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.
Proton motive force during growth of Streptococcus lactis cells   总被引:38,自引:20,他引:18       下载免费PDF全文
Experiments with the aerotolerant anaerobe Streptococcus lactis provide the opportunity for determining the proton motive force (Δp) in dividing cells. The two components of Δp, ΔΨ (the transmembrane potential) and ΔpH (the chemical gradient of H+), were determined by the accumulation of radiolabeled tetraphenylphosphonium (TPP+) and benzoate ions. The ΔΨ was calibrated with the K+ diffusion potential in starved, valinomycin-treated cells. With resting, glycolyzing cells, the Δp was measured also by the accumulation of the non-metabolizable sugar thiomethyl-β-galactoside (TMG). In resting cells the Δp, calculated either by adding ΔΨ and ZΔpH or from the levels of TMG, was relatively constant between pH 5 to 7, decreasing from 160 to 150 mV and decreasing further to 100 mV at pH 8.0. With the TPP+ probe for ΔΨ, we confirmed our previous finding that the K+ ions dissipate ΔΨ and increase ΔpH, whereas Na+ ions have little effect on ΔΨ and no effect on ΔpH. [3H]TPP+ and [14C]benzoate were added during exponential phase to S. lactis cells growing at pH 5 to 7 at 28°C in a defined medium with glucose as energy source. As with resting cells, the ΔpH and ΔΨ were dependent on the pH of the medium. At pH 5.1, the ΔpH was equivalent to 60 mV (alkaline inside) and decreased to 25 mV at pH 6.8. The ΔΨ increased from 83 mV (negative inside) at pH 5.1 to 108 mV at pH 6.8. The Δp, therefore, was fairly constant between pH 5 and 7, decreasing from 143 to 133 mV. The values for Δp in growing cells, just as in resting cells, are consistent with a system in which the net efflux of H+ ions is effected by a membrane-bound adenosine triphosphatase and glycolytically generated adenosine triphosphate. The data suggest that in both growing and resting cells the pH of the medium and its K+ concentration are the two principal factors that determine the relative contribution of ΔpH and ΔΨ to the proton motive force.  相似文献   

2.
Streptococcus thermophilus is unable to metabolize the galactose moiety of lactose. In this paper, we show that a transformant of S. thermophilus SMQ-301 expressing Streptococcus salivarius galK and galM was able to grow on galactose and expelled at least twofold less galactose into the medium during growth on lactose.  相似文献   

3.
When cells of the anaerobic thermophile Clostridium thermoaceticum grow in batch culture and homoferment glucose to acetic acid, the pH of the medium decreases until growth and then acid production cease, at about pH 5. We postulated that the end product of fermentation limits growth by acting as an uncoupling agent. Thus, when the pH of the medium is low, the cytoplasm of the cells becomes acidified below a tolerable pH. We have therefore measured the internal pH of growing cells and compared these values with those of nongrowing cells incubated in the absence of acetic acid. Growing cells maintained an interior about 0.6 pH units more alkaline than the exterior throughout most of batch growth (i.e., ΔpH = 0.6). We also measured the transmembrane electrical potential (ΔΨ), which decreased from 140 mV at pH 7 at the beginning of growth to 80 mV when the medium had reached pH 5. The proton motive force, therefore, was 155 mV at pH 7, decreasing to 120 mV at pH 5. When further fermentation acidified the medium below pH 5, both the ΔpH and the ΔΨ collapsed, indicating that these cells require an internal pH of at least 5.5 to 5.7. Cells harvested from stationary phase and suspended in citrate-phosphate buffer maintained a ΔpH of 1.5 at external pH 5.0. This ΔpH was dissipated by acetic acid (at the concentrations found in the growth medium) and other weak organic acids, as well as by ionophores and inhibitors of glycolysis and of the H+-ATPase. Nongrowing cells had a ΔΨ which ranged from about 116 mV at external pH 7 to about 55 mV at external pH 5 and which also was sensitive to ionophores. Since acetic acid, in its un-ionized form, diffuses passively across the cytoplasmic membrane, it effectively renders the membrane permeable to protons. It therefore seems unlikely that mutations at one or a few loci would result in C. thermoaceticum cells significantly more acetic acid tolerant than their parental type.  相似文献   

4.
The hydrolysis of the dipeptide leucyl-leucine by whole cells of Streptococcus cremoris Wg2 was dependent on the presence of the energy source lactose. Incubation of cells with uncouplers or ATPase inhibitors prevented the increase of peptidase activity upon the addition of lactose. Incubation with the ionophore nigericin resulted in decreased peptide hydrolysis activity, while incubation with valinomycin led to increased hydrolysis activity. In the presence of nigericin the ΔpH component of the proton motive force was decreased, while the electrical potential was increased. With valinomycin, the electrical potential was collapsed and the ΔpH was increased. When the external pH was decreased from 8 to 5, the rate of peptide hydrolyzing activity by whole cells increased with increasing ΔpH component. In contrast, the peptide hydrolyzing activity in the cell extract decreased with decreasing external pH. These results indicate that the ΔpH component of the proton motive force determines the leucyl-leucine hydrolyzing activity in S. cremoris Wg2.  相似文献   

5.
A Thermus thermophilus selector strain for production of thermostable and thermoactive α-galactosidase was constructed. For this purpose, the native α-galactosidase gene (agaT) of T. thermophilus TH125 was inactivated to prevent background activity. In our first attempt, insertional mutagenesis of agaT by using a cassette carrying a kanamycin resistance gene led to bacterial inability to utilize melibiose (α-galactoside) and galactose as sole carbohydrate sources due to a polar effect of the insertional inactivation. A Gal+ phenotype was assumed to be essential for growth on melibiose. In a Gal background, accumulation of galactose or its metabolite derivatives produced from melibiose hydrolysis could interfere with the growth of the host strain harboring recombinant α-galactosidase. Moreover, the AgaT strain had to be Kms for establishment of the plasmids containing α-galactosidase genes and the kanamycin resistance marker. Therefore, a suitable selector strain (AgaT Gal+ Kms) was generated by applying integration mutagenesis in combination with phenotypic selection. To produce heterologous α-galactosidase in T. thermophilus, the isogenes agaA and agaB of Bacillus stearothermophilus KVE36 were cloned into an Escherichia coli-Thermus shuttle vector. The region containing the E. coli plasmid sequence (pUC-derived vector) was deleted before transformation of T. thermophilus with the recombinant plasmids. As a result, transformation efficiency and plasmid stability were improved. However, growth on minimal agar medium containing melibiose was achieved only following random selection of the clones carrying a plasmid-based mutation that had promoted a higher copy number and greater stability of the plasmid.  相似文献   

6.
Fifty-one mutants of Kluyveromyces lactis that cannot grow on lactose (Lac-) were isolated and characterized. All of the mutations are in nuclear genes, are recessive in their wild-type allele and define seven complementation groups, which we designate lac3 through lac9. Strains bearing mutations in lac3, lac5, lac7, lac8 and lac9 are also unable to grow on galactose (Gal-). Since the Gal- and Lac- phenotype co-segregate, they are probably due to a single mutation. Strains bearing mutations in any of the seven complementation groups grow normally on glucose. However, strains bearing mutations in lac3, lac5 and lac6 do not grow on glucose if lactose is also present in the medium. Likewise, strains bearing mutations in lac3 and lac5 do not grow on glucose in the presence of galactose. Complementation groups lac4 and lac5 are loosely linked and map within a cluster of auxotrophic mutations on a chromosome that we designate chromosome 2. The remaining five groups are unlinked. Thus, there is no evidence for clustering of Lac genes into an operon-like regulatory unit.——To further characterize the nature of the Lac- phenotype, the basal and inducible level of β-galactosidase activity were measured. All mutants had nearly normal basal enzyme levels, except those in lac4, which had barely detectable levels. Inducible enzyme levels varied from barely detectable levels in mutants bearing lac4 mutations up to four-fold inducible levels in strains bearing mutations in other complementation groups. In all cases, however, induction levels were below the 30-fold level obtained in wild-type cells. Three strains bearing lac5 mutations contain increased enzyme activity in the absence of inducer, indicating constitutive synthesis of β-galactosidase. In summary, these data indicate that several genes are necessary for synthesis of β-galactosidase activity.  相似文献   

7.
The mechanism responsible for an increased rate of acid production when yogurt starter cultures are grown in milk treated with lactase enzyme was investigated by studying carbohydrate utilization and acid development by a pure culture of Streptococcus thermophilus and a mixed yogurt starter culture consisting of S. thermophilus and Lactobacillus bulgaricus. In milk containing glucose, galactose, and lactose, glucose and lactose (but not free galactose) were fermented. Fermentation of lactose in control milk was accompanied by the release of free galactose, with the result that carbohydrate utilization was less efficient than in treated milk. This phenomenon also occurred when lactose was fermented by S. thermophilus in broth culture. Carbohydrate utilization by the mixed yogurt culture was more rapid when the lactose in milk was partially prehydrolyzed. Our results suggest that the more rapid acid development that took place when a mixed yogurt starter culture was grown in milk containing prehydrolyzed lactose was the result of a more rapid and efficient utilization of carbohydrate by S. thermophilus when free glucose in addition to lactose was available for fermentation. The evidence presented also suggests that uptake and utilization of glucose and lactose by S. thermophilus are different in broth and milk cultures.  相似文献   

8.
Alanyl-alpha-glutamate transport has been studied in Lactococcus lactis ML3 cells and in membrane vesicles fused with liposomes containing beefheart cytochrome c oxidase as a proton-motive-force-generating system. The uptake of Ala-Glu observed in de-energized cells can be stimulated 26-fold upon addition of lactose. No intracellular dipeptide pool could be detected in intact cells. In fused membranes, a 40-fold accumulation of Ala-Glu was observed in response to a proton motive force. Addition of ionophores and uncouplers resulted in a rapid efflux of the accumulated dipeptide, indicating that Ala-Glu accumulation is directly coupled to the proton motive force as a driving force. Ala-Glu uptake is an electrogenic process and the dipeptide is transported in symport with two protons. In both fused membranes and intact cells the same affinity constant (0.70 mM) for Ala-Glu uptake was found. Accumulated Ala-Glu is exchangeable with externally added alanyl-glutamate, glutamyl-glutamate, and leucyl-leucine, while no exchange occurred upon addition of the amino acid glutamate or alanine. These results indicate that the Ala-Glu transport system has a broad substrate specificity.  相似文献   

9.
Exopolysaccharide production by Lactobacillus casei CG11 was studied in basal minimum medium containing various carbon sources (galactose, glucose, lactose, sucrose, maltose, melibiose) at concentrations of 2, 5, 10, and 20 g/liter. L. casei CG11 produced exopolysaccharides in basal minimum medium containing each of the sugars tested; lactose and galactose were the poorest carbon sources, and glucose was by far the most efficient carbon source. Sugar concentrations had a marked effect on polymer yield. Plasmid-cured Muc- derivatives grew better in the presence of glucose and attained slightly higher populations than the wild-type strain. The values obtained with lactose were considerably lower for both growth and exopolysaccharide yield. The level of specific polymer production per cell obtained with glucose was distinctively lower for Muc- derivatives than for the Muc+ strain. The polymer produced by L. casei CG11 in the presence of glucose was different from that formed in the presence of lactose. The polysaccharide produced by L. casei CG11 in basal minimum medium containing 20 g of glucose per liter had an intrinsic viscosity of 1.13 dl/g. It was rich in glucose (76%), which was present mostly as 2- or 3-linked residues along with some 2,3 doubly substituted glucose units, and in rhamnose (21%), which was present as 2-linked or terminal rhamnose; traces of mannose and galactose were also present.  相似文献   

10.
H+ conductance in de-energized cells of Clostridium thermoaceticum and Clostridium acetobutylicum was determined from the rate of realkalinization of the medium after an acid pulse. In both organisms, cell membrane proton permeability was increased by fermentation end products and ionophores. In C. thermoaceticum, H+ conductance was increased by Na+ ions compared with K+ as counterions. In these cells, addition of Na+, but not K+, elicited efflux of H+; H+ efflux was stimulated by SCN and decreased by various ionophores. We concluded that C. thermoaceticum possesses an electrogenic Na+/H+ antiporter. In contrast, C. acetobutylicum cells did not have an electrogenic Na+/H+ antiporter.  相似文献   

11.
Lactose-limited fermentations of 49 dairy Streptococcus thermophilus strains revealed four distinct fermentation profiles with respect to galactose consumption after lactose depletion. All the strains excreted galactose into the medium during growth on lactose, except for strain IMDOST40, which also displayed extremely high galactokinase (GalK) activity. Among this strain collection eight galactose-positive phenotypes sensu stricto were found and their fermentation characteristics and Leloir enzyme activities were measured. As the gal promoter seems to play an important role in the galactose phenotype, the galR-galK intergenic region was sequenced for all strains yielding eight different nucleotide sequences (NS1 to NS8). The gal promoter played an important role in the Gal-positive phenotype but did not determine it exclusively. Although GalT and GalE activities were detected for all Gal-positive strains, GalK activity could only be detected for two out of eight Gal-positive strains. This finding suggests that the other six S. thermophilus strains metabolize galactose via an alternative route. For each type of fermentation profile obtained, a representative strain was chosen and four complete Leloir gene clusters were sequenced. It turned out that Gal-positive strains contained more amino acid differences within their gal genes than Gal-negative strains. Finally, the biodiversity regarding lactose-galactose utilization among the different S. thermophilus strains used in this study was shown by RAPD-PCR. Five Gal-positive strains that contain nucleotide sequence NS2 in their galR-galK intergenic region were closely related.  相似文献   

12.
To study the influence of phosphoglucomutase (PGM) activity on exopolysaccharide (EPS) synthesis in glucose- and lactose-growing Streptococcus thermophilus, a knockout PGM mutant and a strain with elevated PGM activity were constructed. The pgmA gene, encoding PGM in S. thermophilus LY03, was identified and cloned. The gene was functional in Escherichia coli and was shown to be expressed from its own promoter. The pgmA-deficient mutant was unable to grow on glucose, while the mutation did not affect growth on lactose. Overexpression of pgmA had no significant effect on EPS production in glucose-growing cells. Neither deletion nor overexpression of pgmA changed the growth or EPS production on lactose. Thus, the EPS precursors in lactose-utilizing S. thermophilus are most probably formed from the galactose moiety of lactose via the Leloir pathway, which circumvents the need for a functional PGM.  相似文献   

13.
14.
It is possible that the low levels of production of exopolysaccharides (EPSs) by lactic acid bacteria could be improved by altering the levels of enzymes in the central metabolism that influence the production of precursor nucleotide sugars. To test this hypothesis, we identified and cloned the galU gene, which codes for UDP glucose pyrophosphorylase (GalU) in Streptococcus thermophilus LY03. Homologous overexpression of the gene led to a 10-fold increase in GalU activity but did not have any effect on the EPS yield when lactose was the carbon source. However, when galU was overexpressed in combination with pgmA, which encodes phosphoglucomutase (PGM), the EPS yield increased from 0.17 to 0.31 g/mol of carbon from lactose. A galactose-fermenting LY03 mutant (Gal+) with increased activities of the Leloir enzymes was also found to have a higher EPS yield (0.24 g/mol of carbon) than the parent strain. The EPS yield was further improved to 0.27 g/mol of carbon by overexpressing galU in this strain. However, the highest EPS yield, 0.36 g/mol of carbon, was obtained when pgmA was knocked out in the Gal+ strain. Measurements of the levels of intracellular metabolites in the cultures revealed that the Gal+ strains had considerably higher glucose 1-phosphate levels than the other strains, and the strain lacking PGM activity had threefold-higher levels of glucose 1-phosphate than the other Gal+ strains. These results show that it is possible to increase EPS production by altering the levels of enzymes in the central carbohydrate metabolism.  相似文献   

15.
Twelve lactose-assimilating strains of the yeast species Kluyveromyces marxianus and its varieties marxianus, lactis and bulgaricus were studied with respect to transport mechanisms for lactose, glucose and galactose, fermentation of these sugars and the occurrence of extracellular lactose hydrolysis. The strains fell into three groups. Group I (two strains): Fermentation of lactose, glucose and galactose, extracellular lactose hydrolysis, apparent facilitated diffusion of glucose and galactose; Group II (two strains): Lactose not fermented, glucose and galactose fermented and transported by an apparent proton symport, extracellular hydrolysis of lactose present (one strain) or questionable; Group III (eight strains): Lactose, glucose and galactose fermented, lactose transported by an apparent proton symport mechanism, extracellular hydrolysis of lactose and transport modes for glucose and galactose variable.  相似文献   

16.
The components of the proton motive force (Δp), namely, membrane potential (Δψ) and transmembrane pH gradient (ΔpH), were determined in the nitrifying bacteria Nitrosomonas europaea and Nitrobacter agilis. In these bacteria both Δψ and ΔpH were dependent on external pH. Thus at pH 8.0, Nitrosomonas europaea and Nitrobacter agilis had Δψ values of 173 mV and 125 mV (inside negative), respectively, as determined by the distribution of the lipophilic cation [3H]tetraphenyl phosphonium. Intracellular pH was determined by the distribution of two weak acids, 14C-benzoic and 14C-acetyl salicylic, and the weak base [14C]methylamine. Nitrosomonas europaea accumulated 14C-benzoic acid and 14C-acetyl salicylic acid when the external pH was below 7.0 and [14C]methylamine at alkaline pH. Similarly, Nitrobacter agilis accumulated the two weak acids below an external pH of about 7.5 and [14C]methylamine above this pH. As these bacteria grow best between pH 7.5 and 8.0, they do not appear to have a ΔpH (inside alkaline). Thus, above pH 7.0 for Nitrosomonas europaea and pH 7.5 for Nitrobacter agilis, Δψ only contributed to Δp. In Nitrosomonas europaea the total Δp remained almost constant (145 to 135 mV) when the external pH was varied from 6 to 8.5. In Nitrobacter agilis, Δp decreased from 178 mV (inside negative) at pH 6.0 to 95 mV at pH 8.5. Intracellular pH in Nitrosomonas europaea varied from 6.3 at an external pH of 6.0 to 7.8 at external pH 8.5. In Nitrobacter agilis, however, intracellular pH was relatively constant (7.3 to 7.8) over an external pH range of 6 to 8.5. In Nitrosomonas europaea, Δp and its components (Δψ and ΔpH) remained constant in cells at various stages of growth, so that the metabolic state of cells did not affect Δp. Such an experiment was not possible with Nitrobacter agilis because of low cell yields. The effects of protonophores and ATPase inhibitors on ΔpH and Δψ in the two nitrifying bacteria are considered.  相似文献   

17.
《Process Biochemistry》2010,45(9):1494-1503
Lactose, an inexpensive, soluble substrate, offers reasonably good induction for cellulase production by Trichoderma reesei. The fungus does not uptake lactose directly. Lactose is hydrolyzed to extracellular glucose and galactose for subsequent ingestion. The roles of this extracellular hydrolysis step were investigated in this study. Batch and continuous cultures were grown on the following substrates: lactose, lactose–glycerol mixtures, glucose, galactose, and glucose–galactose mixtures. Cell growth, substrate consumption, lactose hydrolysis, and lactase and cellulase production were followed and modeled. Cells grew much faster on glucose than on galactose, but with comparable cell yields. Glucose (at >0.3 g/L) repressed the galactose consumption. Cellulase synthesis was growth-independent while lactase synthesis was growth-dependent, except at D < ∼0.065 h−1 where a basal level lactase production was observed. For cellulase production the optimal D was 0.055–0.065 h−1 where the enzyme activity and productivity were both near maxima. The model suggested that lactase synthesis was subject to weak galactose repression. As the galactose concentration increased at high D (>0.1 h−1), lactase synthesis became repressed. The insufficient lactase synthesis limited the lactose hydrolysis rate. Extracellular lactose hydrolysis was concluded to be the rate-limiting step for growth of T. reesei Rut C30 on lactose.  相似文献   

18.
In Escherichia coli, the small multidrug resistance (SMR) transporter protein EmrE confers host resistance to a broad range of toxic quaternary cation compounds (QCC) via proton motive force in the plasma membrane. Biologically produced QCC also act as EmrE osmoprotectant substrates within the cell and participate in host pH regulation and osmotic tolerance. Although E. coli EmrE is one of the most well-characterized SMR members, it is unclear how the substrates it transports into the periplasm escape across the outer membrane (OM) in Gram-negative bacteria. We tested the hypothesis that E. coli EmrE relies on an unidentified OM protein (OMP) to complete the extracellular release of its QCC. Eleven OMP candidates were screened using an alkaline phenotypic growth assay to identify OMP involvement in EmrE-mediated QCC efflux. E. coli single-gene deletion strains were transformed with plasmid-carried copies of emrE to detect reduced-growth and rescued-growth phenotypes under alkaline conditions. Among the 11 candidates, only the ΔompW strain showed rescued alkaline growth tolerance when transformed with pEmrE, supporting the corresponding protein''s involvement in EmrE osmoprotectant efflux. Coexpression of plasmids carrying the ompW and emrE genes transformed into the E. coli ΔompW and ΔemrE strains demonstrated a functional complementation restoring the original alkaline loss-of-growth phenotype. Methyl viologen drug resistance assays of pEmrE and pOmpW plasmid-complemented E. coli ΔompW and wild-type strains found higher host drug resistance than with other plasmid combinations. This study confirms our hypothesis that the porin OmpW participates in the efflux of EmrE-specific substrates across the OM.  相似文献   

19.
A light-induced proton gradient (ΔpH) increase as exhibited by an increase of 9-aminoacridine fluorescence quenching is demonstrated between the external medium and the interior of the halophytic green alga Dunaliella salina. The formation and maintenance of the ΔpH is sensitive to electron transport inhibitors and to uncouplers. It is inhibited by p-chloromercuribenzenesulfonic acid (50% inhibition at 3 micromolar), which does not affect photosynthetic O2 evolution. It is concluded that the observed ΔpH is located across the plasmalemma or the chloroplast envelope. The formation and maintenance of the light-induced proton gradient requires the presence of Na+. Substitution of NaCl by KCl or glycerol results in inhibition of the ΔpH formation. The proton gradient is also sensitive to ATPase and energy transfer inhibitors. It is suggested that a Na+/H+ pump mechanism may be involved in the formation of the proton gradient in intact Dunaliella cells.  相似文献   

20.
The redox potentials of the oriented films of the wild-type, the E194Q-, E204Q- and D96N-mutated bacteriorhodopsins (bR), prepared by adsorbing purple membrane (PM) sheets or its mutant on a Pt electrode, have been examined. The redox potentials (V) of the wild-type bR were −470 mV for the 13-cis configuration of the retinal Shiff base in bR and −757 mV for the all-trans configuration in H2O, and −433 mV for the 13-cis configuration and −742 mV for the all-trans configuration in D2O. The solvent isotope effect (ΔV=V(D2O)−V(H2O)), which shifts the redox potential to a higher value, originates from the cooperative rearrangements of the extensively hydrogen-bonded water molecules around the protonated CN part in the retinal Schiff base. The redox potential of bR was much higher for the 13-cis configuration than that for the all-trans configuration. The redox potentials for the E194Q mutant in the extracellular region were −507 mV for the 13-cis configuration and −788 mV for the all-trans configuration; and for the E204Q mutant they were −491 mV for the 13-cis configuration and −769 mV for the all-trans configuration. Replacement of the Glu194 or Glu204 residues by Gln weakened the electron withdrawing interaction to the protonated CN bond in the retinal Schiff base. The E204 residue is less linked with the hydrogen-bonded network of the proton release pathway compared with E194. The redox potentials of the D96N mutant in the cytoplasmic region were −471 mV for the 13-cis configuration and −760 mV for the all-trans configuration which were virtually the same as those of the wild-type bR, indicating that the D to N point mutation of the 96 residue had no influence on the interaction between the D96 residue and the CN part in the Schiff base under the light-adapted condition. The results suggest that the redox potential of bR is closely correlated to the hydrogen-bonded network spanning from the retinal Schiff base to the extracellular surface of bR in the proton transfer pathway.  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