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.
Escherichia coli and Salmonella typhimurium preferentially utilize sugar substrates of the phosphoenol-pyruvate:glycose phosphotransferase system (PTS) when the growth medium also contains other sugars. This phenomenon, diauxic growth, is regulated by the crr gene, which encodes the PTS protein IIIGlc (Saffen, D.W., Presper, K.A., Doering, T.L., and Roseman, S. (1987) J. Biol. Chem. 16241-16253). We have proposed that non-PTS permeases are regulated by their interaction with IIIGlc, and in vitro studies from other laboratories have provided support for this model, but the in vivo effects of excess IIIGlc are not known. In the present studies, transformed cells that overproduced IIIGlc 2- and 10-fold, respectively, were constructed from a pts+ strain of E. coli and plasmids containing the crr gene. In the 2-fold overproducer, fermentation of, and growth on the non-PTS carbohydrates glycerol, lactose, maltose, and melibiose was generally more sensitive to the glucose analogue methyl-alpha-D-glucopyranoside than in a control strain containing normal levels of IIIGlc. In addition, inhibition of lactose permease activity by methyl-alpha-glucoside (inducer exclusion) was more effective in the 2-fold overproducer than in the control strain, particularly when the permease activity was high. The 10-fold IIIGlc overproducing strain had a requirement for the amino acids methionine, isoleucine, leucine, and valine that may or may not be related to the increased concentration of IIIGlc. Fermentation of non-PTS carbohydrates was also poor in the latter strain. Finally, lactose permease activity was 50% of that in control cells containing the same levels of beta-galactosidase, and the lactose permease activity in the IIIGlc overproducer was reduced to an extremely low level in the presence of methyl alpha-glucoside. Thus there is an inverse relationship between the cellular concentration of IIIGlc and the ability to metabolize non-PTS substrates. The results are consistent with the model where inducer exclusion is affected by a direct interaction between IIIGlc and a non-PTS transport system.  相似文献   

2.
The solution structure of the complex of enzyme IIA of the N,N'-diacetylchitobiose (Chb) transporter with the histidine phosphocarrier protein HPr has been solved by NMR. The IIA(Chb)-HPr complex completes the structure elucidation of representative cytoplasmic complexes for all four sugar branches of the bacterial phosphoryl transfer system (PTS). The active site His-89 of IIA(Chb) was mutated to Glu to mimic the phosphorylated state. IIA(Chb)(H89E) and HPr form a weak complex with a K(D) of ~0.7 mM. The interacting binding surfaces, concave for IIA(Chb) and convex for HPr, complement each other in terms of shape, residue type, and charge distribution, with predominantly hydrophobic residues, interspersed by some uncharged polar residues, located centrally, and polar and charged residues at the periphery. The active site histidine of HPr, His-15, is buried within the active site cleft of IIA(Chb) formed at the interface of two adjacent subunits of the IIA(Chb) trimer, thereby coming into close proximity with the active site residue, H89E, of IIA(Chb). A His89-P-His-15 pentacoordinate phosphoryl transition state can readily be modeled without necessitating any significant conformational changes, thereby facilitating rapid phosphoryl transfer. Comparison of the IIA(Chb)-HPr complex with the IIA(Chb)-IIB(Chb) complex, as well as with other cytoplasmic complexes of the PTS, highlights a unifying mechanism for recognition of structurally diverse partners. This involves generating similar binding surfaces from entirely different underlying structural elements, large interaction surfaces coupled with extensive redundancy, and side chain conformational plasticity to optimize diverse sets of intermolecular interactions.  相似文献   

3.
4.
In the present studies we report that membrane-associated proteases in Salmonella typhimurium and Escherichia coli catalyze limited proteolysis of IIIGlcSlow. We have previously reported (Meadow, N. D., and Roseman, S. (1982) J. Biol. Chem. 257, 14526-14537) the isolation of two electrophoretically distinguishable forms of IIIGlc, which is a phosphocarrier and regulatory protein of the phosphoenolpyruvate:glycose phosphotransferase system. The two species of IIIGlc were designated IIIGlcFast and IIIGlcSlow; IIIGlcSlow is 7 amino acid residues longer than IIIGlcFast at its NH2 terminus. The majority of the protease activity is located in the outer membrane fraction from both species of bacteria, with the cytoplasmic fraction being devoid of activity. The site of cleavage is at the Lys-Ser bond located at residues 7-8 of IIIGlcSlow. The enzyme is an endopeptidase which liberates the expected heptapeptide (Gly-Leu-Phe-Asp-Lys-Leu-Lys). Both the large fragment of the limited proteolytic reaction, IIIGlcFast, and the small fragment, the heptapeptide, are stable to further proteolysis by membranes for more than 17 h at 37 degrees C. The activity in E. coli membranes has an absolute requirement for divalent metal ion (Mg2+ or Ca2+) and is heat-resistant, whereas the activity in S. typhimurium membranes is stimulated by divalent metal ion and is heat-sensitive. These results suggest significant differences between the two enzymes. The physiological function of the limited proteolysis of IIIGlc is not known.  相似文献   

5.
The relationship of adenylate cyclase, inorganic orthophosphate and the proteins of the phosphoenolpyruvate:glycose phosphotransferase system (PTS) was studied. A strain deleted for the genes for Enzyme I and IIIglc of the PTS was transformed with plasmids expressing either Enzyme I and HPr, IIIglc or all three proteins. The fully reconstituted strain showed a Pi-dependent stimulation of adenylate cyclase activity; in contrast, the strain expressing only IIIglc showed a Pi-dependent inhibition of adenylate cyclase activity.  相似文献   

6.
A phosphoenolpyruvate: dihydroxyacetone phosphotransferase was induced in Escherichia coli grown on dihydroxyacetone as sole carbon source or in its presence. This is the first example of a triose which can be acted upon by the membrane complex to provide a central intermediate in glycolysis. The presence of this system explains the ability of a mutant, in which the ATP-dependent glycerol kinase is genetically replaced by a glycerol: NAD 2-oxidoreductase, to grow on glycerol.  相似文献   

7.
The properties of an N-acetyl-d-glucosamine-transport system have been studied by following the intracellular accumulation of methyl 2-acetamido-2-deoxy-alpha-d-[1-(14)C]glucoside by Escherichia coli. The same analogue was used to assay phosphoenolpyruvate phosphotransferase activity of toluene-treated cells. Transport and phosphorylation are induced by growth on d-glucosamine or N-acetyl-d-glucosamine. Mutants resistant to N-iodoacetyl-d-glucosamine are defective in the uptake and phosphorylation of the labelled glycoside.  相似文献   

8.
9.
The solution structure of trimeric Escherichia coli enzyme IIA(Chb) (34 kDa), a component of the N,N'-diacetylchitobiose/lactose branch of the phosphotransferase signal transduction system, has been determined by NMR spectroscopy. Backbone residual dipolar couplings were used to provide long range orientational restraints, and long range (|i - j| > or = 5 residues) nuclear Overhauser enhancement restraints were derived exclusively from samples in which at least one subunit was 15N/13C/2H/(Val-Leu-Ile)-methyl-protonated. Each subunit consists of a three-helix bundle. Hydrophobic residues lining helix 3 of each subunit are largely responsible for the formation of a parallel coiled-coil trimer. The active site histidines (His-89 from each subunit) are located in three symmetrically placed deep crevices located at the interface of two adjacent subunits (A and C, C and B, and B and A). Partially shielded from bulk solvent, structural modeling suggests that phosphorylated His-89 is stabilized by electrostatic interactions with the side chains of His-93 from the same subunit and Gln-91 from the adjacent subunit. Comparison with the x-ray structure of Lactobacillus lactis IIA(Lac) reveals some substantial structural differences, particularly in regard to helix 3, which exhibits a 40 degrees kink in IIA(Lac) versus a 7 degrees bend in IIA(Chb). This is associated with the presence of an unusually large (230-angstroms3) buried hydrophobic cavity at the trimer interface in IIA(Lac) that is reduced to only 45 angstroms3) in IIA(Chb).  相似文献   

10.
The first two reactions in the phosphotransfer sequence of bacterial phosphoenolpyruvate:glycose phosphotransferase systems are the autophosphorylation of Enzyme I by phosphoenolpyruvate followed by the transfer of the phospho group to the low-molecular weight protein, HPr. Transient state kinetic methods were used to estimate the second-order rate constants for both phosphotransfer reactions. These measurements support previous conclusions that only the dimer of Enzyme I, EI2, is autophosphorylated, and that the rate of formation of dimer is slow compared to the rate of its phosphorylation. The rate constants of the two autophosphorylation reactions of EI2 by PEP are 6.6 x 10(6) M(-1) s(-1), and differ from one another by a factor of less than 3. The rate constant for the transfer reaction between phospho-EI2 and HPr is unusually large for a covalent reaction between two proteins (220 x 10(6) M(-1) s(-1)), while the constant for the reverse reaction is 4.2 x 10(6) M(-1) s(-1). Using the previously reported equilibrium constant for the autophosphorylation reaction, 1.5, the overall equilibrium constant for phosphotransfer from PEP to HPr is 80, somewhat higher than that previously reported. The results also show that EI2 can phosphorylate multiple molecules of HPr without dissociating to a monomer (EI), and that EI can accept a phospho group from phospho-HPr. These results are directly applicable to predicting the rates of phosphoenolpyruvate phosphotransferase system sugar uptake in whole cells.  相似文献   

11.
12.
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   

13.
K G Peri  E B Waygood 《Biochemistry》1988,27(16):6054-6061
In Escherichia coli, N-acetylglucosamine (nag) metabolism is joined to glycolysis via three specific enzymes that are the products of the nag operon. The three genes of the operon, nagA, nagB, and nagE, were found to be carried by a colicin plasmid, pLC5-21, from a genomic library of E. coli [Clarke, L., & Carbon, J. (1976) Cell (Cambridge, Mass.) 9,91-99]. The nagE gene that codes for enzyme IIN-acetylglucosamine of the phosphoenolpyruvate:sugar phosphotransferase system (PTS) was sequenced. The nagE sequence is preceded by a catabolite gene activator protein binding site and ends in a putative rho-independent termination site. The amino acid sequence determined from this DNA sequence shows 44% homology to enzymes IIglucose and IIIglucose of the PTS. Enzyme IIN-acetylglucosamine, which has 648 amino acids and a molecular weight of 68,356, contains a histidine at residue 569 which is homologous to the active site of IIIglc. Sequence homologies with enzymes IIglucose, II beta-glucoside, and IIsucrose indicate that residues His-190, His-213, and His-295 of enzyme IInag are also conserved and that His-190 is probably the second active site histidine. Other sequence homologies among these enzymes II suggest that they contain several sequence transpositions. Preliminary models of the enzymes II are proposed.  相似文献   

14.
15.
The tertiary structure of the histidine-containing phosphocarrier protein (HPr) of the Escherichia coli phosphoenolpyruvate:sugar phosphotransferase system has been determined by x-ray diffraction at 2.8-A resolution. Initially, a partial structure was fitted to the multiple isomorphous replacement map and then least-squares refined by the Konnert/Hendrickson restrained parameter method (Konnert, J. H., and Hendrickson, W. A. (1980) Acta Crystallogr. A36, 344-350) and finally, a subsequent map was computed by use of the phase combination method of Read (Read, R. J. (1986) Acta Crystallogr. A42, 140-149). More of the protein structure was located in the latter map. The procedure of model building, least-squares refinement, and electron density map recalculation was repeated until the tertiary structure of HPr was obtained. The overall structure of HPr consists of four beta-strands, three helical regions, and four beta-turns. At the active center, the His15 imidazole interacts with one oxygen atom of the alpha-carboxyl C terminus of the polypeptide chain; the conserved Arg17 side chain interacts with the other oxygen atom of the alpha-carboxyl C terminus as well as with the side chain of Glu85. This is the first x-ray analysis of a protein of the phosphoenolpyruvate:sugar phosphotransferase system. Furthermore, this work represents a protein structure which has been solved by starting with a model that represented only one-third of the scattering matter.  相似文献   

16.
17.
The kinetic parameters in vitro of the components of the phosphoenolpyruvate:glycose phosphotransferase system (PTS) in enteric bacteria were collected. To address the issue of whether the behavior in vivo of the PTS can be understood in terms of these enzyme kinetics, a detailed kinetic model was constructed. Each overall phosphotransfer reaction was separated into two elementary reactions, the first entailing association of the phosphoryl donor and acceptor into a complex and the second entailing dissociation of the complex into dephosphorylated donor and phosphorylated acceptor. Literature data on the K(m) values and association constants of PTS proteins for their substrates, as well as equilibrium and rate constants for the overall phosphotransfer reactions, were related to the rate constants of the elementary steps in a set of equations; the rate constants could be calculated by solving these equations simultaneously. No kinetic parameters were fitted. As calculated by the model, the kinetic parameter values in vitro could describe experimental results in vivo when varying each of the PTS protein concentrations individually while keeping the other protein concentrations constant. Using the same kinetic constants, but adjusting the protein concentrations in the model to those present in cell-free extracts, the model could reproduce experiments in vitro analyzing the dependence of the flux on the total PTS protein concentration. For modeling conditions in vivo it was crucial that the PTS protein concentrations be implemented at their high in vivo values. The model suggests a new interpretation of results hitherto not understood; in vivo, the major fraction of the PTS proteins may exist as complexes with other PTS proteins or boundary metabolites, whereas in vitro, the fraction of complexed proteins is much smaller.  相似文献   

18.
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   

19.
The protein HPr is a low-molecular-weight phosphocarrier protein of the bacterial phosphoenolpyruvate:glycose phosphotransferase system. We have recently reported the complete primary amino acid sequence of HPr isolated from Salmonella typhimurium (Weigel, N., Powers, D.A., and Roseman, S. (1982) J. Biol. Chem. 257, 14499-14509). This sequence is incorrect at certain residues; the correct primary structure of the protein is presented in this report. The corrected structure generally agrees with the primary sequence predicted for HPr from Escherichia coli (based on the nucleotide sequence of the corresponding ptsH gene). The one apparent ambiguity is at the carboxyl terminus.  相似文献   

20.
Enzyme IIIGlc of the phosphoenolpyruvate: sugar phosphotransferase system (PTS) of Salmonella typhimurium can occur in two forms: phosphorylated and nonphosphorylated. Phosphorylated IIIGlc (P-IIIGlc) has a slightly lower mobility during sodium dodecyl sulphate/polyacrylamide gel electrophoresis than IIIGlc. In bacterial extracts both phosphoenolpyruvate (the physiological phosphoryl donor of the PTS) as well as ATP can phosphorylate IIIGlc. The ATP-catalyzed reaction is dependent on phosphoenolpyruvate synthase, however, and is due to prior conversion of ATP to phosphoenolpyruvate. The phosphoryl group of phosphorylated IIIGlc is hydrolysed after boiling in sodium dodecyl sulfate but phosphorylated IIIGlc can be discriminated from IIIGlc if treated with this detergent at room temperature. We have used the different mobilities of IIIGlc and P-IIIGlc to estimate the proportion of these two forms in intact cells. Wild-type cells contain predominantly P-IIIGlc in the absence of PTS sugars. In an S. typhimurium mutant containing a leaky ptsI17 mutation (0.1% enzyme I activity remaining) both forms of IIIGlc occur in approximately equal amounts. Addition of PTS sugars such as glucose results, both in wild-type and mutant, in a dephosphorylation of P-IIIGlc. This correlates well with the observed inhibition of non-PTS uptake systems by PTS sugars via nonphosphorylated IIIGlc.  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