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 15 毫秒
1.
The fermentative metabolism of Escherichia coli was reengineered to efficiently convert glycerol to succinate under anaerobic conditions without the use of foreign genes. Formate and ethanol were the dominant fermentation products from glycerol in wild-type Escherichia coli ATCC 8739, followed by succinate and acetate. Inactivation of pyruvate formate-lyase (pflB) in the wild-type strain eliminated the production of formate and ethanol and reduced the production of acetate. However, this deletion slowed growth and decreased cell yields due to either insufficient energy production or insufficient levels of electron acceptors. Reversing the direction of the gluconeogenic phosphoenolpyruvate carboxykinase reaction offered an approach to solve both problems, conserving energy as an additional ATP and increasing the pool of electron acceptors (fumarate and malate). Recruiting this enzyme through a promoter mutation (pck*) to increase expression also increased the rate of growth, cell yield, and succinate production. Presumably, the high NADH/NAD+ ratio served to establish the direction of carbon flow. Additional mutations were also beneficial. Glycerol dehydrogenase and the phosphotransferase-dependent dihydroxyacetone kinase are regarded as the primary route for glycerol metabolism under anaerobic conditions. However, this is not true for succinate production by engineered strains. Deletion of the ptsI gene or any other gene essential for the phosphotranferase system was found to increase succinate yield. Deletion of pflB in this background provided a further increase in the succinate yield. Together, these three core mutations (pck*, ptsI, and pflB) effectively redirected carbon flow from glycerol to succinate at 80% of the maximum theoretical yield during anaerobic fermentation in mineral salts medium.Renewable bioenergy offers the potential to solve many environmental problems associated with petroleum-based fuels and chemicals. Biodiesel is produced by reacting vegetable oil or animal fat with alcohol (methanol or ethanol) and used as a transportation fuel in many countries (33). Glycerol is formed as an abundant waste product with limited commercial uses. As the worldwide production of biodiesel continues to increase, the development of effective uses for glycerol may prove essential for the economics and competitiveness of the biodiesel industry. The value of glycerol waste from biodiesel is similar to that of sugars currently used to produce fuel ethanol. Bioconversion of glycerol to higher-value products that replace petroleum, such as polymers, surfactants, solvents, and chemical intermediates, represents an opportunity to decrease waste and improve the economics of the biodiesel industry (5).Many previous investigations have focused on the fermentative production of 1,3-propanediol (1,3-PD) from glycerol (2, 26, 35). Microorganisms including Klebsiella (14), Citrobacter (6), Enterobacter (1), Lactobacillus (29), and Clostridium (10, 28) have the native ability to ferment glycerol into this product. Dupont and Genencor have commercialized a 1,3-PD-based polyester, a condensation product of 1,3-PD and terephthalic acid using glucose as the feedstock. Potential demand for this polymer is estimated to be 1 billion to 2 billion pounds per year over the next 10 years (26). Other investigations of glycerol fermentation have described the production of hydrogen and ethanol (15), polyhydroxyalkanoates (PHAs) (20, 27), glyceric acid (13), and small amounts of succinate (21).Succinic acid is currently used as a specialty chemical in the agricultural, food, and pharmaceutical industries (24, 34). It has also been identified by the U.S. Department of Energy as one of the top 12 building block chemicals (31) because it can be converted into a wide variety of products, including green solvents, pharmaceutical products, and biodegradable plastics (24, 34). Succinate is primarily produced from petroleum-derived maleic anhydride. Recent increases in the petroleum price have generated considerable interest in the fermentative production of succinate from sugars using either natural succinate-producing rumen bacteria or metabolically engineered Escherichia coli strains (24, 36, 38). Succinate can also be produced from glycerol by rumen bacteria, such as Anaerobiospirillum succiniciproducens (21). However, these strains require complex nutrients that increase costs of production, purification, and waste treatment.E. coli has been previously engineered for the commercial production of 1,3-PD from sugars by Dupont and Genecor (26). It is an excellent organism for biotechnology applications but was long thought incapable of anaerobic growth on glycerol (23). Recent studies demonstrated that E. coli can ferment glycerol anaerobically (8, 11, 25, 33), and a new model was proposed for glycerol fermentation (11). In this model, glycerol is metabolized through the glycerol dehydrogenase (encoded by gldA) and dihydroxyacetone kinase (encoded by dhaKLM) pathway with the production of ethanol and acetate as primary fermentation products (11). Small amounts of succinate and 1,2-propanediol were also produced. Native genes encoding glycerol dehydrogenase and dihydroxyacetone kinase were expressed from a plasmid to increase the rates of glycerol metabolism and ethanol production (32). Succinate production has also been increased by expressing Clostridium freundii dihydroxyacetone kinase (encoded by dhaKL) (11). However, neither of these enhanced pathways would appear suitable for efficient succinate production due to the absence of net ATP production and the requirement for phosphoenolpyruvate as a phosphoryl donor for dihydroxyacetone, limiting the carboxylation of this intermediate (Fig. (Fig.11).Open in a separate windowFIG. 1.Glycerol uptake and fermentation by E. coli. (A) Native E. coli pathways. Bold black arrows represent dominant fermentation reactions prior to engineering; thin black arrows represent minor fermentation reactions. GlpK and GlpD are thought to function primarily during aerobic metabolism. Pathways are based on current reviews in EcoSal (3, 4, 22), data available in Ecocyc (19), and primary literature (11, 12, 18, 25, 30). (B) Engineered pathway for the fermentative metabolism of glycerol to succinate. Bold black arrows represent the engineered reactions for glycerol fermentation to succinate as the dominant product; thin black arrows represent minor fermentation reactions in the engineered strain. Dashed arrows represent reactions that are not functional due to deletions in ptsI and pflB. Deleted genes are marked with a black X. In native E. coli strains, phosphoenolpyruvate carboxykinase functions during gluconeogenesis to produce phosphoenolpyruvate. Mutational activation of the pck gene (denoted pck*) allows this enzyme to function in the reverse direction and to serve as the dominant carboxylation step, conserving energy as ATP. With this engineered pathway, competing needs for PEP have been eliminated and net ATP production has been increased. PEP is boxed to indicate a common pool. Abbreviations: DHA, dihydroxyacetone; DHAP, dihydroxyacetone 3-phosphate; PEP, phosphoenolpyruvate; G3P, glycerol 3-phosphate; GA3P, glyceraldehydes 3-phosphate.Previous studies in our laboratory (16, 17, 36, 38) have engineered E. coli ATCC 8739 for the efficient production of succinate from glucose by recruiting genes from alternative pathways (36, 38). In this paper, we report the use of a similar approach to engineer strains for succinate production from glycerol in mineral salts medium.  相似文献   

2.
3.
在甘油发酵初期,采用30℃的温度促进细胞生长;随后采用40℃的温度促进甘油的产出.与传统的35℃恒温发酵相比,在摇瓶及15L气升式反应器中采用变温发酵,最终甘油的产率分别提高28.7%,34.0%.  相似文献   

4.
重组大肠杆菌生物转化甘油生产3-羟基丙酸   总被引:1,自引:0,他引:1  
目的:以甘油为底物构建高效的3-羟基丙酸生产菌株。方法:以自身携带乙醛脱氢酶的E.coli BL21(DE3)plysS作为宿主,异源表达源自Klebsiella pneumoniae的甘油脱水酶基因dhaB。结果:重组菌E.coli HP获得的甘油脱水酶比活力在1.0mmol/L IPTG的诱导下达到了77.2 U/mg,摇瓶条件下,3-HP的最大产量为5.44 g/L,摩尔转化率为53%,该产量比目前报道的最高水平(4.4 g/L)提高了23.6%。结论:重组菌株E.coli HP实现了甘油向3-羟基丙酸(3-HP)的高效生物转化。  相似文献   

5.
Given its availability and low price, glycerol has become an ideal feedstock for the production of fuels and chemicals. We recently reported the pathways mediating the metabolism of glycerol in Escherichia coli under anaerobic and microaerobic conditions. In this work, we engineer E. coli for the efficient conversion of glycerol to d-lactic acid (d-lactate), a negligible product of glycerol metabolism in wild-type strains. A homofermentative route for d-lactate production was engineered by overexpressing pathways involved in the conversion of glycerol to this product and blocking those leading to the synthesis of competing by-products. The former included the overexpression of the enzymes involved in the conversion of glycerol to glycolytic intermediates (GlpK-GlpD and GldA-DHAK pathways) and the synthesis of d-lactate from pyruvate (d-lactate dehydrogenase). On the other hand, the synthesis of succinate, acetate, and ethanol was minimized through two strategies: (i) inactivation of pyruvate-formate lyase (ΔpflB) and fumarate reductase (ΔfrdA) (strain LA01) and (ii) inactivation of fumarate reductase (ΔfrdA), phosphate acetyltransferase (Δpta), and alcohol/acetaldehyde dehydrogenase (ΔadhE) (strain LA02). A mutation that blocked the aerobic d-lactate dehydrogenase (Δdld) also was introduced in both LA01 and LA02 to prevent the utilization of d-lactate. The most efficient strain (LA02Δdld, with GlpK-GlpD overexpressed) produced 32 g/liter of d-lactate from 40 g/liter of glycerol at a yield of 85% of the theoretical maximum and with a chiral purity higher than 99.9%. This strain exhibited maximum volumetric and specific productivities for d-lactate production of 1.5 g/liter/h and 1.25 g/g cell mass/h, respectively. The engineered homolactic route generates 1 to 2 mol of ATP per mol of d-lactate and is redox balanced, thus representing a viable metabolic pathway.Lactic acid (lactate) and its derivatives have many applications in the food, pharmaceutical, and polymer industries (13, 30). An example is polylactic acid, a renewable, biodegradable, and environmentally friendly polymer produced from d- and l-lactate (19). In this context, biological processes have the advantage of being able to produce chirally pure lactate from inexpensive media containing only the carbon source and mineral salts (43). While lactic acid bacteria traditionally have been used in the production of d-lactate from carbohydrate-rich feedstocks, several laboratories recently have reported alternative biocatalysts (13, 30), many of which are engineered Escherichia coli strains that produce d- or l-lactate (4, 8, 50, 51, 52).Unlike the aforementioned reports, which have dealt with the use of carbohydrates, our work focuses on the use of glycerol as a carbon source for the production of d-lactate. Glycerol has become an inexpensive and abundant substrate due to its generation in large amounts as a by-product of biodiesel and bioethanol production (18, 32, 47). The conversion of glycerol to higher-value products has been proposed as a path to economic viability for the biofuels industry (47). One such product is lactate, whose production could be readily integrated into existing biodiesel and bioethanol facilities, thus establishing true biorefineries.Although many microorganisms are able to metabolize glycerol (25), the use of industrial microbes such as E. coli could greatly accelerate the development of platforms to produce fuels and chemicals from this carbon source. We recently reported on the ability of E. coli to metabolize glycerol under either anaerobic or microaerobic conditions and identified the environmental and metabolic determinants of these processes (9, 11, 28). In one of the studies, the pathways involved in the microaerobic utilization of glycerol were elucidated, and they are shown in Fig. Fig.11 (9). A common characteristic of glycerol metabolism under either anaerobic or microaerobic conditions is the generation of ethanol as the primary product and the negligible production of lactate (6, 9, 11, 28). In the work reported here, the knowledge base created by the aforementioned studies was used to engineer E. coli for the efficient conversion of glycerol to d-lactate in minimal medium. The engineered strains hold great promise as potential biocatalysts for the conversion of low-value glycerol streams to a higher-value product like d-lactate.Open in a separate windowFIG. 1.Pathways involved in the microaerobic utilization of glycerol in E. coli (9). Genetic modifications supporting the metabolic engineering strategies employed in this work are illustrated by thicker lines (overexpression of gldA-dhaKLM, glpK-glpD, and ldhA) or cross bars (disruption of pflB, pta, adhE, frdA, and dld). Broken lines illustrate multiple steps. Relevant reactions are represented by the names of the gene(s) coding for the enzymes: aceEF-lpdA, pyruvate dehydrogenase complex; adhE, acetaldehyde/alcohol dehydrogenase; ackA, acetate kinase; dhaKLM, dihydroxyacetone kinase; dld, respiratory d-lactate dehydrogenase; fdhF, formate dehydrogenase, part of the formate hydrogenlyase complex; frdABCD, fumarate reductase; gldA, glycerol dehydrogenase; glpD, aerobic glycerol-3-phosphate dehydrogenase; glpK, glycerol kinase; hycB-I, hydrogenase 3, part of the formate hydrogenlyase complex; ldhA, fermentative d-lactate dehydrogenase; pflB, pyruvate formate-lyase; pta, phosphate acetyltransferase; pykF, pyruvate kinase. Abbreviations: DHA, dihydroxyacetone; DHAP, DHA phosphate; G-3-P, glycerol-3-phosphate; PEP, phosphoenolpyruvate; PYR, pyruvate; P/O, amount of ATP produced in the oxidative phosphorylation per pair of electrons transferred through the electron transport system; QH2, reduced quinones.  相似文献   

6.
琥珀酸是一种具有重要应用价值的生物基平台化合物。对大肠杆菌focA-pflB ldhA突变株QQS101在严格厌氧条件下生长和葡萄糖代谢能力进行了考察,比较分析了葡萄糖与大肠杆菌混合酸发酵产物的单位碳的还原程度,认为非严格厌氧条件有利于QQS101发酵葡萄糖积累琥珀酸,进一步对有氧生长碳源进行了对比试验的结果表明,以木糖支持有氧生长,QQS101摇瓶发酵39 h消耗葡萄糖37.6 g/L,琥珀酸的产量达到31.01 g/L,摩尔产率为1.258 mol Succinate/mol Glucose。发酵过程中,丙氨酸的添加能够提高琥珀酸的摩尔产率。  相似文献   

7.
Glycerol has become a desirable feedstock for the production of fuels and chemicals due to its availability and low price, but many barriers to commercialization remain. Previous investigators have made significant improvements in the yield of ethanol from glycerol. We have developed a fermentation process for the efficient microaerobic conversion of glycerol to ethanol by Escherichia coli that presents solutions to several other barriers to commercialization: rate, titer, specific productivity, use of inducers, use of antibiotics, and safety. To increase the rate, titer, and specific productivity to commercially relevant levels, we constructed a plasmid that overexpressed glycerol uptake genes dhaKLM, gldA, and glpK, as well as the ethanol pathway gene adhE. To eliminate the cost of inducers and antibiotics from the fermentation, we used the adhE and icd promoters from E. coli in our plasmid, and we implemented glycerol addiction to retain the plasmid. To address the safety issue of off-gas flammability, we optimized the fermentation process with reduced-oxygen sparge gas to ensure that the off-gas remained nonflammable. These advances represent significant progress toward the commercialization of an E. coli-based glycerol-to-ethanol process.  相似文献   

8.
产琥珀酸重组大肠杆菌的发酵性能研究   总被引:2,自引:0,他引:2  
研究了重组大肠杆菌JM001(△ppc)/pTrc99a-pck发酵产琥珀酸的性能,结果表明厌氧条件下其耗糖能力和产酸能力分别为对照菌株JM001的4.2倍和15.3倍。进一步优化发酵条件表明:采用接入菌泥的发酵方式比按照10%接种量转接厌氧发酵的效果要好,琥珀酸的对葡萄糖的质量收率提高了约10%,且副产物乙酸的量进一步降低。初始葡萄糖浓度高于60g/L时会对菌株的生长和产酸产生抑制,且浓度越高,抑制作用越明显。7L发酵罐放大实验中,整个厌氧发酵阶段葡萄糖的消耗速率为0.42g/(L.h),琥珀酸对葡萄糖的质量收率为67.75%,琥珀酸的生产强度为0.28g/(L.h)。  相似文献   

9.
The effects of acetate and butyrate during glycerol fermentation to 1,3-propanediol at pH 7.0 by Clostridium butyricum CNCM 1211 were studied. At pH 7.0, the calculated quantities of undissociated acetic and butyric acids were insufficient to inhibit bacterial growth. The initial addition of acetate or butyrate at concentrations of 2.5 to 15 gL−1 had distinct effects on the metabolism and growth of Clostridium butyricum. Acetate increased the biomass and butyrate production, reducing the lag time and 1,3-propanediol production. In contrast, the addition of butyrate induced an increase in 1,3-propanediol production (yield: 0.75 mol/mol glycerol, versus 0.68 mol/mol in the butyrate-free culture), and reduced the biomass and butyrate production. It was calculated that reduction of butyrate production could provide sufficient NADH to increase 1,3-propanediol production. The effects of acetate and butyrate highlight the metabolic flexibility of Cl. butyricum CNCM 1211 during glycerol fermentation. Received: 2 January 2001 / Accepted: 6 February 2001  相似文献   

10.
Six commercial wine yeast strains and three nonindustrial strains (two laboratory strains and one haploid strain derived from a wine yeast strain) were engineered to produce large amounts of glycerol with a lower ethanol yield. Overexpression of the GPD1 gene, encoding a glycerol-3-phosphate dehydrogenase, resulted in a 1.5- to 2.5-fold increase in glycerol production and a slight decrease in ethanol formation under conditions simulating wine fermentation. All the strains overexpressing GPD1 produced a larger amount of succinate and acetate, with marked differences in the level of these compounds between industrial and nonindustrial engineered strains. Acetoin and 2,3-butanediol formation was enhanced with significant variation between strains and in relation to the level of glycerol produced. Wine strains overproducing glycerol at moderate levels (12 to 18 g/liter) reduced acetoin almost completely to 2,3-butanediol. A lower biomass concentration was attained by GPD1-overexpressing strains, probably due to high acetaldehyde production during the growth phase. Despite the reduction in cell numbers, complete sugar exhaustion was achieved during fermentation in a sugar-rich medium. Surprisingly, the engineered wine yeast strains exhibited a significant increase in the fermentation rate in the stationary phase, which reduced the time of fermentation.  相似文献   

11.
Availability, low prices, and a high degree of reduction make glycerol an ideal feedstock to produce reduced chemicals and fuels via anaerobic fermentation. Although glycerol metabolism in Escherichia coli had been thought to be restricted to respiratory conditions, we report here the utilization of this carbon source in the absence of electron acceptors. Cells grew fermentatively on glycerol and exhibited exponential growth at a maximum specific growth rate of 0.040 ± 0.003 h−1. The fermentative nature of glycerol metabolism was demonstrated through studies in which cell growth and glycerol utilization were observed despite blocking several respiratory processes. The incorporation of glycerol in cellular biomass was also investigated via nuclear magnetic resonance analysis of cultures in which either 50% U-13C-labeled or 100% unlabeled glycerol was used. These studies demonstrated that about 20% of the carbon incorporated into the protein fraction of biomass originated from glycerol. The use of U-13C-labeled glycerol also allowed the unambiguous identification of ethanol and succinic, acetic, and formic acids as the products of glycerol fermentation. The synthesis of ethanol was identified as a metabolic determinant of glycerol fermentation; this pathway fulfills energy requirements by generating, in a redox-balanced manner, 1 mol of ATP per mol of glycerol converted to ethanol. A fermentation balance analysis revealed an excellent closure of both carbon (~95%) and redox (~96%) balances. On the other hand, cultivation conditions that prevent H2 accumulation were shown to be an environmental determinant of glycerol fermentation. The negative effect of H2 is related to its metabolic recycling, which in turn generates an unfavorable internal redox state. The implications of our findings for the production of reduced chemicals and fuels were illustrated by coproducing ethanol plus formic acid and ethanol plus hydrogen from glycerol at yields approaching their theoretical maximum.  相似文献   

12.
Cecropin-X发酵过程中工程菌质粒稳定性的研究   总被引:4,自引:0,他引:4  
通过连续斜面转接实验 (5 0次 )检测重组Cecropin X工程菌质粒的结构稳定性 ,采用 30L自动控制发酵罐观察不同培养基 ,有无选择压力和溶氧高低等培养条件对质粒分裂稳定性的影响。结果表明 ,该工程菌质粒具有结构稳定性和分裂不稳定性。当外源蛋白表达时 ,便会出现分裂不稳定性 ,表达量越高 ,质粒丢失情况越严重。经1 2h的培养 ,当采用TB和 2×LB作为发酵培养基时 ,带质粒的菌为 6 8 5 %和 98 0 % ,包涵体得率有很大差异 ,干重分别为 1 2 4和 0 4 0g L。发酵培养基中 (TB)氨苄青霉素浓度为 0和 1 0 0 μg mL时 ,带质粒的菌为 6 8 5 %和 92 0 % ,包涵体得率基本一致 ,干重分别为 1 2 4和 1 2 0g L。通过改变搅拌速度来调节溶氧量 ,当转速为 1 5 0和 2 5 0r min时 ,带质粒的菌为 6 8 5 %和 1 0 0 % ,包涵体得率有较大差异 ,干重分别为 1 2 4和 0 71g L。  相似文献   

13.
Whey, an abundant byproduct of the dairy industry, contains large amounts of protein and lactose which could be used for fuel ethanol production. We have investigated a new organism as a candidate for such fermentations: recombinant Escherichia coli containing the genes encoding the ethanol pathway from Zymomonas mobilis. The highest level of ethanol achieved, 68 g/L, was produced after 108 hours in Luria broth containing 140 g lactose/L. Fermentations of lower lactose concentrations were completed more rapidly with approximately 88% of theoretical yields. Reconstituted sweet whey (60 g lactose/L)was fermented more slowly than lactose in Luria broth requiring 144 hours to produce 26 g ethanol/L. Supplementing sweet whey with a trace metal mix and ammonium sulfate reduced the required fermentation time to 72 hours and increased final ethanol concentration (28 g ethanol/L). By adding proteinases during fermentation, the requirement for ammonia was completely eliminated, and the rate of fermentation further improved (30 g ethanol/L after 48 hours). This latter incresed in rate of ethanol production and ethanol yield are presumed to result from incorporation of amino acids released by hydrolysis of whey proteins. The fermentation of sweet whey by ethanologenic E. coil reduced the nonvolatile residue by approximately 70%. This should reduce biological oxygen demand and reduce the cost of waste treatment. Whey supplemented with trace metals and small amounts of proteinase may represent an economically attractive feedstock for the production of ethanol and other useful chemicals.  相似文献   

14.
Proton motive force (Δp) generation by Escherichia coli wild type cells during glycerol fermentation was first studied. Its two components, electrical—the membrane potential (?φ) and chemical—the pH transmembrane gradient (ΔpH), were established and the effects of external pH (pHex) were determined. Intracellular pH was 7.0 and 6.0 and lower than pHex at pH 7.5 and 6.5, respectively; and it was higher than pHex at pH 5.5. At high pHex, the increase of ?φ (?130 mV) was only partially compensated by a reversed ΔpH, resulting in a low Δp. At low pHex ?φ and consequently Δp were decreased. The generation of Δp during glycerol fermentation was compared with glucose fermentation, and the difference in Δp might be due to distinguished mechanisms for H+ transport through the membrane, especially to hydrogenase (Hyd) enzymes besides the F0F1-ATPase. H+ efflux was determined to depend on pHex; overall and N,N’-dicyclohexylcarbodiimide (DCCD)-inhibitory H+ efflux was maximal at pH 6.5. Moreover, ΔpH was changed at pH 6.5 and Δp was different at pH 6.5 and 5.5 with the hypF mutant lacking all Hyd enzymes. DCCD-inhibited ATPase activity of membrane vesicles was maximal at pH 7.5 and decreased with the hypF mutant. Thus, Δp generation by E. coli during glycerol fermentation is different than that during glucose fermentation. Δp is dependent on pHex, and a role of Hyd enzymes in its generation is suggested.  相似文献   

15.
Infection of stationary-phase Escherichia coli with MS2 bacteriophage results in the production, but not the release, of progeny virus by the host. Substantial protein synthesis in stationary-phase cells indicates that general protein synthesis is not sufficient to assure cell lysis or viral release.  相似文献   

16.
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   

17.
Whole cells of Escherichia coli B 10 having high tryptophan synthetase activity were used directly as an enzyme source to produce L-tryptophan from indole and L- or D,L-serine. This strain is tryptophan auxotrophic, which is tryptophanase negative and, in addition, L- and D-serine deaminase negative under production conditions. To avoid inhibition of tryptophan synthetase by a high concentration of indole, nonaqueous organic solvents, Amberlite XAD-2 adsorbent, and nonionic detergents were used as reservoirs of indole in the reaction mixture for the production of L-tryptophan. As a result, different effects were observed on the production of L-tryptophan. Particularly, among the nonionic detergents, Triton X-100 was very efficient. Using Triton X-100 for production of L-tryptophan from indole and L- or D,L-serine by whole cells of Escherichia coli B 10, 14.14 g/100 mL and 14.2 g/100 mL of L-tryptophan were produced at 37 degrees C for 60 h.  相似文献   

18.
Aerobic growth conditions significantly influenced anaerobic succinate production in two-stage fermentation by Escherichia coli AFP111 with knockouts in rpoS, pflAB, ldhA, and ptsG genes. At a low cell growth rate limited by glucose, enzymes involved in the reductive arm of the tricarboxylic acid cycle and the glyoxylate shunt showed elevated activities, providing AFP111 with intracellular redox balance and increased succinic acid yield and productivity.Succinic acid is valued as one of the key basic chemicals used in the preparation of biodegradable polymers or as raw material for chemicals of the C4 family (8, 19). The fermentative production of succinic acid from renewable resources is environmentally acceptable and sustainable (3). A breakthrough in genetically engineering Escherichia coli (6, 7, 11, 18) for succinate production was the isolation of strain AFP111 (1, 4), a mutant of NZN111 with a spontaneous ptsG mutation (pflAB ldhA double mutant). The process involves a two-stage fermentation, with aerobic cell growth followed by anaerobic conditions for succinate production (16, 21, 22). The aerobically induced enzymes can maintain their activity during the anaerobic phase and significantly affect succinate fermentation (22, 23). Using the best transition time based on the activities of the key enzymes and other physiological states, a two-stage fermentation using the recombinant AFP111 strain harboring pTrc99A-pyc achieved a final succinic acid concentration and productivity of 99.2 g·liter−1 and 1.3 g·liter−1·h−1, respectively (21).Aerobic cell growth is essential for the subsequent anaerobic fermentation. However, few studies have focused on the regulation of aerobic cell growth. As a regulation method, gluconeogenic carbon sources were used instead of glucose for the aerobic growth of Escherichia coli NZN111 and the activities of enzymes that are favorable for the anaerobic synthesis of succinate were enhanced (23, 24). Unfortunately, a gluconeogenic carbon source (e.g., sodium acetate) might increase the osmotic pressure of culture media, which would be detrimental to succinate production (23). As another regulation method, a glucose feeding strategy controlling the glucose concentration at about 0.5 g·liter−1 up to 1 g·liter−1 was reported to prevent excessive formation of acetic acid (16).In this study, we investigated different glucose feeding strategies for the aerobic growth phase of the two-phase process for succinate production by E. coli AFP111. Specifically, we compared several growth rates by using glucose limitation in addition to maximum growth under conditions of excess glucose.E. coli AFP111 [F+ λ rpoS396(Am) rph-1 ΔpflAB::Cam ldhA::Kan ptsG] (4, 16), which was a kind gift from D. P. Clark (Southern Illinois University), was the only strain used in this study. Luria-Bertani (LB) medium (60 ml) was used for inoculum culture in 1,000-ml flasks, and 3 liters of chemically defined medium (13, 14) was used for two-stage culture in a 7-liter fermentor. Two-stage fermentations were divided into three types, based on the glucose feeding strategy used during the aerobic stage. For type I culture, the glucose concentration was maintained at about 20 g·liter−1 during aerobic cell growth. Type II and III cultures comprised a batch process and subsequent glucose-limited fed-batch process (Fig. (Fig.1).1). The batch process initially contained 13 g/liter of glucose. The fed-batch process began when the dry cell weight (DCW) reached about 6 g/liter, with type II and type III cultures using a 600 g/liter glucose feed to achieve cell growth rates of 0.15 h−1 and 0.07 h−1, respectively (10). When the DCW reached 12 g·liter−1, the aerobically grown cells were directly transferred to anaerobic conditions (Fig. (Fig.1).1). For the anaerobic process, oxygen-free CO2 was sparged at 0.5 liter·min−1, the pH was controlled between 6.4 and 6.8 with intermittent supplementation of solid magnesium carbonate hydroxide, and the glucose concentration was maintained at about 20 g·liter−1 by supplying glucose in an 800-g·liter−1 solution.Open in a separate windowFIG. 1.Concentrations of glucose (circles), DCW (triangles), and succinic acid (squares) in the three types of two-stage fermentation by AFP111. μ, growth rate.The optical density at 600 nm was used to monitor cell growth, and this value was correlated to DCW. The concentration of glucose was assayed with an enzyme electrode analyzer, and organic acids were quantified by high-performance liquid chromatography (HPLC). The intracellular concentrations of NADH and NAD+ were assayed with a cycling method (12). The activities of isocitrate lyase (ICL) (20), pyruvate kinase (PYK) (17), phosphoenolpyruvate (PEP) carboxykinase (PCK) (20, 23), PEP carboxylase (PPC) (23), and malate dehydrogenase (MDH) (23) were measured spectrophotometrically at the end of the aerobic phase and 12 h after the onset of the anaerobic phase.All three types of fermentations were terminated when the succinate concentration increased less than 1 g·liter−1 in 5 h. Type III fermentation was terminated at a final succinic acid concentration of 101.2 g·liter−1 and an anaerobic-phase productivity of 1.89 g·liter−1·h−1 (Fig. (Fig.1).1). Trace amounts of by-products (such as acetate, ethanol, and pyruvate) accumulated and did not follow any trend in the anaerobic phase (data not shown).At the end of the aerobic culture phase, the specific enzyme activities of PCK, PYK, and ICL in type III culture were 2.9, 2.5, and 11.4 times higher, respectively, than the activities in type I culture (Table (Table1)1) . This phenomenon is consistent with published reports that suggest that the expression of enzymes involved in anaplerotic metabolism and the glyoxylate shunt (5, 15) is elevated in E. coli grown under glucose-limited conditions. These enzymes maintained their activities in the subsequent anaerobic phase (Table (Table1)1) and would be central to succinate production (22, 23). The elevated levels of PCK and PPC would provide the reductive branch of the tricarboxylic acid (TCA) cycle with oxaloacetate (OAA) at a higher rate (9), thereby supplying both malate and citrate (Table (Table11).

TABLE 1.

Activities of enzymes at the end of the aerobic culture phase and 12 h after the onset of the anaerobic phase
Fermentation typeaStagebMean sp act of enzyme ± SD (U/mg protein)c
PCKPPCMDHPYKICL
IAerobic0.82 ± 0.050.22 ± 0.0521.97 ± 0.151,175 ± 11.380.12 ± 0.00
Anaerobic0.55 ± 0.020.19 ± 0.0018.27 ± 1.05978 ± 12.330.09 ± 0.00
IIAerobic1.46 ± 0.100.23 ± 0.0425.69 ± 0.372,053 ± 3.650.73 ± 0.03
Anaerobic1.09 ± 0.010.20 ± 0.0135.55 ± 0.781,430 ± 13.780.41 ± 0.02
IIIAerobic2.38 ± 0.110.16 ± 0.0023.5 ± 0.132,955 ± 8.771.37 ± 0.00
Anaerobic1.75 ± 0.030.21 ± 0.0143.8 ± 0.622,501 ± 10.151.02 ± 0.01
Open in a separate windowaFermentation types were mentioned in culture conditions section.b“Aerobic” represents the data obtained at the end of aerobic culture; “Anaerobic” represents those obtained 12 h after transition to anaerobic fermentation.cThe standard deviations (SD) were calculated from triplicate samples of the same run.The reductive branch of the TCA cycle consumes 4 mol of electrons to form 2 mol of succinate based on 1 mol of glucose (1, 4). Therefore, the conversion of glucose to succinate through the reductive arm of the TCA cycle alone will lead to an intracellular imbalance of reducing equivalents (2, 18). Fortunately, the glyoxylate shunt (2, 18, 22) is available to provide 10 mol of electrons by converting 1 mol of glucose to 1 mol of succinate and 2 mol of CO2 (22). In the case of the ptsG mutant strain AFP111, when the molar flux at the PEP branch point flowing to OAA versus flowing to pyruvate reaches a ratio of 5:2, the intracellular redox balance is satisfied and the maximum theoretical mass yield of 1.12 g·g−1 succinic acid is achieved (22). Based on the elevated activities of PCK, PYK, and ICL (Table (Table1),1), both pathways leading to succinate were enhanced after glucose-limited growth. The succinic acid yields of 1.03 to 1.07 g·g−1 in the two glucose-limited processes approached the maximum theoretical yield for AFP111 (22), and these yields were about two times greater than the yield in the type I fermentation (Table (Table22).

TABLE 2.

Succinic acid production during anaerobic fermentation phasea
Fermentation typeMean ± SD
Succinic acid (g·liter−1)Yield (g·g−1)Productivity (g·liter−1·h−1)Specific productivity at 12 h (mg·g−1·h−1)NADH at 12 h mmol·(g DCW)−1NADH/NAD+ ratio at 12 h
I35.0 ± 0.740.43 ± 0.050.98 ± 0.04105 ± 150.88 ± 0.070.55 ± 0.08
II74.3 ± 3.241.03 ± 0.011.32 ± 0.05160 ± 81.95 ± 0.111.05 ± 0.10
III101.2 ± 1.041.07 ± 0.021.89 ± 0.07227 ± 111.97 ± 0.151.27 ± 0.13
Open in a separate windowaThe data were calculated only for the anaerobic stage. The standard deviations (SD) were calculated from two independent two-stage fermentations.In addition to differences in succinic acid yields, the glucose-limited and type I fermentations each resulted in significantly different specific succinic acid productivities (Table (Table2).2). A specific succinic acid productivity of 227 mg·g−1·h−1 was obtained at 12 h in type III fermentation. Because two pathways are needed for succinate production due to redox constraints, and enzyme activities in both pathways were elevated by glucose limitation, the results suggest that operating with glucose limitation provides the cells with greater metabolic flexibility to achieve a redox balance. Furthermore, the results suggest that one or more of these enzymes are limiting succinate formation under batch conditions (type I fermentation). Considering the NADH/NAD+ assays (Table (Table2),2), the results would support the hypothesis that succinate production was limited by insufficient NADH (2, 18).In summary, our study presented an efficient method of aerobic cell cultivation for two-stage succinate fermentation by engineered E. coli. Since the physiological state of aerobically grown cells was essential for their subsequent anaerobic succinate fermentation, some other environmental and physiology factors in the aerobic growth phase may also play an important role in improving succinate production.  相似文献   

19.
The Escherichia coli H serogroup is determined by flagellin, which has both H-type-specific and cross-reactive epitopes. The cross-reactive epitopes are responsible for the cross-reaction found in agglutination. To identify the specific epitope in H33 flagellin, the H33 flagellin gene was sequenced and the encoded central variable region (CVR) was determined. Four overlapping fragments of the CVR were prepared and their specificity was verified using different H-type antisera. Short fragments carrying potential H-type-specific determinants were selected, and monoclonal antibodies (MAbs) against these fragments were prepared. A murine MAb of subtype IgG1 showing specificity to H33 flagellin was produced. The epitope of the MAb was mapped to amino acid residues 250-260.  相似文献   

20.
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