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.
Heat shock inLactococcus lactis subsp.lactis may induce as many as 16 proteins after a temperature shift from 30° to 40°C. Five induced proteins were found to be immunologically related to theEscherichia coli GroEL, DnaK, DnaJ, and GrpE proteins, and to theBacillus subtilis 43 factor. From these initial studies we conclude that, inL. lactis subsp.lactis, a heat shock response similar to that known to occur in other prokaryotes might exist.  相似文献   

2.
The lactic acid bacterium,Lactococcus lactis subsp.lactis IL1403 was subjected to defferent cold temperatures for various times. Physiological experiments showed that this strain had an improved survival capacity in stationary phase as the temperature decreased. Two-dimensional electrophoresis of proteins extracted from cold-temperature exposed cultures showed that a dozen proteins are overexpressed up to threefold compared with exposure at 30°C. Most of these proteins are overexpressed first, temporarily and second, in the first 10 h after the transfer to 8°C. These observations indicate that response to cold stress inL. lactis subsp.lactis is an active phenomenon.  相似文献   

3.
A nucleoside N-deoxyribosyltransferase-homologous gene was detected by homological search in the genomic DNA of Lactococcus lactis subsp. lactis. The gene yejD is composed of 477 nucleotides encoding 159 amino acids with only 25% identity, which is low in comparison to the amino acid sequences of the N-deoxyribosyltransferases from other lactic acid bacteria, i.e. Lactobacillus leichmannii and Lactobacillus helveticus. The residues responsible for catalytic and substrate-binding sites in known enzymes are conserved at Gln49, Asp73, Asp93 (or Asp95), and Glu101, respectively. The recombinant YejD expressed in Escherichia coli shows a 2-deoxyribosyl transfer activity to and from both bases of purine and pyrimidine, showing that YejD should be categorized as a class II N-deoxyribosyltransferase. Interestingly, the base-exchange activity as well as the heat stability of YejD was enhanced by the presence of monovalent cations such as K(+), NH(4)(+), and Rb(+), indicating that the Lactococcus enzyme is a K(+)-activated Type II enzyme. However, divalent cations including Mg(2+) and Ca(2+) significantly inhibit the activity. Whether or not the yejD gene product actually participates in the nucleoside salvage pathway of Lc. lactis remains unclear, but the lactic acid bacterium possesses the gene coding for the nucleoside N-deoxyribosyltransferase activated by K(+) on its genome.  相似文献   

4.
Relatedness between Lactococcus lactis subsp. cremoris and L. lactis subsp. lactis was assessed by Southern hybridization analysis, with cloned chromosomal genes as probes. The results indicate that strains of the two subspecies form two distinct groups and that the DNA sequence divergence between L. lactis subsp. lactis and L. lactis subsp. cremoris is estimated to be between 20 and 30%. The previously used phenotypic criteria do not fully discriminate between the groups; therefore, we propose a new classification which is based on DNA homology. In agreement with this revised classification, the L. lactis subsp. lactis and L. lactis subsp. cremoris strains from our collection have distinct phage sensitivities.  相似文献   

5.
Unlike other lactic acid bacteria, Lactococcus lactis subsp. lactis NCDO 2118 was able to grow in a medium lacking glutamate and the amino acids of the glutamate family. Growth in such a medium proceeded after a lag phase of about 2 days and with a reduced growth rate (0.11 h−1) compared to that in the reference medium containing glutamate (0.16 h−1). The enzymatic studies showed that a phosphoenolpyruvate carboxylase activity was present, while the malic enzyme and the enzymes of the glyoxylic shunt were not detected. As in most anaerobic bacteria, no α-ketoglutarate dehydrogenase activity could be detected, and the citric acid cycle was restricted to a reductive pathway leading to succinate formation and an oxidative branch enabling the synthesis of α-ketoglutarate. The metabolic bottleneck responsible for the limited growth rate was located in this latter pathway. As regards the synthesis of glutamate from α-ketoglutarate, no glutamate dehydrogenase was detected. While the glutamate synthase-glutamine synthetase system was detected at a low level, high transaminase activity was measured. The conversion of α-ketoglutarate to glutamate by the transaminase, the reverse of the normal physiological direction, operated with different amino acids as nitrogen donor. All of the enzymes assayed were shown to be constitutive.  相似文献   

6.
7.
Resistance to a broad class of isometric bacteriophages that infect strains of Lactococcus lactis has been engineered into a dairy starter by expression of antisense mRNA targeted against a conserved bacteriophage gene. Maximum protection is obtained only when the entire 1,654-bp coding sequence for a 51-kDa protein is positioned in the antisense orientation with respect to a promoter sequence that functions in L. lactis subsp. lactis. Expression of the antisense mRNA results in more than 99% reduction of the total number of PFU. Plaques that do form are characterized by their relatively small size and irregular shape. A variety of truncated genes, including the open reading frame expressed in the sense orientation, fail to provide any significant measure of resistance as compared with that of the intact open reading frame. Southern hybridization with probes specific for the conserved region reveal that the [ill] plasmid constructs are maintained despite the presence of a large complement of other indigenous plasmids. Strains harboring the antisense mRNA plasmid construct grow and produce acid at a rate equivalent to that of the host strain alone, suggesting that antisense expression is not deleterious to normal cellular metabolism.  相似文献   

8.
Histidine biosynthesis genes in Lactococcus lactis subsp. lactis.   总被引:9,自引:5,他引:4       下载免费PDF全文
The genes of Lactococcus lactis subsp. lactis involved in histidine biosynthesis were cloned and characterized by complementation of Escherichia coli and Bacillus subtilis mutants and DNA sequencing. Complementation of E. coli hisA, hisB, hisC, hisD, hisF, hisG, and hisIE genes and the B. subtilis hisH gene (the E. coli hisC equivalent) allowed localization of the corresponding lactococcal genes. Nucleotide sequence analysis of the 11.5-kb lactococcal region revealed 14 open reading frames (ORFs), 12 of which might form an operon. The putative operon includes eight ORFs which encode proteins homologous to enzymes involved in histidine biosynthesis. The operon also contains (i) an ORF encoding a protein homologous to the histidyl-tRNA synthetases but lacking a motif implicated in synthetase activity, which suggests that it has a role different from tRNA aminoacylation, and (ii) an ORF encoding a protein that is homologous to the 3'-aminoglycoside phosphotransferases but does not confer antibiotic resistance. The remaining ORFs specify products which have no homology with proteins in the EMBL and GenBank data bases.  相似文献   

9.
A plasmid known to be associated with mupirocin resistance of Staphylococcus aureus has been isolated and a restriction enzyme map constructed. An EcoRI fragment of 4.05 kb from this plasmid has been cloned into an Escherichia coli-Staphylococcus aureus shuttle vector and shown to carry the gene for resistance to mupirocin. The DNA sequence of a small section of the gene has been determined and the derived amino acid sequence compared with a data bank. The amino acid sequence is identical for eight amino acids with the sequence of isoleucyl tRNA synthetase of E. coli. This finding adds to the evidence that mupirocin resistance is the result of a modified isoleucyl tRNA synthetase.  相似文献   

10.
Carbohydrate-starved cultures of Lactococcus lactis subsp. lactis IL1403 showed enhanced resistance to heat, ethanol, acid, osmotic, and oxidative stresses. This cross-protection seems to be established progressively during the transitional growth phase, with maximum resistance occurring when cells enter the stationary phase. Chloramphenicol or rifamycin treatment does not abolish the development of a tolerant cell state but, on the contrary, seems to provoke this response in L. lactis subsp. lactis.  相似文献   

11.
The occurrence of the acmA gene, encoding the lactococcal N-acetylmuramidase in new lactococcal isolates from raw milk cheeses, has been determined. Isolates were genotypically identified to the subspecies level with a PCR technique. On the basis of PCR amplification of the acmA gene, the presence or absence of an additional amplicon of approximately 700 bp correlated with Lactococcus lactis subspecies. L. lactis subsp. lactis exhibits both the expected 1,131-bp product and the additional amplicon, whereas L. lactis subsp. cremoris exhibits a single 1,131-bp fragment.  相似文献   

12.
DNA fragments with promoter activity were isolated from the chromosome of Lactococcus lactis subsp. lactis. For the isolation, a promoter probe vector based on the cat gene was constructed, which allowed direct selection with chloramphenicol in Bacillus subtilis and L. lactis. Four of the putative promoters (P1, P2, P10, and P21) were analyzed further by sequencing, mapping of the 5' end of the mRNA, Northern (RNA blot) hybridization, and chloramphenicol acetyltransferase activity measurements. From these fragments, -10 and -35 regions resembling the consensus Escherichia coli sigma 70 and B. subtilis sigma 43 promoters were identified. Another set of promoters, together with a signal sequence, were also isolated from the same organism. These fragments promoted secretion of TEM beta-lactamase from L. lactis. When the two sets of promoters were compared, it was found that the ones isolated with the cat vector were more efficient (produced more mRNA). By changing the promoter part of the promoter-signal sequence fragment giving the best TEM beta-lactamase secretion into a more efficient one (P2), a 10-fold increase in enzyme production was obtained.  相似文献   

13.
The lactic acid tolerance response (LATR) of the lactic acid bacterium Lactococcus lactis subsp. lactis has been studied. A dramatic increase in survival to a severe acid stress (pH 3.9) was obtained by preexposing the cells for 30 min to a mildly acid shock at pH 5.5. Whole-cell protein extract analysis revealed that during the acid tolerance response 33 polypeptides are induced over the level of naive cells. Among these are the major heat shock proteins DnaK and GroEL. In conjunction with a previous report (Hartke et al. 1994), the results establish that L. lactis can adapt to lactic acid exposure in two different ways: a logarithmic phase LATR, which may be activated by protons, and a stationary-phase LATR, which needs no activation by protons. Both systems are independent of de novo protein synthesis. Received: 8 February 1996 / Accepted: 11 March 1996  相似文献   

14.
DNA fragments with promoter activity were isolated from the chromosome of Lactococcus lactis subsp. lactis. For the isolation, a promoter probe vector based on the cat gene was constructed, which allowed direct selection with chloramphenicol in Bacillus subtilis and L. lactis. Four of the putative promoters (P1, P2, P10, and P21) were analyzed further by sequencing, mapping of the 5' end of the mRNA, Northern (RNA blot) hybridization, and chloramphenicol acetyltransferase activity measurements. From these fragments, -10 and -35 regions resembling the consensus Escherichia coli sigma 70 and B. subtilis sigma 43 promoters were identified. Another set of promoters, together with a signal sequence, were also isolated from the same organism. These fragments promoted secretion of TEM beta-lactamase from L. lactis. When the two sets of promoters were compared, it was found that the ones isolated with the cat vector were more efficient (produced more mRNA). By changing the promoter part of the promoter-signal sequence fragment giving the best TEM beta-lactamase secretion into a more efficient one (P2), a 10-fold increase in enzyme production was obtained.  相似文献   

15.
J L Steele  L L McKay 《Plasmid》1989,22(1):32-43
Conjugal transfer of genetic material by Lactococcus lactis subsp. lactis 11007 was examined. A plasmid of 88 MDa (pJS88) was identified in addition to the previously reported conjugally transferred plasmids of 32 (pKB32) and 4.8 MDa. Proteinase activity, reduced bacteriophage sensitivity, bacteriocin resistance, and conjugal transfer ability were encoded by pJS88. The ability to metabolize lactose (Lac+) was encoded by pKB32, and the 4.8-MDa plasmid was cryptic. When a strain containing both pKB32 and pJS88 was mated with a recipient deficient in host-mediated homologous recombination (Rec-), a plasmid of 40 MDa (pJS40) was observed in approximately 50% of the Lac+ transconjugants. DNA-DNA hybridization results indicated that pJS40 contained homology with both pKB32 and pJS88. These results indicated that pKB32 was conjugally transferred via conduction and suggested that pJS40 is a deletion derivative of a pKB32::pJS88 cointegrate. A Rec- strain containing pKB32 and pJS88 mediated Lac+ conjugal transfer, suggesting that the pKB32::pJS88 cointegrate could form via a rec-independent event. Resolution of the pKB32::pJS88 cointegrate was observed in both Rec- and Rec+ hosts. Cointegrate formation and resolution via rec-independent mechanisms suggest the involvement of a transposable element in the Tn3 family.  相似文献   

16.
Bacterial isolates from bean-sprouts were screened for anti- Listeria monocytogenes bacteriocins using a well diffusion method. Thirty-four of 72 isolates inhibited the growth of L.monocytogenes Scott A. One, HPB 1688, which had the biggest inhibition zone against L.monocytogenes Scott A, was selected for subsequent analysis. Both ribotyping and DNAsequencing of 16S ribosomal RNA gene demonstrated that the isolate was Lactococcus lactis subsp. lactis . Polymerase chain reaction and nucleotide sequencing revealed that thegenomic DNA of the bean-sprout isolates contained a nisin Z structural gene. In MRS broth,bean-sprout isolate HPB 1688 survived at 3–4·5°C for at least 20 d, grew at 4°Cand produced anti-listerial compoundsat 5°C. When co-cultured with L. monocytogenes in MRS broth, the isolate inhibited thegrowth of L. monocytogenes at 4°C after 14d and at 10°C after 2 d. When co-inoculatedwith 102cells g−1 of L.monocytogenes on fresh-cut ready-to-eat Caesar salad, L. lactis subsp. lactis (108cells g−1) was able to reduce the number of L. monocytogenes by 1–1·4 logs after storage for 10 d at 7° and 10°C. A bacteriocin-producing Enterococcusfaecium was also able to reduce the numbers of L. monocytogenes onCaesar salad, butdid not act synergistically when co-inoculated with L. lactis subsp. lactis .  相似文献   

17.
The heat shock response in Lactococcus lactis subsp. lactis was characterized with respect to synthesis of a unique set of proteins induced by thermal stress. A shift in temperature from 30 to 42°C was sufficient to arrest the growth of L. lactis subsp. lactis, but growth resumed after a shift back to 30°C. Heat shock at 50°C reduced the viable cell population by 103; however, pretreatment of the cells at 42°C made them more thermoresistant to exposure at 50°C. The enhanced synthesis of approximately 13 proteins was observed in cells labeled with 35S upon heat shock at 42°C. Of these heat shock-induced proteins, two appeared to be homologs of GroEL and DnaK, based on their molecular weights and reactivity with antiserum against the corresponding Escherichia coli proteins. Therefore, we conclude that L. lactis subsp. lactis displays a heat shock response similar to that observed in other mesophilic bacteria.  相似文献   

18.
It is now generally recognized that cell growth conditions in nature are often suboptimal compared to controlled conditions provided in the laboratory. Natural stresses like starvation and acidity are generated by cell growth itself. Other stresses like temperature or osmotic shock, or oxygen, are imposed by the environment. It is now clear that defense mechanisms to withstand different stresses must be present in all organisms. The exploration of stress responses in lactic acid bacteria has just begun. Several stress response genes have been revealed through homologies with known genes in other organisms. While stress response genes appear to be highly conserved, however, their regulation may not be. Thus, search of the regulation of stress response in lactic acid bacteria may reveal new regulatory circuits. The first part of this report addresses the available information on stress response in Lactococcus lactis.Acid stress response may be particularly important in lactic acid bacteria, whose growth and transition to stationary phase is accompanied by the production of lactic acid, which results in acidification of the media, arrest of cell multiplication, and possible cell death. The second part of this report will focus on progress made in acid stress response, particularly in L. lactis and on factors which may affect its regulation. Acid tolerance is presently under study in L. lactis. Our results with strain MG1363 show that it survives a lethal challenge at pH 4.0 if adapted briefly (5 to 15 minutes) at a pH between 4.5 and 6.5. Adaptation requires protein synthesis, indicating that acid conditions induce expression of newly synthesized genes. These results show that L. lactis possesses an inducible response to acid stress in exponential phase.To identify possible regulatory genes involved in acid stress response, we determined low pH conditions in which MG1363 is unable to grow, and selected at 37°C for transposition insertional mutants which were able to survive. About thirty mutants resistant to low pH conditions were characterized. The interrupted genes were identified by sequence homology with known genes. One insertion interrupts ahrC, the putative regulator of arginine metabolism; possibly, increased arginine catabolism in the mutant produces metabolites which increase the pH. Several other mutations putatively map at some step in the pathway of (p)ppGpp synthesis. Our results suggest that the stringent response pathway, which is involved in starvation and stationary phase survival, may also be implicated in acid pH tolerance.  相似文献   

19.
To apply recombinant DNA techniques for genetic manipulation of the industrially important lactococci, an efficient and reliable high-frequency transformation system must be available. High-voltage electric pulses have been demonstrated to enhance uptake of DNA into protoplasts and intact cells of numerous gram-negative and gram-positive microorganisms. The objective of this study was to develop a system for electroporating intact cells of Lactococcus lactis subsp. lactis LM0230 (previously designated Streptococcus lactis LM0230) with a commercially available electroporation unit (BTX Transfector 100; BTX, Inc., San Diego, Calif.). Parameters which influenced the efficiency of transformation included growth phase and final concentration of cells, ionic strength of the suspending medium, concentration of plasmid DNA, and the amplitude and duration of the pulse. Washed suspensions of intact cells suspended in deionized distilled water were subjected to one high-voltage electric pulse varying in voltage (300 to 900 V corresponding to field strengths of 5 to 17 kV/cm) and duration (100 microseconds to 1 s). Transformation efficiencies of 10(3) transformants per microgram of DNA were obtained when dense suspensions (final concentration, 5 x 10(10) CFU/ml) of stationary-phase cells were subjected to one pulse with a peak voltage of 900 V (field strength, 17 kV/cm) and a pulse duration of 5 ms in the presence of plasmid DNA. Dilution of porated cells in broth medium followed by an expression period of 2 h at 30 degrees C was beneficial in enhancing transformation efficiencies. Plasmids ranging in size from 9.8 to 30.0 kilobase pairs could be transformed by this procedure.  相似文献   

20.
Lactococcus lactis subsp.lactis 484 produced a proteinaceous antibacterial substance designated as lactococcin capable of inhibiting members of theLactococcus group,Bacillus cereus, Staphylococcus aureus, andSalmonella typhi. Growth of this culture in the presence of 2–30 g/ml of ethidium bromide or acriflavin or novobiocin, and at elevated temperature (39° and 41°C), could not produce any lactococcin-negative (Lap) variants. However, protoplast-induced curing with lysozyme was successful in developing Lap derivatives. Two types of cured derivatives, namely Lac Lap+ and Lac Lap, were obtained. Lap variants were also lacking sucrose-fermenting ability (Suc+) and lactococcin resistance (Lapr). The lactose-negative (Lac) variants and Lap+ were clearly lacking the largest (65 Md) plasmid. However, Lap Suc Laps variants lost a 2 Md plasmid.L. lactis subsp.lactis 484 transferred lactose-fermenting ability as well as Lap+ Suc+ Lapr phenotypes simultaneously toL. lactis subsp.lactis LM 2306 and LM 0230 by surface mating at a frequency of 10–4 and 10–1 per donor respectively. However, cured Lac Lap transconjugants could not transfer Lac+ Lap+ Suc+ Lapr phenotypes to any of these recipient strains. Our results indicate that Lac+ and Lap+ Suc+ Lapr phenotypes are associated with 65 Md and 2 Md plasmids respectively. Conjugal transfer of 2 Md plasmid is possible only in the presence of a conjugative 65 Md plasmid.  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