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.
Summary When continuous, steady-state, glucose-limited cultures ofClostridium acetobutylicum were sparged with CO, the completely or almost completely acidogenic fermentations became solventogenic. Alcohol (butanol and ethanol) and lactate production at very high specific production rates were initiated and sustained without acetone, and little or no acetate and butyrate formation. In one fermentation, strong butyrate uptake without acetone formation was observed. Growth could be sustained even with 100% inhibition of H2 formation. Although CO gasing inhibited growth up to 50%, and H2 formation up to 100%, it enhanced the rate of glucose uptake up to 300%. TheY ATP was strongly affected and mostly reduced with respect to its steady-state value. The results support the hypothesis that solvent formation is triggered by an altered electron flow.  相似文献   

2.
Extracts prepared from non-solvent-producing cells of Clostridium acetobutylicum contained methyl viologen-linked hydrogenase activity (20 U/mg of protein at 37°C) but did not display carbon monoxide dehydrogenase activity. CO addition readily inhibited the hydrogenase activity of cell extracts or of viable metabolizing cells. Increasing the partial pressure of CO (2 to 10%) in unshaken anaerobic culture tube headspaces significantly inhibited (90% inhibition at 10% CO) both growth and hydrogen production by C. acetobutylicum. Growth was not sensitive to low partial pressures of CO (i.e., up to 15%) in pH-controlled fermentors (pH 4.5) that were continuously gassed and mixed. CO addition dramatically altered the glucose fermentation balance of C. acetobutylicum by diverting carbon and electrons away from H2, CO2, acetate, and butyrate production and towards production of ethanol and butanol. The butanol concentration was increased from 65 to 106 mM and the butanol productivity (i.e., the ratio of butanol produced/total acids and solvents produced) was increased by 31% when glucose fermentations maintained at pH 4.5 were continuously gassed with 85% N2-15% CO versus N2 alone. The results are discussed in terms of metabolic regulation of C. acetobutylicum saccharide fermentations to achieve maximal butanol or solvent yield.  相似文献   

3.
Jong Jin Lim 《Biopolymers》1976,15(12):2371-2383
The transition temperatures tt and enthalpy changes ΔH in the helix–coil transition of solid tendon collagen soaked in a solution containing one of the following stabilizing or destabilizing agents, HCHO, NaF, NaCl, NaI, NaBr, NaOH, NH2CONH2, CaCl2, MgCl2, were measured as a function of molar concentration by a calorimetric method. The temperature and the enthalpy changes accompanying the transition behaved in a similar manner: when the tt was depressed by the presence of ions, similar behaviour was observed in ΔH. Both parameters (tt and ΔH) increased for HCHO, and decreased for NaF and NaCl at concentrations lower than 0.2 M. Above 0.2 M they increased for NaF and NaCl, and decreased in the presence of the other reagents listed above. The average tt and the ΔH observed in collagen soaked in water were 63.5°C and 12.3 cal/g, respectively. In addition to the parameters mentioned above, the molar effectiveness of the various reagents was obtained for the cases where there was a linear relationship between the tt and molar concentration of the reagent in the solution. Since both the tt and the ΔH were observed to vary, the entropy change (ΔS) accompanying the transition was calculated using thermodynamic relations. In order to explain the ΔS observed as a function of ionic concentration, the thermodynamic relationships have been obtained from a partition function under suitable assumptions. Since the partition function is dependent on the number of hydrogen bonds responsible for collagen stability, the result obtained has been compared with the values predicted by the two most quoted models for collagen. The present study is in accordance with the Ramachandran model for collagen structure, which predicts more than one hydrogen bond per three residues.  相似文献   

4.
Summary The effect of product gases, H2 and CO2, on solvent production was studied using a continuous culture of alginate-immobilized Clostridium acetobutylicum. Initially, in order to find the optimum dilution rate for aceton--butanol production in this system, fermentations were carried out at various dilution rates. With 10% H2 and 10% CO2 in the sparging gas, a dilution rate of 0.07 h–1 was found to maximize volumetric productivity (0.58 g·l–1·h–1), while the maximum specific productivity of 0.27 g·h–1 occured at 0.12 h–1. Continuous cultures with vigorous sparging of N2 produced only acids. It was concluded that in the case of continuous fermentation H2 is essential for good solvent production, although good solvent production is possible in an H2-absent environment in the case of batch fermentations. When the fermentation was carried out at atmospheric pressure under H2-enriched conditions, the presence of CO2 in the sparging gas did not slow down glucose metabolism; rather it changed the direction of the phosphoroclastic reaction and as a result increased the butanol/acetone ratio.  相似文献   

5.
Acetogens share the capacity to convert H2 and CO2 into acetate for energy conservation (ATP synthesis). This reaction is attractive for applications, such as gas fermentation and microbial electrosynthesis. Different H2 partial pressures prevail in these distinctive applications (low concentrations during microbial electrosynthesis [<40 Pa] vs. high concentrations with gas fermentation [>9%]). Strain selection thus requires understanding of how different acetogens perform under different H2 partial pressures. Here, we determined the H2 threshold (H2 partial pressure at which acetogenesis halts) for eight different acetogenic strains under comparable conditions. We found a three orders of magnitude difference between the lowest and highest H2 threshold (6 ± 2 Pa for Sporomusa ovata vs. 1990 ± 67 Pa for Clostridium autoethanogenum), while Acetobacterium strains had intermediate H2 thresholds. We used these H2 thresholds to estimate ATP gains, which ranged from 0.16 to 1.01 mol ATP per mol acetate (S. ovata vs. C. autoethanogenum). The experimental H2 thresholds thus suggest strong differences in the bioenergetics of acetogenic strains and possibly also in their growth yields and kinetics. We conclude that no acetogen is equal and that a good understanding of their differences is essential to select the most optimal strain for different biotechnological applications.  相似文献   

6.
Summary H2 is a central metabolite in the process of methane digestion. In this study, the partial pressure of H2 was decreased by sparging the gas phase of the digester through an auxiliary reactor in which a Rhodomicrobium vaniellii culture or a mixed culture of sulfate-reducing bacteria was allowed to develop at the expense of H2 and CO2 present in the biogas. The decrease of the H2 concentration in the gas phase was significant. A 18–23 percent increase of the gas production rate and a concomitantly improved removal of volatile fatty acids from the mixed liquor was obtained. The sulfate-reducing bacteria appeared to be slightly more effective than the phototrophs. The results suggest that the increased biogas production rate is due to the decrease of propionic acid formation and the concomitant stimulation of propionate degradation.Abbreviations CODt Total chemical oxygen demand - CODs Soluble chemical oxygen demand - SS Suspended solids - DM Dry matter - VFA Volatile fatty acids  相似文献   

7.
Processes for the biotechnological production of kerosene and diesel blendstocks are often economically unattractive due to low yields and product titers. Recently, Clostridium acetobutylicum fermentation products acetone, butanol, and ethanol (ABE) were shown to serve as precursors for catalytic upgrading to higher chain-length molecules that can be used as fuel substitutes. To produce suitable kerosene and diesel blendstocks, the butanol:acetone ratio of fermentation products needs to be increased to 2–2.5:1, while ethanol production is minimized. Here we show that the overexpression of selected proteins changes the ratio of ABE products relative to the wild type ATCC 824 strain. Overexpression of the native alcohol/aldehyde dehydrogenase (AAD) has been reported to primarily increase ethanol formation in C. acetobutylicum. We found that overexpression of the AADD485G variant increased ethanol titers by 294%. Catalytic upgrading of the 824(aadD485G) ABE products resulted in a blend with nearly 50 wt%≤C9 products, which are unsuitable for diesel. To selectively increase butanol production, C. beijerinckii aldehyde dehydrogenase and C. ljungdhalii butanol dehydrogenase were co-expressed (strain designate 824(Cb ald-Cl bdh)), which increased butanol titers by 27% to 16.9 g L−1 while acetone and ethanol titers remained essentially unaffected. The solvent ratio from 824(Cb ald-Cl bdh) resulted in more than 80 wt% of catalysis products having a carbon chain length≥C11 which amounts to 9.8 g L−1 of products suitable as kerosene or diesel blendstock based on fermentation volume. To further increase solvent production, we investigated expression of both native and heterologous chaperones in C. acetobutylicum. Expression of a heat shock protein (HSP33) from Bacillus psychrosaccharolyticus increased the total solvent titer by 22%. Co-expression of HSP33 and aldehyde/butanol dehydrogenases further increased ABE formation as well as acetone and butanol yields. HSP33 was identified as the first heterologous chaperone that significantly increases solvent titers above wild type C. acetobutylicum levels, which can be combined with metabolic engineering to further increase solvent production.  相似文献   

8.
Corncob is a potential feedstock in Thailand that can be used for fermentable sugar production through dilute sulfuric acid pretreatment and enzymatic hydrolysis. To recover high amounts of monomeric sugars from corncob, the sulfuric pretreatment conditions were optimized by using response surface methodology with three independent variables: sulfuric acid concentration, temperature, and time. The highest response of total sugars, 48.84 g/L, was found at 122.78°C, 4.65 min, and 2.82% (v/v) H2SO4. With these conditions, total sugars from the confirmation experiment were 46.29 g/L, with 5.51% error from the predicted value. The hydrolysate was used as a substrate for acetone–butanol–ethanol fermentation to evaluate its potential for microbial growth. The simultaneous saccharification and fermentation (SSF) showed that C. beijerinckii TISTR 1461 can generate acetone–butanol–ethanol products at 11.64 g/L (5.29 g/L acetone, 6.26 g/L butanol, and 0.09 g/L ethanol) instantly using sugars from the hydrolysed corncob with Novozymes 50013 cellulase enzyme without an overliming process.  相似文献   

9.
When the partial pressure of H2 was decreased by lowering the total pressure in the headspace of the reactor in a batch fermentation process from 760 mm Hg to 380 mm Hg containing Enterobacter cloacae, the molar yield of H2 increased from 1.9 mol to 3.9 mol H2/mol glucose. The maximum production rate was 0.017 mmol H2/h l at 380 mm Hg. The lag period as well as total batch time of H2 production decreased using a decreased partial pressure.  相似文献   

10.
Summary Carbon distribution from substrates to products in Clostridium acetobutylicum ATCC 824 was investigated by adding 14C-labeled substrates as tracers. Comparison of carbon conversion between chloramphenicol (CAP)-treated and untreated cultures was also studied. The percentage of 14C recovery in butanol, acetone and ethanol from uniformly labeled [14C]glucose was increased by 17, 25 and 30%, respectively, after CAP addition. The incorporation of 14C in solvents from 14C-labeled acetate and butyrate was also increased by the antibiotic treatment. A total 14C recovery of 12% in all the products from added [14C]Na2CO3 indicates significant heterotrophic CO2 fixation in this microorganism. The ratio of carbon in butanol derived from glucose, acetate and butyrate was about 71:6:18, and this ratio was unchanged by CAP treatment.This paper represents contribution No. 2685 of the Rhode Island Agricultural Experimental StationCorrespondence to: R. W. Traxler  相似文献   

11.
Summary The metabolite pattern of batch cultures ofLactobacillus casei LMG 6400,Clostridium butyricum LMG 1213t1 andEscherichia coli LMG 2093 was effected only for the latter organism when the H2 partial pressure was below 1 atmosphere: high hydrogen partial pressures increased the formate formation, low pressures gave rise to increased acetate production and higher cell yields.  相似文献   

12.
A possible way to improve the economic efficacy of acetone–butanol–ethanol fermentation is to increase the butanol ratio by eliminating the production of other by-products, such as acetone. The acetoacetate decarboxylase gene (adc) in the hyperbutanol-producing industrial strain Clostridium acetobutylicum EA 2018 was disrupted using TargeTron technology. The butanol ratio increased from 70% to 80.05%, with acetone production reduced to approximately 0.21 g/L in the adc-disrupted mutant (2018adc). pH control was a critical factor in the improvement of cell growth and solvent production in strain 2018adc. The regulation of electron flow by the addition of methyl viologen altered the carbon flux from acetic acid production to butanol production in strain 2018adc, which resulted in an increased butanol ratio of 82% and a corresponding improvement in the overall yield of butanol from 57% to 70.8%. This study presents a general method of blocking acetone production by Clostridium and demonstrates the industrial potential of strain 2018adc.  相似文献   

13.
Summary When Clostridium acetobutylicum was grown in continuous culture under phosphate limitation (0.74 mM) at a pH of 4.3, glucose was fermented to butanol, acetone and ethanol as the major products. At a dilution rate of D=0.025 h–1 and a glucose concentration of 300 mM, the maximal butanol and acetone concentrations were 130 mM and 74 mM, respectively. 20% of the glucose remained in the medium. On the basis of these results a two-stage continuous process was developed in which 87.5% of the glucose was converted into butanol, acetone and ethanol. The cells and minor amounts of acetate and butyrate accounted for the remaining 12.5% of the substrate. The first stage was run at D=0.125 h–1 and 37° C and the second stage at D=0.04 h–1 and 33° C. High yields of butanol and acetone were also obtained in batch culture under phosphate limitation.  相似文献   

14.
A mass spectrometry (MS) membrane sensor was developed and applied to on-line product measurement in acetone-butanol fermentation. The sensor facilitated the monitoring of acetone, butanol, ethanol, H2 and CO2, and single-compound calibration curves for both acetone and butanol showed a linear relationship between the product concentration and the MS response. However, when an actual fermentation was monitored, the product concentration calculated from the MS response was smaller than the concentration determined by gas chromatography, and the relationship between the response and the product concentration was nonlinear. It was found that large amounts of gases (H2, CO2) entering the MS analyzation chamber were causing a ‘space charge effect’, which resulted in an MS response ceiling. The problem could be resolved by reducing the surface area of the sensor membrane. Under some fermentation conditions, a by-product, n-butyl butyrate, was produced, and this interfered with the measurement of butanol due to a peak overlapping effect. However, it was found that this could be compensated for by using an empirical equation. Application of the MS membrane sensor in a fed batch culture of acetone-butanol fermentation resulted in successful control of the butanol concentration.  相似文献   

15.
The objective of the present study was to evaluate a comprehensive set of urinary biomarkers for oxidative damage to lipids, proteins and DNA, in man. Eighteen moderately trained males (mean age 24.6±0.7) exercised 60?min at 70% of maximal O2 uptake on a cycle ergometer. Urine fractions for 12?h were collected 1 day before, and for 3 consecutive days after exercise.

As biomarkers of lipid peroxidation, 8 aldehydes (i.e. propanal, butanal, pentanal, hexanal, heptanal, octanal, nonanal and malondialdehyde—MDA)and acetone were analyzed in urines by gas chromatography with electron capture detection (GC-ECD). As a biomarker of protein oxidation, o,o′-dityrosine was analyzed in urine samples by a recently developed isotope dilution HPLC-atmospheric pressure chemical ionization (APCI)-tandem-mass spectrometry (HPLC-APCI-MS/MS) methodology. As a biomarker of oxidative DNA damage, urinary excretion of 8-hydroxy-2′-deoxyguanosine (8-OHdG) was measured by an ELISA method.

On the day of exercise, significant increases were observed in urinary excretions of acetone (?p<0.025, n=18) and butanal (?p<0.01, n=18) in the 12?h daytime fractions compared to the daytime fraction before exercise. The urinary acetone excretion was also significantly (?p<0.05) increased on the 1st day after exercise. Octanal and nonanal were increased in the daytime urine fraction on the 2nd day after exercise. However, these increases were of borderline significance (?p=0.09 and p=0.07, respectively).

Significantly elevated urinary o,o′-dityrosine amounts were observed in the daytime fraction on the day of exercise (?p<0.025) and on the 1st day after exercise (?p=0.07) compared to the before exercise daytime fraction.

Excretion of urinary 8-OHdG was statistically significantly increased in the daytime fractions on the day of exercise (?p=0.07) and on the 1st day after exercise (?p<0.025) compared to before exercise daytime fraction.

Increases in urinary excretions of acetone, propanal, pentanal, MDA and 8-OHdG significantly correlated with training status (hours of exercise/week) of the volunteers, while o,o′-dityrosine did not.

To our knowledge, the present study is the first to evaluate a multi-parameter non-invasive biomarker set for damage to three main cellular targets of ROS. It shows that 1?h of exercise may already induce oxidative damage in moderately trained individuals and that the chosen urinary biomarkers are sensitive enough to monitor such damage.  相似文献   

16.
Li X  Wang Y  Zhang S  Chu J  Zhang M  Huang M  Zhuang Y 《Bioresource technology》2011,102(2):1142-1148
The effects of light/dark cycle, mixing pattern and partial pressure of H2 on the growth and hydrogen production of Rhodobacter sphaeroides ZX-5 were investigated. The results from light/dark cycle culture showed that little or no hydrogen production was observed during the dark periods, and the hydrogen production immediately recovered once illumination was resumed. Also, it was found that the optimum condition of shaking velocity was 120 rpm for hydrogen photo-fermentation. Meanwhile, shaking during H2 production phase (i.e., cell growth stationary phase) of photo-fermentation played a crucial role on effectively enhancing the phototrophic hydrogen production, rather than that during cell exponential growth phase. The other factor evaluated was hydrogen partial pressure in the culture system. The substrate conversion efficiency increased from 86.07% to 95.56% along with the decrease of the total pressure in the photobioreactor from 1.082 × 105 to 0.944 × 105 Pa, which indicated that reduction of H2 partial pressure by lowering the operating pressure substantially improved H2 production in an anaerobic, photo-fermentation process.  相似文献   

17.
Heavy water (H218O) has been used to label DNA of soil microorganisms in stable isotope probing experiments, yet no measurements have been reported for the 18O content of DNA from soil incubated with heavy water. Here we present the first measurements of atom% 18O for DNA extracted from soil incubated with the addition of H218O. Four experiments were conducted to test how the atom% 18O of DNA, extracted from Ponderosa Pine forest soil incubated with heavy water, was affected by the following variables: (1) time, (2) nutrients, (3) soil moisture, and (4) atom% 18O of added H2O. In the time series experiment, the atom% 18O of DNA increased linearly (R 2 = 0.994, p < 0.01) over the first 72 h of incubation. In the nutrient addition experiment, there was a positive correlation (R 2 = 0.991, p = 0.006) between the log10 of the amount of tryptic soy broth, a complex nutrient broth, added to soil and the log10 of the atom% 18O of DNA. For the experiment where soil moisture was manipulated, the atom% 18O of DNA increased with higher soil moisture until soil moisture reached 30%, above which 18O enrichment of DNA declined as soils became more saturated. When the atom% 18O for H2O added was varied, there was a positive linear relationship between the atom% 18O of the added water and the atom% 18O of the DNA. Results indicate that quantification of 18O incorporated into DNA from H218O has potential to be used as a proxy for microbial growth in soil.  相似文献   

18.
NAD(P)H:H2 pathways are theoretically predicted to reach equilibrium at very low partial headspace H2 pressure. An evaluation of the directionality of such near‐equilibrium pathways in vivo, using a defined experimental system, is therefore important in order to determine its potential for application. Many anaerobic microorganisms have evolved NAD(P)H:H2 pathways; however, they are either not genetically tractable, and/or contain multiple H2 synthesis/consumption pathways linked with other more thermodynamically favourable substrates, such as pyruvate. We therefore constructed a synthetic ferredoxin‐dependent NAD(P)H:H2 pathway model system in Escherichia coli BL21(DE3) and experimentally evaluated the thermodynamic limitations of nucleotide pyridine‐dependent H2 synthesis under closed batch conditions. NADPH‐dependent H2 accumulation was observed with a maximum partial H2 pressure equivalent to a biochemically effective intracellular NADPH/NADP+ ratio of 13:1. The molar yield of the NADPH:H2 pathway was restricted by thermodynamic limitations as it was strongly dependent on the headspace : liquid ratio of the culture vessels. When the substrate specificity was extended to NADH, only the reverse pathway directionality, H2 consumption, was observed above a partial H2 pressure of 40 Pa. Substitution of NADH with NADPH or other intermediates, as the main electron acceptor/donor of glucose catabolism and precursor of H2, is more likely to be applicable for H2 production.  相似文献   

19.
Cymbidium goeringii (Orchidaceae). Mean observed population heterozygosity (H op=0.181), expected heterozygosity (H eP=0.240), and total genetic diversity (H T =0.351) were all higher than average values for species with similar life-history characteristics. A considerable deficit of heterozygotes relative to Hardy–Weinberg expectations was detected (77% of fixation indices were positive) with a mean F IS of 0.278. On average, 90% of the total genetic diversity was found within populations (mean G ST = 0.098). An indirect estimate of the number of migrants per generation (Nm=2.30, calculated from G ST , Nm=8.48, calculated from the frequencies of unique alleles) indicated that gene flow has been extensive in C. goeringii. Results of a spatial autocorrelation analysis based on allele frequencies of 16 populations revealed a trend with respect to the distance classes (0<63 km, six significant positive values; beyond that distance, 12 significant negative values). It is highly probable that C. goeringii has a history of relative large, continuous populations that had greater chance for gene movement among adjacent populations via large numbers of small seeds, following the last Ice Age. Factors contributing to the high levels of genetic diversity found within populations of C. goeringii include its large and continuous populations, its long-lived perennial habit, its widespread geographical distribution, and its ability for relatively long distance seed movement by wind. Received 18 June 1998/ Accepted in revised form 18 November 1998  相似文献   

20.
Summary An overflow filtration unit for cell recycle with Clostridium acetobutylicum was developed. A cellulose-triacetate ultrafiltration membrane with a cut-off volume of 20 000 MW was found to work best. C. acetobutylicum was grown in continuous culture under phosphate limitation (0.74 mM) at a pH value of 4.4 with cell recycle, the cell dry weight in the culture vessel reached 13.1 g/l at a dilution rate of D=0.10 h-1 and 37°C. 377 mM of glucose were fermented to 190 mM butanol, 116.2 mM acetone and 25.8 mM ethanol. Total acids were 47.6 mM. The butanol productivity was 1.41 g/l/h. At a dilution rate of 0.40 h-1 the butanol productivity was increased to 4.1 g/l/h but glucose consumption was decreased to 285 mM and butanol, acetone and ethanol production to 138.2, 97.5, 16.5 mM, respectively.  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