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.
2.
The protein, HPr, a necessary component of the phosphoenolpyruvate phosphotransferase system (PTS) in bacteria, was purified from Streptococcus salivarius by column chromatography. The purified preparation gave only one band when analyzed by sodium dodecylsulfate gel electrophoresis or by isoelectric focusing in polyacrylamide gel (pI = 4.85). However, electrophoresis in Tris-containing buffers under non-denaturing conditions revealed 2 bands that could be phosphorylated by PEP in the presence of enzyme I of the PTS or by ATP with the HPr kinase. Homogeneous preparations of these 2 forms could be obtained by preparative electrophoresis. Each preparation exhibited only 1 band when analyzed by electrophoresis under non-denaturing conditions, indicating that the doublet observed before preparative electrophoresis was not an electrophoretic artefact. The electrophoretic mobility of each protein was not modified following heat-treatment at 100 degrees C for 20 min or storage at -40 degrees C for several months. Both HPr proteins catalyzed in vitro the PEP-dependent phosphorylation of glucose, but at a rate slightly lower than that observed with a preparation of HPr containing both forms of the protein. Both forms were also able to transfer the phosphate group from PEP to the other specific PTS proteins known in S salivarius. Rabbit polyclonal antibodies directed against each form reacted with both proteins. The presence of the 2 forms of HPr was detected in fresh cellular extracts of S salivarius; however, their intracellular ratio varied according to growth conditions. A doublet was also found in many other streptococcal species tested (S mutans, S sobrinus, S sanguis, S thermophilus, S bovis, S rattus) and also in L lactis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
We have reported previously that the phosphoenolpyruvate:mannose phosphotransferase system (mannose PTS) of Streptococcus salivarius, consisting of an Enzyme II mannose (EIIman) and two forms of Enzyme III mannose (IIIman) with Mr values of 38,900 and 35,200, respectively, concomitantly transports and phosphorylates mannose, as well as glucose and fructose. In this paper, we report the presence, in S. salivarius, of alternative specific fructose and glucose PTSs encoded by inducible and cryptic genes, respectively. Protein phosphorylation experiments conducted with [32P]phosphoenolpyruvate have allowed us to identify by SDS-PAGE and autoradiography the EII fructose (EIIfru) (Mr 57,500) and the EII glucose (EIIglc) (Mr 58,700). No proteins corresponding to IIIfru or IIIglc could be detected. EIIfru phosphorylated fructose on the C-1 position rather than, as with the constitutive mannose PTS, on the C-6 position. Growth on fructose resulted in the induction of EIIfru as well as an increase of 1-phosphofructokinase activity. Nevertheless, the genes encoding these proteins were independently regulated. Studies carried out with spontaneous mutants lacking the low-molecular-mass form of IIIman (mutants A37, G29 and B31) showed that EIIfru was expressed in glucose-grown cells of strains G29 and B31, but not in strain A37, whereas the cryptic gene encoding EIIglc was activated in all three mutant strains. The results obtained with the mutants suggest that the three spontaneous mutants were not all mutated on the gene encoding IIIman although all of them lacked IIIman.  相似文献   

4.
In most streptococci, glucose is transported by the phosphoenolpyruvate (PEP):glucose/mannose phosphotransferase system (PTS) via HPr and IIAB(Man), two proteins involved in regulatory mechanisms. While most strains of Streptococcus thermophilus do not or poorly metabolize glucose, compelling evidence suggests that S. thermophilus possesses the genes that encode the glucose/mannose general and specific PTS proteins. The purposes of this study were to determine (i) whether these PTS genes are expressed, (ii) whether the PTS proteins encoded by these genes are able to transfer a phosphate group from PEP to glucose/mannose PTS substrates, and (iii) whether these proteins catalyze sugar transport. The pts operon is made up of the genes encoding HPr (ptsH) and enzyme I (EI) (ptsI), which are transcribed into a 0.6-kb ptsH mRNA and a 2.3-kb ptsHI mRNA. The specific glucose/mannose PTS proteins, IIAB(Man), IIC(Man), IID(Man), and the ManO protein, are encoded by manL, manM, manN, and manO, respectively, which make up the man operon. The man operon is transcribed into a single 3.5-kb mRNA. To assess the phosphotransfer competence of these PTS proteins, in vitro PEP-dependent phosphorylation experiments were conducted with purified HPr, EI, and IIAB(Man) as well as membrane fragments containing IIC(Man) and IID(Man). These PTS components efficiently transferred a phosphate group from PEP to glucose, mannose, 2-deoxyglucose, and (to a lesser extent) fructose, which are common streptococcal glucose/mannose PTS substrates. Whole cells were unable to catalyze the uptake of mannose and 2-deoxyglucose, demonstrating the inability of the S. thermophilus PTS proteins to operate as a proficient transport system. This inability to transport mannose and 2-deoxyglucose may be due to a defective IIC domain. We propose that in S. thermophilus, the general and specific glucose/mannose PTS proteins are not involved in glucose transport but might have regulatory functions associated with the phosphotransfer properties of HPr and IIAB(Man).  相似文献   

5.
Recently we reported the phosphoenolpyruvate (PEP)-dependent phosphorylation of a 55-kilodalton protein of Streptococcus faecalis catalyzed by enzyme I and histidine-containing protein (HPr) of the phosphotransferase system (J. Deutscher, FEMS Microbiol. Lett. 29:237-243, 1985). The purified 55-kilodalton protein was found to exhibit dihydroxyacetone kinase activity. Glycerol was six times more slowly phosphorylated than dihydroxyacetone. The Kms were found to be 0.7 mM for ATP, 0.45 mM for dihydroxyacetone, and 0.9 mM for glycerol. PEP-dependent phosphorylation of dihydroxyacetone kinase stimulated phosphorylation of both substrates about 10-fold. Fructose 1,6-diphosphate at concentrations higher than 2 mM inhibited the activity of phosphorylated and unphosphorylated dihydroxyacetone kinase in a noncompetitive manner. The rate of PEP-dependent phosphorylation of dihydroxyacetone kinase was about 200-fold slower than the phosphorylation rate of III proteins (also called enzyme III or factor III), which so far have been considered the only phosphoryl acceptors of histidyl-phosphorylated HPr. P-Dihydroxyacetone kinase was found to be able to transfer its phosphoryl group in a backward reaction to HPr. Following [32P]PEP-dependent phosphorylation and tryptic digestion of dihydroxyacetone kinase, we isolated a labeled peptide composed of 37 amino acids, as determined by amino acid analysis. The single histidyl residue of this peptide most likely carries the phosphoryl group in phosphorylated dihydroxyacetone kinase.  相似文献   

6.
The oral bacterium Streptococcus salivarius takes up lactose via a transporter called LacS that shares 95% identity with the LacS from Streptococcus thermophilus, a phylogenetically closely related organism. S. thermophilus releases galactose into the medium during growth on lactose. Expulsion of galactose is mediated via LacS and stimulated by phosphorylation of the transporter by HPr(His approximately P), a phosphocarrier of the phosphoenolpyruvate:sugar phosphotransferase transport system (PTS). Unlike S. thermophilus, S. salivarius grew on lactose without expelling galactose and took up galactose and lactose concomitantly when it is grown in a medium containing both sugars. Analysis of the C-terminal end of S. salivarius LacS revealed a IIA-like domain (IIA(LacS)) almost identical to the IIA domain of S. thermophilus LacS. Experiments performed with purified proteins showed that S. salivarius IIA(LacS) was reversibly phosphorylated on a histidine residue at position 552 not only by HPr(His approximately P) but also by HPr(Ser-P)(His approximately P), a doubly phosphorylated form of HPr present in large amounts in rapidly growing S. salivarius cells. Two other major S. salivarius PTS proteins, IIAB(L)(Man) and IIAB(H)(Man), were unable to phosphorylate IIA(LacS). The effect of LacS phosphorylation on growth was studied with strain G71, an S. salivarius enzyme I-negative mutant that cannot synthesize HPr(His approximately P) or HPr(Ser-P)(His approximately P). These results indicated that (i) the wild-type and mutant strains had identical generation times on lactose, (ii) neither strain expelled galactose during growth on lactose, (iii) both strains metabolized lactose and galactose concomitantly when grown in a medium containing both sugars, and (iv) the growth of the mutant was slightly reduced on galactose.  相似文献   

7.
Column chromatography of the Escherichia coli mannitol permease (mannitol-specific enzyme II of the phosphotransferase system) in the presence of deoxycholate has revealed that the active permease can exist in at least two association states with apparent molecular weights consistent with a monomer and a dimer. The monomeric conformation is favored by the presence of mannitol and by the phosphoenolpyruvate (PEP)-dependent phosphorylation of the protein. The dimer is stabilized by inorganic phosphate (Pi), which also stimulates phospho-exchange between mannitol and mannitol 1-phosphate (a partial reaction in the overall PEP-dependent phosphorylation of mannitol). Kinetic analysis of the phospho-exchange reaction revealed that Pi stimulates phospho-exchange by increasing the Vmax of the reaction. A kinetic model for mannitol permease function is presented involving both conformations of the permease. The monomer (or a less-stable conformation of the dimer) is hypothesized to be involved in the initial mannitol-binding and PEP-dependent phosphorylation steps, while the stably associated dimer is suggested to participate in later steps involving direct phosphotransfer between the permease, mannitol and mannitol 1-phosphate.  相似文献   

8.
9.
Sugar transport via the phosphoenolpyruvate (PEP) phosphotransferase system involves PEP-dependent phosphorylation of the general phosphotransferase system protein, HPr, at histidine 15. However, gram-positive bacteria can also carry out ATP-dependent phosphorylation of HPr at serine 46 by means of (Ser)HPr kinase. In this study, we demonstrate that (Ser)HPr kinase in crude preparations of Streptococcus mutans Ingbritt and Streptococcus salivarius ATCC 25975 is membrane associated, with pH optima of 7.0 and 7.5, respectively. The latter organism possessed 7- to 27-fold-higher activity than S. mutans NCTC 10449, GS-5, and Ingbritt strains. The enzyme in S. salivarius was activated by fructose-1,6-bisphosphate (FBP) twofold with 0.05 mM ATP, but this intermediate was slightly inhibitory with 1.0 mM ATP at FBP concentrations up to 10 mM. Similar inhibition was observed with the enzyme from S. mutans Ingbritt. A variety of other glycolytic intermediates had no effect on kinase activity under these conditions. The activity and regulation of (Ser)HPr kinase were assessed in vivo by monitoring P-(Ser)-HPr formation in steady-state cells of S. mutans Ingbritt grown in continuous culture with limiting glucose (10 and 50 mM) and with excess glucose (100 and 200 mM). All four forms of HPr [free HPr, P approximately (His)-HPr, P-(Ser)-HPr, and P approximately (His)-P-(Ser)-HPr] could be detected in the cells; however, significant differences in the intracellular levels of the forms were apparent during growth at different glucose concentrations. The total HPr pool increased with increasing concentrations of glucose in the medium, with significant increases in the P-(Ser)-HPr and P approximately HHis)-P-(Ser)-HPr concentrations. For example, while total PEP-dependent phosphorylation [P approximately(His)-HPr plus P approximately (His)-P-(Ser)-HPr] varied only from 21.5 to 52.5 microgram mg of cell protein (-1) in cells grown at the four glucose concentrations, the total ATP-dependent phosphorylation [P-(Ser)-HPr plus P approximately (His)-P-(Ser)-HPr] increased 12-fold from the 10 mM glucose-grown cells (9.1 microgram mg of cell protein (-1) to 106 and 105 microgram mg(-1) in the 100 and 200 mM glucose-grown cultures, respectively. (Ser)HPr kinase activity in membrane preparations of the cells varied little between the 10, 50, and 100 mM glucose-grown cells but increased threefold in the 200 mM glucose-grown cells. The intracellular levels of ATP, glucose-6-phosphate, and FBP increased with external glucose concentration, with the level of FBP being 3.8-fold higher for cells grown with 200 mM glucose than for those grown with 10 mM glucose. However, the variation in the intracellular levels of FBP, particularly between cells grown with 100 and 200 mM glucose, did not correlate with the extent of P-(Ser)-HPr formation, suggesting that the activity of (Ser)HPr kinase is not critically dependent on the availability of intracellular FBP.  相似文献   

10.
Glucose uptake by Listeria monocytogenes Scott A was inhibited by the bacteriocin pediocin JD and by the protonophore carbonyl cyanide m-chlorophenyhydrazone. Experiments with monensin, nigericin, chlorhexidine diacetate, dinitrophenol, and gramicidin, however, showed that glucose uptake could occur in the absence of a proton motive force. L. monocytogenes cell extracts phosphorylated glucose when phosphoenolpyruvate (PEP) was present in the assay mixture, and whole cells incubated with 2-deoxyglucose accumulated 2-deoxyglucose-6-phosphate, indicating the presence of a PEP-dependent phosphotransferase system in this organism. Glucose phosphorylation also occurred when ATP was present, suggesting that a proton motive force-mediated glucose transport system may also be present. We conclude that L. monocytogenes Scott A accumulates glucose by phosphotransferase and proton motive force-mediated systems, both of which are sensitive to pediocin JD.  相似文献   

11.
The phosphoenolpyruvate:sugar phosphotransferase system (PTS) is made of several proteins. Two of them are designated general proteins because they are required for the transport and phosphorylation of all sugars of the PTS. These two proteins are found in the soluble fraction of cellular extracts and are termed HPr and enzyme I (EI). We reported in this work the purification and the characterization of these two proteins from Streptococcus salivarius ATCC 25975. HPr was purified by DEAE-cellulose chromatography, molecular sieving on Ultrogel AcA44, and carboxymethylcellulose chromatography. Sodium dodecyl sulfate electrophoresis in the presence of urea revealed a single band with a molecular weight of 6700. The protein contained no tryptophan and had a pI of 4.8. The purification scheme of EI was as follows: DEAE-cellulose chromatography, hydroxylapatite chromatography, DEAE-Sephadex A-50 chromatography, preparative electrophoresis, and molecular sieving on Ultrogel AcA34. The five-step purification for EI produced a 199-fold purified preparation with a specific activity of 530 mumol of HPr phosphorylated per minute per milligram of protein at 37 degrees C. The fraction obtained after filtration on Ultrogel AcA34 gave one band (68 000) on sodium dodecyl sulfate - polyacrylamide gel electrophoresis. The molecular weight of the native enzyme determined by gel filtration at 4 degrees C was 135 000, suggesting that it was a dimer. Enzyme I had a pI of 4.2, a pH optimum of 6.7, a Km for HPr of about 27 microM, a Km for phosphoenolpyruvate of 0.48 mM, and kinetics that were consistent with a Ping-Pong mechanism. Evidence had been obtained which indicated that S. salivarius enzyme I was antigenically very similar to enzyme I from various strains of Streptococcus mutans, but not to the enzyme from Bacillus subtilis, Staphylococcus aureus, Streptococcus faecalis, and Escherichia coli.  相似文献   

12.
Spore coat proteins obtained by extraction with sodium dodecylsulfate/dithiothreitol from six Bacillus spores were compared by immunoblot analysis using antibodies to spore coat proteins from two strains of B. megaterium. Although the extract from spores of each strain had heterogenous proteins with various molecular weights, there were some bands which cross-reacted with specific antibodies from B. megaterium spores. Specific antibody to 48K protein from B. megaterium ATCC 12872 cross-reacted with 17K protein from B. megaterium ATCC 19213, 13K protein from B. cereus and 50K protein from B. subtilis 60015 and B. subtilis NRRL B558. Also, specific antibody to 22K protein from the same strain cross-reacted with 22K and 17K proteins from B. megaterium ATCC 19213 and 13K protein from B. cereus T. Specific antibody to 17K protein from B. megaterium ATCC 19213 reacted with 22K and 19K proteins in addition to 17K protein of own strain, and it was cross-reactive with 16K protein from B. megaterium ATCC 12872, 19K and 27K proteins from B. thiaminolyticus, 13K protein from B. cereus.  相似文献   

13.
Six species of ruminal bacteria were surveyed for the phosphoenolpyruvate (PEP)-dependent phosphorylation of glucose. Selenomonas ruminantium HD4, Streptococcus bovis JB1, and Megasphaera elsdenii B159 all showed significant activity, but Butyrivibrio fibrisolvens 49, Bacteroides succinogenes S85, and Bacteroides ruminicola B1(4) showed low rates of PEP-dependent phosphorylation and much higher rates in the presence of ATP. S. ruminantium HD4, S. bovis JB1, and M. elsdenii B159 also used PEP to phosphorylate the nonmetabolizable glucose analog 2-deoxy-D-glucose (2-DG). Rates of 2-DG phosphorylation with ATP were negligible for S. bovis JB1 and M. elsdenii B159, but toluene-treated cells of S. ruminantium HD4 phosphorylated 2-DG in the presence of ATP as well as PEP. Cell-free extracts of S. ruminantium HD4 used ATP but not PEP to phosphorylate glucose and 2-DG. Since PEP could serve as a phosphoryl donor in toluene-treated cells but not in cell-free extracts, there was evidence for membrane and hence phosphotransferase system involvement in the PEP-dependent activity. The ATP-dependent phosphorylating enzymes from S. ruminantium HD4 and S. bovis JB1 had molecular weights of approximately 48,000 and were not inhibited by glucose 6-phosphate. Based on these criteria, they were glucokinases rather than hexokinases. The S. ruminantium HD4 glucokinase was competitively inhibited by 2-DG and mannose, sugars that differ from glucose in the C-2 position. Since 2-DG was a competitive inhibitor of glucose, the same enzyme probably phosphorylates both sugars. The S. bovis JB1 glucokinase was not inhibited by either 2-DG or mannose and had a higher Km and Vmax for glucose.  相似文献   

14.
15.
Starved cells of Streptococcus lactis ML3 grown previously on lactose, galactose, or maltose were devoid of adenosine 5'-triphosphate contained only three glycolytic intermediates: 3-phosphoglycerate, 2-phosphoglycerate, and phosphoenolpyruvate (PEP). The three metabolites (total concentration, ca 40 mM) served as the intracellular PEP potential for sugar transport via PEP-dependent phosphotransferase systems. When accumulation of [14C]lactose by iodoacetate-inhibited starved cells was abolished within 1 s of commencement of transport, a phosphorylated disaccharide was identified by autoradiography. The compound was isolated by ion-exchange (borate) chromatography, and enzymatic analysis showed that the derivative was 6-phosphoryl-O-beta-D-galactopyranosyl (1 leads to 4')-alpha-D-glucopyranose (lactose 6-phosphate). After maximum lactose uptake (ca. 15 mM in 15 s) the cells were collected by membrane filtration and extracted with trichloroacetic acid. Neither free nor phosphorylated lactose was detected in cell extracts, but enzymatic analysis revealed high levels of galactose 6-phosphate and glucose 6-phosphate. The starved organisms rapidly accumulated glucose, 2-deoxy-D-glucose, methyl-beta-D-thiogalactopyranoside, and o-nitrophenyl-beta-D-galactopyranoside in phosphorylated form to intracellular concentrations of 32, 32, 42, and 38.5 mM, respectively. In contrast, maximum accumulation of lactose (ca. 15 mM) was only 40 to 50% that of the monosaccharides. From the stoichiometry of PEP-dependent lactose transport and the results of enzymatic analysis, it was concluded that (i) ca. 60% of the PEP potential was utilized via the lactose phosphotransferase system for phosphorylation of the galactosyl moiety of the disaccharide, and (ii) the residual potential (ca. 40%) was consumed during phosphorylation of the glucose moiety.  相似文献   

16.
Energy transduction and solute transport in streptococci   总被引:8,自引:1,他引:7  
Metabolic energy in lactic streptococci can be generated by substrate level phosphorylation and by efflux of end-products in symport with protons. During growth on lactose or glucose Streptococcus cremoris maintains a high proton motive force and phosphate potential. Both energy intermediates dissipate rapidly when the energy supply stops. In the initial phase of starvation the internal phosphoenolpyruvate (PEP) pool increases rapidly and this enables the organism for a prolonged period during starvation to accumulate the energy source via a PEP-dependent uptake system.  相似文献   

17.
Glucose uptake and accumulation by Clostridium beijerinckii BA101, a butanol hyperproducing mutant, were examined during various stages of growth. Glucose uptake in C. beijerinckii BA101 was repressed 20% by 2-deoxyglucose and 25% by mannose, while glucose uptake in C. beijerinckii 8052 was repressed 52 and 28% by these sugars, respectively. We confirmed the presence of a phosphoenolpyruvate (PEP)-dependent phosphotransferase system (PTS) associated with cell extracts of C. beijerinckii BA101 by glucose phosphorylation by PEP. The PTS activity associated with C. beijerinckii BA101 was 50% of that observed for C. beijerinckii 8052. C. beijerinckii BA101 also demonstrated lower PTS activity for fructose and glucitol. Glucose phosphorylation by cell extracts derived from both C. beijerinckii BA101 and 8052 was also dependent on the presence of ATP, a finding consistent with the presence of glucokinase activity in C. beijerinckii extracts. ATP-dependent glucose phosphorylation was predominant during the solventogenic stage, when PEP-dependent glucose phosphorylation was dramatically repressed. A nearly twofold-greater ATP-dependent phosphorylation rate was observed for solventogenic stage C. beijerinckii BA101 than for solventogenic stage C. beijerinckii 8052. These results suggest that C. beijerinckii BA101 is defective in PTS activity and that C. beijerinckii BA101 compensates for this defect with enhanced glucokinase activity, resulting in an ability to transport and utilize glucose during the solventogenic stage.  相似文献   

18.
Polyclonal antibodies were elicited against seven of the 33 different proteins of the large subunit of the chloroplast ribosome from Chlamydomonas reinhardtii. Three of these proteins are synthesized in the chloroplast and four are made in the cytoplasm and imported. In western blots, six of the seven antisera are monospecific for their respective large subunit ribosomal proteins, and none of these antisera cross-reacted with any chloroplast small subunit proteins from C. reinhardtii. Antisera to the three chloroplast-synthesized ribosomal proteins cross-reacted with specific Escherichia coli large subunit proteins of comparable charge and molecular weight. Only one of the four antisera to the chloroplast ribosomal proteins synthesized in the cytoplasm cross-reacted with an E. coli large subunit protein. None of the antisera cross-reacted with any E. coli small subunit proteins. On the assumption of a procaryotic, endosymbiotic origin for the chloroplast, those chloroplast ribosomal proteins still synthesized within the organelle appear to have retained more antigenic sites in common with E. coli ribosomal proteins than have those which are now the products of cytoplasmic protein synthesis. Antisera to this cytoplasmically synthesized group of chloroplast ribosomal proteins did not recognize any antigenic sites among C. reinhardtii cytoplasmic ribosomal proteins, suggesting that the genes for the cytoplasmically synthesized chloroplast ribosomal proteins either are not derived from the cytoplasmic ribosomal protein genes or have evolved to a point where no antigenic similarities remain.   相似文献   

19.
20.
The mannose permease of Escherichia coli is a component of the phosphotransferase system. It transports mannose and related hexoses by a mechanism that couples sugar transport with sugar phosphorylation. It is a complex consisting of two transmembrane subunits (II-PMan and II-MMan) and a hydrophilic subunit (IIIMan). IIIMan also exists in a soluble form as dimer in the cytoplasm. Each monomer of IIIMan consists of two structurally and functionally distinct domains which are linked by a flexible hinge of the sequence KAAPAPAAAAPKAAPTPAKP. Both domains are transiently phosphorylated. The NH2-terminal domain (P13) is phosphorylated at N-3 of His-10 by the cytoplasmic phosphorylcarrier protein phospho-HPr. The COOH-terminal domain (P20) is phosphorylated by P13 at N-1 of His-175. Phosphoryltransfer occurs not only between P13 and P20 on the same IIIMan subunit but also between isolated domains and between domains on different subunits of the dimer. In the presence of the IIMan subunits, the phosphoryl group is directly transferred from His-175 of P20 to the sugar substrates of the permease. The P13 domain contains the contact sites for dimerization of IIIMan. The P20 domain contains the contact sites for interaction with the IIMan subunits. By reconstructing the ptsL gene, the two domains were expressed as individual polypeptides and the length of the hinge between P13 and P20 was changed. The in vivo and in vitro activities of mutant IIIMan were little affected by these modifications. The hinge is highly sensitive to proteolytic cleavage in vitro and its specificity for proteases can be modified by introducing the appropriate specificity determinants.  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