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.
To explain the competition for nitrogenous nutrients observed in mixed strain cultures of Lactococcus lactis and Leuconostoc mesenteroides, the utilization of peptides as a source of essential amino acids for growth in a chemically defined medium was compared in 12 strains of dairy origin. Both species were multiple amino acid auxotrophs and harboured a large set of intracellular peptidases. Lactococcus lactis can use a wide variety of peptides up to 13 amino acid residues whereas Leuc. mesenteroides assimilated only shorter peptides containing up to seven amino acids. Growth was limited by the transport of peptides and not by their hydrolysis. The nutritional value of peptides varied with the strains and the composition of the peptides, L. lactis being advantaged over Leuc. mesenteroides.  相似文献   

2.
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   

3.
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   

4.
In cultures of L. lactis ssp. cremoris SK110, phage SK11G-resistant through the presence of pSK112, phage-sensitive variants segregated spontaneously that lacked the plasmid. In overnight batch culture these comprised up to 1% of the total population. Upon prolonged incubation in chemostat culture, a further loss of resistance was observed after a lag period. At high growth rates (0.7 h-1) this period amounted to approximately 35 generations, whereas cultures grown at rates of 0.4 and 0.1 h-1 remained resistant for 55 and 70 generations, respectively. At average-to-high growth rate, characteristics of the partially mixed populations that evolved were comparable to those of pure cultures of L. lactis ssp. cremoris SK110. However, in the culture fluid of the mixed populations that occurred at growth rate 0.1 h-1, higher acetate and formate concentrations were found than in the fluid of pure cultures of L. lactis ssp. cremoris SK110. This indicated that the former metabolized lactose more efficiently. Competition experiments between the resistant strain and a cured, sensitive derivative, L. lactis ssp. cremoris SK112, gave stable mixed populations. It is concluded that at average-to-high growth rates, loss of resistance from cultures of L. lactis ssp. cremoris SK110 had occurred due to instability of the plasmid and not to a competitive disadvantage of the resistant strain towards emerging sensitive variants.  相似文献   

5.
The maximum specific growth rate of Streptococcus lactis and Streptococcus cremoris on synthetic medium containing glutamate but no glutamine decreases rapidly above pH 7. Growth of these organisms is extended to pH values in excess of 8 in the presence of glutamine. These results can be explained by the kinetic properties of glutamate and glutamine transport (B. Poolman, E. J. Smid, and W. N. Konings, J. Bacteriol. 169:2755-2761, 1987). At alkaline pH the rate of growth in the absence of glutamine is limited by the capacity to accumulate glutamate due to the decreased availability of glutamic acid, the transported species of the glutamate-glutamine transport system. Kinetic analysis of leucine and valine transport shows that the maximal rate of uptake of these amino acids by the branched-chain amino acid transport system is 10 times higher in S. lactis cells grown on synthetic medium containing amino acids than in cells grown in complex broth. For cells grown on synthetic medium, the maximal rate of transport exceeds by about 5 times the requirements at maximum specific growth rates for leucine, isoleucine, and valine (on the basis of the amino acid composition of the cell). The maximal rate of phenylalanine uptake by the aromatic amino acid transport system is in small excess of the requirement for this amino acid at maximum specific growth rates. Analysis of the internal amino acid pools of chemostat-grown cells indicates that passive influx of (some) aromatic amino acids may contribute to the net uptake at high dilution rates.  相似文献   

6.
The presence and the nucleotide sequence of four multidrug resistance genes, lmrA, lmrP, lmrC, and lmrD, were investigated in 13 strains of Lactococcus lactis ssp. lactis, four strains of Lactococcus lactis ssp. cremoris, two strains of Lactococcus plantarum, and two strains of Lactococcus raffinolactis. Multidrug resistance genes were present in all L. lactis isolates tested. However, none of them could be detected in the strains belonging to the species L. raffinolactis and L. plantarum, suggesting a different set of multidrug resistance genes in these species. The analysis of the four deduced amino acid sequences established two different variants depending on the subspecies of L. lactis. Either lmrA, or lmrP, or both were found naturally disrupted in five strains, while full-length lmrD was present in all strains.  相似文献   

7.
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   

8.
The peptides released from beta-casein by the action of PI-type proteinase (PrtP) from Lactococcus lactis subsp. cremoris Wg2 have been identified by on-line coupling of liquid chromatography to mass spectrometry. After 24 h of incubation of beta-casein with purified PrtP, a stable mixture of peptides was obtained. The trifluoroacetic acid-soluble peptides of this beta-casein hydrolysate were fractionated by high-performance liquid chromatography and introduced into the liquid chromatography-ion spray mass spectrometry interface. Multiply charged ions were generated from trifluoroacetic acid-soluble peptides under low nozzle voltage conditions, yielding the MH+ mass of each eluted peptide. All peptides corresponding to each of the MH+ calculated masses were determined. In those cases in which different peptides were possible, further identification was achieved by collision-induced dissociation under higher nozzle voltage conditions. Hydrolysis of beta-casein by PrtP was observed to proceed much further than reported previously. More than 40% of the peptide bonds are cleaved by PrtP, resulting in the formation of more than 100 different oligopeptides. With the exception of Phe, significant release of amino acids or di- and tripeptides could not be observed. Interestingly, one-fifth of the identified oligopeptides are small enough to be taken up by the oligopeptide transport system. Uptake of these peptides could supply L. lactis with all amino acids, including the essential ones, indicating that growth of L. lactis might be possible on peptides released from beta-casein by proteinase only.  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.
AIMS: The effects of medium-composition and fermentation parameters on the properties of mixed mesophilic starters were studied. The starter was composed of Lactococcus lactis ssp. lactis (L. lactis), Lactococcus lactis ssp. cremoris (L. cremoris), Lactobacillus rhamnosus (Lact. rhamnosus) and Leuconostoc mesenteroides ssp. cremoris (Leuc. cremoris). METHODS AND RESULTS: The media used were reconstituted skim milk (RSM), and whey-based media with either citrate or phosphate buffers. The fermentation parameters were incubation temperature (22 degrees C or 32 degrees C), no pH control, and pH control in pH zones of either pH 6.0-5.8 or pH 6.0-5.2. The starter properties were strain ratio, specific acidifying activity (SAA), total population, residual carbohydrates and organic acids produced. The growth of L. lactis was favoured under pH control in whey-based media. High concentrations of Lact. rhamnosus were favoured in whey-based media prepared at 32 degrees C. The highest contents of Leuc. cremoris were obtained in starters prepared in RSM at 22 degrees C without pH control. Starters prepared under pH control gave the highest populations and made it possible for significantly lower inoculation rates (IR) to be used to carry out subsequent milk fermentations. However, the SAA of starters prepared under pH control were lower than the SAA of starters grown without any pH control. CONCLUSIONS: None of the conditions enabled the strain ratio at inoculation to be maintained. The data show that it is possible to prepare a mesophilic starter that has a significant probiotic Lact. rhamnosus content; this starter could be used in the preparation of probiotic-containing cheeses or in Leuc. cremoris for aroma production in fermented milks. SIGNIFICANCE AND IMPACT OF THE STUDY: This study provides data on what should be expected with respect to strain ratios and IR if cheesemakers decide to shift their aroma-producing starter production method from the traditional 'milk-based without pH control' method to whey-based media used with pH-zone control strategies.  相似文献   

11.
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   

12.
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   

13.
Lactococcus lactis strains from the subsp. cremoris are described as more sensitive to osmotic stress than subsp. lactis strains. We examined the relation between osmotic tolerance and the activity of the betaine transporter BusA among 34 strains of L. lactis. The cremoris strains that showed reduced growth at high osmolality failed to accumulate betaine. The nature of the defect was found to vary among cremoris strains: lack of the busA encoding region, absence of synthesis or synthesis of an inactive form of BusA. The results suggest that the selection of strains well fitted to the dairy production lead to the loss of an otherwise efficient adaptation mechanism.  相似文献   

14.
The stability of plasmids in Lactococcus lactis ssp. lactis strains C2 and ML3, and L. lactis ssp. cremoris strains ML1 and SC607, was investigated by extended incubation of bacterial cells in low nutrient media under acidic conditions. Strains were grown overnight (16-18 h) in skim milk and unbuffered medium (M17-) at 32 degrees C and subsequently held at that temperature for extended periods (greater than or equal to 96 h). Lac- variants were obtained from each strain in milk and (M17-) broth. The plasmid profiles of Lac- variants when compared with their parental Lac+ strains showed loss of one or more plasmid bands. None of the Lac- mutants showed loss of smaller plasmids (less than 5 MDa) indicating that smaller plasmids in lactococci are more stable under these conditions than larger plasmids (greater than 10 MDa). Concomitant loss of the Lac+ phenotype and plasmids by the method used in the present investigation may have application for isolating mutants devoid of one or more plasmids.  相似文献   

15.
Quantitative analyses of the utilization of amino acids by Lactococcus lactis subsp. cremoris FD1 in yeast extract medium (YE) and in casein peptone medium (CP) have been performed. Both free and peptide-bound amino acids were measured. In the CP most amino acids are peptide-bound and some amino acids are virtually only present in peptides. Thirty-six per cent of all peptide bonds in CP are hydrolysed during fermentation (6·3 mmol peptide bonds per gram biomass formed) and there is a transition of the growth rate related ATP consumption Y xATP (mmol ATP g biomass-1) from 25 mmol g-1 to 71 mmol g-1 coincident with a decrease of the peptide consumption. In YE most of the amino acids are on the free form and only 26% of the peptide bonds are hydrolysed during fermentation (1·5 mmol peptide bonds per gram biomass formed). A constant Y xATP= 38 mmol g-1 prevails throughout the fermentation in YE.  相似文献   

16.
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   

17.
Peptide transport in microorganisms is important for nutrition of the cell and various signalling processes including regulation of gene expression, sporulation, chemotaxis, competence and virulence development. Peptide transport is mediated via different combinations of ion-linked and ATP-binding cassette (ABC) transporters, the latter utilizing single or multiple peptide-binding proteins with overlapping specificities. The paradigm for research on peptide transport is Lactococcus lactis, in which the uptake of peptides containing essential amino acids is vital for growth on milk proteins. Differential expression and characteristics of peptide-binding proteins in several Lactococcus lactis strains resulted in apparent conflicts with older literature. Recent developments and new data now make the pieces of the puzzle fall back into place again and confirm the view that the oligopeptide-binding proteins determine the uptake selectivity of their cognate ABC transporters. Besides reviewing the current data on binding specificity and transport selectivity of peptide transporters in L. lactis, the possible implications for peptide utilization by other bacterial species are discussed.  相似文献   

18.
Peptidases and amino acid catabolism in lactic acid bacteria   总被引:28,自引:0,他引:28  
The conversion of peptides to free amino acids and their subsequent utilization is a central metabolic activity in prokaryotes. At least 16 peptidases from lactic acid bacteria (LAB) have been characterized biochemically and/or genetically. Among LAB, the peptidase systems of Lactobacillus helveticus and Lactococcus lactis have been examined in greatest detail. While there are homologous enzymes common to both systems, significant differences exist in the peptidase complement of these organisms. The characterization of single and multiple peptidase mutants indicate that these strains generally exhibit reduced specific growth rates in milk compared to the parental strains. LAB can also catabolize amino acids produced by peptide hydrolysis. While the catabolism of amino acids such as Arg, Thr, and His is well understood, few other amino acid catabolic pathways from lactic acid bacteria have been characterized in significant detail. Increasing research attention is being directed toward elucidating these pathways as well as characterizing their physiological and industrial significance.  相似文献   

19.
The cell envelope-associated proteinases (CEPs) of the lactobacilli have key roles in bacterial nutrition and contribute to the development of the organoleptic properties of fermented milk products as well, as they can release bioactive health-beneficial peptides from milk proteins. The influence of the peptide supply, carbohydrate source, and osmolites on the CEP activity of the cheese starter Lactobacillus delbrueckii subsp. lactis CRL 581 was investigated. The CEP activity levels were controlled by the peptide content of the growth medium. The maximum activity was observed in a basal minimal defined medium, whereas in the presence of Casitone, Casamino Acids, or yeast extract, the synthesis of CEP was inhibited 99-, 70-, and 68-fold, respectively. The addition of specific di- or tripeptides containing branched-chain amino acids, such as leucylleucine, prolylleucine, leucylglycylglycine, or leucylproline, to the growth medium negatively affected CEP activity, whereas dipeptides without branched-chain amino acids had no effect on the enzyme's production. The carbon source and osmolites did not affect CEP activity. The CEP of L. delbrueckii subsp. lactis CRL 581 exhibited a mixed-type CEP(I/III) variant caseinolytic specificity. Mass-spectrometric screening of the main peptide peaks isolated by reverse-phase high-pressure liquid chromatography allowed the identification of 33 and 32 peptides in the alpha(s1)- and beta-casein hydrolysates, respectively. By characterizing the peptide sequence in these hydrolysates, a pattern of alpha(s1)- and beta-casein breakdown was defined and is reported herein, this being the first report for a CEP of L. delbrueckii subsp. lactis. In this pattern, a series of potentially bioactive peptides (antihypertensive and phosphopeptides) which are encrypted within the precursor protein could be visualized.  相似文献   

20.
Regulation of Proteolytic Enzyme Activity in Lactococcus lactis   总被引:3,自引:1,他引:2       下载免费PDF全文
Two different Lactococcus lactis host strains, L. lactis subsp. lactis MG1363 and L. lactis subsp. cremoris SK1128, both containing plasmid pNZ521, which encodes the extracellular serine proteinase (PrtP) from strain SK110, were used to study the medium and growth-rate-dependent activity of three different enzymes involved in the proteolytic system of lactococci. The activity levels of PrtP and both the intracellular aminopeptidase PepN and the X-prolyl-dipeptidyl aminopeptidase PepXP were studied during batch and continuous cultivation. In both strains, the PrtP activity level was regulated by the peptide content of the medium. The highest activity level was found during growth in milk, and the lowest level was found during growth in the peptide-rich laboratory medium M17. Regulation of the intracellular peptidase activity appeared to be a strain-dependent phenomenon. In cells of strain MG1363, the activity levels of PepN and PepXP were regulated in a similar way to that observed for PrtP. In cells of strain SK1128, the levels of both peptidases were not significantly influenced by the peptide content of the medium. The presence of specific concentrations of the dipeptide prolylleucine could mimic the low activity levels of the regulated proteolytic enzymes, even to the activity level found on M17 medium. The effect of the presence of the dipeptide prolylleucine in the medium on the activity level of the regulated proteolytic enzymes was confirmed at fixed growth rates in chemostat cultures.  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