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 0 毫秒
1.
A promoter-like mutation, ptsP160, has been identified which drastically reduces expression of the genes specifying two proteins, HPr and enzyme I, of the phosphoenolpyruvate:sugar phosphotransferase system (PTS) in Salmonella typhimurium. This mutation lies between trzA, a gene specifying susceptibility to 1,2,4-triazole, and ptsH, the structural gene for HPr. It leads to a loss of active transport of those sugars that require the PTS for entry into the cell. Pseudorevertants of strains carrying this promoter-like mutation have additional lesions very closely linked to ptsP160 by transduction analysis and are noninducible for HPr and enzyme I above a basal level. Presumably, strains carrying ptsP160 are defective in the normal induction mechanism for HPr and enzyme I, and the pseudorevertants derived from them result from second-site initiation signals within or near this promoter-like element. The induction of HPr and enzyme I above their noninduced levels apparently is not required for transport of at least one PTS sugar, methyl alpha-d-glucopyranoside, since this sugar is taken up by the pseudorevertants at the same rate as by the wild type. The existence of a promoter-like element governing the coordinate inducibility of both HPr and enzyme I suggests that ptsH and ptsI constitute an operon. Wild-type levels of a sugar-specific PTS protein, factor III, are synthesized in response to the crr(+) gene in both a ptsP160 strain and its pseudorevertants; this suggests that the crr(+) gene has its own promoter distinct from ptsP.  相似文献   

2.
The activity of adenylate cyclase of Escherichia coli measured in toluene-treated cells under standard conditions is subject to control by the phosphoenolpyruvate:sugar phosphotransferase system (PTS). Sugars such as glucose, which are transported by the PTS, will inhibit adenylate cyclase provided the PTS is functional. An analysis was made of the properties of E. coli strains carrying mutations in PTS proteins. Leaky mutants in the PTS protein HPr are similar to wild-type strains with respect to cAMp regulation; adenylate cyclase activity in toluene-treated cells and intracellular cAMP levels are in the normal range. Furthermore, adenylate cyclase in toluene-treated cells of leaky HPr mutants is inhibited by glucose. In contrast, mutations in the PTS protein Enzyme I result in abnormalities in cAMP regulation. Enzyme I mutants generally have low intracellular cAMP levels. Leaky Enzyme I mutants show an unusual phosphoenolpyruvate-dependent activation of adenylate cyclase that is not seen in Enzyme I+ revertants or in Enzyme I deletions. A leaky Enzyme I mutant exhibits changes in the temperature-activity profile for adenylate cyclase, indicating that adenylate cyclase activity is controlled by Enzyme I. Temperature-shift studies suggest a functional complex between adenylate cyclase and a regulator protein at 30 °C that can be reversibly dissociated at 40 °C. These studies further support the model for adenylate cyclase activation that involves phosphoenolpyruvate-dependent phosphorylation of a PTS protein complexed to adenylate cyclase.  相似文献   

3.
The inducible, mannitol-specific Enzyme II of the phosphoenolpyruvate:sugar phosphotransferase system has been purified approximately 230-fold from Escherichia coli membranes. The enzyme, initially solubilized with deoxycholate, was first subjected to hydrophobic chromatography on hexyl agarose and then purified by several ion exchange steps in the presence of the nonionic detergent, Lubrol PX. The purified protein appears homogeneous by several criteria and probably consists of a single kind of polypeptide chain with a molecular weight of 60,000 (+/- 5%). In addition to catalyzing phosphoenolpyruvate-dependent phosphorylation of mannitol in the presence of the soluble enzymes of the phosphotransferase system, the purified Enzyme II also catalyzes mannitol 1-phosphate:mannitol transphosphorylation in the absence of these components. A number of other physical and catalytic properties of the enzyme are described. The availability of a stable, homogeneous Enzyme II should be invaluable for studying the mechanism of sugar translocation and phosphorylation catalyzed by the bacterial phosphotransferase system.  相似文献   

4.
Studies on the reversion characteristics of Escherichia coli strains carrying various mutations in the pts region have led to the recognition of a mutation, suc-1, with a previously undescribed phenotype. Strains carrying the suc-1 mutation grow normally on most sources of carbon but are unable to utilize succinate effectively. The suc-1 mutation can be separated genetically from the tightly linked ptsI6 mutation. Reversion of suc-1 mutants for growth on succinate yields interesting classes of suppressor mutations.  相似文献   

5.
6.
Sugars transported by a bacterial phosphoenolpyruvate:sugar phosphotransferase system (PTS) require two soluble proteins: HPr, a low-molecular-weight phosphate-carrier protein, and enzyme I. The structural genes coding for HPr (ptsH) and Enzyme I (ptsI) are shown to be cotransducible in Salmonella typhimurium. The gene order of this region of the Salmonella chromosome is cysA-trzA-ptsH-ptsI...(crr). A method for the isolation of trzA-pts deletion is described. One class of pts deletions extends through ptsH and into ptsI; a second class includes both ptsH and ptsI and extends into or through the crr gene. The crr gene either codes for or regulates the synthesis of a third PTS protein (factor III) which is sugar-specific. A hypothesis is presented for a mechanism of deletion formation.  相似文献   

7.
Histidine-containing phosphocarrier protein (HPr) is common to all of the phosphoenolpyruvate:sugar phosphotransferase systems (PTS) in Escherichia coli and Salmonella typhimurium, except the fructose-specific PTS. Strains which lack HPr activity (ptsH) have been characterized in the past, and it has proved difficult to delineate between tight and leaky mutants. In this study four different parameters of ptsH strains were measured: in vitro sugar phosphorylation activity of the mutant HPr; detection of 32P-labeled P-HPr; ability of monoclonal antibodies to bind mutant HPr; and sensitivity of ptsH strains to fosfomycin. Tight ptsH strains could be defined; they were fosfomycin resistant and produced no HPr protein or completely inactive mutant HPr. All leaky ptsH strains were fosfomycin sensitive, usually produced normal amounts of mutant HPr protein, and had low but measurable activity, and HPr was detectable as a phosphoprotein. This indicates that the regulatory functions of the PTS require a very low level of HPr activity (about 1%). The antibodies used to detect mutant HPr in crude extracts were two monoclonal immunoglobulin G antibodies Jel42 and Jel44. Both antibodies, which have different pIs, inhibited PTS sugar phosphorylation assays, but the antibody-HPr complex could still be phosphorylated by enzyme I. Preliminary evidence suggests that the antibodies bind to two different epitopes which are in part located in a beta-sheet structure.  相似文献   

8.
Wild-type glycerol kinase of Escherichia coli is inhibited by both nonphosphorylated enzyme IIIGlc of the phosphoenolpyruvate:carbohydrate phosphotransferase system and fructose 1,6-diphosphate. Mutant glycerol kinase, resistant to inhibition by fructose 1,6-diphosphate, was much less sensitive to inhibition by enzyme IIIGlc. The difference between the wild-type and mutant enzymes was even greater when inhibition was measured in the presence of both enzyme IIIGlc and fructose 1,6-diphosphate. The binding of enzyme IIIGlc to glycerol kinase required the presence of the substrate glycerol.  相似文献   

9.
Purified IIIGlc of the phosphoenolpyruvate:sugar phosphotransferase system of Salmonella typhimurium inhibits glycerol kinase. Phosphorylation of IIIGlc via phosphoenolpyruvate, enzyme I, and HPr abolishes this inhibition. The glycerol facilitator is not inhibited by IIIGlc. It is proposed that regulation of glycerol metabolism by the phosphoenolpyruvate:sugar phosphotransferase system is at the level of glycerol kinase.  相似文献   

10.
HPr is a protein of the bacterial phosphoenolpyruvate:sugar phosphotransferase transport system (PTS). In Gram-positive bacteria, HPr can be phosphorylated on Ser(46) by HPr(Ser) kinase/phosphorylase (HPrK/P) and on His(15) by enzyme I (EI) of the PTS. In vitro studies have shown that phosphorylation on one residue greatly inhibits the second phosphorylation. However, streptococci contain significant amounts of HPr(Ser-P)(His approximately P) during exponential growth, and recent studies suggest that phosphorylation of HPr(Ser-P) by EI is involved in the recycling of HPr(Ser-P)(His approximately P). We report in this paper a study on the phosphorylation of Streptococcus salivarius HPr, HPr(Ser-P), and HPr(S46D) by EI. Our results indicate that (i) the specificity constant (k(cat)/K(m)) of EI for HPr(Ser-P) at pH 7.9 was approximately 5000-fold smaller than that observed for HPr, (ii) no metabolic intermediates were able to stimulate HPr(Ser-P) phosphorylation, (iii) the rate of HPr phosphorylation decreased at pHs below 6.5, while that of HPr(Ser-P) increased and was almost 10-fold higher at pH 6.1 than at pH 7.9, (iv) HPr(S46D), a mutated HPr alleged to mimic HPr(Ser-P), was also phosphorylated more efficiently under acidic conditions, and, lastly, (v) phosphorylation of Bacillus subtilis HPr(Ser-P) by B. subtilis EI was also stimulated at acidic pH. Our results suggest that the high levels of HPr(Ser-P)(His approximately P) in streptococci result from the combination of two factors, a high physiological concentration of HPr(Ser-P) and stimulation of HPr(Ser-P) phosphorylation by EI at acidic pH, an intracellular condition that occurs in response to the acidification of the external medium during growth of the culture.  相似文献   

11.
A rapid, high-yield procedure has been developed for the purification of HPr from the Escherichia coli phosphoenolpyruvate dependent phosphotransferase system. During this procedure, the protein copurifies with a 2500-dalton homopolysaccharide which we have identified as alpha 1-6 glucan. The results of steady-state kinetic measurements of the phosphotransferase activity demonstrate that the polysaccharide works as an activator of the phosphotransferase system probably at the level of the HPr:P-E1 complex or the P-HPr:E11 complex.  相似文献   

12.
Histidine-containing protein, HPr, of the Escherichia coli phosphoenolpyruvate:sugar phosphotransferase system has an active site that involves His-15, which is phosphorylated to form a N delta 1-P-histidine, Arg-17, and the carboxy-terminal residue Glu-85. Mutant HPrs with alterations to the three C-terminal residues, Glu-85, Leu-84, and Glu-83, were produced by site-directed mutagenesis. The properties of these mutants were assessed by kinetic analysis of enzyme I, enzyme IImannose, enzyme IIN-acetylglucosamine, and enzyme IImannitol, and the phosphohydrolysis properties of the HPr mutants. The results show that it is the C-terminal alpha-carboxyl of Glu-85 that is involved in the active site, and this involvement may be restricted to the phosphoryl donor action of HPr. The contribution of this alpha-carboxyl group is modest as the deletion of Glu-85 resulted in the reduction of the enzyme II activity (kcat/Km) to about 33%. Removal of both Glu-85 and Leu-84 yields an HPr that is an impaired substrate of both the enzyme I and enzyme II reactions. Glu-83 appears to have no role in the active site.  相似文献   

13.
The bacterial PEP:sugar phosphotransferase system couples the phosphorylation and translocation of specific sugars across the membrane. The activity of the first protein in this pathway, enzyme I (EI), is regulated by a monomer-dimer equilibrium where a Mg(2+)-dependent autophosphorylation by PEP requires the dimer. Dimerization constants for dephospho- and phospho-EI and inactive mutants EI(H189E) and EI(H189A) (in which Glu or Ala is substituted for the active site His189) have been measured under a variety of conditions by sedimentation equilibrium at pH 7.5 and 4 and 20 degrees C. Concurrently, thermal unfolding of these forms of EI has been monitored by differential scanning calorimetry and by changes in the intrinsic tryptophanyl residue fluorescence. Phosphorylated EI and EI(H189E) have 10-fold increased dimerization constants [ approximately 2 x 10(6) (M monomer)(-1)] compared to those of dephospho-EI and EI(H189A) at 20 degrees C. Dimerization is strongly promoted by 1 mM PEP with 2 mM MgCl(2) [K(A)' > or = 10(8) M(-1) at 4 or 20 degrees C], as demonstrated with EI(H189A) which cannot undergo autophosphorylation. Together, 1 mM PEP and 2 mM Mg(2+) also markedly stabilize and couple the unfolding of C- and N-terminal domains of EI(H189A), increasing the transition temperature (T(m)) for unfolding the C-terminal domain by approximately 18 degrees C and that for the N-terminal domain by approximately 9 degrees C to T(max) congruent with 63 degrees C, giving a value of K(D)' congruent with 3 microM PEP at 45 degrees C. PEP alone also promotes the dimerization of EI(H189A) but only increases T(m) approximately 5 degrees C for C-terminal domain unfolding without affecting N-terminal domain unfolding, giving an estimated value of K(D)' congruent with 0.2 mM for PEP dissociation in the absence of Mg(2+) at 45 degrees C. In contrast, the dimerization constant of phospho-EI at 20 degrees C is the same in the absence and presence of 5 mM PEP and 2 mM MgCl(2). Thus, the separation of substrate binding effects from those of phosphorylation by studies with the inactive EI(H189A) has shown that intracellular concentrations of PEP and Mg(2+) are important determinants of both the conformational stability and dimerization of dephospho-EI.  相似文献   

14.
Enzyme IIA(Glc), encoded by the crr gene of the phosphoenolpyruvate:sugar phosphotransferase system, plays an important role in regulating intermediary metabolism in Escherichia coli ("catabolite repression"). One function involves inhibition of inducible transport systems ("inducer exclusion"), and with lactose permease, a galactoside is required for unphosphorylated IIA(Glc) binding to cytoplasmic loops IV/V and VI/VII [Sondej, M., Sun, J. et al. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 3525-3530]. With inside-out membrane vesicles containing the permease, [(125)I]IIA(Glc) binding promoted by melibiose exhibits an affinity (K(D)(IIA)) of approximately 1 microM and a stoichiometry of one mole of IIA(Glc) per six moles of lactose permease. Both the quantity of [(125)I]IIA(Glc) bound and the sugar concentration required for half-maximal IIA(Glc) binding (K(0.5)(IIA)(sug)) was measured for eight permease substrates. Differences in maximal IIA(Glc) binding are observed, and the K(0.5)(IIA)(sug) does not correlate with the affinity of LacY for sugar. Furthermore, K(0.5)(IIA)(sug) does not correlate with sugar affinities for various permease mutants. IIA(Glc) does not bind to a mutant (Cys154 --> Gly), which is locked in an outwardly facing conformation, binds with increased stoichiometry to mutant Lys131 --> Cys, and binds only weakly to two other mutants which appear to be predominantly in either an outwardly or an inwardly facing conformation. When the latter two mutations are combined, sugar-dependent IIA(Glc) binding returns to near wild-type levels. The findings suggest that binding of various substrates to lactose permease results in a collection of unique conformations, each of which presents a specific surface toward the inner face of the membrane that can interact to varying degrees with IIA(Glc).  相似文献   

15.
The solution structure of the second protein-protein complex of the Escherichia coli phosphoenolpyruvate: sugar phosphotransferase system, that between histidine-containing phosphocarrier protein (HPr) and glucose-specific enzyme IIA(Glucose) (IIA(Glc)), has been determined by NMR spectroscopy, including the use of dipolar couplings to provide long-range orientational information and newly developed rigid body minimization and constrained/restrained simulated annealing methods. A protruding convex surface on HPr interacts with a complementary concave depression on IIA(Glc). Both binding surfaces comprise a central hydrophobic core region surrounded by a ring of polar and charged residues, positive for HPr and negative for IIA(Glc). Formation of the unphosphorylated complex, as well as the phosphorylated transition state, involves little or no change in the protein backbones, but there are conformational rearrangements of the interfacial side chains. Both HPr and IIA(Glc) recognize a variety of structurally diverse proteins. Comparisons with the structures of the enzyme I-HPr and IIA(Glc)-glycerol kinase complexes reveal how similar binding surfaces can be formed with underlying backbone scaffolds that are structurally dissimilar and highlight the role of redundancy and side chain conformational plasticity.  相似文献   

16.
In enteric bacteria, chromosomally encoded permeases specific for lactose, maltose, and melibiose are allosterically regulated by the glucose-specific enzyme IIA of the phosphotransferase system. We here demonstrate that the plasmid-encoded raffinose permease of enteric bacteria is similarly subject to this type of inhibition.  相似文献   

17.
18.
Proteins of the phosphoenolpyruvate:sugar phosphotransferase system (PTS) of Bacillus subtilis were overexpressed, purified to near homogeneity, and characterized. The proteins isolated include Enzyme I, HPr, the glucose-specific IIA domain of the glucose-specific Enzyme II (IIAglc), and the mannitol-specific IIA protein, IIAmtl. Site specific mutant proteins of IIAglc and HPr were also overexpressed and purified, and their properties were compared with those of the wild type proteins. These proteins and their phosphorylated derivatives were characterized with respect to their immunological cross-reactivities employing the Western blot technique and in terms of their migratory behavior during sodium dodecyl sulfate-gel electrophoresis, nondenaturing gel electrophoresis, and isoelectric focusing. The interactions between homologous and heterologous Enzymes I and HPrs, between homologous and heterologous HPrs and the IIAglc proteins, and between homologous and heterologous IIAglc proteins and IIBCscr of B. subtilis as well as IICBglc of Escherichia coli were defined and compared kinetically. The mutant HPrs and IIAglc proteins were also characterized kinetically as PTS phosphocarrier proteins and/or as inhibitors of the phosphotransferase reactions of the PTS. These studies revealed that complexation of IIAglc with the mutant form of HPr in which serine 46 was replaced by aspartate (S46D) did not increase the rate of phosphoryl transfer from phospho Enzyme I to S46D HPr more than when IIAmtl was complexed to S46D HPr. These findings do not support a role for HPr(Ser-P) in the preferential utilization of one PTS carbohydrate relative to another. Functional analyses in E. coli established that IIAglc of B. subtilis can replace IIAglc of E. coli with respect both to sugar transport and to regulation of non-PTS permeases, catabolic enzymes, and adenylate cyclase. Site-specific mutations in histidyl residues 68 and 83 (H68A and H83A) inactivated IIAglc of B. subtilis with respect to phosphoryl transfer and its various regulatory roles.  相似文献   

19.
We investigated the claim (J. Daniel, J. Bacteriol. 157:940-941, 1984) that nonphosphorylated enzyme IIIGlc of the phosphoenolpyruvate:sugar phosphotransferase system is required for full synthesis of bacterial cyclic AMP (cAMP). In crp strains of Salmonella typhimurium, cAMP synthesis by intact cells was regulated by the phosphorylation state of enzyme IIIGlc. Introduction of either a pstHI deletion mutation or a crr::Tn10 mutation resulted in a low level of cAMP synthesis. In contrast, crp strains containing a leaky pstI mutation exhibited a high level of cAMP synthesis which was inhibited by phosphotransferase system carbohydrates. From these results, we conclude that phosphorylated enzyme IIIGlc rather than nonphosphorylated enzyme IIIGlc is required for full cAMP synthesis.  相似文献   

20.
We present the cloning and sequencing of the ptsI gene, encoding enzyme I (EI) of the phosphoenolpyruvate (PEP): sugar phosphotransferase (PTS) transport system from Streptococcus salivarius. The ptsI gene corresponds to an open reading frame of 1731 nucleotides, which translates into a putative 577-amino acid (aa) protein with a M(r) of 62,948 and a pI of 4.49. The EI was produced in Escherichia coli under the control of its own promoter located immediately upstream of ptsI, a situation never previously reported for any other gene coding for an EI. The deduced aa sequence of the S. salivarius EI shows a high degree of similarity with the E. coli EI and the EI moiety of the multiphosphoryl transfer protein from Rhodobacter capsulatus. The S. salivarius EI also shares a highly conserved aa cluster with a non-PTS protein, the maize pyruvate:orthophosphate dikinase. The conserved cluster is located in a domain which is hypothesized to be the PEP-binding site.  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