to influence pyruvate carboxylase activity (21, 24). Finally, oxygen transfer rates in shake flasks are often poor compared to those in stirred (laboratory) bioreactors. The formation of significant concentrations (25 g liter−1) of glycerol, a well-known redox sink in S. cerevisiae (41), in shake flask cultures of the engineered malate-producing strain (51) was a strong indication of oxygen limitation.Initial experiments in aerobic, pH-controlled bioreactor cultures of the malate- and succinate-producing Saccharomyces cerevisiae strain RWB525 yielded only low concentrations of these C4 dicarboxylic acids. The goal of the present study was to identify process parameters that explain the different production levels in shake flask and bioreactor cultures. To this end, we analyzed, both separately and in combination, the impact of culture pH and concentrations of calcium, carbon dioxide, and oxygen on the production of malate and succinate.  相似文献   

11.
12.
13.
14.
15.
The cell energy fraction that powered maintenance and expression of genes encoding pro-phage elements, pta-ack cluster, early sporulation, sugar ABC transporter periplasmic proteins, 6-phosphofructokinase, pyruvate kinase, and fructose-1,6-disphosphatase in acetogen Clostridium sp. MT871 was re-directed to power synthetic operon encoding isobutanol biosynthesis at the expense of these genes achieved via their elimination. Genome tailoring decreased cell duplication time by 7.0 ± 0.1 min (p < 0.05) compared to the parental strain, with intact genome and cell duplication time of 68 ± 1 min (p < 0.05). Clostridium sp. MT871 with tailored genome was UVC-mutated to withstand 6.1 % isobutanol in fermentation broth to prevent product inhibition in an engineered commercial biocatalyst producing 5 % (674.5 mM) isobutanol during two-step continuous fermentation of CO2/H2 gas blend. Biocatalyst Clostridium sp. MT871RG11IBR6 was engineered to express six copies of synthetic operon comprising optimized synthetic format dehydrogenase, pyruvate formate lyase, acetolactate synthase, acetohydroxyacid reductoisomerase, 2,3-dihydroxy-isovalerate dehydratase, branched-chain alpha-ketoacid decarboxylase gene, aldehyde dehydrogenase, and alcohol dehydrogenase, regaining cell duplication time of 68 ± 1 min (p < 0.05) for the parental strain. This is the first report on isobutanol production by an engineered acetogen biocatalyst suitable for commercial manufacturing of this chemical/fuel using continuous fermentation of CO2/H2 blend thus contributing to the reversal of global warming.  相似文献   

16.
《Insect Biochemistry》1984,14(6):639-643
Extracts of tissues of the lower termites, Reticulitermes flavipes and Coptotermes lacteus, and the higher termite, Nasutitermes exitiosus, possess acetyl-CoA synthetase and all the enzymes of the tricarboxylic acid cycle and are thus able to oxidize acetate to CO2. The specific activities of these enzymes in R. flavipes are sufficient to cope with the rate of acetogenesis by the gut microbiota. The presence of the malic enzyme and malate dehydrogenase, but not pyruvate carboxylase or phosphoenolpyruvate carboxykinase, indicates that they may be important as anaplerotic enzymes for the conversion of pyruvate to oxalacetate. An apparent absence of pyruvate dehydrogenase in all termites suggests that they do not convert pyruvate to acetyl-CoA, but rather convert acetate (transported from the hindgut) to this compound. All the enzymes of glycolysis were present in termite extracts. Thus any glucose absorbed from the midgut, and originating from hydrolysis of cellulose by salivary or midgut enzymes, can be metabolized by termites as an energy source.  相似文献   

17.
Corynebacterium glutamicum lacking the succinate dehydrogenase complex can produce succinate aerobically with acetate representing the major byproduct. Efforts to increase succinate production involved deletion of acetate formation pathways and overexpression of anaplerotic pathways, but acetate formation could not be completely eliminated. To address this issue, we constructed a pathway for recycling wasted carbon in succinate-producing C. glutamicum. The acetyl-CoA synthetase from Bacillus subtilis was heterologously introduced into C. glutamicum for the first time. The engineered strain ZX1 (pEacsA) did not secrete acetate and produced succinate with a yield of 0.50 mol (mol glucose)−1. Moreover, in order to drive more carbon towards succinate biosynthesis, the native citrate synthase encoded by gltA was overexpressed, leading to strain ZX1 (pEacsAgltA), which showed a 22% increase in succinate yield and a 62% decrease in pyruvate yield compared to strain ZX1 (pEacsA). In fed-batch cultivations, strain ZX1 (pEacsAgltA) produced 241 mM succinate with an average volumetric productivity of 3.55 mM h−1 and an average yield of 0.63 mol (mol glucose) −1, making it a promising platform for the aerobic production of succinate at large scale.  相似文献   

18.

Background

Pyruvate kinase is an important element in flux control of the intermediate metabolism. It catalyzes the irreversible conversion of phosphoenolpyruvate into pyruvate and is under allosteric control. In Corynebacterium glutamicum, this enzyme was regarded as promising target for improved production of lysine, one of the major amino acids in animal nutrition. In pyruvate kinase deficient strains the required equimolar ratio of the two lysine precursors oxaloacetate and pyruvate can be achieved through concerted action of the phosphotransferase system (PTS) and phosphoenolpyruvate carboxylase (PEPC), whereby a reduced amount of carbon may be lost as CO2 due to reduced flux into the tricarboxylic acid (TCA) cycle. In previous studies, deletion of pyruvate kinase in lysine-producing C. glutamicum, however, did not yield a clear picture and the exact metabolic consequences are not fully understood.

Results

In this work, deletion of the pyk gene, encoding pyruvate kinase, was carried out in the lysine-producing strain C. glutamicum lysCfbr, expressing a feedback resistant aspartokinase, to investigate the cellular response to deletion of this central glycolytic enzyme. Pyk deletion was achieved by allelic replacement, verified by PCR analysis and the lack of in vitro enzyme activity. The deletion mutant showed an overall growth behavior (specific growth rate, glucose uptake rate, biomass yield) which was very similar to that of the parent strain, but differed in slightly reduced lysine formation, increased formation of the overflow metabolites dihydroxyacetone and glycerol and in metabolic fluxes around the pyruvate node. The latter involved a flux shift from pyruvate carboxylase (PC) to PEPC, by which the cell maintained anaplerotic supply of the TCA cycle. This created a metabolic by-pass from PEP to pyruvate via malic enzyme demonstrating its contribution to metabolic flexibility of C. glutamicum on glucose.

Conclusion

The metabolic flux analysis performed illustrates the high flexibility of the metabolic network of C. glutamicum to compensate for external perturbation. The organism could almost maintain its growth and production performance through a local redirection of the metabolic flux, thereby fulfilling all anabolic and catabolic needs. The formation of the undesired overflow metabolites dihydroxyacetone and glycerol, in the deletion mutant, however, indicates a limiting capacity of the metabolism down-stream of their common precursor glyceraldehyde 3-phosphate and opens possibilities for further strain engineering.  相似文献   

19.
20.
Sabatini A  Vacca A  Iotti S 《PloS one》2012,7(1):e29529
A novel procedure is presented which, by balancing elements and electric charge of biochemical reactions which occur at constant pH and pMg, allows assessing the thermodynamics properties of reaction ΔrG ′0, ΔrH ′0, ΔrS ′0 and the change in binding of hydrogen and magnesium ions of these reactions. This procedure of general applicability avoids the complex calculations required by the use of the Legendre transformed thermodynamic properties of formation ΔfG ′0, ΔfH ′0 and ΔfS ′0 hitherto considered an obligatory prerequisite to deal with the thermodynamics of biochemical reactions. As a consequence, the term “conditional” is proposed in substitution of “Legendre transformed” to indicate these thermodynamics properties. It is also shown that the thermodynamic potential G is fully adequate to give a criterion of spontaneous chemical change for all biochemical reactions and then that the use of the Legendre transformed G′ is unnecessary. The procedure proposed can be applied to any biochemical reaction, making possible to re-unify the two worlds of chemical and biochemical thermodynamics, which so far have been treated separately.  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