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.
Membrane vesicles of Streptococcus lactis were used to characterize a novel anion exchange involving phosphate and sugar 6-phosphates. For vesicles loaded with 50 mM phosphate at pH 7, homologous phosphate:phosphate exchange had a maximal rate of 130 nmol/min/mg of protein and a Kt of 0.21 mM external phosphate; among phosphate analogues tested, only arsenate replaced phosphate. Heterologous exchange was studied by 2-deoxyglucose 6-phosphate entry into phosphate-loaded vesicles; this reaction had a maximal velocity of 31 nmol/min/mg of protein and a Kt of 26 microM external substrate. Sugar phosphate moved intact during this exchange, since its entry led to loss of internal 32Pi without transfer of 32P to sugar phosphate. Inhibitions of phosphate exchange suggested that the preferred sugar phosphate substrates were (Kiapp): glucose, 2-deoxyglucose, and mannose 6-phosphates (approximately 20 microM) greater than fructose 6-phosphate (150 microM) greater than glucosamine 6-phosphate (420 microM) greater than alpha-methylglucoside 6-phosphate (740 microM). Stoichiometry for phosphate:2-deoxyglucose 6-phosphate antiport was 2:1 at pH 7, and since initial rates of exchange were unaffected by charge carrying ionophores (gramicidin, valinomycin, a protonophore), this unequal stoichiometry indicated the electroneutral exchange of two monovalent phosphates for a single divalent sugar phosphate.  相似文献   

2.
An intracellular hexose 6-phosphate:phosphohydrolase (EC 3.1.3.2) has been purified from Streptococcus lactis K1. Polyacrylamide disc gel electrophoresis of the purified enzyme revealed one major activity staining protein and one minor inactive band. The Mr determined by gel permeation chromatography was 36,500, but sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed a single polypeptide of apparent Mr 60,000. The enzyme exhibited a marked preference for hexose 6-phosphates, and the rate of substrate hydrolysis (at 5 mM concentration) decreased in the order, galactose 6-phosphate greater than 2-deoxy-D-glucose 6-phosphate greater than fructose 6-phosphate greater than mannose 6-phosphate greater than glucose 6-phosphate. Hexose 1-phosphates, p-nitrophenylphosphate, pyrophosphate, and nucleotides were not hydrolyzed at a significant rate. In addition, the glycolytic intermediates comprising the intracellular phosphoenolpyruvate potential in the starved cells (phosphoenolpyruvate and 2- and 3-phosphoglyceric acids) were not substrates for the phosphatase. Throughout the isolation, the hexose 6-phosphate:phosphohydrolase was stabilized by Mn2+ ion, and the purified enzyme was dependent upon Mn2+, Mg2+, Fe2+, or Co2+ for activation. Other divalent metal ions including Pb2+, Cu2+, Zn2+, Cd2+, Ca2+, Ba2+, Sr2+, and Ni2+ were unable to activate the enzyme, and the first four cations were potent inhibitors. Enzymatic hydrolysis of 2-deoxy-D-glucose 6-phosphate was inhibited by fluoride when Mg2+ was included in the assay, but only slight inhibition occurred in the presence of Mn2+, Fe2+, or Co2+. The inhibitory effect of Mg2+ plus fluoride was specifically and completely reversed by Fe2+ ion. The hexose 6-phosphate:phosphohydrolase catalyzes the in vivo hydrolysis of 2-deoxy-D-glucose 6-phosphate in stage II of the phosphoenolpyruvate-dependent futile cycle in S. lactis (J. Thompson and B. M. Chassy, J. Bacteriol. 151:1454-1465, 1982).  相似文献   

3.
The addition of 2-deoxy-D-glucose to cultures of Streptococcus lactis 133 that were growing exponentially on sucrose or lactose reduced the growth rate by ca. 95%. Inhibition did not occur with glucose or mannose as the growth sugar. The reduction in growth rate was concomitant with rapid accumulation of the analog in phosphorylated form (2-deoxy-D-glucose 6-phosphate) via the phosphoenolpyruvate-dependent mannose:phosphotransferase system. Within 5 min the intracellular 2-deoxy-D-glucose 6-phosphate concentration reached a steady-state level of greater than 100 mM. After maximum accumulation of the sugar phosphate, the rate of sucrose metabolism (glycolysis) decreased by only 30%, but the cells were depleted of fructose-1,6-diphosphate. The addition of glucose to 2-deoxy-D-glucose 6-phosphate preloaded cells caused expulsion of 2-deoxy-D-glucose and a resumption of normal growth. S. lactis 133 contained an intracellular Mg2+-dependent, fluoride-sensitive phosphatase which hydrolyzed 2-deoxy-D-glucose 6-phosphate (and glucose 6-phosphate) to free sugar and inorganic phosphate. Because of continued dephosphorylation and efflux of the non-metabolizable analog, the maintenance of the intracellular 2-deoxy-D-glucose 6-phosphate pool during growth stasis was dependent upon continued glycolysis. This steady-state condition represented a dynamic equilibrium of: (i) phosphoenolpyruvate-dependent accumulation of 2-deoxy-D-glucose 6-phosphate, (ii) intracellular dephosphorylation, and (iii) efflux of free 2-deoxy-D-glucose. This sequence of events constitutes a futile cycle which promotes the dissipation of phosphoenolpyruvate. We conclude that 2-deoxy-D-glucose functions as an uncoupler by dissociating energy production from growth in S. lactis 133.  相似文献   

4.
Starved cells of Streptococcus lactis ML3 (grown previously on galactose, lactose, or maltose) accumulated methyl-beta-D-thiogalactopyranoside (TMG) by the lactose:phosphotransferase system. More than 98% of accumulated sugar was present as a phosphorylated derivative, TMG-6-phosphate (TMG-6P). When a phosphotransferase system sugar (glucose, mannose, 2-deoxyglucose, or lactose) was added to the medium simultaneously with TMG, the beta-galactoside was excluded from the cells. Galactose enhanced the accumulation of TMG-6P. Glucose, mannose, lactose, or maltose plus arginine, was added to a suspension of TMG-6P-loaded cells of S. lactis ML3, elicited rapid expulsion of intracellular solute. The material recovered in the medium was exclusively free TMG. Expulsion of galactoside required both entry and metabolism of an appropriate sugar, and intracellular dephosphorylation of TMG-6P preceded efflux of TMG. The rate of dephosphorylation of TMG-6P by permeabilized cells was increased two-to threefold by adenosine 5'-triphosphate but was strongly inhibited by fluoride. S. lactis ML3 (DGr) was derived from S. lactis ML3 by positive selection for resistance to 2-deoxy-D-glucose and was defective in the enzyme IIMan component of the glucose:phosphotransferase system. Neither glucose nor mannose excluded TMG from cells of S. lactic ML3 (DGr), and these two sugars failed to elicit TMG expulsion from preloaded cells of the mutant strain. Accumulation of TMG-6P by S. lactis ML3 can be regulation by two independent mechanisms whose activities promote exclusion or expulsion of galactoside from the cell.  相似文献   

5.
beta-D-Phosphogalactoside galactohydrolase (beta-Pgal) was examined in a number of lactic streptococci by use of the chromogenic substrate o-nitrophenyl-beta-D-galactopyranoside-6-phosphate. Specific activity of beta-Pgal ranged from 0.563 units/mg of protein in Streptococcus lactis UN, to 0.120 in S. diacetilactics 18-16. Essentially no beta-D-galactoside galactohydrolase (beta-gal) was found in these organisms when o-nitrophenyl-beta-D-galactopyranoside served as the chromogenic substrate. S. lactis 7962 was the one exception found. This organism contained rather high levels of beta-gal, and very little beta-Pgal could be detected. beta-Pgal activity was examined in streptococci that differed widely in both their proteolytic ability and rates of lactic acid production during growth in milk. Differences in proteolytic ability did not influence beta-Pgal synthesis; also, the rate of lactic acid production was independent of the level of beta-Pgal present in the cell, since the rate of lactic acid production could be increased approximately fourfold without changing the amount of beta-Pgal present in the cell. Various carbohydrates were tested as potential inducers of the enzyme. Although galactose, either as the free sugar or combined with glucose in lactose, was the only inducer, noninducing sugars such as mannose or glucose showed some ability to cause fluctuations in the basal level of beta-Pgal. Cells growing in mannose or glucose exhibited about 30% of the maximal enzyme levels found in cells growing in lactose or galactose. No gratuitous inducers were found.  相似文献   

6.
Hydrolysis of sugar phosphates by crude and purified preparations of periplasmic hexose phosphatase from Salmonella typhimurium followed Michaelis-Menten kinetics. The enzyme bound glucose 1-phosphate with high affinity (Km = 10 microM) but bound glucose 6-phosphate with low affinity (Km = 2,000 microM). The order of substrate affinities was glucose 1-phosphate greater than mannose 1-phosphate = galactose 1-phosphate greater than fructose 1-phosphate greater than glucose 6-phosphate. These results and others suggest that the physiological function of the enzyme is the periplasmic hydrolysis of hexose 1-phosphates.  相似文献   

7.
Uptake and metabolism of sucrose by Streptococcus lactis   总被引:15,自引:11,他引:4       下载免费PDF全文
Transport and metabolism of sucrose in Streptococcus lactis K1 have been examined. Starved cells of S. lactis K1 grown previously on sucrose accumulated [14C]sucrose by a phosphoenolpyruvate-dependent phosphotransferase system (PTS) (sucrose-PTS; Km, 22 microM; Vmax, 191 mumol transported min-1 g of dry weight of cells-1). The product of group translocation was sucrose 6-phosphate (6-O-phosphoryl-D-glucopyranosyl-1-alpha-beta-2-D-fructofuranoside). A specific sucrose 6-phosphate hydrolase was identified which cleaved the disaccharide phosphate (Km, 0.10 mM) to glucose 6-phosphate and fructose. The enzyme did not cleave sucrose 6'-phosphate(D-glucopyranosyl-1-alpha-beta-2-D-fructofuranoside-6'-phosphate). Extracts prepared from sucrose-grown cells also contained an ATP-dependent mannofructokinase which catalyzed the conversion of fructose to fructose 6-phosphate (Km, 0.33 mM). The sucrose-PTS and sucrose 6-phosphate hydrolase activities were coordinately induced during growth on sucrose. Mannofructokinase appeared to be regulated independently of the sucrose-PTS and sucrose 6-phosphate hydrolase, since expression also occurred when S. lactis K1 was grown on non-PTS sugars. Expression of the mannofructokinase may be negatively regulated by a component (or a derivative) of the PTS.  相似文献   

8.
Thermodynamics of the enzyme-catalyzed (alkaline phosphatase, EC 3.1.3.1) hydrolysis of glucose 6-phosphate, mannose 6-phosphate, fructose 6-phosphate, ribose 5-phosphate, and ribulose 5-phosphate have been investigated using microcalorimetry and, for the hydrolysis of fructose 6-phosphate, chemical equilibrium measurements. Results of these measurements for the processes sugar phosphate2- (aqueous) + H2O (liquid) = sugar (aqueous) + HPO2++-(4) (aqueous) at 25 degrees C follow: delta Ho = 0.91 +/- 0.35 kJ.mol-1 and delta Cop = -48 +/- 18 J.mol-1.K-1 for glucose 6-phosphate; delta Ho = 1.40 +/- 0.31 kJ.mol-1 and delta Cop = -46 +/- 11 J.mol-1.dK-1 for mannose 6-phosphate; delta Go = -13.70 +/- 0.28 kJ.mol-1, delta Ho = -7.61 +/- 0.68 kJ.mol-1, and delta Cop = -28 +/- 42 J.mol-1.K-1 for fructose 6-phosphate; delta Ho = -5.69 +/- 0.52 kJ.mol-1 and delta Cop = -63 +/- 37 J.mol-1.K-1 for ribose 5-phosphate; and delta Ho = -12.43 +/- 0.45 kJ.mol-1 and delta Cop = -84 +/- 30 J.mol-1.K-1 for the hydrolysis of ribulose 5-phosphate. The standard state is the hypothetical ideal solution of unit molality. Estimates are made for the equilibrium constants for the hydrolysis of ribose and ribulose 5-phosphates. The effects of pH, magnesium ion concentration, and ionic strength on the thermodynamics of these reactions are considered.  相似文献   

9.
The inorganic pyrophosphate-requiring 6-phosphofructokinase of Entamoeba histolytica has been further investigated. The molecular weight of the enzyme is approximately 83,000 and its isoelectric point occurs at pH 5.8 to 6.0. The divalent cation requirement for reaction was explored. In the direction of fructose 6-phosphate formation half-maximal rate required 500 muM magnesium ion; in the direction of fructose bisphosphate formation 8 muM magnesium ion sufficed. ATP, PPi, polyphosphate, acetyl phosphate, or carbamyl phosphate cannot replace PPi as phosphate donor for the conversion of fructose 6-phosphate to fructose bisphosphate. In the direction of fructose 6-phosphate formation arsenate can replace orthophosphate. Isotope exchange studies indicate that little or no exchange occurs between Pi and PPi or between fructose 6-phosphate and fructose bisphosphate in the absence of a third substrate. These findings appear to rule out phosphoenzyme formation and a ping-pong reaction mechanism. PPi, Pi, and fructose bisphosphate are competitive inhibitors of fructose bisphosphate, PPi, and fructose 6-phosphate, respectively. This argues against an ordered mechanism and suggests a random mechanism. Fructose 6-phosphate and Pi were noncompetitive with respect to each other indicating the formation of a dead end complex. These product inhibition relationships are in accord with a Random Bi Bi mechanism.  相似文献   

10.
The aim of this work was to investigate the contribution of fructose 2,6-bisphosphate to the regulation of carbohydrate metabolism under phosphate stress. The study exploited heterotrophic tobacco callus lines expressing a modified mammalian 6-phosphofructo-2-kinase/fructose 2,6-bisphosphatase that increased the fructose 2,6-bisphosphate content of the tissue. The phosphate status of two transgenic and one untransformed cell line was perturbed by incubation with 2-deoxyglucose, a phosphate-sequestering agent, and by growth of callus on phosphate-depleted media. 31P-NMR spectroscopy confirmed that both treatments decreased cellular levels of inorganic phosphate and phosphorylated metabolites. Despite large decreases in the amounts of phosphate esters, UDPglucose and adenylates in response to phosphate deficiency, the fructose 2,6-bisphosphate content of each line was unaffected by 2-deoxyglucose and increased during growth on phosphate-limited media. Short-term treatment of callus with 2-deoxyglucose had only minor effects on the carbohydrate status of each line, whereas long-term phosphate deficiency caused an increase in starch and a decrease in soluble sugar content in both transgenic and control lines. There were no consistent differences between the three callus lines in metabolism of [U-14C]glucose in response to incubation with 2-deoxyglucose. In contrast, there was a decrease in partitioning of label into glycolytic products (particularly organic acids) in untransformed callus during growth on phosphate-depleted medium. This decrease was greatly attenuated in the transgenic lines with increased fructose 2,6-bisphosphate content. This suggests that the conversion of hexose phosphates to triose phosphates is constrained under phosphate-deficient conditions, and that this restriction can be relieved by activation of pyrophosphate:fructose-6-phosphate 1-phosphotransferase. However, since the transgenic and control lines did not differ in the extent to which the carbohydrate content changed in response to growth on phosphate-depleted media, it is concluded that an increase in flux through pyrophosphate:fructose-6-phosphate 1-phosphotransferase is not a major component of the metabolic response of heterotrophic tobacco cells to phosphate deficiency.  相似文献   

11.
Streptococcus lactis K1 has the capacity to grow on many sugars, including sucrose and lactose, in the presence of high levels (greater than 500 mM) of 2-deoxy-D-glucose. Initially, growth of the organism was transiently halted by the addition of comparatively low concentrations (less than 0.5 mM) of the glucose analog to the culture. Inhibition was coincident with (i) rapid accumulation of 2-deoxy-D-glucose 6-phosphate (ca. 120 mM) and preferential utilization of phosphoenolpyruvate via the mannose:phosphotransferase system, (ii) depletion of phosphorylated glycolytic intermediates, and (iii) a 60% reduction in intracellular ATP concentration. During the 5- to 10-min period of bacteriostasis the intracellular concentration of 2-deoxy-D-glucose 6-phosphate rapidly declined, and the concentrations of glycolytic intermediates were restored to near-normal levels. When growth resumed, the cell doubling time (Td) and the steady-state levels of 2-deoxy-D-glucose 6-phosphate maintained by the cells were dependent upon the medium concentration of 2-deoxy-D-glucose. Resistance of S. lactis K1 to the potentially toxic analog was a consequence of negative regulation of the mannose:phosphotransferase system by two independent mechanisms. The first, short-term response occurred immediately after the initial "overshoot" accumulation of 2-deoxy-D-glucose 6-phosphate, and this mechanism reduced the activity (fine control) of the mannose:phosphotransferase system. The second, long-term mechanism resulted in repression of synthesis (coarse control) of enzyme IImannose. The two regulatory mechanisms reduced the rate of 2-deoxy-D-glucose translocation via the mannose:phosphotransferase system and minimized the activity of the phosphoenolpyruvate-dependent futile cycle of the glucose analog (J. Thompson and B. M. Chassy, J. Bacteriol. 151:1454-1465, 1982). Phosphoenolpyruvate was thus conserved for transport of the growth sugar and for generation of ATP required for biosynthetic and work functions of the growing cell.  相似文献   

12.
Growth of galactose-adapted cells of Streptococcus lactis ML(3) in a medium containing a mixture of glucose, galactose, and lactose was characterized initially by the simultaneous metabolism of glucose and lactose. Galactose was not significantly utilized until the latter sugars had been exhausted from the medium. The addition of glucose or lactose to a culture of S. lactis ML(3) growing exponentially on galactose caused immediate inhibition of galactose utilization and an increase in growth rate, concomitant with the preferential metabolism of the added sugar. Under nongrowing conditions, cells of S. lactis ML(3) grown previously on galactose metabolized the three separate sugars equally rapidly. However, cells suspended in buffer containing a mixture of glucose plus galactose or lactose plus galactose again consumed glucose or lactose preferentially. The rate of galactose metabolism was reduced by approximately 95% in the presence of the inhibitory sugar, but the maximum rate of metabolism was resumed upon exhaustion of glucose or lactose from the system. When presented with a mixture of glucose and lactose, the resting cells metabolized both sugars simultaneously. Lactose, glucose, and a non-metabolizable glucose analog (2-deoxy-d-glucose) prevented the phosphoenolpyruvate-dependent uptake of thiomethyl-beta-d-galactopyranoside (TMG), but the accumulation of TMG, like galactose metabolism, commenced immediately upon exhaustion of the metabolizable sugars from the medium. Growth of galactose-adapted cells of the lactose-defective variant S. lactis 7962 in the triple-sugar medium was characterized by the sequential metabolism of glucose, galactose, and lactose. Growth of S. lactis ML(3) and 7962 in the triple-sugar medium occurred without apparent diauxie, and for each strain the patterns of sequential sugar metabolism under growing and nongrowing conditions were identical. Fine control of the activities of preexisting enzyme systems by catabolite inhibition may afford a satisfactory explanation for the observed sequential utilization of sugars by these two organisms.  相似文献   

13.
Membranes of Streptococcus lactis were solubilized with 1.1% octyl-beta-D-glucopyranoside in the presence of 0.37% acetone/ether-washed phospholipid from several sources. After adding excess Escherichia coli phospholipid as bath-sonicated liposomes, phosphate:sugar phosphate antiport was reconstituted in proteoliposomes by a 25-fold dilution in 0.1 M KPi (pH 7). Assays of 32Pi:Pi exchange showed that antiport was subject to an inactivation which varied in severity according to the lipid present at solubilization. Recovery of Pi-linked exchange was improved by the presence of 10-20% glycerol or other osmolyte during extraction. The osmolytes tested in this regard have included polyols (glycerol, erythritol, xylitol, sorbitol), sugars (glucose, trehalose), and two amino acids (glycine, proline). Each gave 10--20-fold increased recoveries of 32Pi:Pi antiport compared to controls using only detergent and lipid; these precautions were not required for the efficient reconstitution of F0F1-ATPase. Antiport in the artificial system was studied most carefully when glycerol was the stabilizing additive. For that case, the Kt values for Pi or 2-deoxyglucose 6-phosphate transport (275 and 25 microM, respectively) were the same as in native membranes. Maximal rates of Pi and 2-deoxyglucose 6-phosphate transport (200 and 42 nmol/min/mg of protein, respectively) and the turnover number for Pi exchange (25--50/s) suggested that antiporters were recovered without loss of activity. We conclude that the quantitative aspects of bacterial anion exchange are amenable to study in an artificial system, and that the use of osmolytes as general stabilants can be a valuable adjunct to current techniques for reconstitution of integral membrane transport proteins.  相似文献   

14.
Membrane protein solubilized by octyl-beta-D-glucopyranoside in the presence of dispersed phospholipid was incubated with bath-sonicated liposomes and additional detergent. The proteoliposomes formed on dilution showed transport and exchange properties consistent with a reconstitution of phosphate:sugar 6-phosphate antiport. Thus, phosphate self-exchange was found only when protein from induced cells was used; this exchange was blocked by a sugar 6-phosphate, not by a sugar 1-phosphate; and proteoliposomes supported an accumulation of 2-deoxyglucose 6-phosphate with no added source of energy. Solubilization and reconstitution of protein was most effective when performed in the presence of gram-positive phospholipids.  相似文献   

15.
1. The isolation and properties of a mutant of Escherichia coli devoid of fructose 1-phosphate kinase activity are described. 2. This mutant grew in media containing any one of a variety of substances, including hexoses, hexose 6-phosphates, sugar acids and glucogenic substrates, at rates not significantly different from those at which the parent organism grew on these substrates. However, only the parent grew on fructose or fructose 1-phosphate. 3. Fructose and fructose 1-phosphate inhibit the growth of the mutant, but not of its parent, on other carbon sources. 4. Even though not previously exposed to fructose, the mutant took up [(14)C]fructose rapidly but to only a small extent: [(14)C]fructose 1-phosphate was identified as the predominant labelled product. In contrast, the equally rapid but more extensive uptake of [(14)C]fructose by the parent organism required prior growth in the presence of fructose.  相似文献   

16.
Mutants of Escherichia coli devoid of the membrane-spanning proteins PtsG and PtsMP, which are components of the phosphoenolpyruvate-dependent phosphotransferase system (PTS) and which normally effect the transport into the cells of glucose and mannose, do not grow upon or take up either sugar. Pseudorevertants are described that take up, and grow upon, mannose at rates strongly dependent on the mannose concentration in the medium (apparent Km > 5 mM); such mutants do not grow upon glucose but are derepressed for the components of the fructose operon. Evidence is presented that mannose is now taken up via the fructose-PTS to form mannose 6-phosphate, which is further utilized for growth via fructose 6-phosphate and fructose 1,6-bisphosphate.  相似文献   

17.
The mannose selection system employs the phosphomannose isomerase (PMI) gene as selectable gene and mannose, converted to mannose-6-phosphate by endogenous hexokinase, as selective agent. The transgenic PMI-expressing cells have acquired the ability to convert mannose-6-phosphate to fructose-6-phosphate, while the non-transgenic cells accumulate mannose-6-phosphate with a concomitant consumption of the intracellular pools of phosphate and ATP. Thus, certain steps of mannose selection depend on the cells’ own metabolism which may be affected by a number of factors, some of which are studied here using Agrobacterium tumefaciens-mediated gene transfer to sugar beet cotyledonary explants. Four frequently employed saccharides (sucrose, glucose, fructose, and maltose) were tested at various concentrations and were found to interact strongly with the phytotoxic effect of mannose, glucose being able to counteract nearly 100% of an almost complete mannose-induced growth inhibition. Sucrose, maltose, and fructose also alleviated significantly the mannose-induced growth inhibition, but were 4-, 5-, and 7-fold less potent than glucose, respectively (calculated as hexose equivalents). The transformation frequencies were also dependent on the nature and concentration of the added carbohydrates, but in this respect sucrose resulted in the highest transformation frequencies, about 1.0%, while glucose and fructose gave significantly lower frequencies. The selection efficiencies were highest in the presence of maltose where no non-transgenic escapes were found over a range of concentrations. The effect of the light intensity was also investigated and the transformation frequencies were positively correlated to light intensity, although the relative impact of light on growth in the presence of mannose appeared not to be dependent on the mannose concentration. Additional phosphate in the selection media had a strong positive effect on the transformation frequencies, suggesting phosphate limitation during selection. The mannose selection system was found to be relatively genotype-independent, provided a slight optimization of the mannose concentrations during selection. Analysis of F1-offspring showed that all studied primary transformants resulted in PMI-expressing plantlets and that the segregational patterns were in accordance with expectations in at least 50% of the transformants, confirming the stable and active inheritance of the PMI-gene.  相似文献   

18.
Resting cells of Staphylococcus aureus displayed a phosphate (Pi) exchange that was induced by growth with glucose 6-phosphate (G6P) or sn-glycerol 3-phosphate (G3P). Pi-loaded membrane vesicles from these cells accumulated 32Pi, 2-deoxyglucose 6-phosphate (2DG6P) or G3P by an electroneutral exchange that required no external source of energy. On the other hand, when vesicles were loaded with morpholinopropane sulfonic acid (MOPS), only transport of 32Pi (and L-histidine) was observed, and in that case transport depended on addition of an oxidizable substrate (DL-lactate). In such MOPS-loaded vesicles, accumulation of the organic phosphates, 2DG6P and G3P, could not be observed until vesicles were preincubated with both Pi and DL-lactate to establish an internal pool of Pi. This trans effect demonstrates that movement of 2DG6P or G3P is based on an antiport (exchange) with internal Pi. Reconstitution of membrane protein allowed a quantitative analysis of Pi-linked exchange. Pi-loaded proteoliposomes and membrane vesicles had comparable activities for the homologous 32Pi: Pi exchange (Kt's of 2.2 and 1.4 mM; Vmax's of 180 and 83 nmol Pi/min per mg protein), indicating that the exchange reaction was recovered intact in the artificial system. Other work showed that heterologous exchange from either G6P- or G3P-grown cells had a preference for 2DG6P (Kt = 27 microM) over G3P (Kt = 1.3 mM) and Pi (Kt = 2.2 mM), suggesting that the same antiporter was induced in both cases. We conclude that 32Pi: Pi exchange exhibited by resting cells reflects operation of an antiporter with high specificity for sugar 6-phosphate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Summary Mutual correction of co-cultivated fibroblasts from patients with Hunter's and Hurler's syndrome could be inhibited by either fructose 1-phosphate or mannose 6-phosphate. In the presence of fructose 1-phosphate a 50% mixture of fibroblasts from a patient with Hunter's syndrome and a normal homozygous individual showed an increased35S-sulphate incorporation into acid mucopolysaccharides. When fibroblast cultures from one obligate and two possible carriers of Hunter's syndrome were tested for35S-sulphate incorporation, the cultures showed either twice the normal35S-sulphate incorporation into acid mucopolysaccharides in the presence of fructose 1-phosphate or an abnormally high incorporation in the presence as well as in the absence of the sugar phosphate.  相似文献   

20.
All of the lactic streptococci examined except Streptococcus lactis ML8 fermented galactose to lactate, formate, acetate, and ethanol. The levels of pyruvate-formate lyase and lactate dehydrogenase were elevated and reduced, respectively, in galactose-grown cells compared with glucose- or lactose-grown cells. Reduced intracellular levels of both the lactate dehydrogenase activator (fructose, 1,6-diphosphate) and pyruvate-formate lyase inhibitors (triose phosphates) appeared to be the main factors involved in the diversion of lactate to the other products. S. lactis ML8 produced only lactate from galactose, apparently due to the maintenance of high intracellular levels of fructose 1,6-diphosphate and triose phosphates. The growth rates of all 10 Streptococcus cremoris strains examined decreased markedly with galactose concentrations below about 30 mM. This effect appeared to be correlated with uptake predominantly by the low-affinity galactose phosphotransferase system and initial metabolism via the D-tagatose 6-phosphate pathway. In contrast, with four of the five S. lactis strains examined, galactose uptake and initial metabolism involved more extensive use of the high-affinity galactose permease and Leloir pathway. With these strains the relative flux of galactose through the alternate pathways would depend on the exogenous galactose concentration.  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