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 31 毫秒
1.
Proline, which is the most abundant residue in beta-casein, stimulates growth of Lactococcus lactis in a proline-requiring strain (Lactococcus lactis subsp. cremoris Wg2) and in a proline-prototrophic strain (Lactococcus lactis subsp. lactis ML3). Both strains lack a proline-specific uptake system, and free proline can enter the cell only by passive diffusion across the cytoplasmic membrane. On the other hand, lactococci can actively take up proline-containing peptides via the lactococcal di- and tripeptide transport system, and these peptides are the major source of proline. Consequently, lactococcal growth on amino acid-based media is highly stimulated by the addition of proline-containing di- and tripeptides. Growth of L. lactis subsp. lactis ML3 on chemically defined media supplemented with casein does not appear proline limited. Addition of dipeptides (including proline-containing peptides) severely inhibits growth on a casein-containing medium, which indicates that the specific growth rate is determined by the balanced supply of different di- or tripeptides which compete for the same di- and tripeptide transport system.  相似文献   

2.
A proton motive force-driven di-tripeptide carrier protein (DtpT) and an ATP-dependent oligopeptide transport system (Opp) have been described for Lactococcus lactis MG1363. Using genetically well-defined mutants in which dtpT and/or opp were inactivated, we have now established the presence of a third peptide transport system (DtpP) in L. lactis. The specificity of DtpP partially overlaps that of DtpT. DtpP transports preferentially di- and tripeptides that are composed of hydrophobic (branched-chain amino acid) residues, whereas DtpT has a higher specificity for more-hydrophilic and charged peptides. The toxic dipeptide L-phenylalanyl-beta-chloro-L-alanine has been used to select for a di-tripeptide transport-negative mutant with the delta dtpT strain as a genetic background. This mutant is unable to transport di- and tripeptides but still shows uptake of amino acids and oligopeptides. The DtpP system is induced in the presence of di- and tripeptides containing branched-chain amino acids. The use of ionophores and metabolic inhibitors suggests that, similar to Opp, DtpP-mediated peptide transport is driven by ATP or a related energy-rich phosphorylated intermediate.  相似文献   

3.
The gene encoding the di- and tripeptide transport protein (DtpT) of Lactobacillus helveticus (DtpTLH) was cloned with the aid of the inverse PCR technique and used to complement the dipeptide transport-deficient and proline-auxotrophic Escherichia coli E1772. Functional expression of the peptide transporter was shown by the uptake of prolyl-[14C] alanine in whole cells and membrane vesicles. Peptide transport via DtpT in membrane vesicles is driven by the proton motive force. The system has specificity for di- and tripeptides but not for amino acids or tetrapeptides. The dtpTLH gene consists of 1,491 bp, which translates into a 497-amino-acid polypeptide. DtpTLH shows 34% identity to the di- and tripeptide transport protein of Lactococcus lactis and is also homologous to various peptide transporters of eukaryotic origin, but the similarity between these proteins is confined mainly to the N-terminal halves.  相似文献   

4.
Peptide utilization by group N streptococci.   总被引:3,自引:0,他引:3  
The rate of glycylleucine uptake by Group N streptococci varied widely. One strain of Streptococcus cremoris did not transport the dipeptide or utilize tripeptides. In peptide-utilizing strains, amino acid, dipeptide and tripeptide transport were distinct, although dipeptides inhibited tripeptide utilization. Specificity determinants for peptide transport and utilization were similar to those reported in Gram-negative bacteria. Peptide utilization in S. lactis was not completely dependent on the transport of intact peptides.  相似文献   

5.
Listeria monocytogenes takes up di- and tripeptides via a proton motive force-dependent carrier protein. This peptide transport system resembles the recently cloned and sequenced secondary di- and tripeptide transport system of Lactococcus lactis (A. Hagting, E. R. S. Kunji, K. J. Leenhouts, B. Poolman, and W. N. Konings, J. Biol. Chem. 269:11391-11399, 1994). The peptide permease of L. monocytogenes has a broad substrate specificity and allows transport of the nonpeptide substrate 5-aminolevulinic acid, the toxic di- and tripeptide analogs, alanyl-beta-chloroalanine and alanyl-alanyl-beta-chloroalanine, and various di- and tripeptides. No extracellular peptide hydrolysis was detected, indicating that peptides are hydrolyzed after being transported into the cell. Indeed, peptidase activities in response to various synthetic substrates were detected in cell extracts obtained from L. monocytogenes cells grown in brain heart infusion broth or defined medium. The di- and tripeptide permease can supply L. monocytogenes with essential amino acids for growth and might contribute to growth of this pathogen in various foods where peptides are supplied by proteolytic activity of other microorganisms present in these foods. Possible roles of this di- and tripeptide transport system in the osmoregulation and virulence of L. monocytogenes are discussed.  相似文献   

6.
The gene encoding a tripeptidase (pepT) of Lactococcus lactis subsp. cremoris (formerly subsp. lactis) MG1363 was cloned from a genomic library in pUC19 and subsequently sequenced. The tripeptidase of L. lactis was shown to be homologous to PepT of Salmonella typhimurium with 47.4% identity in the deduced amino acid sequences. L. lactis PepT was enzymatically active in Escherichia coli and allowed growth of a peptidase-negative leucine-auxotrophic E. coli strain by liberation of Leu from a tripeptide. Using a two-step integration-excision system, a pepT-negative mutant of L. lactis was constructed. No differences between the growth of the mutant and that of the wild-type strain in milk or in chemically defined medium with casein as the sole source of essential amino acids were observed.  相似文献   

7.
Lactococcus lactis ML3 possesses two different peptide transport systems of which the substrate size restriction and specificity have been determined. The first system is the earlier-described proton motive force-dependent di-tripeptide carrier (E. J. Smid, A. J. M. Driessen, and W. N. Konings, J. Bacteriol. 171:292-298, 1989). The second system is a metabolic energy-dependent oligopeptide transport system which transports peptides of four to at least six amino acid residues. The involvement of a specific oligopeptide transport system in the utilization of tetra-alanine and penta-alanine was established in a mutant of L. lactis MG1363 that was selected on the basis of resistance to toxic analogs of alanine and alanine-containing di- and tripeptides. This mutant is unable to transport alanine, dialanine, and trialanine but still shows uptake of tetra-alanine and penta-alanine. The oligopeptide transport system has a lower activity than the di-tripeptide transport system. Uptake of oligopeptides occurs in the absence of a proton motive force and is specifically inhibited by vanadate. The oligopeptide transport system is most likely driven by ATP or a related energy-rich, phosphorylated intermediate.  相似文献   

8.
A lysine antimetabolite, L-4-oxalysine [H2NCH2CH2OCH2CH(NH2)COOH], and oxalysine-containing di-, tri-, tetra- and pentapeptides inhibited growth of Candida albicans H317. Micromolar amounts of amino acids were found to overcome ammonium repression of the di- and tripeptide transport system(s) in strain H317. Several amino acids increased the toxicity of oxalysine-containing di- and tripeptides for C. albicans with little or no increase in toxicity of oxalysine or oxalysine-containing tetra- and pentapeptides. L-Lysine completely reversed the toxicity of oxalysine by competing with the transport of oxalysine into the cells. In contrast, L-lysine increased the toxicity of oxalysine-containing di- and tripeptides, but had no effect on the toxicity of oxalysine-containing tetra- and pentapeptides. Incubation of cells with L-lysine for 4 h resulted in a 15-fold increase in the rate of transport of radiolabelled dileucine, indicating that increased sensitivity of C. albicans to some toxic peptides in the presence of L-lysine may be attributed to an increased rate of transport of these peptides. Our results indicate that the dipeptide and tripeptide transport system(s) of C. albicans are regulated by micromolar amounts of amino acids in a similar fashion to the regulation of peptide transport in Saccharomyces cerevisiae and that multiple peptide transport systems differentially regulated by various nitrogen sources and amino acids exist in C. albicans.  相似文献   

9.
Utilization of Oligopeptides by Listeria monocytogenes Scott A   总被引:1,自引:0,他引:1       下载免费PDF全文
For effective utilization of peptides, Listeria monocytogenes possesses two different peptide transport systems. The first one is the previously described proton motive force (PMF)-driven di- and tripeptide transport system (A. Verheul, A. Hagting, M.-R. Amezaga, I. R. Booth, F. M. Rombouts, and T. Abee, Appl. Environ. Microbiol. 61:226–233, 1995). The present results reveal that L. monocytogenes possesses an oligopeptide transport system, presumably requiring ATP rather than the PMF as the driving force for translocation. Experiments to determine growth in a defined medium containing peptides of various lengths suggested that the oligopeptide permease transports peptides of up to 8 amino acid residues. Peptidase activities towards several oligopeptides were demonstrated in cell extract from L. monocytogenes, which indicates that upon internalization, the oligopeptides are hydrolyzed to serve as sources of amino acids for growth. The peptide transporters of the nonproteolytic L. monocytogenes might play an important role in foods that harbor indigenous proteinases and/or proteolytic microorganisms, since Pseudomonas fragi as well as Bacillus cereus was found to enhance the growth of L. monocytogenes to a large extent in a medium in which the milk protein casein was the sole source of nitrogen. In addition, growth stimulation was elicited in this medium when casein was hydrolyzed by using purified protease from Bacillus licheniformis. The possible contribution of the oligopeptide transport system in the establishment of high numbers of L. monocytogenes cells in fermented milk products is discussed.  相似文献   

10.
We have isolated and characterized the Saccharomyces cerevisiae PTR3 gene by functional complementation of a mutant deficient for amino acid-inducible peptide transport. PTR3 is predicted to encode a protein of 678 amino acids that exhibits no similarity to any other protein in the database. Deletion of the PTR3 open reading frame pleiotropically reduced the sensitivity to toxic peptides and amino acid analogues. Initial rates of radiolabelled dipeptide uptake demonstrated that elimination of PTR3 resulted in the loss of amino acid-induced levels of peptide transport. PTR3 was required for amino acid-induced expression of PTR2 , the gene encoding the dipeptide/tripeptide transport protein, but was not necessary for nitrogen catabolite repression of peptide import or PTR2 expression. It was determined that PTR3 also modulates expression of BAP2 , the gene encoding the branched-amino acid permease. Furthermore, we present genetic evidence that suggests that PTR3 functions within a novel regulatory pathway that facilitates amino acid induction of the PTR system.  相似文献   

11.
A slowly milk-coagulating variant (Fmc(-)) of Lactobacillus helveticus CRL 1062, designated S1, was isolated and characterized. Strain S1 possessed all the known essential components required to utilize casein as a nitrogen source, which include functional proteinase and peptidase activities as well as functional amino acid, di- and tripeptide, and oligopeptide transport systems. The amino acid requirements of strain S1 were similar to those of the parental strain. However, on a purine-free, chemically defined medium, the growth rate of the Fmc(-) strain was threefold lower than that of the wild-type strain. L. helveticus S1 was found to be defective in IMP dehydrogenase activity and therefore was deficient in the ability to synthesize XMP and GMP. This conclusion was further supported by the observation that the addition of guanine or xanthine to milk, a substrate poor in purine compounds, restored the Fmc(+) phenotype of L. helveticus S1.  相似文献   

12.
13.
The consumption of amino acids and peptides was monitored during growth in milk of proteinase-positive (Prt+) and -negative (Prt-) strains of Lactococcus lactis. The Prt- strains showed monophasic exponential growth, while the Prt+ strains grew in two phases. The first growth phases of the Prt+ and Prt- strains were in same, and no hydrolysis of casein was observed. Also, the levels of consumption of amino acids and peptides in the Prt+ and Prt- strains were similar. At the end of this growth phase, not all free amino acids and peptides were used, indicating that the remaining free amino acids and peptides were unable to sustain growth. The consumption of free amino acids was very low (about 5 mg/liter), suggesting that these nitrogen sources play only a minor role in growth. Oligopeptide transport-deficient strains (Opp-) of L. lactis were unable to utilize oligopeptides and grew poorly in milk. However, a di- and tripeptide transport-deficient strain (DtpT-) grew exactly like the wild type (Opp+ Dtpt+) did. These observations indicate that oligopeptides represent the main nitrogen source for growth in milk during the first growth phase. In the second phase of growth of Prt+ strains, milk proteins are hydrolyzed to peptides by the proteinase. Several of the oligopeptides formed are taken up and hydrolyzed internally by peptidases to amino acids, several of which are subsequently released into the medium (see also E.R.S. Kunji, A. Hagting, C.J. De Vries, V. Juillard, A.J. Haandrikman, B. Poolman, and W.N. Konings, J. Biol. Chem. 270:1569-1574, 1995).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The specific oligopeptide transport system Opp is essential for growth of Lactococcus lactis in milk. We examined the biodiversity of oligopeptide transport specificity in the L. lactis species. Six strains were tested for (i) consumption of peptides during growth in a chemically defined medium and (ii) their ability to transport these peptides. Each strain demonstrated some specific preferences for peptide utilization, which matched the specificity of peptide transport. Sequencing of the binding protein OppA in some strains revealed minor differences at the amino acid level. The differences in specificity were used as a tool to unravel the role of the binding protein in transport specificity. The genes encoding OppA in four strains were cloned and expressed in L. lactis MG1363 deleted for its oppA gene. The substrate specificity of these engineered strains was found to be similar to that of the L. lactis MG1363 parental strain, whichever oppA gene was expressed. In situ binding experiments demonstrated the ability of OppA to interact with non-transported peptides. Taken together, these results provide evidence for a new concept. Despite that fact that OppA is essential for peptide transport, it is not the (main) determinant of peptide transport specificity in L. lactis.  相似文献   

15.
Despite research into the nutritional requirements of Helicobacter pylori, little is known regarding its use of complex substrates, such as peptides. Analysis of genome sequences revealed putative ABC-type transporter genes for dipeptide (dppABCDF) and oligopeptide (oppABCD) transport. Genes from each system were PCR amplified, cloned, and disrupted by cassette insertion either individually (dppA, dppB, dppC, oppA, oppB, and oppC) or to create double mutants (dppA oppA, dppB oppB, dppB dppC, and oppB oppC). Peptide-utilizing abilities of the strains were assessed by monitoring growth in a chemically defined medium where the only source of the essential amino acid isoleucine was from peptides of various lengths (two to nine amino acids long). The dipeptide system mutants lacked the ability to use certain dipeptides, hexapeptides, and nonapeptides. However, these mutants retained some ability to grow with other dipeptides, tripeptides, and tetrapeptides. Of the oligopeptide mutants, only the oppB strain differed significantly from the wild type. This strain showed a wild-type phenotype for growth with longer peptides (hexa- and nonapeptides) while having a decreased ability to utilize di-, tri-, and tetrapeptides. The dppA oppA and dppB oppB mutants showed similar phenotypes to those of the dppA and dppB mutants, respectively. Peptide digestion by metalloproteases was ruled out as the cause for residual peptide transport by growing mutant strains in the presence of either EDTA or EGTA. Degradation products associated with a fluorescein isothiocyanate-labeled hexapeptide (plus cells) were minimal. An as yet unidentified peptide transport system(s) in H. pylori is proposed to be responsible for the residual transport.  相似文献   

16.
A mutant of fast milk-coagulating (Fmc+) Lactococcus lactis subsp. lactis C2, designated L. lactis KB4, was identified. Although possessing the known components essential for utilizing casein as a nitrogen source, which include functional proteinase (PrtP) activity and oligopeptide, di- and tripeptide, and amino acid transport systems, KB4 exhibited a slow milk coagulation (Fmc) phenotype. When the amino acid requirements of L. lactis C2 were compared with those of KB4 by use of a chemically defined medium, it was found that KB4 was unable to grow in the absence of aspartic acid. This aspartic acid requirement could also be met by aspartate-containing peptides. The addition of aspartic acid to milk restored the Fmc+ phenotype of KB4. KB4 was found to be defective in pyruvate carboxylase and thus was deficient in the ability to form oxaloacetate and hence aspartic acid from pyruvate and carbon dioxide. The results suggest that when lactococci are propagated in milk, aspartate derived from casein is unable to meet fully the nutritional demands of the lactococci, and they become dependent upon aspartate biosynthesis.  相似文献   

17.
Aminotransferases, which catalyze the last step of biosynthesis of most amino acids and the first step of their catabolism, may be involved in the growth of Lactococcus lactis in milk. Previously, we isolated two aminotransferases from L. lactis, AraT and BcaT, which are responsible for the transamination of aromatic amino acids, branched-chain amino acids, and methionine. In this study, we demonstrated that double inactivation of AraT and BcaT strongly reduced the growth of L. lactis in milk. Supplementation of milk with amino acids and keto acids that are substrates of both aminotransferases did not improve the growth of the double mutant. On the contrary, supplementation of milk with isoleucine or a dipeptide containing isoleucine almost totally inhibited the growth of the double mutant, while it did not affect or only slightly affected the growth of the wild-type strain. These results suggest that AraT and BcaT play a major role in the growth of L. lactis in milk by degrading the intracellular excess isoleucine, which is responsible for the growth inhibition. The growth inhibition by isoleucine is likely to be due to CodY repression of the proteolytic system, which is necessary for maximal growth of L. lactis in milk, since the growth of the CodY mutant was not affected by addition of isoleucine to milk. Moreover, we demonstrated that AraT and BcaT are part of the CodY regulon and therefore are regulated by nutritional factors, such as the carbohydrate and nitrogen sources.  相似文献   

18.
Listeria monocytogenes is a gram-positive intracellular pathogen responsible for opportunistic infections in humans and animals. Here we identified and characterized the dtpT gene (lmo0555) of L. monocytogenes EGD-e, encoding the di- and tripeptide transporter, and assessed its role in growth under various environmental conditions as well as in the virulence of L. monocytogenes. Uptake of the dipeptide Pro-[14C]Ala was mediated by the DtpT transporter and was abrogated in a DeltadtpT isogenic deletion mutant. The DtpT transporter was shown to be required for growth when the essential amino acids leucine and valine were supplied as peptides. The protective effect of glycine- and proline-containing peptides during growth in defined medium containing 3% NaCl was noted only in L. monocytogenes EGD-e, not in the DeltadtpT mutant strain, indicating that the DtpT transporter is involved in salt stress protection. Infection studies showed that DtpT contributes to pathogenesis in a mouse infection model but has no role in bacterial growth following infection of J774 macrophages. These studies reveal that DptT may contribute to the virulence of L. monocytogenes.  相似文献   

19.
AIMS: To study the production of hydrophilic and hydrophobic peptides in cheese by 32 wild Lactococcus lactis strains of different RAPD patterns and to compare them with the peptides produced by lactococcal cells incubated with whole casein. METHOD AND RESULTS: Chromatograms of peptides from cheeses made using each strain as single starter culture were divided into five regions, and strains were classified in three groups by hierarchical cluster analysis of region areas. Thirty out of the 32 wild L. lactis strains produced higher levels of hydrophobic peptides in cheese than on whole casein. CONCLUSIONS: Cheese was a more favourable substrate than whole casein for hydrophobic peptide formation by L. lactis strains. SIGNIFICANCE AND IMPACT OF THE STUDY: New strains of lactococci should be screened for bitterness under cheese conditions, as the formation of hydrophobic peptides may be underestimated in assays with casein as substrate.  相似文献   

20.
A slowly milk-coagulating variant (Fmc) of Lactobacillus helveticus CRL 1062, designated S1, was isolated and characterized. Strain S1 possessed all the known essential components required to utilize casein as a nitrogen source, which include functional proteinase and peptidase activities as well as functional amino acid, di- and tripeptide, and oligopeptide transport systems. The amino acid requirements of strain S1 were similar to those of the parental strain. However, on a purine-free, chemically defined medium, the growth rate of the Fmc strain was threefold lower than that of the wild-type strain. L. helveticus S1 was found to be defective in IMP dehydrogenase activity and therefore was deficient in the ability to synthesize XMP and GMP. This conclusion was further supported by the observation that the addition of guanine or xanthine to milk, a substrate poor in purine compounds, restored the Fmc+ phenotype of L. helveticus S1.  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