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.
Tripeptidases from Lactococcus lactis subsp. lactis (L9PepTR), L. lactis subsp. cremoris (L6PepTR), and L. lactis subsp. hordniae (hTPepTR) were cloned, overexpressed, purified, and characterized. Although these enzymes contained three to seven naturally occurring amino acid differences, both metal-binding and catalytic sites were highly conserved. The k(cat) values of hTPepTR were approximately 1.5- to 2-fold higher than those of L9PepTR, while, for L6PepTR, they were approximately 0.8- to 1.4-times the L9PepTR values. The K(m) of tripeptidase from subsp. lactis (L9PepTR) was considerably larger when glycine was the amino acid located at both the N- and C-terminus of the peptide substrate. In addition, the K(m) values of L9PepTR increased in the following order for YGG, LGG, FGG, SGG, and alpha-aminoisobutyrylglycylglycine, while the k(cat)/K(m) decreased in the same order. These results suggest that the dipole moment and steric hindrance of the N-terminal amino acid side chain may be the most important factors controlling substrate specificity.  相似文献   

2.
A 96 kDa aminopeptidase was purified from Streptococcus salivarius subsp. thermophilus NCDO 573. The enzyme had similar properties to aminopeptidases isolated from lactococci and lactobacilli and showed a high degree of N -terminal amino acid sequence homology to aminopeptidase N from Lactococcus lactis subsp. cremoris. It catalysed the hydrolysis of a range of aminoacyl 4-nitroanilides and 7-amido-4-methylcoumarin derivatives, dipeptides, tripeptides and oligopeptides. In common with aminopeptidases from other lactic acid bacteria, the enzyme from Strep. salivarius subsp. thermophilus showed highest activity with lysyl derivatives but was also very active with arginyl and leucyl derivatives. Relative activity with alanyl, phenylalanyl, tyrosyl, seryl and valyl derivatives was considerably lower and with glycyl, glutamyl and prolyl derivatives almost negligible. The aminopeptidase also catalysed the hydrolysis of dipeptides and tripeptides but mostly at rates much less than that with L-lysyl-4-nitroanilide and oligopeptides. The enzyme catalysed the successive hydrolysis of various amino acid residues from the N -terminus of several oligopeptides but it was unable to cleave peptide bonds on the N -terminal side of a proline residue.  相似文献   

3.
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   

4.
Tripeptidase (PepT) and dipeptidase (PepV), the enzymes located in the final stage of the intracellular proteolytic system, were demonstrated to be distributed widely in lactic acid bacteria, especially in lactococci. Both the tripeptidase genes (pepT) and dipeptidase genes (pepV) of 15 lactococcal strains consisting of the type and domestic strains were cloned and sequenced using normal and TAIL PCR methods. Amino acid sequences of these enzymes were highly conserved among strains. Evolutionary distance trees based on the sequence of 1239 nucleotides of pepT and 1416 nucleotide of pepV showed a similar cluster as that obtained from the 1499 fragment of the 16S rRNA. Based on this profile, the species Lactococcus lactis is reasonably divided into three subspecies groups, subsp. lactis, cremoris, and hordniae, as in the current classification. Figure of trees from pepT and pepV were essentially identical to each other and slightly more intricate than that from 16S rRNA. The K nuc values obtained from pepT and pepV genes were approximately ten times as high as that from 16S rRNA. Considering these results, phylogenetic analysis based on pepT and pepV genes may aid in a more precise index of classification of L. lactis subspecies. PepT and PepV seem to have evolved in similar directions in lactococci.  相似文献   

5.
The proteolytic system of Bifidobacterium animalis subsp. lactis was analyzed, and an intracellular endopeptidase (PepO) was identified and characterized. This work reports the first complete cloning, purification, and characterization of a proteolytic enzyme in Bifidobacterium spp. Aminopeptidase activities (general aminopeptidases, proline iminopeptidase, X-prolyl dipeptidylaminopeptidase) found in cell extracts of B. animalis subsp. lactis were higher for cells that had been grown in a milk-based medium than for those grown in MRS. A high specific proline iminopeptidase activity was observed in B. animalis subsp. lactis. Whole cells and cell wall-bound protein fractions showed no caseinolytic activity; however, the combined action of intracellular proteolytic enzymes could hydrolyze casein fractions rapidly. The endopeptidase activity of B. animalis subsp. lactis was examined in more detail, and the gene encoding an endopeptidase O in B. animalis subsp. lactis was cloned and overexpressed in Escherichia coli. The deduced amino acid sequence for B. animalis subsp. lactis PepO indicated that it is a member of the M13 peptidase family of zinc metallopeptidases and displays 67.4% sequence homology with the predicted PepO protein from Bifidobacterium longum. The recombinant enzyme was shown to be a 74-kDa monomer. Activity of B. animalis subsp. lactis PepO was found with oligopeptide substrates of at least 5 amino acid residues, such as met-enkephalin, and with larger substrates, such as the 23-amino-acid peptide alpha s1-casein(f1-23). The predominant peptide bond cleaved by B. animalis subsp. lactis PepO was on the N-terminal side of phenylalanine residues. The enzyme also showed a post-proline secondary cleavage site.  相似文献   

6.
The fifth phage resistance factor from the prototype phage-insensitive strain Lactococcus lactis subsp. lactis ME2 has been characterized and sequenced. The genetic determinant for Prf (phage resistance five) was subcloned from the conjugative plasmid pTN20, which also encodes a restriction and modification system. Typical of other abortive resistance mechanisms, Prf reduces the efficiency of plaquing to 10(-2) to 10(-3) and decreases the plaque size and burst size of the small isometric-headed phage p2 in L. lactis subsp. lactis LM0230. However, normal-size plaques occurred at a frequency of 10(-4) and contained mutant phages that were resistant to Prf, even after repeated propagation through a sensitive host. Prf does not prevent phage adsorption or promote restriction and modification activities, but 90% of Prf+ cells infected with phage p2 die. Thus, phage infections in Prf+ cells are aborted. Prf is effective in both L. lactis subsp. lactis and L. lactis subsp. cremoris strains against several small isometric-headed phages but not against prolate-headed phages. The Prf determinant was localized by Tn5 mutagenesis and subcloning. DNA sequencing identified a 1,056-nucleotide structural gene designated abiC. Prf+ expression was obtained when abiC was subcloned into the lactococcal expression vector pMG36e. abiC is distinct from two other lactococcal abortive phage resistance genes, abiA (Hsp+, from L. lactis subsp. lactis ME2) and abi416 (Abi+, from L. lactis subsp. lactis IL416). Unlike abiA, the action of abiC does not appear to affect DNA replication. Thus, abiC represents a second abortive system found in ME2 that acts at a different point of the phage lytic cycle.  相似文献   

7.
The diversity of a collection of 102 lactococcus isolates including 91 Lactococcus lactis isolates of dairy and nondairy origin was explored using partial small subunit rRNA gene sequence analysis and limited phenotypic analyses. A subset of 89 strains of L. lactis subsp. cremoris and L. lactis subsp. lactis isolates was further analyzed by (GTG)(5)-PCR fingerprinting and a novel multilocus sequence analysis (MLSA) scheme. Two major genomic lineages within L. lactis were found. The L. lactis subsp. cremoris type-strain-like genotype lineage included both L. lactis subsp. cremoris and L. lactis subsp. lactis isolates. The other major lineage, with a L. lactis subsp. lactis type-strain-like genotype, comprised L. lactis subsp. lactis isolates only. A novel third genomic lineage represented two L. lactis subsp. lactis isolates of nondairy origin. The genomic lineages deviate from the subspecific classification of L. lactis that is based on a few phenotypic traits only. MLSA of six partial genes (atpA, encoding ATP synthase alpha subunit; pheS, encoding phenylalanine tRNA synthetase; rpoA, encoding RNA polymerase alpha chain; bcaT, encoding branched chain amino acid aminotransferase; pepN, encoding aminopeptidase N; and pepX, encoding X-prolyl dipeptidyl peptidase) revealed 363 polymorphic sites (total length, 1,970 bases) among 89 L. lactis subsp. cremoris and L. lactis subsp. lactis isolates with unique sequence types for most isolates. This allowed high-resolution cluster analysis in which dairy isolates form subclusters of limited diversity within the genomic lineages. The pheS DNA sequence analysis yielded two genetic groups dissimilar to the other genotyping analysis-based lineages, indicating a disparate acquisition route for this gene.  相似文献   

8.
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   

9.
In Gouda and Cheddar type cheeses the amino acid conversion to aroma compounds, which is a major process for aroma formation, is essentially due to lactic acid bacteria (LAB). In order to evaluate the respective role of starter and nonstarter LAB and their interactions in cheese flavor formation, we compared the catabolism of phenylalanine, leucine, and methionine by single strains and strain mixtures of Lactococcus lactis subsp. cremoris NCDO763 and three mesophilic lactobacilli. Amino acid catabolism was studied in vitro at pH 5.5, by using radiolabeled amino acids as tracers. In the presence of alpha-ketoglutarate, which is essential for amino acid transamination, the lactobacillus strains degraded less amino acids than L. lactis subsp. cremoris NCDO763, and produced mainly nonaromatic metabolites. L. lactis subsp. cremoris NCDO763 produced mainly the carboxylic acids, which are important compounds for cheese aroma. However, in the reaction mixture containing glutamate, only two lactobacillus strains degraded amino acids significantly. This was due to their glutamate dehydrogenase (GDH) activity, which produced alpha-ketoglutarate from glutamate. The combination of each of the GDH-positive lactobacilli with L. lactis subsp. cremoris NCDO763 had a beneficial effect on the aroma formation. Lactobacilli initiated the conversion of amino acids by transforming them mainly to keto and hydroxy acids, which subsequently were converted to carboxylic acids by the Lactococcus strain. Therefore, we think that such cooperation between starter L. lactis and GDH-positive lactobacilli can stimulate flavor development in cheese.  相似文献   

10.
A novel pepX gene was cloned from isolated DNA of Lactococcus lactis by PCR. The deduced amino acid sequence of the 89-kDa protein showed 94, 93, 65, and 44% identity with the pepX protein from Lactococcus lactis subsp. cremoris, Lactococcus lactis subsp. lactis, Lactobacillus delbruecki subsp. bulgaricus, and Lactobacillus helveticus, respectively, and contained a serine protease G-K-S-Y-L-G consensus motif. The pepX gene has been cloned into pET17b and was expressed at a high level in Escherichia coli BL21 (DE3) LysS. PepX was purified to approximate homogeneity with ammonium sulfate precipitation and DEAE Sephadex A-50 chromatography. Optimal pepX activity was observed at pH 8.0 and 37 degrees C. According to SDS-PAGE analysis, pepX has a molecular mass of approximately 89 kDa. The peptidase can remove completely the unwanted X-Pro from the N-terminal of the target protein, releasing the naturally active protein and peptide, revealing a prospective application of pepX in large-scale production of pharmaceutical protein and peptide products.  相似文献   

11.
The uptake of the basic amino acids arginine, ornithine, and lysine was studied in membrane vesicles derived from cells of Lactococcus lactis which were fused with liposomes in which beef heart mitochondrial cytochrome c oxidase was incorporated as a proton motive force (PMF)-generating system. In the presence of ascorbate N,N,N'N'-tetramethylphenylenediamine-cytochrome c as the electron donor, these fused membranes accumulated lysine but not ornithine or arginine under aerobic conditions. The mechanism of energy coupling to lysine transport was examined in membrane vesicles of L. lactis subsp. cremoris upon imposition of an artificial electrical potential (delta psi) or pH gradient or both and in fused membranes of these vesicles with cytochrome c oxidase liposomes in which the delta psi and delta pH were manipulated with ionophores. Lysine uptake was shown to be coupled to the PMF and especially to the delta psi, suggesting a proton symport mechanism. The lysine carrier appeared to be specific for L and D isomers of amino acids with a guanidine or NH2 group at the C6 position of the side chain. Uptake of lysine was blocked by p-chloromercuribenzene sulfonic acid but not by maleimides. Counterflow of lysine could not be detected in L. lactis subsp. cremoris, but in the arginine-ornithine antiporter-containing L. lactis subsp. lactis, rapid counterflow occurred. Homologous exchange of lysine and heterologous exchange of arginine and lysine were mediated by this antiporter. PMF-driven lysine transport in these membranes was noncompetitively inhibited by arginine, whereas the uptake of arginine was enhanced by lysine. These observations are compatible with a model in which circulation of lysine via the lysine carrier and the arginine-ornithine antiporter leads to accumulation of arginine.  相似文献   

12.
The gene specifying an endopeptidase of Lactococcus lactis, named pepO, was cloned from a genomic library of L. lactis subsp. cremoris P8-2-47 in lambda EMBL3 and was subsequently sequenced. pepO is probably the last gene of an operon encoding the binding-protein-dependent oligopeptide transport system of L. lactis. The inferred amino acid sequence of PepO showed that the lactococcal endopeptidase has a marked similarity to the mammalian neutral endopeptidase EC 3.4.24.11 (enkephalinase), whereas no obvious sequence similarity with any bacterial enzyme was found. By means of gene disruption, a pepO-negative mutant was constructed. Growth and acid production of the mutant strain in milk were not affected, indicating that the endopeptidase is not essential for growth of L. lactis in milk.  相似文献   

13.
Lactococcus lactis subsp. lactis NCDO 763 (also designated ML3) possesses an X-prolyl dipeptidyl aminopeptidase (X-PDAP; EC 3.4.14.5). X-PDAP mutants were selected by an enzymatic plate assay on the basis of their inability to hydrolyze an L-phenylalanyl-L-proline-beta-naphthylamide substrate. A DNA bank from L. lactis subsp. lactis NCDO 763 was constructed in one of these X-PDAP mutants, and one clone in which the original X-PDAP phenotype was restored was detected by the enzymatic plate assay. The X-PDAP gene, designated pepXP, was further subcloned and sequenced. It codes for a protein containing 763 residues. Comparison of the amino-terminal sequence of the X-PDAP enzyme with the amino acid sequence deduced from the pepXP gene indicated that the enzyme is not subjected to posttranslational modification or exported via processing of a signal peptide. The pepXP gene from L. lactis subsp. lactis NCDO 763 in more than 99% homologous to the pepXP gene from L. lactis subsp. cremoris P8-2-47 described elsewhere (B. Mayo, J. Kok, K. Venema, W. Bockelmann, M. Teuber, H. Reinke, and G. Venema, Appl. Environ. Microbiol. 57:38-44, 1991) and is also conserved in other lactococcal strains.  相似文献   

14.
Lactococcus lactis subsp. lactis NCDO 763 (also designated ML3) possesses an X-prolyl dipeptidyl aminopeptidase (X-PDAP; EC 3.4.14.5). X-PDAP mutants were selected by an enzymatic plate assay on the basis of their inability to hydrolyze an L-phenylalanyl-L-proline-beta-naphthylamide substrate. A DNA bank from L. lactis subsp. lactis NCDO 763 was constructed in one of these X-PDAP mutants, and one clone in which the original X-PDAP phenotype was restored was detected by the enzymatic plate assay. The X-PDAP gene, designated pepXP, was further subcloned and sequenced. It codes for a protein containing 763 residues. Comparison of the amino-terminal sequence of the X-PDAP enzyme with the amino acid sequence deduced from the pepXP gene indicated that the enzyme is not subjected to posttranslational modification or exported via processing of a signal peptide. The pepXP gene from L. lactis subsp. lactis NCDO 763 in more than 99% homologous to the pepXP gene from L. lactis subsp. cremoris P8-2-47 described elsewhere (B. Mayo, J. Kok, K. Venema, W. Bockelmann, M. Teuber, H. Reinke, and G. Venema, Appl. Environ. Microbiol. 57:38-44, 1991) and is also conserved in other lactococcal strains.  相似文献   

15.
Cells of Streptococcus salivarius subsp. thermophilus and Lactococcus lactis subsp. lactis entrapped in k-carrageenan-locust bean gum gel performed similarly to free cells in the conversion of lactose to lactic acid. Bead diameter influenced the fermentation rate. Cells entrapped in smaller beads (0.5 to 1.0 mm) showed higher release rates, higher lactose, glucose, and formic acid utilization, higher galactose accumulation, and higher lactic acid production than did cells entrapped in larger beads (1.0 to 2.0 mm). Values for smaller beads were comparable with those for free cells. Immobilization affected the fermentation rate of lactic acid bacteria, especially Lactobacillus delbrueckii subsp. bulgaricus. Entrapped cells of L. delbrueckii subsp. bulgaricus demonstrated a lower lactic acid production than did free cells in batch fermentation. The kinetics of the production of formic and pyruvic acids by L. lactis subsp. lactis and S. salivarius subsp. thermophilus are presented.  相似文献   

16.
17.
H Holo  O Nilssen    I F Nes 《Journal of bacteriology》1991,173(12):3879-3887
A new bacteriocin, termed lactococcin A (LCN-A), from Lactococcus lactis subsp. cremoris LMG 2130 was purified and sequenced. The polypeptide contained no unusual amino acids and showed no significant sequence similarity to other known proteins. Only lactococci were killed by the bacteriocin. Of more than 120 L. lactis strains tested, only 1 was found resistant to LCN-A. The most sensitive strain tested, L. lactis subsp. cremoris NCDO 1198, was inhibited by 7 pM LCN-A. By use of a synthetic DNA probe, lcnA was found to be located on a 55-kb plasmid. The lcnA gene was cloned and sequenced. The sequence data revealed that LCN-A is ribosomally synthesized as a 75-amino-acid precursor including a 21-amino-acid N-terminal extension. An open reading frame encoding a 98-amino-acid polypeptide was found downstream of and in the same operon as lcnA. We propose that this open reading frame encodes an immunity function for LCN-A. In Escherichia coli lcnA did not cause an LCN-A+ phenotype. L. lactis subsp. lactis IL 1403 produced small amounts of the bacteriocin and became resistant to LCN-A after transformation with a recombinant plasmid carrying lcnA. The other lactococcal strains transformed with the same recombinant plasmid became resistant to LCN-A but did not produce any detectable amount of the bacteriocin.  相似文献   

18.
19.
An endopeptidase gene (pepE) was isolated from a previously constructed genomic library of Lactobacillus helveticus CNRZ32. The pepE gene consisted of a 1,314-bp open reading frame encoding a putative peptide of 52.1 kDa. Significant identity was found between the deduced amino acid sequence of pepE and the sequences for aminopeptidase C from Lactobacillus delbrueckii subsp. lactis DSM7290, L. helveticus CNRZ32, Streptococcus thermophilus CNRZ302, and Lactococcus lactis subsp. cremoris AM2. A recombinant PepE fusion protein containing an N-terminal six-histidine tag was constructed and purified to electrophoretic homogeneity. Characterization of PepE revealed that it was a thiol-dependent protease having a monomeric mass of 50 kDa, with optimum temperature, NaCl concentration, and pH for activity at 32 to 37 degrees C, 0.5%, and 4.5, respectively. PepE had significant activity under conditions which simulate those of ripening cheese (10 degrees C, 4% NaCl, pH 5.1). PepE hydrolyzed internal peptide bonds in Met-enkephalin and bradykinin; however, hydrolysis of alpha-, beta-, and kappa-caseins was not detected.  相似文献   

20.
The peptidases of thermophilic lactic acid bacteria have a key role in the proteolysis of Swiss cheeses during warm room ripening. To compare their peptidase activities toward a dairy substrate, a tryptic/chymotryptic hydrolysate of purified beta-casein was used. Thirty-four peptides from 3 to 35 amino acids, including three phosphorylated peptides, constitute the beta-casein hydrolysate, as shown by tandem mass spectrometry. Cell extracts prepared from Lactobacillus helveticus ITG LH1, ITG LH77, and CNRZ 32, Lactobacillus delbrueckii subsp. lactis ITG LL14 and ITG LL51, L. delbrueckii subsp. bulgaricus CNRZ 397 and NCDO 1489, and Streptococcus thermophilus CNRZ 385, CIP 102303, and TA 060 were standardized in protein. The peptidase activities were assessed with the beta-casein hydrolysate as the substrate at pH 5.5 and 24 degrees C (conditions of warm room ripening) by (i) free amino acid release, (ii) reverse-phase chromatography, and (iii) identification of undigested peptides by mass spectrometry. Regardless of strain, L. helveticus was the most efficient in hydrolyzing beta-casein peptides. Interestingly, cell extracts of S. thermophilus were not able to release a significant level of free proline from the beta-casein hydrolysate, which was consistent with the identification of numerous dipeptides containing proline. With the three lactic acid bacteria tested, the phosphorylated peptides remained undigested or weakly hydrolyzed indicating their high intrinsic resistance to peptidase activities. Finally, several sets of peptides differing by a single amino acid in a C-terminal position revealed the presence of at least one carboxypeptidase in the cell extracts of these species.  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