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.
The growth characteristics ofCandida blankii as a function of temperature and pH in a simulated bagasse hemicellulose hydrolysate were determined in chemostat culture. The highest maximum specific growth rate of 0.44h–1 was reached at 38°C and at pH 5.5, with a sharp decrease in growth rate on either side of this temperature. Growth occurred at 46°C but not at 48°C. The protein and cell yields varied little below 40°C and the respective values were 0.22 and 0.5 g/g at 38°C. At the lower pH values, a severe linear decrease in cell and protein yields occurred, whereas a small increase in these yields at decreasing pH values was found when acetic acid was omitted from the medium. In the presence of acetic acid, a very sharp decrease in the growth rate at pH values below pH 4.5 was noted, despite the very low residual acetic acid concentrations, of less than 50 mg/l, in the culture.  相似文献   

2.
3.
Measurement of the flux through the citrate fermentation pathway in resting cells of Lactococcus lactis CRL264 grown in a pH-controlled fermentor at different pH values showed that the pathway was constitutively expressed, but its activity was significantly enhanced at low pH. The flux through the citrate-degrading pathway correlated with the magnitude of the membrane potential and pH gradient that were generated when citrate was added to the cells. The citrate degradation rate and proton motive force were significantly higher when glucose was metabolized at the same time, a phenomenon that could be mimicked by the addition of lactate, the end product of glucose metabolism. The results clearly demonstrate that citrate metabolism in L. lactis is a secondary proton motive force-generating pathway. Although the proton motive force generated by citrate in cells grown at low pH was of the same magnitude as that generated by glucose fermentation, citrate metabolism did not affect the growth rate of L. lactis in rich media. However, inhibition of growth by lactate was relieved when citrate also was present in the growth medium. Citrate did not relieve the inhibition by other weak acids, suggesting a specific role of the citrate transporter CitP in the relief of inhibition. The mechanism of citrate metabolism presented here provides an explanation for the resistance to lactate toxicity. It is suggested that the citrate metabolic pathway is induced under the acidic conditions of the late exponential growth phase to make the cells (more) resistant to the inhibitory effects of the fermentation product, lactate, that accumulates under these conditions.  相似文献   

4.
A potential-sensitive fluorescent probe, 3,3-dipropylthiadicarbocyanine iodide, was used to analyze, at pH 7.5 and 10.5, the effects of Bacillus thuringiensis toxins on the membrane potential generated by the efflux of K+ ions from brush border membrane vesicles purified from the midgut of the tobacco hornworm, Manduca sexta. Fluorescence levels were strongly influenced by the pH and ionic strength of the media. Therefore, characterization of the effects of the toxins was conducted at constant pH and ionic strength. Under these conditions, the toxins had little effect on the fluorescence levels measured in the presence or absence of ionic gradients, indicating that the ionic selectivity of their pores is similar to that of the intact membrane. Valinomycin greatly increased the potential generated by the diffusion of K+ ions although membrane permeability to the other ions used to maintain the ionic strength constant also influenced fluorescence levels. In the presence of valinomycin, active toxins (Cry1Aa, Cry1Ab, Cry1Ac, Cry1C and Cry1E) efficiently depolarized the membrane at pH 7.5 and 10.5.  相似文献   

5.
The mechanism of metabolic energy production by malolactic fermentation in Lactococcus lactis has been investigated. In the presence of L-malate, a proton motive force composed of a membrane potential and pH gradient is generated which has about the same magnitude as the proton motive force generated by the metabolism of a glycolytic substrate. Malolactic fermentation results in the synthesis of ATP which is inhibited by the ionophore nigericin and the F0F1-ATPase inhibitor N,N-dicyclohexylcarbodiimide. Since substrate-level phosphorylation does not occur during malolactic fermentation, the generation of metabolic energy must originate from the uptake of L-malate and/or excretion of L-lactate. The initiation of malolactic fermentation is stimulated by the presence of L-lactate intracellularly, suggesting that L-malate is exchanged for L-lactate. Direct evidence for heterologous L-malate/L-lactate (and homologous L-malate/L-malate) antiport has been obtained with membrane vesicles of an L. lactis mutant deficient in malolactic enzyme. In membrane vesicles fused with liposomes, L-malate efflux and L-malate/L-lactate antiport are stimulated by a membrane potential (inside negative), indicating that net negative charge is moved to the outside in the efflux and antiport reaction. In membrane vesicles fused with liposomes in which cytochrome c oxidase was incorporated as a proton motive force-generating mechanism, transport of L-malate can be driven by a pH gradient alone, i.e., in the absence of L-lactate as countersubstrate. A membrane potential (inside negative) inhibits uptake of L-malate, indicating that L-malate is transported an an electronegative monoanionic species (or dianionic species together with a proton). The experiments described suggest that the generation of metabolic energy during malolactic fermentation arises from electrogenic malate/lactate antiport and electrogenic malate uptake (in combination with outward diffusion of lactic acid), together with proton consumption as result of decarboxylation of L-malate. The net energy gain would be equivalent to one proton translocated form the inside to the outside per L-malate metabolized.  相似文献   

6.
In Leuconostoc mesenteroides subsp. mesenteroides 19D, citrate is transported by a secondary citrate carrier (CitP). Previous studies of the kinetics and mechanism of CitP performed in membrane vesicles of L. mesenteroides showed that CitP catalyzes divalent citrate HCit2-/H+ symport, indicative of metabolic energy generation by citrate metabolism via a secondary mechanism (C. Marty-Teysset, J. S. Lolkema, P. Schmitt, C. Divies, and W. N. Konings, J. Biol. Chem. 270:25370-25376, 1995). This study also revealed an efficient exchange of citrate and D-lactate, a product of citrate/carbohydrate cometabolism, suggesting that under physiological conditions, CitP may function as a precursor/product exchanger rather than a symporter. In this paper, the energetic consequences of citrate metabolism were investigated in resting cells of L. mesenteroides. The generation of metabolic energy in the form of a pH gradient (delta pH) and a membrane potential (delta psi) by citrate metabolism was found to be largely dependent on cometabolism with glucose. Furthermore, in the presence of glucose, the rates of citrate utilization and of pyruvate and lactate production were strongly increased, indicating an enhancement of citrate metabolism by glucose metabolism. The rate of citrate metabolism under these conditions was slowed down by the presence of a membrane potential across the cytoplasmic membrane. The production of D-lactate inside the cell during cometabolism was shown to be responsible for the enhancement of the electrogenic uptake of citrate. Cells loaded with D-lactate generated a delta psi upon dilution in buffer containing citrate, and cells incubated with citrate built up a pH gradient upon addition of D-lactate. The results are consistent with an electrogenic citrate/D-lactate exchange generating in vivo metabolic energy in the form of a proton electrochemical gradient across the membrane. The generation of metabolic energy from citrate metabolism in L. mesenteroides may contribute significantly to the growth advantage observed during cometabolism of citrate and glucose.  相似文献   

7.
A facultative psychrophilic bacterium, strain L-2, that grows at 0 and 5°C as minimum growth temperatures in complex and defined media, respectively, was isolated. On the basis of taxonomic studies, strain L-2 was identified as Cobetia marina. The adaptability of strain L-2 to cold temperature was higher than that of the type strain and of other reported strains of the same species. When the bacterium was grown at 5–15°C in a defined medium, it produced a high amount of trans-unsaturated fatty acids. By contrast, in a complex medium in the same temperature range it produced a low amount of trans-unsaturated fatty acids. In the complex medium at 5°C, the bacterium exhibited a three-fold higher growth rate than that obtained in the defined medium. Following a temperature shift from 11 to 5°C, strain L-2 grew better in complex than in defined medium. Furthermore, when the growth temperature was shifted from 0 to 5°C both the growth rate and the yield of strain L-2 growing in complex medium was markedly enhanced. These phenomena suggest that an upshift of the growth temperature had a positive effect on metabolism. The effects of adding complex medium components to the defined medium on bacterial growth rate and fatty acid composition at 5°C were also studied. The addition of yeast extract followed by peptone was effective in promoting rapid growth, while glutamate addition was less effective, resulting in a cis-unsaturated fatty acid ratio similar to that of cells grown in the complex medium. These results suggest that the rapid growth of strain L-2 at low temperatures requires a high content of various amino acids rather than the presence of a high ratio of cis-unsaturated fatty acids in the cell membrane.  相似文献   

8.
This work was designed to determine the role of the acidity and aluminium stress in the selection of partners in the Acacia symbioses with relevance to the persistence of the microsymbiont Bradyrhizobium in the soil and the growth and nodulation of the host plant respectively. Fifteen strains of Bradyrhizobium from Acacia mangium and Faidherbia albida formed a very homogenous acid tolerant group as indicated by their ability to grow better in a medium at pH 4.5 than in a medium at pH 6.8. By contrast, a growth experiment using an acid liquid media (pH 4.5), containing different concentrations of aluminium successfully identified strains sensitive to aluminium toxicity and those able to grow even in the presence of 100 M AlCl3.Our results suggest that high amounts of aluminium in the soil rather than acidity (pH 4.5) were a major soil factor for selection of Bradyrhizobium strains capable of establishing a permanently high population under natural conditions.Unlike the behaviour of the microsymbiont, growth and nodulation of Acacia mangium and Faidherbia albida were not affected by aluminium, even at 100 M, but they might be significantly affected by medium acidity (pH 4.5) depending on plant provenances. It is therefore suggested that ability of the host plant to tolerate acidity stress should be taken into account first when screening effective Acacia-Bradyrhizobium combinations for use in afforestation trials.  相似文献   

9.
Summary Fifteen organic acids were examined for their abilities to support the growth and anthocyanin accumulation by suspension cultures of wild carrot (Daucus carota L.) using ammonium as the sole nitrogen source. Glutarate, adipate, pimelate, azelate, cinnamate, and phthalate were toxic to the culture. They prevented growth and anthocyanin accumulation at 5 mM or less in media that were otherwise adequate for growth. Succinate, fumarate, malate, α-ketoglutarate, glutamate, maleate, malonate, tartarate, and citrate all supported growth and anthocyanin accumulation but in varying amounts. The growth achieved in medium containing 20 mM acid was higher at an initial pH of 5.5 than at an initial pH of 4.5. The growth achieved was dependent on the organic acid used, its concentration, and the initial pH of the medium. When growth occurred the final pH was higher than the initial pH with most of the acids. Anthocyanin accumulation was greatest with succinate at 14 to 20 mM at an initial pH of 4.2 or 4.3 and declined when the initial pH was above 4.3. These studies were supported by grants from General Foods Corporation and the W. Alton Jones Foundation.  相似文献   

10.
Growth and differentiation of mycelial strands in Rigidoporus lignosus have been shown to depend on suitable combinations of the pH of the media and the nature of the nitrogen and carbon sources. Amino acids as sole nitrogen sources gave rise to vegetative mycelium. At pH 4.5, growth and mycelial strand differentiation required asparagine, as the fungus failed to grow in the absence of this amino acid. However, at pH 6, differentiation of strands occurred appreciably in asparagine-deficient media, suggesting a close balance between pH and amino acid requirements. Ammonium was required for strand differentiation, while nitrate, as a sole nitrogen source, maintained the fungus undifferentiated. Of the carbohydrates tested, only glucose, fructose and mannose supported strand differentiation. Starch was found to be particularly effective in promoting growth of vegetative mycelium. Strand differentiation required more specific conditions than growth of the vegetative mycelium.  相似文献   

11.
In cells of Leuconostoc oenos, the fermentation of L-malic acid generates both a transmembrane pH gradient, inside alkaline, and an electrical potential gradient, inside negative. In resting cells, the proton motive force ranged from -170 mV to -88 mV between pH 3.1 and 5.6 in the presence Of L-malate. Membrane potentials were calculated by using a model for probe binding that accounted for the different binding constants at the different pH values at the two faces of the membrane. The delta psi generated by the transport of monovalent malate, H-malate-, controlled the rate of fermentation. The fermentation rate significantly increased under conditions of decreased delta psi, i.e., upon addition of the ionophore valinomycin in the presence of KCl, whereas in a buffer depleted of potassium, the addition of valinomycin resulted in a hyperpolarization of the cell membrane and a reduction of the rate of fermentation. At the steady state, the chemical gradient for H-malate- was of the same magnitude as delta psi. Synthesis of ATP was observed in cells performing malolactic fermentation.  相似文献   

12.
This study examines the interactions that occur between Saccharomyces cerevisiae and Oenococcus oeni strains during the process of winemaking. Various yeast/bacteria pairs were studied by applying a sequential fermentation strategy which simulated the natural winemaking process. First, four yeast strains were tested in the presence of one bacterial strain leading to the inhibition of the bacterial component. The extent of inhibition varied widely from one pair to another and closely depended on the specific yeast strain chosen. Inhibition was correlated to weak bacterial growth rather than a reduction in the bacterial malolactic activity. Three of the four yeast strains were then grown with another bacteria strain. Contrary to the first results, this led to the bacterial stimulation, thus highlighting the importance of the bacteria strain. The biochemical profile of the four yeast fermented media exhibited slight variations in ethanol, SO(2) and fatty acids produced as well as assimilable consumed nitrogen. These parameters were not the only factors responsible for the malolactic fermentation inhibition observed with the first bacteria strain. The stimulation of the second has not been reported before in such conditions and remains unexplained.  相似文献   

13.
The effects of the Ca2+/H+ exchanger A23187 and the K+/H+ exchanger nigericin on the growth of Neurospora crassa were analyzed. Both ionophores had the same effects on the fungus. They both inhibited growth in liquid media, apical extension being more affected than protein synthesis. A sudden challenge to either ionophore on solid media rapidly stopped hyphal extension. Additionally, both ionophores induced profuse mycelium branching and upward hyphal growth. Hyphae growing on nigericin-containing media also burst at the apex. Both ionophores caused a rapid inhibition in the apically-occurring synthesis of structural wall polysaccharides, but they did not affect mitochondrial energy conservation. With the use of DiBAC, a membrane-potential sensitive fluorophore, it was excluded that their effects were due to depletion of the plasma membrane potential. Considering that both ionophores exchange H+ for different metallic ions, we concluded that their effect was due to dissipation of a proton gradient, which is directly or indirectly involved in the apical growth of the fungus. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

14.
Thiobacillus ferro-oxidans is capable of using the oxidation of Fe2+ by O2 at pH 2.0 as the sole source of energy for growth and CO2 fixation. The bacterium maintains an intracellular pH of 6.5 over a range of external pH from 1.0 to 8.0, as measured by [14C]acetate and [3H]methylamine distribution. The membrane potential was estimated by the distribution of the lipid-soluble cation dibenzyldimethylammonium and the anion SCN-. At pH 2.0 (the pH of growth) during Fe2+ oxidation the transmembrane pH gradient is 4.5 units with an opposing membrane potential of -10mV, giving a proton electrochemical gradient of +256mV. This gradient is actively maintained.  相似文献   

15.
Low-acid foods (pH greater than or equal to 4.5) are not sufficiently acidic to prevent growth of Clostridium botulinum in otherwise optimal conditions. The combination of sub-optimal pH and sub-optimal temperature may, however, result in a very significant reduction in the risk of growth of this bacterium compared with the risk in optimal conditions. The combined effect of incubation temperatures of 12 degrees and 16 degrees C and pH values between 5.2 and 5.5 on growth and toxin production from spores of Cl. botulinum during incubation for 28 d has been investigated. Growth and formation of toxin (type B) were detected only in medium at pH 5.5 and incubated at 16 degrees C, corresponding to a probability of growth from a single spore within 14 d of 1.6 x 10(-5). The probability of growth in 28 d in the remaining conditions was less than 9 x 10(-6). After transfer of inoculated media from 12 degrees to 30 degrees C growth occurred at pH 5.2-5.5 within 19 d. After transfer of inoculated media from 12 degrees to 20 degrees C growth occurred at pH 5.5 and 5.4 but not at pH 5.3 or 5.2 in 40 d. Growth at pH 5.2-5.5 was accompanied by formation of toxin, in most cases of types A or B. In addition to the effect of sub-optimal temperature and pH, chelation of divalent metal ions by citrate may have contributed to inhibition.  相似文献   

16.
David G. Pope 《Planta》1978,140(2):137-142
Growth of Triticum aestivum L. cv. Cappelle Desprez coleoptiles is promoted by 5.7×10–5 M indole acetic acid (IAA) as effectively in pH 3.4 buffer as in water, but IAA is not effective in the presence of buffer at pH 3.0 or 3.2 A combination of 5.7×10–5 M IAA and pH 3.4 buffer promotes growth to a greater extent than pH 3.2 buffer alone, which is optimal for acid-induced growth. IAA employed at 10–7 M is still effective at promoting growth in the presence of pH 3.4 buffer, moreover, IAA at 10–7 M interacts synergistically with the acidic buffer to promote growth. It is concluded that IAA and acid promote growth via separate mechanisms, and that IAA does not promote cell wall loosening by rendering the cell wall more acid.Abbreviation IAA Indoleacetic acid  相似文献   

17.
The expression of genes encoding sodium-translocating NADH:quinone oxidoreductase (Na+-NQR) was studied in the marine bacterium Vibrio harveyi and in the enterobacterium Klebsiella pneumoniae. It has been shown that such parameters as NaCl concentration, pH value, and presence of an uncoupler in the growth media do not influence significantly the level of nqr expression. However, nqr expression depends on the growth substrates used by these bacteria. Na+-NQR is highly repressed in V. harveyi during anaerobic growth, and nqr expression is modulated by electron acceptors and values of their redox potentials. The latter effect was shown to be independent of the ArcAB regulatory system. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users. Accession number: EF394942 (Vibrio harveyi arcB gene, partial cds).  相似文献   

18.
Propionic acid and its sodium salt have long been used as additives in poultry feed to reduce microbial populations, including Salmonella spp. Propionic acids in poultry feed may have a potential role in inhibiting growth of Salmonella in the chicken intestine. In this study, we determined growth response of a Salmonella typhimurium poultry isolate to propionic acid and sodium propionate under aerobic and anaerobic conditions. Growth rate consistently decreased with the addition of greater concentrations of either propionic acid or sodium propionate. The extent of growth inhibition was much greater with propionic acid than the sodium form. Media pH decreased only with addition of propionic acid. Growth inhibition was more effective under anaerobic growth conditions with either propionic acid or sodium propionate. When determined at the same pH level, growth rate was significantly lowered by addition of 25 mM of either propionate or sodium propionate alone, and also by the decrease in pH levels (P<0.05). These results showed that growth inhibition of S. typhimurium by propionic acid or sodium propionate is greatly enhanced by pH decrease, and to lesser extent by anaerobiosis. We also found that sodium propionate was more inhibitory for growth of S. typhimurium than propionic acid when compared at the same pH levels.  相似文献   

19.
A psychrophilic bacterium, Pseudomonas syringae (Lz4W) from Antarctica, was used as a model system to establish a correlation, if any, between thermal adaptation, trans-fatty acid content and membrane fluidity. In addition, attempts were made to clone and sequence the cti gene of P. syringae (Lz4W) so as to establish its characteristics with respect to the cti of other Pseudomonas spp. and also to in vitro mutagenize the cti gene so as to generate a cti null mutant. The bacterium showed increased proportion of saturated and trans-monounsaturated fatty acids when grown at 28°C compared to cells grown at 5°C, and the membrane fluidity decreased with growth temperature. In the mutant, the trans-fatty acid was not synthesized, and the membrane fluidity also decreased with growth temperature, but the decrease was not to the extent that was observed in the wild-type cells. Thus, it would appear that synthesis of trans-fatty acid and modulation of membrane fluidity to levels comparable to the wild-type cells is essential for growth at higher temperatures since the mutant exhibits growth arrest at 28°C. In fact, the cti null mutant-complemented strain of P. syringae (Lz4W-C30b) that was capable of synthesizing the trans-fatty acid was indeed capable of growth at 28°C, thus confirming the above contention. The cti gene of P. syringae (Lz4W) that was cloned and sequenced exhibited high sequence identity with the cti of other Pseudomonas spp. and exhibited all the conserved features.  相似文献   

20.
Clostridium thermobutyricum produces butyrate as the main fermentation product from glucose, and from yeast extract, which is required for substantial growth. After sequential transfer in the presence of increasing butyrate concentrations, strain JW 171 K grew in the presence of up to 350 mM butyrate either at pH 5.5 or at pH 8.0 and at 40 degrees C as well as at 60 degrees C. This result indicated that butyrate-dependent growth inhibition was independent from the concentration of undissociated butyric acid. Increased butyrate concentration decreased the level of tolerated glucose from above 15% to below 10%. At 0.05 and 2.0% (wt/vol) yeast extract, the Y(Glucose) was 30 and 55 g dry weight cells per mole glucose, respectively. Y(ATP) values between 18 and 21 g weight cells per mole ATP, obtained after growth in the presence of 2% yeast extract, indicate that the butyrate fermentation under thermophilic growth conditions is as energy efficient as it is under mesophilic conditions. Externally added acetate stimulated the production of butyrate. Supplemented 14C-acetate was converted to butyrate, resulting in the formation of 44% labeled butyrate (i.e. formed from 14C-acetate) and 56% unlabeled butyrate (formed from glucose and yeast extract). Continuous removal of H2 in batch cultures led to a shift in the fermentation products from more butyrate to the more oxidized and more energy yielding acetate.  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