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.
2.
3.
4.
Liu CQ  Charoechai P  Khunajakr N  Deng YM  Widodo  Dunn NW 《Gene》2002,297(1-2):241-247
A plasmid-borne copper resistance operon (lco) was identified from Lactococcus lactis subsp. lactis LL58-1. The lco operon consists of three structural genes lcoABC. The predicted products of lcoA and lcoB were homologous to chromosomally encoded prolipoprotein diacylglyceral transferases and two uncharacterized proteins respectively, and the product of lcoC is similar to several multicopper oxidases, which are generally plasmid-encoded. This genetic organization represents a new combination of genes for copper resistance in bacteria. The three genes are co-transcribed from a copper-inducible promoter, which is controlled by lcoRS encoding a response regulator and a kinase sensor. The five genes are flanked by two insertion sequences, almost identical to IS-LL6 from L. lactis. Transposon mutagenesis and subcloning analysis indicated that the three structural genes were all required for copper resistance. Copper assay results showed that the extracellular concentration of copper of L. lactis LM0230 containing the lco operon was significantly higher than that of the host strain when copper was added at concentrations from 2 to 3 mM. The results suggest that the lco operon conferred copper resistance by reducing the intracellular accumulation of copper ions in L. lactis.  相似文献   

5.
Dihydroorotate dehydrogenases (DHODs) catalyze the oxidation of dihydroorotate to orotate in the only redox reaction in pyrimidine biosynthesis. The pyrimidine binding sites are very similar in all structurally characterized DHODs, suggesting that the prospects for identifying a class-specific inhibitor directed against this site are poor. Nonetheless, two compounds that bind specifically to the Class 1A DHOD from Lactococcus lactis, 3,4-dihydroxybenzoate (3,4-diOHB) and 3,5-dihydroxybenzoate (3,5-diOHB), have been identified [Palfey et al. (2001) J. Med. Chem. 44, 2861-2864]. The mechanism of inhibitor binding to the Class 1A DHOD from L. lactis has now been studied in detail and is reported here. Titrations showed that 3,4-diOHB binds more tightly at higher pH, whereas the opposite is true for 3,5-diOHB. Isothermal titration calorimetry and absorbance spectroscopy showed that 3,4-diOHB ionizes to the phenolate upon binding to the enzyme, but 3,5-diOHB does not. The charge-transfer band that forms in the 3,4-diOHB complex allowed the kinetics of binding to be observed in stopped-flow experiments. Binding was slow enough to observe from pH 6 to pH 8 and was (minimally) a two-step process consisting of the rapid formation of a complex that isomerized to the final charge-transfer complex. Orotate and 3,5-diOHB bind too quickly to follow directly, but their dissociation kinetics were studied by competition and described adequately with a single step. Crystal structures of both inhibitor complexes were determined, showing that 3,5-diOHB binds in the same orientation as orotate. In contrast, 3,4-diOHB binds in a twisted orientation, enabling one of its phenolic oxygens to form a very strong hydrogen bond to an asparagine, thus stabilizing the phenolate and causing charge-transfer interactions with the pi-system of the flavin, resulting in a green color.  相似文献   

6.
The function of the Lactococcus lactis bacteriophage bIL66 middle time-expressed operon (M-operon), involved in sensitivity to the abortive infection mechanism AbiD1, was examined. Expression of the M-operon is detrimental to Escherichia coli cells, induces the SOS response and is lethal to recA and recBC E. coli mutants, which are both deficient in recombinational repair of chromosomal double-stranded breaks (DSBs). The use of an inducible expression system allowed us to demonstrate that the M-operon-encoded proteins generate a limited number of randomly distributed chromosomal DSBs that are substrates for ExoV-mediated DNA degradation. DSBs were also shown to occur upstream of the replication initiation point of unidirectionally theta-replicating plasmids. The characteristics of the DSBs lead us to propose that the endonucleolytic activity of the M-operon is not specific to DNA sequence, but rather to branched DNA structures. Genetic and physical analysis performed with different derivatives of the M-operon indicated that two orf s ( orf2 and orf3 ) are needed for nucleolytic activity. The orf3 product has amino acid homology with the E. coli RuvC Holliday junction resolvase. By site-specific mutagenesis, we have shown that one of the amino acid residues constituting the active centre of RuvC enzyme (Glu-66) and conserved in ORF3 (Glu-67) is essential for the nucleolytic activity of the M-operon gene product(s). We therefore propose that orf2 and orf3 of the M-operon code for a structure-specific endonuclease (M-nuclease), which might be essential for phage multiplication.  相似文献   

7.
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   

8.
9.
10.
11.
Summary Cell wall-associated proteinases were isolated from Lactococcus lactis subsp. cremoris AC1 and subsp. lactis NCDO 763 in order to compare their specificities towards different caseins. Two purification strategies were applied. Cells grown in casein-free M17 medium were a suitable starting material for purification, since electrophoretic purity could be achieved after one chromatographic step. Both enzymes has an apparent molecular mass of about 145000 daltons as judged by sodium dodecyl sulphate-polyacrylamide gel electrophoresis. Electrophoresis and reversed phase HPLC patterns of hydrolysates of s1-, s2-, -, and K-caseins indicated that both proteinases had a similar specificity. The enzyme of L. lactis subsp. lactis split s1- and s2-caseins more extensively than that of L. lactis subsp. cremoris.  相似文献   

12.
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   

13.
Summary The cell wall proteinases of Lactococcus lactis subsp. lactis NCDO 763 and L. lactis subsp. cremoris AC1 hydrolyse -casein with a similar specificity even though some quantitative differences can be observed for a few degradation products analysed by reverse phase HPLC and sodium dodecyl sulphate-polyacrylamide gel electrophoresis. The main peptides soluble in 1.1% trifluoroacetic acid and liberated by the two proteinases were identified and have been found to be the same for the two enzymes. They are located in two areas of the -casein sequence (53–93 and the C-terminal part: 129–209) and they include bitter tasting or physiologically active fragments. No narrow specificity was observed for these proteinases. However, glutamine and serine residues are more frequently encountered in position P1 and P1 of the sensitive peptide bond and the close environment (position P2 to P4 and P2 to P4) of the cleaved bond is mainly hydrophobic.  相似文献   

14.
15.
The Lactococcus lactis sex-factor aggregation gene cluA   总被引:4,自引:0,他引:4  
A gene, cluA, was cloned from the chromosomally located sex factor of Lactococcus lactis MG1363. Sequence analysis revealed significant homology with previously described aggregation proteins in Enterococcus and Streptococcus species. The possibility that cluA was an equivalent protein involved in cell aggregation between donor and recipient bacteria during lactococcal conjugation was confirmed by its expression under the control of a heterologous promoter in L. lactis. Analysis of the homology between the CluA protein and the related proteins of Enterococcus and Streptococcus allowed a common structure for these proteins to be postulated. This consisted of five domains. Functionally conserved domains I and V act respectively as a secretary leader and C-terminal membrane anchor. Domains II and IV are conserved at the amino acid level and probably have common structural roles whereas domain III is variable and may control binding specificity.  相似文献   

16.
dTDP-rhamnose is an important precursor of cell wall polysaccharides and rhamnose-containing exopolysaccharides (EPS) in Lactococcus lactis. We cloned the rfbACBD operon from L. lactis MG1363, which comprises four genes involved in dTDP-rhamnose biosynthesis. When expressed in Escherichia coli, the lactococcal rfbACBD genes could sustain heterologous production of the Shigella flexneri O antigen, providing evidence of their functionality. Overproduction of the RfbAC proteins in L. lactis resulted in doubled dTDP-rhamnose levels, indicating that the endogenous RfbAC activities control the intracellular dTDP-rhamnose biosynthesis rate. However, RfbAC overproduction did not affect rhamnose-containing B40-EPS production levels. A nisin-controlled conditional RfbBD mutant was unable to grow in media lacking the inducer nisin, indicating that the rfb genes have an essential role in L. lactis. Limitation of RfbBD activities resulted in the production of altered EPS. The monomeric sugar of the altered EPS consisted of glucose, galactose, and rhamnose at a molar ratio of 1:0.3:0.2, which is clearly different from the ratio in the native sugar. Biophysical analysis revealed a fourfold-greater molecular mass and a twofold-smaller radius of gyration for the altered EPS, indicating that these EPS are more flexible polymers with changed viscosifying properties. This is the first indication that enzyme activity at the level of central carbohydrate metabolism affects EPS composition.  相似文献   

17.
Phage bIL66 is unable to grow on Lactococcus lactis cells harboring the abortive infection gene abiD1. Spontaneous phage mutants able to grow on AbiD1 cells were used to study phage-Abi interaction. A 1.33-kb DNA segment of a mutant phage allowed growth of AbiD1s phages in AbiD1 cells when present in trans. Sequence analysis of this segment revealed an operon composed of four open reading frames, designated orf1 to orf4. The operon is transcribed 10 min after infection from a promoter presenting an extended -10 consensus sequence but no -35 sequence. Analysis of four independent AbiD1r mutants revealed a different point mutation localized in orf1, implying that this open reading frame is needed for sensitivity to AbiD1. However, the sensitivity is partly suppressed when orf3 is expressed in trans on a high-copy-number plasmid, suggesting that AbiD1 acts by decreasing the concentration of an available orf3 product.  相似文献   

18.
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   

19.
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   

20.
Characterization by partial 16S rRNA gene sequencing, ribotyping, and green fluorescent protein-based nisin bioassay revealed that 6 of 20 human milk samples contained nisin-producing Lactococcus lactis bacteria. This suggests that the history of humans consuming nisin is older than the tradition of consuming fermented milk products.  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