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.
Transport of acidic amino acids in Bacillus subtilis is an electrogenic process in which L-glutamate or L-aspartate is symported with at least two protons. This is shown by studies of transport in membrane vesicles in which a proton motive force is generated by oxidation of ascorbate-phenazine methosulfate or by artificial ion gradients. An inwards-directed sodium gradient had no (stimulatory) effect on proton motive force-driven L-glutamate uptake. The transporter is specific for L-glutamate and L-aspartate. L-Glutamate transport is inhibited by beta-hydroxyaspartate and cysteic acid but not by alpha-methyl-glutamate. The gene encoding the L-glutamate transport protein of B. subtilis (gltPBsu) was cloned by complementation of Escherichia coli JC5412 for growth on glutamate as the sole source of carbon, energy, and nitrogen, and its nucleotide sequence was determined. Putative promoter, terminator, and ribosome binding site sequences were found in the flanking regions. UUG is most likely the start codon. gltPBsu encodes a polypeptide of 414 amino acid residues and is homologous to several proteins that transport glutamate and/or structurally related compounds such as aspartate, fumarate, malate, and succinate. Both sodium- and proton-coupled transporters belong to this family of dicarboxylate transporters. Hydropathy profiling and multiple alignment of the family of carboxylate transporters suggest that each of the proteins spans the cytoplasmic membrane 12 times with both the amino and carboxy termini on the inside.  相似文献   

2.
In order to clarify contradictory reports on the electrogenicity of sodium/L-glutamate cotransport, this cotransport was studied using brush-border membrane vesicles isolated from rabbit renal cortex. Beforehand, the claim that the symport of L-glutamate with Na+ is linked to simultaneous antiport with K+ has been confirmed by the demonstration that equilibrium exchange of L-glutamate is inhibited by potassium. Concerning the electrogenicity of the system, the following results are reported: net uptake of sodium-dependent L-glutamate uptake was stimulated when the transmembranal electrical potential difference was increased by replacing a sodium sulfate gradient by a sodium nitrate gradient. At 100 mM Na+ the 'relative electrogenicity' of the initial uptake in the presence of intravesicular potassium was 2-times higher than in its absence. At a sodium concentration of 20 mM, when overall uptake was reduced, the relative electrogenicity in the presence of K+ was even 3-fold higher than in K+-free media. The relative electrogenicity of sodium/D-glucose cotransport measured under the same experimental conditions was not affected by K+. These results are discussed in terms of a model where the apparent electrogenicity of a cotransport system is dependent on the extent to which the charge translocating step is rate limiting ('rate limitancy'). It is proposed that potassium antiport, while decreasing charge stoichiometry of Na+/glutamate transport, increases the relative rate limitancy of the transport step translocating three cations (probably two Na+, one H+) together with one glutamate. Thereby the positive electrogenicity of glutamate uptake increases, in complete contrast to what would be expected from simple considerations of charge stoichiometry.  相似文献   

3.
We have compared the recF genes from Escherichia coli K-12, Salmonella typhimurium, Pseudomonas putida, and Bacillus subtilis at the DNA and amino acid sequence levels. To do this we determined the complete nucleotide sequence of the recF gene from Salmonella typhimurium and we completed the nucleotide sequence of recF gene from Pseudomonas putida begun by Fujita et al. (1). We found that the RecF proteins encoded by these two genes contain respectively 92% and 38% amino acid identity with the E. coli RecF protein. Additionally, we have found that the S. typhimurium and P. putida recF genes will complement an E. coli recF mutant, but the recF gene from Bacillus subtilis [showing about 20% identity with E. coli (2)] will not. Amino acid sequence alignment of the four proteins identified four highly conserved regions. Two of these regions are part of a putative phosphate binding loop. In one region (position 36), we changed the lysine codon (which is essential for ATPase, GTPase and kinase activity in other proteins having this phosphate binding loop) to an arginine codon. We then tested this mutation (recF4101) on a multicopy plasmid for its ability to complement a recF chromosomal mutation and on the E. coli chromosome for its effect on sensitivity to UV irradiation. The strain with recF4101 on its chromosome is as sensitive as a null recF mutant strain. The strain with the plasmid-borne mutant allele is however more UV resistant than the null mutant strain. We conclude that lysine-36 and possibly a phosphate binding loop is essential for full recF activity. Lastly we made two chimeric recF genes by exchanging the amino terminal 48 amino acids of the S. typhimurium and E. coli recF genes. Both chimeras could complement E. coli chromosomal recF mutations.  相似文献   

4.
The driving forces for L-glutamate transport were determined in purified canalicular (cLPM) and basolateral (i.e. sinusoidal and lateral; blLPM) rat liver plasma membrane vesicles. Initial rates of L-glutamate uptake in cLPM vesicles were stimulated by a Na+ gradient (Na+o greater than Na+i), but not by a K+ gradient. Stimulation of L-glutamate uptake was specific for Na+, temperature sensitive, and independent of nonspecific binding. Sodium-dependent L-glutamate uptake into cLPM vesicles exhibited saturation kinetics with an apparent Km of 24 microM, and a Vmax of 21 pmol/mg X min at an extravesicular sodium concentration of 100 mM. Specific anionic amino acids inhibited L-[3H]glutamate uptake and accelerated the exchange diffusion of L-[3H]glutamate. An outwardly directed K+ gradient (K+i greater than K+o) further increased the Na+ gradient (Na+o greater than Na+i)-dependent uptake of L-glutamate in cLPM vesicles, resulting in a transient accumulation of L-glutamate above equilibrium values (overshoot). The K+ effect had an absolute requirement for Na+. In contrast, in blLPM the initial rates of L-glutamate uptake were only minimally stimulated by a Na+ gradient, an effect that could be accounted for by contamination of the blLPM vesicles with cLPM vesicles. These results indicate that hepatic Na+ gradient-dependent transport of L-glutamate occurs at the canalicular domain of the plasma membrane, whereas transport of L-glutamate across sinusoidal membranes results mainly from passive diffusion. These findings provide an explanation for the apparent discrepancy between the ability of various in vitro liver preparations to transport glutamate and suggest that a canalicular glutamate transport system may serve to reabsorb this amino acid from bile.  相似文献   

5.
Separate proteins for proton-linked transport of D-xylose, L-arabinose, D-galactose, L-rhamnose and L-fucose into Escherichia coli are being studied. By cloning and sequencing the appropriate genes, the amino acid sequences of proteins for D-xylose/H+ symport (XylE), L-arabinose/H+ symport (AraE), and part of the protein for D-galactose/H+ symport (GalP) have been determined. These are homologous, with at least 28% identical amino acid residues conserved in the aligned sequences, although their primary sequences are not similar to those of other E. coli transport proteins for lactose, melibiose, or D-glucose. However, they are equally homologous to the passive D-glucose transport proteins from yeast, rat brain, rat adipocytes, human erythrocytes, human liver, and a human hepatoma cell line. The substrate specificity of GalP from E. coli is similar to that of the mammalian glucose transporters. Furthermore, the activities of GalP, AraE and the mammalian glucose transporters are all inhibited by cytochalasin B and N-ethylmaleimide. Conserved residues in the aligned sequences of the bacterial and mammalian transporters are identified, and the possible roles of some in sugar binding, cation binding, cytochalasin binding, and reaction with N-ethylmaleimide are discussed. Each protein is independently predicted to form 12 hydrophobic, membrane-spanning alpha-helices with a central hydrophilic segment, also comprised of alpha-helix. This unifying structural model of the sugar transporters shares features with other ion-linked transport proteins for citrate or tetracycline.  相似文献   

6.
Amino acid transport in membrane vesicles of Bacillus stearothermophilus was studied. A relatively high concentration of sodium ions is needed for uptake of L-alanine (Kt = 1.0 mM) and L-leucine (Kt = 0.4 mM). In contrast, the Na(+)-H(+)-L-glutamate transport system has a high affinity for sodium ions (Kt less than 5.5 microM). Lithium ions, but no other cations tested, can replace sodium ions in neutral amino acid transport. The stimulatory effect of monensin on the steady-state accumulation level of these amino acids and the absence of transport in the presence of nonactin indicate that these amino acids are translocated by a Na+ symport mechanism. This is confirmed by the observation that an artificial delta psi and delta mu Na+/F but not a delta pH can act as a driving force for uptake. The transport system for L-alanine is rather specific. L-Serine, but not L-glycine or other amino acids tested, was found to be a competitive inhibitor of L-alanine uptake. On the other hand, the transport carrier for L-leucine also translocates the amino acids L-isoleucine and L-valine. The initial rates of L-glutamate and L-alanine uptake are strongly dependent on the medium pH. The uptake rates of both amino acids are highest at low external pH (5.5 to 6.0) and decline with increasing pH. The pH allosterically affects the L-glutamate and L-alanine transport systems. The maximal rate of L-glutamate uptake (Vmax) is independent of the external pH between pH 5.5 and 8.5, whereas the affinity constant (Kt) increases with increasing pH. A specific transport system for the basic amino acids L-lysine and L-arginine in the membrane vesicles has also been observed. Transport of these amino acids occurs most likely by a uniport mechanism.  相似文献   

7.
Bacillus subtilis has an alkaline phosphatase multigene family. Two members of this gene family, phoAIII and phoAIV, were cloned, taking advantage of in vitro constructed strains containing a plasmid insertion within one or the other of the structural genes. The DNA sequences of the two genes showed approximately 64% identity at the DNA level and 63% identity in the deduced primary amino acid sequences. The phoAIII and phoAIV genes code for predicted proteins of 47,149 and 45,935 Da, respectively. Comparison of the deduced primary amino acid sequence of the mature proteins with other sequenced alkaline phosphatases from Escherichia coli, yeast, and humans shows 25-30% identity. Based on the refined crystal structure of E. coli alkaline phosphatase, it appears that the active site and the core of the structure are retained in both Bacillus alkaline phosphatases. However, both proteins are truncated at the amino terminus compared with other mature alkaline phosphatases, three sizable surface loops of E. coli are deleted, and a minidomain is replaced with a larger domain in the model. Neither Bacillus alkaline phosphatase sequenced contains any cysteine residues, an amino acid implicated in intrachain disulfide bond formation in other alkaline phosphatases.  相似文献   

8.
E Shechter 《Biochimie》1986,68(3):357-365
Secondary active transport is defined as the transport of a solute in the direction of its increasing electrochemical potential coupled to the facilitated diffusion of a second solute (usually an ion) in the direction of its decreasing electrochemical potential. The coupling agents are membrane proteins (carriers), each of which catalyzes simultaneously the facilitated diffusion of the driving ion and the active transport of a given solute. The review starts with some considerations on the energetics followed by a presentation of the kinetics of secondary active transport. Examples of information which may be gained by such studies are discussed. In the second part, some examples of secondary transport are given; we also describe the characteristics of the corresponding carriers. The various transport systems presented are: the D-glucose/Na+ symport in brush-border membranes, the lactose/H+ symport in E. coli, the Na+/H+ antiport, the different transport systems in the inner mitochondrial membrane.  相似文献   

9.
In the presence of electrochemical energy, several branched-chain neutral and acidic amino acids were found to accumulate in membrane vesicles of Bacillus stearothermophilus. The membrane vesicles contained a stereo-specific transport system for the acidic amino acids L-glutamate and L-aspartate, which could not translocate their respective amines, L-glutamine and L-asparagine. The transport system was thermostable (Ti = 70 degrees C) and showed highest activities at elevated temperatures (60 to 65 degrees C). The membrane potential or pH gradient could act as the driving force for L-glutamate uptake, which indicated that the transport process of L-glutamate is electrogenic and that protons are involved in the translocation process. The electrogenic character implies that the anionic L-glutamate is cotransported with at least two monovalent cations. To determine the mechanistic stoichiometry of L-glutamate transport and the nature of the cotranslocated cations, the relationship between the components of the proton motive force and the chemical gradient of L-glutamate was investigated at different external pH values in the absence and presence of ionophores. In the presence of either a membrane potential or a pH gradient, the chemical gradient of L-glutamate was equivalent to that specific gradient at different pH values. These results cannot be explained by cotransport of L-glutamate with two protons, assuming thermodynamic equilibrium between the driving force for uptake and the chemical gradient of the substrate. To determine the character of the cotranslocated cations, L-glutamate uptake was monitored with artificial gradients. It was established that either the membrane potential, pH gradient, or chemical gradient of sodium ions could act as the driving force for L-glutamate uptake, which indicated that L-glutamate most likely is cotranslocated in symport with one proton and on sodium ion.  相似文献   

10.
The glutamyl-tRNA synthetase (GluRS) of Bacillus subtilis 168T aminoacylates with glutamate its homologous tRNA(Glu) and tRNA(Gln) in vivo and Escherichia coli tRNA(1Gln) in vitro (Lapointe, J., Duplain, L., and Proulx, M. (1986) J. Bacteriol. 165, 88-93). The gltX gene encoding this enzyme was cloned and sequenced. It encodes a protein of 483 amino acids with a Mr of 55,671. Alignment of the amino acid sequences of four bacterial GluRSs (from B. subtilis, Bacillus stearothermophilus, E. coli, and Rhizobium meliloti) gives 20% identity and reveals the presence of several short highly conserved motifs in the first two thirds of these proteins. Conserved motifs are found at corresponding positions in several other aminoacyl-tRNA synthetases. The only sequence similarity between the GluRSs of these Bacillus species and the E. coli glutaminyl-tRNA synthetase (GlnRS), which has no counterpart in the E. coli GluRS, is in a segment of 30 amino acids in the last third of these synthetases. In the three-dimensional structure of the E. coli tRNA(Gln).GlnRS.ATP complex, this conserved peptide is near the anticodon of tRNA(Gln) (Rould, M. A., Perona, J. J., S?ll, D., and Steitz, T. A. (1989) Science 246, 1135-1142), suggesting that this region is involved in the specific interactions between these enzymes and the anticodon regions of their tRNA substrates.  相似文献   

11.
Previous work with L-[3H]glutamate transport by lobster (Homarus americanus) hepatopancreatic brush border membrane vesicles (BBMV) indicated that the transport of this amino acid was stimulated by the presence of both Na+ and Cl- ions in the external medium, however, the specific catalytic or energetic role of each monovalent ion in amino acid transfer was not established (Ahearn and Clay (1987) J. Exp. Biol. 130, 175-191). The present study employs a variety of experimental treatments with this membrane preparation to clarify the nature of the ion dependency in the cotransport process. A zero-trans time course experiment using inwardly-directed transmembrane Na+ or Cl- gradients led to similar transient accumulations of the amino acid above equilibrium values in the presence of equilibrated concentrations of the respective counterions. The uptake overshoots observed in the presence of single ion gradients were significantly increased when gradients of both Na+ and Cl- were used simultaneously. When vesicles were pre-equilibrated with L-[3H]glutamate and either of the monovalent ions, an inwardly-directed gradient of each counterion led to the transient accumulation of additional labelled amino acid above its equilibrium concentration, indicating that either ion gradient was capable of energizing the net flow of L-glutamate. A cotransport stoichiometry of 1 Na+/1 Cl-/1 L-glutamate was established using the Static Head analysis where a balance of ion and amino acid driving forces were attained with a 7:1 Na+ or Cl- gradient (o greater than i) against a 7:1 L-glutamate gradient (i greater than o).  相似文献   

12.
pH dependence of the Coxiella burnetii glutamate transport system.   总被引:10,自引:2,他引:8  
The transport of glutamate, apparently a primary energy source for Coxiella burnetii, has been examined. C. burnetii is shown to possess a pH-dependent active transport system for L-glutamate with an apparent Kt of 61.1 microM and Vmax of 8.33 pmol/s per mg at pH 3.5. Both L-glutamine and L-asparagine competitively inhibited transport of glutamate, but D-glutamate, L-aspartate, L-glutamate-gamma-methyl ester, methionine sulfoximine, or alpha-ketoglutarate did not compete. This transport system is both temperature and energy dependent. Uptake of glutamate is highly sensitive to uncouplers of oxidative phosphorylation such as 2,4-dinitrophenol and carbonyl cyanide-m-chlorophenyl hydrazone that decrease the proton motive force across the cytoplasmic membrane. ATPase inhibitors such as dicyclohexylcarbodiimide or metabolic poisons such as KCN, NaF, or arsenite were much less effective as inhibitors of glutamate transport. Uptake of glutamate did not appear to be coupled to Na+ symport as in Escherichia coli since no monovalent cation requirement could be demonstrated. Instead, the Vmax of glutamate transport showed good correlation with the transmembrane pH gradient (delta pH). From these results, we propose that L-glutamate transport by C. burnetii is energized via a proton motive force.  相似文献   

13.
14.
Illumination of cell envelope vesicles from H. halobium causes the development of protonmotive force and energizes the uphill transport of glutamate. Although the uncoupler, p-trifluoromethoxycarbonyl cyanide phenylhydrazone (FCCP), and the membrane-permeant cation, triphenylmethylphosphonium (TPMP+), are inhibitory to the effect of light, the time course and kinetics of the production of the energized state for transport, and its rate of decay after illumination, are inconsistent with the idea that glutamate accumulation is driven directly by the protonmotive force. Similarities between the light-induced transport and the Na+-gradient-induced transport of glutamate in these vesicles suggest that the energized state for the amino acid uptake in both cases consists of a transmembrane Na+ gradient (Na+out/Na+in greater than 1). Rapid efflux of 22Na from the envelope vesicles is induced by illumination. FCCP and TPMP+ inhibit the light-induced efflux of Na+ but accelerate the post-illumination relaxation of the Na+ gradient created, suggesting electrogenic antiport of Na+ with another cation, or electrogenic symport with an anion. The light-induced protonmotive force in the H. halobium cell envelope vesicles is thus coupled to Na+ efflux and thereby indirectly to glutamate uptake as well.  相似文献   

15.
The glutamate transporters GltP(Ec) from Escherichia coli and GltP(Ph) from Pyrococcus horikoshii were overexpressed in E. coli and purified to homogeneity with a yield of 1-2 mg/L of culture. Single-particle analysis and electron microscopy indicate that GltP(Ph) is a trimer in detergent solution. Electron microscopy of negatively stained GltP(Ph) two-dimensional crystals shows that the transporter is a trimer also in the membrane. Gel filtration of GltP(Ec) indicates a reversible equilibrium of two oligomeric states in detergent solution that we identified as a trimer and hexamer by blue-native gel electrophoresis and cross-linking. The purified transporters were fully active upon reconstitution into liposomes, as demonstrated by the uptake of radioactively labeled L-aspartate or L-glutamate. L-aspartate/L-glutamate transport of GltP(Ec) involves the cotransport of protons and depends only on pH, whereas GltP(Ph) catalyzes L-glutamate transport with a cotransport of H+ or Na+. L-glutamate induces a fast transient current in GltP(Ph) proteoliposomes coupled to a solid supported membrane (SSM). We show that the electric signal depends on the concentration of Na+ or H+ outside the proteoliposomes and that GltP(Ph) does not require K+ inside the proteoliposomes. In addition, the electrical currents are inhibited by TBOA and HIP-B. The half-saturation concentration for activation of GltP(Ph) glutamate transport (K0.5(glut)) is 194 microM.  相似文献   

16.
Citrate transport via CitS of Klebsiella pneumoniae has been shown to depend on the presence of Na+. This transport system has been expressed in Escherichia coli, and uptake of citrate in E. coli membrane vesicles via this uptake system was found to be an electrogenic process, although the pH gradient is the main driving force for citrate uptake (M. E. van der Rest, R. M. Siewe, T. Abee, E. Schwartz, D. Oesterhelt, and W. N. Konings, J. Biol. Chem. 267:8971-8976, 1992). Analysis of the affinity constants for the different citrate species at different pH values of the medium indicates that H-citrate2- is the transported species. Since the electrical potential across the membrane is a driving force for citrate transport, this indicates that transport occurs in symport with at least three monovalent cations. Citrate efflux is stimulated by Na+ concentrations of up to 5 mM but inhibited by higher Na+ concentrations. Citrate exchange, however, is stimulated by all Na+ concentrations, indicating sequential events in which Na+ binds before citrate for translocation followed by a release of Na+ after release of citrate. CitS has, at pH 6.0 and in the presence of 5 mM citrate on both sides of the membrane, an apparent affinity (K(app)) for Na+ of 200 microM. The Na+/citrate stoichiometry was found to be 1. It is postulated that H-citrate2- is transported via CitS in symport with one Na+ and at least two H+ ions.  相似文献   

17.
l -Glutamate transport by the H+-glutamate and Na+-glutamate symport proteins of Escherichia coli K-12 (GltPEc and GltSEc, respectively) and the Na+-H+-glutamate symport proteins of Bacillus stearothermophilus (GltTBs) and Bacillus caldotenax (GltTBc) was studied in membrane vesicles derived from cells in which the proteins were either homologously or heterologously expressed. Substrate and inhibitor specificity studies indicate that GltPEc, GltTBs and GltTBc fall into the same group of transporters, whereas GltSEc is distinctly different from the others. Also, the cation specificity of GltSEc is different; GltSEc transported l -glutamate with (at least) two Na+, whereas GltPEc, GltTBs and GltTBc catalysed an electrogenic symport of l -glutamate with ≥two H+, i.e. when the proteins were expressed in E. coli Surprisingly studies in membrane vesicles of B. stearothermophilus and B. caldotenax indicated a Na+-H+- l -glutamate symport for both GltTBs and GltTBc. The Na+ dependency of the GltT transporters in the Bacillus strains increased with temperature. These observations suggest that the conformation of the transport proteins in the E. coli and the Bacillus membranes differs, which influences the coupling ion selectivity.  相似文献   

18.
A segment of Bacillus subtilis chromosomal DNA homologous to the Escherichia coli spc ribosomal protein operon was isolated using cloned E. coli rplE (L5) DNA as a hybridization probe. DNA sequence analysis of the B. subtilis cloned DNA indicated a high degree of conservation of spc operon ribosomal protein genes between B. subtilis and E. coli. This fragment contains DNA homologous to the promoter-proximal region of the spc operon, including coding sequences for ribosomal proteins L14, L24, L5, S14, and part of S8; the organization of B. subtilis genes in this region is identical to that found in E. coli. A region homologous to the E. coli L16, L29 and S17 genes, the last genes of the S10 operon, was located upstream from the gene for L14, the first gene in the spc operon. Although the ribosomal protein coding sequences showed 40-60% amino acid identity with E. coli sequences, we failed to find sequences which would form a structure resembling the E. coli target site for the S8 translational repressor, located near the beginning of the L5 coding region in E. coli, in this region or elsewhere in the B. subtilis spc DNA.  相似文献   

19.
Two chemotaxis-defective mutants of Pseudomonas aeruginosa, designated PC3 and PC4, were selected by the swarm plate method after N-methyl-N'-nitro-N-nitrosoguanidine mutagenesis. These mutants were not complemented by the P. aeruginosa cheY and cheZ genes, which had been previously cloned (Masduki et al., J. Bacteriol., 177, 948-952, 1995). DNA sequences downstream of the cheY and cheZ genes were able to complement PC3 but not PC4. Sequence analysis of a 9.7-kb region directly downstream of the cheZ gene found three chemotaxis genes, cheA, cheB, and cheW, and seven unknown open reading frames (ORFs). The predicted translation products of the cheA, cheB, and cheW genes showed 33, 36, and 31% amino acid identity with Escherichia coli CheA, CheB, and CheW, respectively. Two of the unknown ORFs, ORF1 and ORF2, encoded putative polypeptides that resembled Bacillus subtilis MotA (40% amino acid identity) and MotB (34% amino acid identity) proteins, respectively. Although P. aeruginosa was found to have proteins similar to the enteric chemotaxis proteins CheA, CheB, CheW, CheY, and CheZ, the gene encoding a CheR homologue did not reside in the chemotaxis gene cluster. The P. aeruginosa cheR gene could be cloned by phenotypic complementation of the PC4 mutant. This gene was located at least 1,800 kb away from the chemotaxis gene cluster and encoded a putative polypeptide that had 32% amino acid identity with E. coli CheR.  相似文献   

20.
A gene has been cloned from a DNA library from alkaliphilic Bacillus firmus OF4 that functionally complements a mutant strain of Escherichia coli, NM81, that carries a deletion for one of that strain's Na+/H+ antiporter genes (delta nhaA). The cloned alkaliphile gene restored to NM81 the ability to grow at pH 7.5 in the presence of 0.6 M NaCl and on 100 mM Li+ in the presence of melibiose, and concomitantly led to an increase in the membrane associated Na+/H+ antiport activity. The biologically active alkaliphile DNA was identified as an incomplete open reading frame, the sequence of which would encode a hydrophobic protein. The insert was used to isolate clones containing the complete open reading frame, which would be predicted to encode a protein with a molecular weight of 42,960 and multiple membrane spanning regions. When the open reading frame was expressed under the control of the T7 promoter, the gene product was localized in the membrane. Southern analysis indicated no homology between the alkaliphile gene, which we propose to call nhaC, and the nhaA gene of Escherichia coli, nor with other genes in digests of DNA from E. coli, Bacillus subtilis, or Bacillus alcalophilus. Although there was also no significant similarity between the deduced protein products of the alkaliphile gene and the nhaA gene of E. coli, there was a small region of significant similarity between the deduced alkaliphile gene product and the protein encoded by a human Na+/H+ antiporter gene (Sardet, C., Franchi, A., and Pouyssegur, J. (1989) Cell 56, 271-280).  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