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 46 毫秒
1.
Aims: To search for new bacteria for efficient production of polyhydroxyalkanoates (PHAs) from glycerol. Methods and Results: Samples were taken from different environments in Germany and Egypt, and bacteria capable of growing in mineral salts medium with glycerol as sole carbon source were enriched. From a wastewater sediment sample in Egypt, a Gram‐negative bacterium (strain MW1) was isolated that exhibited good growth and that accumulated considerable amounts of polyhydroxybutyrate (PHB) from glycerol and also from other carbon sources. The 16S rRNA gene sequence of this isolate exhibited 98·5% and 96·2% similarity to Zobellella denitrificans strain ZD1 and to Zobellella taiwanensis strain ZT1 respectively. The isolate was therefore affiliated as strain MW1 of Z. denitrificans. Strain MW1 grows optimally on glycerol at 41°C and pH 7·3 and accumulated PHB up to 80·4% (w/w) of cell dry weight. PHB accumulation was growth‐associated. Although it was not an absolute requirement, 20 g l?1 sodium chloride enhanced both growth (5 g cell dry weight per litre) and PHB content (87%, w/w). Zobellella denitrificans strain MW1 is also capable to accumulate the poly(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) copolymer if sodium propionate was used as cosubstrate in addition to glycerol. Conclusions: A new PHB‐accumulating strain was isolated and identified. This strain is able to utilize glycerol for growth and PHB accumulation to high content especially in the presence of NaCl that will enable the utilization of waste glycerol from biodiesel industry. Significance and Impact of the Study: This study is the first report on accumulation of PHA in a member of the new genus Zobellella. Furthermore, utilization of glycerol as the sole carbon source for fast growth and PHB biosynthesis, growth in the presence of NaCl and high PHB contents of the cells will make this newly isolated bacterium a potent candidate for industrial production of PHB from crude glycerol occurring as byproduct during biodiesel production.  相似文献   

2.
Glycerol is considered as an ideal feedstock for producing bioplastics via bacterial fermentation due to its ubiquity, low price, and high degree of reduction substrate. In this work, we study the yield and cause of limitation in poly(3‐hydroxybutyrate) (PHB) production from glycerol. Compared to glucose‐based PHB production, PHB produced by Cupriavidus necator grown on glycerol has a low productivity (0.92 g PHB/L/h) with a comparably low maximum specific growth rate of 0.11 h?1. We found that C. necator can synthesize glucose from glycerol and that the lithotrophical utilization of glycerol (non‐fermentative substrate) or gluconeogenesis is an essential metabolic pathway for biosynthesis of cellular components. Here, we show that gluconeogenesis affects the reduction of cell mass, the productivity of biopolymer product, and the molecular chain size of intracellular PHB synthesized from glycerol by C. necator. We use NMR spectroscopy to show that the isolated PHB is capped by glycerol. We then characterized the physical properties of the isolated glycerol‐based PHB with differential scanning calorimetry and tensile tests. We found that although the final molecular weight of the glycerol‐based PHB is lower than those of glucose‐based and commercial PHB, the thermal and mechanical properties of the biopolymers are similar. Biotechnol. Bioeng. 2012; 109: 2808–2818. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
Glycerol, a byproduct of the biodiesel industry, can be used by bacteria as an inexpensive carbon source for the production of value‐added biodegradable polyhydroxyalkanoates (PHAs). Burkholderia cepacia ATCC 17759 synthesized poly‐3‐hydroxybutyrate (PHB) from glycerol concentrations ranging from 3% to 9% (v/v). Increasing the glycerol concentration results in a gradual reduction of biomass, PHA yield, and molecular mass (Mn and Mw) of PHB. The molecular mass of PHB produced utilizing xylose as a carbon source is also decreased by the addition of glycerol as a secondary carbon source dependent on the time and concentration of the addition. 1H‐NMR revealed that molecular masses decreased due to the esterification of glycerol with PHB resulting in chain termination (end‐capping). However, melting temperature and glass transition temperature of the end‐capped polymers showed no significant difference when compared to the xylose‐based PHB. The fermentation was successfully scaled up to 200 L for PHB production and the yield of dry biomass and PHB were 23.6 g/L and 7.4 g/L, respectively. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

4.
In a limited-scale survey, 55 soil streptomycetes were screened for the accumulation of poly (3-hydroxybutyrate) [PHB]. Only 18% of the isolates accumulated PHB ranging between 1.9–7.8% of the dry biomass. The promising isolate DBCC-719, identified as Streptomyces griseorubiginosus, accumulated PHB amounting to 9.5% of the mycelial dry mass in the early stationary phase when grown in chemically defined medium with 2% (wt/vol) glucose as the sole source of carbon. Nitrogen-limiting conditions were inhibitory to growth and PHB accumulation. The isolated polymer was highly soluble in chloroform, gave a sharp peak at 235 nm on digestion with concentrated H2SO4, and had a characteristic infrared spectrum. Received: 26 March 1999 / Accepted: 3 May 1999  相似文献   

5.
The PHB production by Cupriavidus necator H16 depends on the type and concentration of stress factors and on the time of stress application. Hydrogen peroxide and ethanol significantly enhanced PHB accumulation in C. necator cells. Improved yields (10.9 g/L PHB) were observed after exposure of bacterial culture to 0.5 mmol/L H2O2 at the beginning of cultivation and to additional peroxide stress (5 mmol/L H2O2) after 60 h of cultivation (beginning of the stationary phase). Production was then ≈28 % higher than in control (8.50 g/L PHB). The highest yields (11.2 g/L PHB) were observed when ethanol (0.5 %) was applied at the beginning of stationary phase. An application of exogenous stress could thus be used as a simple strategy for a significant improvement of PHB production in C. necator.  相似文献   

6.
Salt‐affected farmlands are increasingly burdened by chlorides, carbonates, and sulfates of sodium, calcium, and magnesium. Intriguingly, the underlying physiological processes are studied almost always under NaCl stress. Two faba bean cultivars were subjected to low‐ and high‐salt treatments of NaCl, Na2SO4, and KCl. Assimilation rate and leaf water vapor conductance were reduced to approximately 25–30% without biomass reduction after 7 days salt stress, but this did not cause severe carbon shortage. The equimolar treatments of Na+, K+, and Cl? showed comparable accumulation patterns in leaves and roots, except for SO42? which did not accumulate. To gain a detailed understanding of the effects caused by the tested ion combinations, we performed nontargeted gas chromatography–mass spectrometry‐based metabolite profiling. Metabolic responses to various salts were in part highly linearly correlated, but only a few metabolite responses were common to all salts and in both cultivars. At high salt concentrations, only myo‐inositol, allantoin, and glycerophosphoglycerol were highly significantly increased in roots under all tested conditions. We discovered several metabolic responses that were preferentially associated with the presence of Na+, K+, or Cl?. For example, increases of leaf proline and decreases of leaf fumaric acid and malic acid were apparently associated with Cl? accumulation.  相似文献   

7.
From cell yields of Thiomicrospira denitrificans grown in the chemostat at different growth rates under anaerobic conditions a value of 1.4mm S2O inf3 sup= per g dry wt and per h could be calculated for maintenance energy requirements, and of 5.65 g dry wt per mole S2O inf3 sup= for the true growth yield.Cell yields of Thiomicrospira denitrificans appeared to be almost half of those of Thiobacillus denitrificans. Though in Thiobacillus denitrificans at D=0.03 h-1 under anaerobic conditions a value was found of 11.60 g dry wt per mole of thiosulphate used for energetic purposes, a value of 5.72 g dry wt per mole of thiosulphate was found under comparable conditions in Thiomicrospira denitrificans. Under aerobic conditions at D=0.03 h-1 values of 18.54 g dry wt per mole of thiosulphate were found in Thiobacillus denitrificans whereas Thiomicrospira denitrificans yielded only 9.38 g dry wt per mole of thiosulphate.As in Thiobacillus denitrificans anaerobic cell yields on sulphide were comparable to those on thiosulphate.Calculations have been made which indicate that the biosynthetic efficiency of Thiomicrospira denitrificans is lower than that of Thiobacillus denitrificans. This can only partly be explained by the absence of adenosine-phosphosulphate (APS) reductase.  相似文献   

8.
The paper reports a study involving the use of Halomonas boliviensis, a moderate halophile, for co-production of compatible solute ectoine and biopolyester poly(3-hydroxybutyrate) (PHB) in a process comprising two fed-batch cultures. Initial investigations on the growth of the organism in a medium with varying NaCl concentrations showed the highest level of intracellular accumulation of ectoine (0.74 g L−1) at 10–15% (w/v) NaCl, while at 15% (w/v) NaCl, the presence of hydroxyectoine (50 mg L−1) was also noted. On the other hand, the maximum cell dry weight and PHB concentration of 10 and 5.8 g L−1, respectively, were obtained at 5–7.5% (w/v) NaCl. A process comprising two fed-batch cultivations was developed—the first culture aimed at obtaining high cell mass and the second for achieving high yields of ectoine and PHB. In the first fed-batch culture, H. boliviensis was grown in a medium with 4.5% (w/v) NaCl and sufficient levels of monosodium glutamate, NH4+, and PO43−. In the second fed-batch culture, the NaCl concentration was increased to 7.5% (w/v) to trigger ectoine synthesis, while nitrogen and phosphorus sources were fed only during the first 3 h and then stopped to favor PHB accumulation. The process resulted in PHB yield of 68.5 wt.% of cell dry weight and volumetric productivity of about 1 g L−1 h−1 and ectoine concentration, content, and volumetric productivity of 4.3 g L−1, 7.2 wt.%, and 2.8 g L−1 day−1, respectively. At salt concentration of 12.5% (w/v) during the second cultivation, the ectoine content was increased to 17 wt.% and productivity to 3.4 g L−1 day−1.  相似文献   

9.
High poly(3-hydroxybutyrate) (PHB) content and volumetric productivity were achieved by fed-batch culture of Halomonas boliviensis using a defined medium. Initial shake flask cultivations in a minimal medium revealed that the growth of H. boliviensis was supported only when the medium was supplemented with aspartic acid, glycine, or glutamine. Addition of 0.1% (w/v) glutamine in the medium resulted in the highest cell dry weight (CDW; 3.9 g l−1). Glutamine was replaced by the less expensive monosodium glutamate (MSG) in the medium without any notable change in the final cell density. Effect of initial concentrations of NH4Cl and K2HPO4 on cell growth and PHB accumulation by H. boliviensis was then analyzed using a fed-batch fermentation system. The best conditions for PHB production by H. boliviensis were attained using 0.4% (w/v) NH4Cl and 0.22% (w/v) K2HPO4 and adding MSG intermittently to the fermentor. Poly(3-hydroxybutyrate) content and CDW reached 90 wt.% and 23 g l−1, respectively, after 18 h of cultivation. In order to increase CDW and PHB content, MSG, NH4Cl, and K2HPO4 were initially fed to the fermentor to maintain their concentrations at 2%, 0.4%, and 0.22% (w/v), respectively, and subsequently their feed was suppressed. This resulted in a CDW of 44 g l−1, PHB content of 81 wt.%, and PHB volumetric productivity of 1.1 g l−1 h−1.  相似文献   

10.
Methylobacterium rhodesianum and Ralstonia eutropha were cultivated to produce polyhydroxybutyrate (PHB) using media which contained glycerol and casein hydrolysates as C/N-substrates. In these media the pH had not to be regulated during the fermentations. The first strain accumulated an average of 39% PHB during 92 h of cultivation in flasks and 50% PHB during 45 h of cultivation in fermenters. The second one yielded an average of 47% PHB during 67 h of cultivation using casein peptone and 65% PHB during 45 h of cultivation using Casamino acids in the medium. Calculated N-balances showed that about 65% of the supplied nitrogen was used for growth of non-PHB cell dry mass. The conversion of glycerol to PHB was 17% (w/w).  相似文献   

11.
Industrial production of biodegradable polyesters such as polyhydroxyalkanoates is hampered by high production costs, among which the costs for substrates and for downstream processing represent the main obstacles. Inexpensive fermentable raw materials such as crude glycerol, an abundant by-product of the biodiesel industry, have emerged to be promising carbon sources for industrial fermentations. In this study, Zobellella denitrificans MW1, a recently isolated bacterium, was used for the production of poly(3-hydroxybutyrate) (PHB) from glycerol as the sole carbon source. Pilot-scale fermentations (42-liter scale) were conducted to scale up the high PHB accumulation capability of this strain. By fed-batch cultivation, at first a relatively high cell density (29.9 ± 1.3 g/liter) was obtained during only a short fermentation period (24 h). However, the PHB content was relatively low (31.0% ± 4.2% [wt/wt]). Afterwards, much higher concentrations of PHB (up to 54.3 ± 7.9 g/liter) and higher cell densities (up to 81.2 ± 2.5 g/liter) were obtained by further fed-batch optimization in the presence of 20 g/liter NaCl, with optimized feeding of glycerol and ammonia to support both cell growth and polymer accumulation over a period of 50 h. A high specific growth rate (0.422/h) and a short doubling time (1.64 h) were attained. The maximum PHB content obtained was 66.9% ± 7.6% of cell dry weight, and the maximum polymer productivity and substrate yield coefficient were 1.09 ± 0.16 g/liter/h and 0.25 ± 0.04 g PHB/g glycerol, respectively. Furthermore, a simple organic solvent extraction process was employed for PHB recovery during downstream processing: self-flotation of cell debris after extraction of PHB with chloroform allowed a convenient separation of a clear PHB-solvent solution from the cells. Maximum PHB recovery (85.0% ± 0.10% [wt/wt]) was reached after 72 h of extraction with chloroform at 30°C, with a polymer purity of 98.3% ± 1.3%.Polyhydroxybutyrate (PHB) is the best-studied example of biodegradable polyesters belonging to the group of polyhydroxyalkanoates (PHAs), which are synthesized by many bacteria and archaea as intracellular carbon and energy reserves (1, 40, 43). In the last decades, these biopolymers have received great attention due to their properties which resemble those of conventional petrochemical-based polymers (49). For instance, PHB is very similar to thermoplastic polypropylene (17). Their production from renewable resources and their complete biodegradability give PHAs promising advantages from an environmental point of view (6). In addition to their special physical traits, such as the elasticity of medium-chain-length PHAs and the high crystallization rate of PHB, PHAs are biocompatible, water resistant, oxygen impermeable, and enantiomerically pure; all of these characteristics broaden the scope of their applications in industry and medicine.So far, higher production costs than those of petrochemical plastics have hindered the successful commercialization of PHB (9). Many efforts have been devoted to reducing the production costs by developing superior microbial strains capable of utilizing cheap substrates and also by applying more efficient fermentation strategies and economical recovery processes (10).Fed-batch fermentation regimens are usually applied to achieve a high cell density, which is necessary for a high productivity and yield, in particular in cases of intracellular products, by frequent or continuous feeding of nutrients when growth proceeds (46). Several fed-batch fermentation processes have been reported for PHA production (21, 28). There are two prevalent cultivation modes for PHB production that are imposed on the microorganisms being used. The more frequently used mode is realized by a complex two-stage cultivation process. In this mode, all nutrients needed for growth to a high cell density are provided during the first phase of the process. In the second phase, imbalanced growth conditions are enforced by providing growth-limiting amounts of nutrients such as nitrogen, phosphate, or oxygen to trigger PHA biosynthesis and accumulation. The model organism for this mode is Ralstonia eutropha (formerly known as Alcaligenes eutrophus and recently reclassified as Cupriavidus necator) (26, 27). In the other cultivation mode, PHB is accumulated concurrently with growth, and therefore a single-stage process is applicable. A well-known example of this mode is PHB production by Alcaligenes latus (18, 47).Although several new downstream processes for the extraction of PHA have been reported as being economically effective, such as the application of surfactants and hypochlorite (9, 38), dispersions of hypochlorite solution and chloroform (14, 15), and the selective dissolution of cell mass by proteolytic enzymes (25) or by sulfuric acid and hypochlorite (48), solvent extraction methods are still regarded as an adequate way to gain intact polymers with a high purity and recovery yield (39). However, there is still a need to develop and improve these extraction methods further to make the entire process much simpler and cheaper (22).In addition to increased prices for crude oil, the abundance of inexpensive raw materials from agriculture and industry as cheap substrates for microbial fermentations, such as crude glycerol from the biodiesel industry, could make the production of PHA from renewable resources more competitive with common plastics (32). Due to increased glycerol production by the growing biodiesel industry, the prices for glycerol became low enough to make this residual compound a cheap carbon source for several industrial fermentation processes, especially for the production of microbial polyesters (11, 34). However, the various amounts of actual fermentable substrates and the presence of other nonfermentable components in feedstock, such as the various concentrations of glycerol and salts in biodiesel coproducts, hinder their use (42). Therefore, tolerant bioprocesses and/or strains tolerant to such variable factors are required.The production of PHA from glycerol has been investigated in only a few studies (4, 12, 24, 33, 42). In a recent study (32), crude glycerol from different biodiesel manufacturers was examined for suitability as a substrate for PHB production. However, significant decreases in PHB productivity and product yields were recorded when NaCl-contaminated crude glycerol was used.Recently, a newly isolated bacterium, Zobellella denitrificans MW1, was characterized as producing large amounts of PHB from glycerol, with enhanced growth and polymer productivity in the presence of NaCl (20). The present study aimed at developing a strategy to improve the volumetric production of PHB by Z. denitrificans MW1, using glycerol as the sole carbon source. For this purpose, several fed-batch cultivations were set up to steadily improve the nutrient supply to attain a high cell density and high PHB productivity. Moreover, the conventional organic solvent extraction method was modified with regard to an economically more feasible large-scale PHB extraction, achieving a high purity and recovery of the polymer.  相似文献   

12.
Alligator weed (Alternanthera philoxeroides) is a stoloniferous, amphibious and perennial herb which has invaded many parts of the world and led to serious environmental and ecological problems. In order to exploit cheap carbon source for poly(3-hydroxybutyrate) (PHB) production, alligator weed hydrolysates were prepared by acid and enzyme treatment and used for PHB production via Cupriavidus necator. The bacterium utilized alligator weed enzymatic hydrolysate and produced the PHB concentration of 3.8 ± 0.2 g/L at the conditions of pH 7.0, 27.5°C, 1.5 g/L of nitrogen source, and 25 g/L of carbon source, this exceeded the value of 2.1 ± 0.1 g/L from acid hydrolysate media at the same conditions. In order to obtain the optimum conditions of PHB production, response surface methodology was employed which improved PHB content. The optimum conditions for PHB production are as follows: carbon source, 34 g/L; nitrogen source, 2 g/L; pH, 7; temperature, 28°C. After 72 hr of incubation, the bacterium produced 8.5 g/L of dry cell weight and 4.8 g/L of PHB. The PHB was subjected to Fourier transform infrared (FTIR) spectroscopy, differential scanning calorimetry (DSC), and Molecular weight analysis and found the melting temperature, number average molecular mass, and polydispersity were 168.20°C, 185 kDa, and 2.1, respectively.  相似文献   

13.
The effects of the new de‐icer potassium carbonate on the growth and species composition of a mixed sward and on the pH and electrical conductivity of the soil were examined in a four‐year field trial and compared with sodium chloride. There were small positive effects of K2CO3 on the biomass yield of the sward at application rates up to 200 g m?2 yr?1, while effects were less negative than those of NaCl at annual application rates of 200 – 400 g m?2 yr?1. The species composition of the sward changed considerably with K2CO3 application, Agropyron repens becoming the dominant species. The pH was elevated slightly more by K2CO3 than by NaCl application. Electrical conductivity as an index of soil salinity was increased by both de‐icers. K2CO3 was more adsorbed to soil colloids than NaCl, so that the rise in E.C. resulting from application of the former was restricted to the uppermost 10 cm of soil.  相似文献   

14.
以披针叶黄华(Thermopsis lanceolata)试管苗为材料,通过组培方法研究其在0、0.2%、0.4%、0.6%、0.8%和1.0%NaCl和Na2SO4胁迫30d后的生长、有机渗透调节物质和无机渗透调节物质(Na+、K+和Ca2+)含量的变化,以探讨其耐盐性机制。结果显示:(1)随NaCl和Na2SO4胁迫浓度的增加,披针叶黄华试管苗叶片脯氨酸和可溶性糖含量均显著持续增加,且NaCl胁迫下脯氨酸上升的幅度均大于相同浓度Na2SO4胁迫下的增幅,而可溶性糖上升的幅度却小于相同浓度Na2SO4胁迫下的幅度;可溶性蛋白含量随NaCl浓度的增大呈先升高后降低的趋势,但随Na2SO4浓度的增加呈持续上升的趋势。(2)随NaCl和Na2SO4浓度的增加,披针叶黄华试管苗Na+含量呈增加趋势且各处理均显著高于对照,Ca2+含量和叶片K+含量却呈逐渐减少趋势且各处理均显著低于对照,而根系K+含量呈先降后升的趋势;Na2SO4胁迫下披针叶黄华试管苗叶片Na+含量上升幅度以及K+和Ca2+含量下降幅度均明显低于相同浓度NaCl胁迫组;而Na+/K+和Na+/Ca2+比值随NaCl和Na2SO4浓度增加而升高;NaCl胁迫下,叶片Na+/K+和Na+/Ca2+高于相同浓度Na2SO4胁迫下的比值,而根系Na+/K+和Na+/Ca2+却低于相同浓度Na2SO4胁迫下的比值。研究表明,盐胁迫下,披针叶黄华试管苗通过抑制叶片中Na+积累并增加可溶性糖和可溶性蛋白含量,在根系中维持较高K+和Ca2+含量以及较低水平Na+/K+和Na+/Ca2+比,以降低披针叶黄华细胞渗透势来适应盐渍环境;披针叶黄华对NaCl胁迫的调节能力弱于Na2SO4。  相似文献   

15.
16.
Summary A method to estimate the glucose concentration in the culture broth using CO2 evolution rate (CER) data from a mass spectrometer was developed.Alcaligenes eutrophus was cultivated to produce poly(3-hydroxybutyric acid) (PHB) from tapioca hydrolysate using this method. Thek value (g glucose/mol CO2), defined as the glucose consumption per CO2 evolution, decreased with culture time and was automatically changed using CER data. The glucose concentration in the culture broth could be controlled at 10 to 20 g/L. A final cell concentration of 106 g/L, PHB concentration of 61 g/L. and PHB content of 58 % of dry cell weight were obtained after 59 h of cultivation.  相似文献   

17.
SYNOPSIS. An osmotic growth requirement for the trypanosomatid Crithidia fasciculata became conspicuous at 32 C. Osmo-supportive compounds were surveyed at 28 vs. 33 C, in “low-osmotic” defined medium. Effective osmotic support was provided by many compounds, e.g., glycerophosphate, sorbitol, mannitol, glycerol, Na isethionate, glycine, arginine, KCl, NaCl, NH4Cl, and K2SO4. The nonspecificity of the requirement was thus evident, but inactivity of the presumably poorly adsorbable pentaerythritol indicated that osmotic pressure was a likely but insufficient condition for satisfying the temperature-enhanced growth requirement most clearly expressed as a need for osmotic support. Fortification of the medium with a combination of glycerophosphate, glycine, glycerol, and Tween 80 permitted good growth at 35 C. Possible relations between the temperature-enhanced osmotic requirement cell membrane damage, and morphological phases of Trypanosoma and Leishmania are discussed.  相似文献   

18.
Poly(3-hydroxybutyrate) (PHB) synthesis was analyzed under microaerobic conditions in a recombinant Escherichia coli arcA mutant using glycerol as the main carbon source. The effect of several additives was assessed in a semi-synthetic medium by the ‘one-factor-at-a-time’ technique. Casein amino acids (CAS) concentration was an important factor influencing both growth and PHB accumulation. Three factors exerting a statistically significant influence on PHB synthesis were selected by using a Plackett–Burman screening design [glycerol, CAS, and initial cell dry weight (CDW) concentrations] and then optimized through a Box–Wilson design. Under such optimized conditions (22.02 g l−1 glycerol, 1.78 g l−1 CAS, and 1.83 g l−1 inoculum) microaerobic batch cultures gave rise to 8.37 g l−1 CDW and 3.52 g l−1 PHB in 48 h (PHB content of 42%) in a benchtop bioreactor. Further improvements in microaerobic PHB accumulation were obtained in fed-batch cultures, in which glycerol was added to maintain its concentration above 5 g l−1. After 60 h, CDW and PHB concentration reached 21.17 and 10.81 g l−1, respectively, which results in a PHB content of 51%. Microaerobic fed-batch cultures allowed a 2.57-fold increase in volumetric productivity when compared with batch cultures. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users. An erratum to this article can be found at  相似文献   

19.
Bacillus mycoides strain RIJ B-017, a growth-associated poly-3-hydroxybutyrate (PHB) producer was grown on sucrose-containing media. PHB accumulated in cells up to 72% of dry cell mass. The overall maximum value of PHB yield (Y p/s) and productivities (Q p andq p) 250 mgp/gs, 120 mgp L−1 h−1 and 30 mgp gx −1 h−1, respectively, were obtained at 15 g/L sucrose. Differential scanning calorimeter heating curve showed two peaks, one at 95.9 °C and another at 165.4°C with a shoulder around 154.6 °C. The viscosity-average molar mass in chloroform at 27°C was 505 kDa. The carbon content of PHB was 55.4% of the mass.  相似文献   

20.
Aim: To develop an approach to enhance polyhydroxybutyrate (PHB) production via the coexpressed phaCAB and vgb genes controlled by arabinose PBAD promoter in Escherichia coli. Method and Results: The polyhydroxyalkanoates (PHAs) synthesis operon, (phaCAB), from Ralstonia eutropha was overexpressed under the regulation of the arabinose PBAD promoter in Escherichia coli, and the vgb gene encoding bacterial haemoglobin from Vitreoscilla stercoraria (VHb) was further cloned at downstream of phaCAB to form an artificial operon. The cell dry weight (CDW), PHB content and PHB concentration were enhanced around 1·23‐, 1·57‐, and 1·93‐fold in the engineered cell harbouring phaCAB–vgb (SY‐2) upon 1% arabinose induction compared with noninduction (0% arabinose). Furthermore, by using a recombinant strain harbouring PBAD promoter‐vgb along with native promoter‐phaCAB construction, the effect of vgb expression level on PHB biosynthesis was positive correlation. Conclusions: The results exploit the possibility to improve the PHB production by fusing the genes phaCAB–vgb from different species under the arabinose regulation system in E. coli. It also demonstrates that increase in VHb level enhances the PHB production. Significance and Impact of the Study: We were successful in providing a new coexpressed system for PHB synthesis in E. coli. This coexpressed system could be regulated by arabinose inducer, and is more stable and cheaper than other induced systems (e.g. IPTG). Furthermore, it could be applied in many biotechnology or fermentation processes.  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