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.
Specialized lambda-transducing phages that carry the Escherichia coli genes ptsH, ptsI, crr, cysM, and cysA have been isolated, and the genes were subcloned in plasmid pBR322. Subcloning and restriction mapping data gave the following clockwise order of genes located at about 52 min on the E. coli genetic map: lig, cysK, ptsH, ptsI, crr, cysM, cysA. The nucleotide sequences of ptsH, ptsI, and crr and the corresponding flanking regions have been determined. These genes encode three cytoplasmic proteins of the phosphoenol-pyruvate:glycose phosphotransferase system: HPr, Enzyme I, and IIIGlc, respectively. The deduced amino acid sequences are consistent with amino acid composition and Edman degradation analyses obtained with the purified proteins. The calculated subunit molecular weight values (9,109 for HPr, 63,489 for Enzyme I, and 18,099 for IIIGlc) also agree well with values obtained with the proteins. Results of gamma delta-transposon insertional studies provided definitive evidence that IIIGlc is the gene product of crr, and therefore that IIIGlc plays a critical role in regulating the metabolism and uptake of certain non-PTS sugars (see accompanying papers: Mitchell, W.J., Saffen, D.W., and Roseman, S. (1987) J. Biol. Chem. 16254-16260; Misko, T.P., Mitchell, W.J., Meadow, N.D., and Roseman, S. (1987) J. Biol. Chem. 16261-16266). The gamma delta transposon studies also suggest that crr is transcribed from an independent promoter located within the ptsI gene. Putative regulatory sequence features include a catabolite gene activator protein-cAMP-binding site and two regions of 2-fold rotational symmetry adjacent to the potential promoter upstream from the HPr structural gene, several ribosome-binding sites, and a rho-independent RNA polymerase termination site downstream from crr. In addition, the ptsI gene contains two highly conserved direct repeats. The significance of these sequence features is discussed with respect to possible multiple forms of pts regulation.  相似文献   

3.
In this paper we report the amino acid sequence of pyruvate phosphate dikinase (PPDK) from Bacteroides symbiosus as determined from the nucleotide sequence of the PPDK gene. Comparison of the B. symbiosus PPDK amino acid sequence with that of the maize PPDK [Matsuoka, M., Ozeki, Y., Yamamoto, N., Hirano, H., Kamo-Murakami, Y., & Tanaka, Y. (1988) J. Biol. Chem. 263, 11080] revealed long stretches of homologous sequence (greater than 70% identity), which contributed to an overall sequence identity of 53%. The circular dichrosim spectra, hydropathy profiles, and calculated secondary structural elements of the two dikinases suggest that they may have very similar tertiary structures as well. A comparison made between the amino acid sequence of the maize and B. symbiosus dikinase with other known protein sequences revealed homology, concentrated in three stretches of sequences, to a mechanistically related enzyme, enzyme I of the Escherichia coli PEP: sugar phosphotransferase system [Saffen, D. W., Presper, K. A., Doering, T. L., Roseman, S. (1987) J. Biol. Chem. 262, 16241]. It is proposed that (i) these three stretches of sequence constitute the site for PEP binding and catalysis and a possible site for the regulation of enzymatic activity and (ii) the conserved sequences exist in a third mechanistically related enzyme, PEP synthase.  相似文献   

4.
An Escherichia coli strain which overproduces the lactose permease was used to investigate the mechanism of allosteric regulation of this permease and those specific for melibiose, glycerol, and maltose by the phosphoenolpyruvate-sugar phosphotransferase system (PTS). Thio-beta-digalactoside, a high affinity substrate of the lactose permease, released the glycerol and maltose permeases from inhibition by methyl-alpha-d-glucoside. Resumption of glycerol uptake occurred immediately upon addition of the galactoside. The effect was not observed in a strain which lacked or contained normal levels of the lactose permease, but growth of wild-type E. coli in the presence of isopropyl-beta-thiogalactoside plus cyclic AMP resulted in enhanced synthesis of the lactose permease so that galactosides relieved inhibition of glycerol uptake. Thiodigalactoside also relieved the inhibition of glycerol uptake caused by the presence of other PTS substrates such as fructose, mannitol, glucose, 2-deoxyglucose, and 5-thioglucose. Inhibition of adenylate cyclase activity by methyl-alpha-glucoside was also relieved by thiodigalactoside in E. coli T52RT provided that the lactose permease protein was induced to high levels. Cooperative binding of sugar and enzyme III(Glc) to the melibiose permease in Salmonella typhimurium was demonstrated, but no cooperativity was noted with the glycerol and maltose permeases. These results are consistent with a mechanism of PTS-mediated regulation of the lactose and melibiose permeases involving a fixed number of allosteric regulatory proteins (enzyme III(Glc)) which may be titrated by the increased number of substrate-activated permease proteins. This work suggests that the cooperativity in the binding of sugar substrate and enzyme III(Glc) to the permease, demonstrated previously in in vitro experiments, has mechanistic significance in vivo. It substantiates the conclusion that PTS-mediated regulation of non-PTS permease activities involves direct allosteric interaction between the permeases and enzyme III(Glc), the postulated regulatory protein of the PTS.  相似文献   

5.
We have previously reported that wild type strains of Escherichia coli grow on the chitin disaccharide N,N'-diacetylchitobiose, (GlcNAc)(2), as the sole source of carbon (Keyhani, N. O., and Roseman, S. (1997) Proc. Natl. Acad. Sci., U. S. A. 94, 14367-14371). A nonhydrolyzable analogue of (GlcNAc)(2,) methyl beta-N, N'-[(3)H]diacetylthiochitobioside ([(3)H]Me-TCB), was used to characterize the disaccharide transport process, which was found to be mediated by the phosphoenolpyruvate:glycose phosphotransferase system (PTS). Here and in the accompanying papers (Keyhani, N. O., Boudker, O., and Roseman, S. (2000) J. Biol. Chem. 275, 33091-33101; Keyhani, N. O., Bacia, K., and Roseman, S. (2000) J. Biol. Chem. 275, 33102-33109; Keyhani, N. O., Rodgers, M., Demeler, B., Hansen, J., and Roseman, S. (2000) J. Biol. Chem. 275, 33110-33115), we report that transport of [(3)H]Me-TCB and (GlcNAc)(2) involves a specific PTS Enzyme II complex, requires Enzyme I and HPr of the PTS, and results in the accumulation of the sugar derivative as a phosphate ester. The phosphoryl group is linked to the C-6 position of the GlcNAc residue at the nonreducing end of the disaccharide. The [(3)H]Me-TCB uptake system was induced only by (GlcNAc)(n), n = 2 or 3. The apparent K(m) of transport was 50-100 micrometer, and effective inhibitors of uptake included (GlcNAc)(n), n = 2 or 3, cellobiose, and other PTS sugars, i.e. glucose and GlcNAc. Presumably the PTS sugars inhibit by competing for PTS components. Kinetic properties of the transport system are described.  相似文献   

6.
7.
J J Ye  J W Neal  X Cui  J Reizer    M H Saier  Jr 《Journal of bacteriology》1994,176(12):3484-3492
Lactobacillus brevis takes up glucose and the nonmetabolizable glucose analog 2-deoxyglucose (2DG), as well as lactose and the nonmetabolizable lactose analoge thiomethyl beta-galactoside (TMG), via proton symport. Our earlier studies showed that TMG, previously accumulated in L. brevis cells via the lactose:H+ symporter, rapidly effluxes from L. brevis cells or vesicles upon addition of glucose and that glucose inhibits further accumulation of TMG. This regulation was shown to be mediated by a metabolite-activated protein kinase that phosphorylase serine 46 in the HPr protein. We have now analyzed the regulation of 2DG uptake and efflux and compared it with that of TMG. Uptake of 2DG was dependent on an energy source, effectively provided by intravesicular ATP or by extravesicular arginine which provides ATP via an ATP-generating system involving the arginine deiminase pathway. 2DG uptake into these vesicles was not inhibited, and preaccumulated 2DG did not efflux from them upon electroporation of fructose 1,6-diphosphate or gluconate 6-phosphate into the vesicles. Intravesicular but not extravesicular wild-type or H15A mutant HPr of Bacillus subtilis promoted inhibition (53 and 46%, respectively) of the permease in the presence of these metabolites. Counterflow experiments indicated that inhibition of 2DG uptake is due to the partial uncoupling of proton symport from sugar transport. Intravesicular S46A mutant HPr could not promote regulation of glucose permease activity when electroporated into the vesicles with or without the phosphorylated metabolites, but the S46D mutant protein promoted regulation, even in the absence of a metabolite. The Vmax but not the Km values for both TMG and 2DG uptake were affected. Uptake of the natural, metabolizable substrates of the lactose, glucose, mannose, and ribose permeases was inhibited by wild-type HPr in the presence of fructose 1,6-diphosphate or by S46D mutant HPr. These results establish that HPr serine phosphorylation by the ATP-dependent, metabolite-activated HPr kinase regulates glucose and lactose permease activities in L. brevis and suggest that other permeases may also be subject to this mode of regulation.  相似文献   

8.
Glucose is taken up in Bacillus subtilis via the phosphoenolpyruvate:glucose phosphotransferase system (glucose PTS). Two genes, orfG and ptsX, have been implied in the glucose-specific part of this PTS, encoding an Enzyme IIGlc and an Enzyme IIIGlc, respectively. We now show that the glucose permease consists of a single, membrane-bound, polypeptide with an apparent molecular weight of 80,000, encoded by a single gene which will be designated ptsG. The glucose permease contains domains that are 40-50% identical to the IIGlc and IIIGlc proteins of Escherichia coli. The B. subtilis IIIGlc domain can replace IIIGlc in E. coli crr mutants in supporting growth on glucose and transport of methyl alpha-glucoside. Mutations in the IIGlc and IIIGlc domains of the B. subtilis ptsG gene impaired growth on glucose and in some cases on sucrose. ptsG mutants lost all methyl alpha-glucoside transport but retained part of the glucose-transport capacity. Residual growth on glucose and transport of glucose in these ptsG mutants suggested that yet another uptake system for glucose existed, which is either another PT system or regulated by the PTS. The glucose PTS did not seem to be involved in the regulation of the uptake or metabolism of non-PTS compounds like glycerol. In contrast to ptsl mutants in members of the Enterobacteriaceae, the defective growth of B. subtilis ptsl mutants on glycerol was not restored by an insertion in the ptsG gene which eliminated IIGlc. Growth of B. subtilis ptsG mutants, lacking IIGlc, was not impaired on glycerol. From this we concluded that neither non-phosphorylated nor phosphorylated IIGlc was acting as an inhibitor or an activator, respectively, of glycerol uptake and metabolism.  相似文献   

9.
By subjecting the lac y gene of Escherichia coli to oligonucleotide-directed, site-specific mutagenesis, Cys148 in the lac permease has been replaced with a Gly residue [Trumble, W. R., Viitanen, P. V., Sarkar, H. K., Poonian, M. S., & Kaback, H. R. (1984) Biochem. Biophys. Res. Commun. 119, 860]. Recombinant plasmids bearing wild-type or mutated lac y were constructed and used to transform E. coli T184. Steady-state levels of lactose accumulation, the apparent Km for lactose under energized conditions, and the KD for p-nitrophenyl alpha-D-galactopyranoside are comparable in right-side-out vesicles containing wild-type or mutant permease. In contrast, the Vmax for lactose transport in vesicles containing mutant permease is significantly decreased. Although antibody binding studies reveal that vesicles from the mutant contain almost as much permease as wild-type vesicles, surprisingly only about one-fourth of the altered molecules bind p-nitrophenyl alpha-D-galactopyranoside with high affinity. Mutant permease is less sensitive to inactivation by N-ethylmaleimide, although the alkylating agent is still capable of completely inhibiting transport activity. Importantly, beta-galactosyl 1-thio-beta-D-galactopyranoside affords complete protection of wild-type permease against N-ethylmaleimide but has no protective effect whatsoever in the mutant. The rate of inactivation of wild-type and mutant permeases by N-ethylmaleimide is increased at alkaline pH and by the presence of a proton electrochemical gradient (interior negative and alkaline), and these phenomena are exaggerated in vesicles containing mutant permease. Finally, p-(chloromercuri)benzenesulfonate, which completely displaces bound p-nitrophenyl alpha-D-galactopyranoside from wild-type permease, does not affect binding in the mutant.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The role of the phosphoenolpyruvate-dependent phosphotransferase system (PTS) in the phenomenon of inducer exclusion was examined in whole cells of Salmonella typhimurium which carried the genes of the Escherichia coli lactose operon on an episome. In the presence of the PTS substrate methyl alpha-D-glucopyranoside, the extent of accumulation of the lactose analog methyl beta-D-thiogalactopyranoside was reduced. A strain carrying a mutation in the gene for Enzyme I was hypersensitive to the PTS effect, while a crr mutant strain was completely resistant. Influx, efflux, and exchange of galactosides via the lactose "permease" were inhibited by methyl alpha-glucoside. This inhibition occurred in the presence of metabolic energy poisons, and therefore does not involve either the generation of metabolic energy or energy-coupling to the lactose transport system. When the cellular content of the lactose permease was increased by induction with isopropyl beta-D-thiogalactopyranoside, cells gradually became less sensitive to inducer exclusion. The extent of inhibition of methyl beta-thiogalactoside accumulation by methyl alpha-glucoside was shown to be dependent on the relative cellular content of the PTS and lactose system. The data were consistent with an hypothesis involving partial inactivation of galactoside transport due to interaction between a component of the PTS and the lactose permease. By examination of the effects of the PTS and lactose uptake and melibiose permease-mediated uptake of methyl beta-thiogalactoside, it was further shown that the manner in which inducer exclusion is expressed is independent on the routes available to the non-PTS sugar for exit from the cell.  相似文献   

11.
E Diaz  D L Anton 《Biochemistry》1991,30(16):4078-4081
S-Adenosylmethionine decarboxylase from Escherichia coli is a member of a small class of enzymes that uses a pyruvoyl prosthetic group. The pyruvoyl group is proposed to form a Schiff base with the substrate and then act as an electron sink facilitating decarboxylation. We have previously shown that once every 6000-7000 turnovers the enzyme undergoes an inactivation that results in a transaminated pyruvoyl group and the formation of an acrolein-like species from the methionine moiety. The acrolein then covalently alkylates the enzyme [Anton, D. L., & Kutny, R. (1987) Biochemistry 26, 6444]. After reduction of the alkylated enzyme with NaBH4, a tryptic peptide with the sequence Ala-Asp-Ile-Glu-Val-Ser-Thr-[S-(3-hydroxypropyl)Cys]-Gly-Val-Ile-Ser-Pro - Leu-Lys was isolated. This corresponds to acrolein alkylation of a cysteine residue in the second tryptic peptide from the NH2 terminal of the alpha-subunit [Anton, D. L., & Kutny, R. (1987) J. Biol. Chem. 262, 2817-2822]. The modified residue derived is from Cys-140 of the proenzyme [Tabor, C. W., & Tabor, H. (1987) J. Biol. Chem. 262, 16037-16040] and lies in the only sequence conserved between rat liver and E. coli S-adenosylmethionine decarboxylase [Pajunen et al. (1988) J. Biol. Chem. 263, 17040-17049]. We suggest that the alkylated Cys residue could have a role in the catalytic mechanism.  相似文献   

12.
The N-terminal six-transmembrane domain (TM) bundle of lactose permease of Escherichia coli is uniformly inverted when assembled in membranes lacking phosphatidylethanolamine (PE). Inversion is dependent on the net charge of cytoplasmically exposed protein domains containing positive and negative residues, net charge of the membrane surface, and low hydrophobicity of TM VII acting as a molecular hinge between the two halves of lactose permease (Bogdanov, M., Xie, J., Heacock, P., and Dowhan, W. (2008) J. Cell Biol. 182, 925-935). Net neutral lipids suppress the membrane translocation potential of negatively charged amino acids, thus increasing the cytoplasmic retention potential of positively charged amino acids. Herein, TM organization of sucrose permease (CscB) and phenylalanine permease (PheP) as a function of membrane lipid composition was investigated to extend these principles to other proteins. For CscB, topological dependence on PE only becomes evident after a significant increase in the net negative charge of the cytoplasmic surface of the N-terminal TM bundle. High negative charge is required to overcome the thermodynamic block to inversion due to the high hydrophobicity of TM VII. Increasing the positive charge of the cytoplasmic surface of the N-terminal TM hairpin of PheP, which is misoriented in PE-lacking cells, favors native orientation in the absence of PE. PheP and CscB also display co-existing dual topologies dependent on changes in the charge balance between protein domains and the membrane lipids. Therefore, the topology of both permeases is dependent on PE. However, CscB topology is governed by thermodynamic balance between opposing lipid-dependent electrostatic and hydrophobic interactions.  相似文献   

13.
The 18.1-kDa protein IIIGlc from Escherichia coli acts as both a phosphocarrier protein in the phosphoenolpyruvate:glycose phosphotransferase system (PTS) and as a signal-transducing protein with respect to the uptake of non-PTS sugars. Phosphorylation of IIIGlc at the N epsilon (N3) position of His-90 was effected through a regeneration system that included MgCl2, DTT, excess PEP, and catalytic amounts of Enzyme I and HPr. NH, 15N, and 13C alpha signal assignments for P-IIIGlc were made through comparison of 15N-1H correlation spectra (HSQC) of uniformly 15N-labeled preparations of phosphorylated and unphosphorylated protein and through analysis of three-dimensional triple-resonance HNCA spectra of P-IIIGlc uniformly labeled with both 15N and 13C. Backbone and side-chain 1H and 13C beta signals were assigned using 3D heteronuclear HCCH-COSY and HCCH-TOCSY spectra of P-IIIGlc. Using this approach, the assignments were made without reference to nuclear Overhauser effect data or assumptions regarding protein structure. The majority of NH, 15N, H alpha, and 13C alpha chemical shifts measured for P-IIIGlc were identical to those obtained for the unphosphorylated protein [Pelton, J. G., Torchia, D. A., Meadow, N. D., Wong, C.-Y., & Roseman, S. (1991) Biochemistry 30, 10043]. Those signals that exhibited shifts corresponded to residues within four segments (1) Leu-87-Gly-100, (2) Val-36-Val-46, (3) His-75-Ser-78, and (4) Ala-131-Val-138. These four segments are in close proximity to the active site residues His-75 and His-90 in the unphosphorylated protein [Worthylake, D., Meadow, N. D., Roseman, S., Liao, D., Hertzberg, O., & Remington, S.J. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 10382], and the chemical shift data provide strong evidence that if any structural changes accompany phosphorylation, they are confined to residues in these four segments. This conclusion is confirmed by comparing NOEs observed in 3D 15N/13C NOESY-HMQC spectra of the two forms of the protein. No NOE differences are seen for residues having the same chemical shifts in IIIGlc and P-IIIGlc. Furthermore, with the exception of residues Ala-76, Asp-94, and Val-96, the NOEs of residues (in the four segments) which exhibited chemical shift differences also had the same NOEs in IIIGlc and P-IIIGlc. In the case of residues Ala-76, Asp-94, and Val-96, minor differences in NOEs, corresponding to interproton distances changes of less than 1.5 A, were observed.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
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   

15.
The major facilitator superfamily (MFS) of transport proteins, which includes the lactose permease of Escherichia coli, contains a conserved motif G-X-X-X-D/E-R/K-X-G-R/K-R/K in the loops that connect transmembrane segments 2 and 3, and transmembrane segments 8 and 9. In three previous studies (Jessen-Marshall, A.E., & Brooker, R.J. 1996. J. Biol. Chem. 271:1400–1404; Jessen-Marshall, A.E., Parker, N., & Brooker, R.J. 1997. J. Bacteriol. 179:2616–2622; and Pazdernik, N., Cain, S.M., & Brooker, R.J. 1997. J. Biol. Chem. 272:26110–26116), suppressor mutations at twenty different sites were identified which restore function to mutant permeases that have deleterious mutations in the conserved loop 2/3 or loop 8/9 motif. In the current study, several of these second-site suppressor mutations have been separated from the original mutation in the conserved motif. The loop 2/3 suppressors were then coupled to a loop 8/9 mutation (P280L) and the loop 8/9 suppressors were coupled to a loop 2/3 mutation (i.e., G64S) to determine if the suppressors could restore function only to a loop 2/3 mutation, a loop 8/9 mutation, or both. The single parent mutations changing the first position in loop 2/3 (i.e., G64S) and loop 8/9 (i.e., P280L) had less than 4% lactose transport activity. Interestingly, most of the suppressors were very inhibitory when separated from the parent mutation. Two suppressors, A50T and G370V, restored substantial transport activity when individually coupled to the mutation in loop 2/3 and also when coupled to the corresponding mutation in loop 8/9. In other words, these suppressors could alleviate a defect imposed by mutations in either half of the permease. From a kinetic analysis, these suppressors were shown to exert their effects by increasing the V max values for lactose transport compared with the single G64S and P280L strains. These results are discussed within the context of our model in which the two halves of the lactose permease interact at a rotationally symmetrical interface, and that lactose transport is mediated by conformational changes at the interface. Received: 18 November 1999/Revised: 11 April 2000  相似文献   

16.
The main mechanism causing catabolite repression by glucose and other carbon sources transported by the phosphotransferase system (PTS) in Escherichia coli involves dephosphorylation of enzyme IIAGlc as a result of transport and phosphorylation of PTS carbohydrates. Dephosphorylation of enzyme IIAGlc leads to 'inducer exclusion': inhibition of transport of a number of non-PTS carbon sources (e.g. lactose, glycerol), and reduced adenylate cyclase activity. In this paper, we show that the non-PTS carbon source glucose 6-phosphate can also cause inducer exclusion. Glucose 6-phosphate was shown to cause inhibition of transport of lactose and the non-metabolizable lactose analogue methyl-β- D -thiogalactoside (TMG). Inhibition was absent in mutants that lacked enzyme IIAGlc or were insensitive to inducer exclusion because enzyme IIAGlc could not bind to the lactose carrier. Furthermore, we showed that glucose 6-phosphate caused dephosphorylation of enzyme IIAGlc. In a mutant insensitive to enzyme IIAGlc-mediated inducer exclusion, catabolite repression by glucose 6-phosphate in lactose-induced cells was much weaker than that in the wild-type strain, showing that inducer exclusion is the most important mechanism contributing to catabolite repression in lactose-induced cells. We discuss an expanded model of enzyme IIAGlc-mediated catabolite repression which embodies repression by non- PTS carbon sources.  相似文献   

17.
B Erni 《Biochemistry》1986,25(2):305-312
The glucose-specific membrane permease (IIGlc) of the bacterial phosphoenolpyruvate-dependent phosphotransferase system (PTS) mediates active transport and concomitant phosphorylation of glucose. The purified permease has been phosphorylated in vitro and has been isolated (P-IIGlc). A phosphate to protein stoichiometry of between 0.6 and 0.8 has been measured. Phosphoryl transfer from P-IIGlc to glucose has been demonstrated. This process is, however, slow and accompanied by hydrolysis of the phosphoprotein unless IIIGlc, the cytoplasmic phosphoryl carrier protein specific to the glucose permease (IIGlc) of the PTS, is added. Addition of unphosphorylated IIIGlc resulted in rapid formation of glucose 6-phosphate with almost no hydrolysis of P-IIGlc accompanying the process. A complex of IIGlc and IIIGlc could be precipitated from bacterial cell lysates with monoclonal anti-IIGlc immunoglobulin. The molar ratio of IIGlc:IIIGlc in the immunoprecipitate was approximately 1:2. Analytical equilibrium centrifugation as well as chemical cross-linking showed that purified IIGlc itself is a dimer (106 kDa), consisting of two identical subunits. These results suggest that the functional glucose-specific permease complex comprises a membrane-spanning homodimer of IIGlc to which four molecules of IIIGlc are bound on the cytoplasmic face.  相似文献   

18.
In an extension of earlier studies on the Escherichia coli plasmid-encoded single-stranded DNA-binding proteins pIP71a SSB, F SSB and R64 SSB [Khamis, M. I., Casas-Finet, J. R., Maki, A. H., Ruvolo, P. P. & Chase, J. W. (1987) Biochemistry 26, 3347-3354; Casas-Finet, J. R., Khamis, M. I., Maki, A. H., Ruvolo, P. P. & Chase, J. W. (1987) J. Biol. Chem. 262, 8574-8593], we have investigated the binding of pIP231a SSB to natural and heavy-atom-derivatized single-stranded homopolynucleotides. Fluorimetric equilibrium binding isotherms indicate that pIP231a SSB has a greater solubility at low ionic strength than any other plasmid SSB protein investigated. Furthermore, its complex with mercurated poly(uridylic acid) [poly(Hg5U)] shows a greater resistance to disruption by salt than the other plasmid SSB complexes. Essentially complete binding of pIP231a SSB to poly(Hg5U) could be achieved, and time-resolved optically detected triplet-state magnetic resonance (ODMR) techniques could be applied to the complex. These methods allowed complete resolution of the three Trp chromophores of pIP231a SSB. Comparison of wavelength-selected ODMR results with those obtained for the poly(Hg5U) complex of a point-mutated chromosomal ssb gene product (Eco SSB) carrying substitutions of Phe for Trp [Khamis, M. I., Casas-Finet, J. R., Maki, A. H., Murphy, J. B. & Chase, J. W. (1987) J. Biol. Chem. 262, 10938-10945] confirm that Trp40 and Trp54 of pIP231a SSB are stacked in the complex, while Trp88 is not. This is the same distribution of stacked Trp residues found in Eco SSB. These results are confirmed further by specific effects observed on the ODMR signals of pIP231a SSB upon binding to poly(Br5U) and poly(dT), which are known to be caused by the stacking of Trp54 with nucleic acid bases.  相似文献   

19.
Enzyme I is the first protein of the phospho transfer sequence in the bacterial phosphoenolpyruvate:glycose phosphotransferase system. This protein exhibits a temperature-dependent monomer/dimer equilibrium. The nucleotide sequence of Escherichia coli ptsI indicates four -SH residues per subunit (Saffen, D. W., Presper, K. A., Doering, T. L., and Roseman, S. (1987) J. Biol. Chem. 262, 16241-16253). In the present experiments, the sulfhydryl groups of the E. coli enzyme were studied with various -SH-specific reagents. Titration of Enzyme I with 5,5'-dithiobis-2-nitrobenzoic acid also revealed four reacting -SH groups. The kinetics of the 5,5'-dithiobis-2-nitrobenzoic acid reaction with Enzyme I exhibit biphasic character, with pseudo-first order rate constants of 2.3 x 10(-2)/s and 2.3 x 10(-3)/s at pH 7.5, at room temperature. Fractional amplitudes associated with the rate constants were 25 +/- 5% for the fast and 75 +/- 5% for the slow rate. The "slow" rate was influenced by ligands that react with Enzyme I (the protein HPr, Mg2+, Mg2+ plus P-enolpyruvate), and also by temperature (at the temperature range where the monomer/dimer association occurs). The fractional ratio of the two rates remained at 1:3 under these conditions. Thus, under all conditions tested, two classes of -SH groups were detected, one reacting more rapidly than the other three -SH groups. Modification of the "fast" -SH group results in an active enzyme capable of forming dimer, whereas modification of the slow -SH groups results in inactive and monomeric Enzyme I. The enzyme was labeled with pyrene maleimide under conditions where only the more reactive sulfhydryl group was derivatized. Hydrolysis by trypsin followed by reverse-phase high performance liquid chromatography analysis of the peptide mixture resulted in only one fluorescent peak. This peak was not observed when the more reactive sulfhydryl residue was protected prior to pyrene maleimide labeling. Amino acid sequencing of the fluorescent peak indicated that the more reactive residue is the C-terminal amino acid residue, cysteine 575. The results provide a means for selectively labeling Enzyme I with a fluorophore at a single site while retaining full catalytic activity.  相似文献   

20.
H Tokuda  H R Kaback 《Biochemistry》1977,16(10):2130-2136
Membrane vesicles isolated from Salmonella typhimurium G-30 grown in the presence of melibiose catalyze methyl 1-thio-beta-D-galactopyranoside (TMG) transport in the presence of sodium or lithium, as shown initially with intact cells by Stock and Roseman (Stock, J., and Roseman, S. (1971), Biochem. Biophys. Res. Commun. 44, 132). TMG-dependent sodium uptake is also observed, but only when a potassium diffusion potential (interior negative) is induced across the vesicle membrane. Cation-dependent TMG accumulation varies with the electrochemical gradient of protons generated as a result of D-lactate oxidation, and the vesicles catalyze D-lactate-dependent sodium efflux in a manner which is consistent with the operation of a proton-sodium exchange mechanism. Although the stoichiometry between sodium and TMG appears to be 1:1 when transport is induced by a potassium diffusion potential, evidence is presented which indicates that the relationship may exceed unity under certain conditions. The results are explained in terms of a model in which TMG-sodium (lithium) symport is driven by an electrochemical gradient of protons which functions to maintain a low intravesicular sodium or lithium concentration through proton--sodium (lithium) antiport.  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