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.
Bacillus licheniformis WBL-3, one of poly-γ-glutamic acid (γ-PGA) producers, depends on the existence of glutamate in the medium. In this paper, γ-PGA synthetase complex gene (pgsBCA) was cloned from Bacillus licheniformis WBL-3. pgsBCA gene of B. licheniformis WBL-3 was highly homologous with pgs-BCA gene of B. licheniformis 14580. The similarity was 97%, but the similarity of pgsBCA gene between B. licheniformis WBL-3 and Bacillus subtilis IFO3336 was only 74%. However, when pgsBCA was expressed in Escherichia coli, the E. coli clone produced γ-PGA extracellularly. The yield of γ-PGA was 8.624 g/l. This result infers that B. licheniformis and B. subtilis has the similar γ-PGA biosynthesis mechanism, namely, glutamic acid is catalyzed by an ATP-dependent amide ligase to synthesize γ-PGA.  相似文献   

2.

Background

Bacillus licheniformis is a Gram-positive, spore-forming soil bacterium that is used in the biotechnology industry to manufacture enzymes, antibiotics, biochemicals and consumer products. This species is closely related to the well studied model organism Bacillus subtilis, and produces an assortment of extracellular enzymes that may contribute to nutrient cycling in nature.

Results

We determined the complete nucleotide sequence of the B. licheniformis ATCC 14580 genome which comprises a circular chromosome of 4,222,336 base-pairs (bp) containing 4,208 predicted protein-coding genes with an average size of 873 bp, seven rRNA operons, and 72 tRNA genes. The B. licheniformis chromosome contains large regions that are colinear with the genomes of B. subtilis and Bacillus halodurans, and approximately 80% of the predicted B. licheniformis coding sequences have B. subtilis orthologs.

Conclusions

Despite the unmistakable organizational similarities between the B. licheniformis and B. subtilis genomes, there are notable differences in the numbers and locations of prophages, transposable elements and a number of extracellular enzymes and secondary metabolic pathway operons that distinguish these species. Differences include a region of more than 80 kilobases (kb) that comprises a cluster of polyketide synthase genes and a second operon of 38 kb encoding plipastatin synthase enzymes that are absent in the B. licheniformis genome. The availability of a completed genome sequence for B. licheniformis should facilitate the design and construction of improved industrial strains and allow for comparative genomics and evolutionary studies within this group of Bacillaceae.  相似文献   

3.
Yeast identification in grape juice concentrates from Argentina   总被引:1,自引:0,他引:1  
Aims: The purpose of this study was to identify yeast species present in spoiled and unspoiled grape juice concentrates from Argentine industries. Methods and Results: Osmophilic and osmotolerant yeasts were isolated from spoiled – visually effervescent – and unspoiled – without any visible damage – grape juice concentrates by the spread‐plate technique in two culture media. Yeast identification was done by classical and molecular methods. Zygosaccharomyces rouxii was the only species isolated from spoiled grape juice concentrates. In unspoiled samples, five different species were identified: Z. rouxii was isolated at a higher frequency, followed in decreasing order by Saccharomyces cerevisiae, Schizosaccharomyces pombe, Pichia anomala and Kluyveromyces delphensis. Conclusions: Yeasts isolated from grape juice concentrates were characterized by a limited taxonomic diversity, where Z. rouxii was the main species isolated. Significance and Impact of the Study: Grape production in Argentina is mainly devoted to the industry where wine and grape juice concentrates represent major types of commercial products. Little information on common yeast contaminants is available for grape juice concentrates. This study constitutes the first report of osmophilic yeast species present in spoiled and unspoiled grape juice concentrates elaborated in Argentina.  相似文献   

4.
Bacilli of the species Bacillus subtilis, B. pumilus, B. mycoides, B. marinus and B. licheniformis (a total of 53 strains) were isolated from 15 invertebrate species and the water of the Vostok Bay, Peter the Great Bay, Sea of Japan. Bacilli were most often isolated from bivalves (22.7%) and sea cucumbers (18.9%); they occurred less frequently in sea urchins and starfish (13.2 and 7.5%, respectively). Most of bacilli strains were isolated from invertebrates inhabiting silted sediments. No Bacillus spp. strains were isolated from invertebrates inhabiting stony and sandy environments. The species diversity of bacilli isolated from marine objects under study was low. Almost all bacterial isolates were resistant to lincomycin. Unlike B. pumilus, B. subtilis isolates were mostly resistant to benzylpenicillin and ampicillin. Antibiotic sensitivity of B. licheniformis strains was variable (two strains were resistant to benzylpenicillin and oxacillin, while one was sensitive). A significant fraction of isolated bacilli contained pigments. Pigmented strains were more often isolated from seawater samples, while colorless ones predominated within hydrobionts. B. subtilis colonies had the broadest range of colors. In the Bacillus strains obtained, DNase, RNase, phosphatase, elastolytic, chitinase, and agarolytic activity was detected. Bacilli strains with hydrolytic activity occurred in invertebrates more often than in seawater.  相似文献   

5.
Conjugative shuttle vectors of the pKVM series, based on an IncP transfer origin and the pMAD vector with a temperature sensitive replication were constructed to establish a markerless gene deletion protocol for Bacilli without natural competence such as the exoenzyme producer Bacillus licheniformis. The pKVM plasmids can be conjugated to strains of B. licheniformis and B. subtilis. For chromosomal gene deletion, regions flanking the target gene are fused and cloned in a pKVM vector prior to conjugative transfer from Escherichia coli to B. licheniformis. Appropriate markers on the vector backbone allow for the identification of the integration at the target locus and thereafter the vector excision, both events taking place via homologous recombination. The functionality of the deletion system was demonstrated with B. licheniformis by a markerless 939 bp in-frame deletion of the yqfD gene and the deletion of a 31 kbp genomic segment carrying a PBSX-like prophage.  相似文献   

6.
To obtain large quantities of glutamic acid-specific protease isolated originally from Bacillus licheniformis (BLase), an expression plasmid was constructed by inserting the BLase gene into a plasmid vector (pUB110) for Bacillus subtilis. B. subtilis strain ISW1214 harboring the resultant recombinant plasmid containing the coding and 5′-promoter and 3′-terminator regions of BLase gene secreted approximately 0.25 g/l of BLase in a culture medium contained in a 90-l jar fermentor, corresponding to nearly 10 times the natural production level and resulting in a stable large-scale production. The amount of BLase in the culture medium accounted for roughly 60% of the total extracellular proteins secreted from the recombinant strain, simplifying enzyme purification.  相似文献   

7.
2,3-Butanediol (2,3-BD) is a valuable bulk chemical with particular use in industry. 2,3-BD has a potential as solvent and fuel additive, as carrier for pharmaceuticals, or as feedstock for the production of synthetic rubber. Until now, the highest 2,3-BD concentrations were obtained with risk group 2 microorganisms (e.g., Klebsiella oxytoca). In this study, the nonpathogenic bacterium Bacillus licheniformis DSM 8785 was used for 2,3-BD production from glucose. In batch experiments, a maximum 2,3-BD concentration of 72.6 g/L was reached from 180 g/L glucose after 86 h. The yield was 0.42 g/g glucose and the productivity was 0.86 g/(L h). During fed-batch cultivation, 2,3-BD production could be increased up to 144.7 g/L, with a productivity of 1.14 g/(L h). Additionally, repeated batch/fed-batch experiments were conducted using immobilized B. licheniformis in the form of LentiKats®. Results showed a high activity and stability of the immobilizates even after multiple medium replacements, as well as 2,3-BD concentrations, yields, and productivities similar to those obtained with free cells. To our knowledge, these results show the highest 2,3-BD concentration reported so far using a risk group 1 microorganism in general and B. licheniformis in particular. Furthermore, productivity lies in the same range with data reported from risk group 2 strains, which makes B. licheniformis DSM 8785 a suitable candidate for large-scale fermentation processes.  相似文献   

8.
A purified preparation of Bacillus licheniformis α-amylase was immunologeeally and electrophoretically compared with commercial crystalline α-amylase of Bacillus subtilis. The former enzyme reacted completely with rabbit antiserum to the same enzyme showing a single precipitin band, and moved toward the cathode in immuno-electrophoresis on agarose at pH 9.6. On the contrary, crystalline α-amylase of Bacillus subtilis migrated to the anode in immunoelectrophoresis at pH 8.6, though it weakly cross-reacted with the antiserum, suggesting that amylases of Bacillus licheniformis and Bacillus subtilis are not identical. In addition, the neutralization test of amylase activity showed that α-amylase of Bacillus licheniformis was much more susceptible to inhibition by the serum than was Bacillus subtilis α-amylase. Each of four species of Bacillus licheniformis α-amylase extracted from the sliced discs after disc electrophoresis on polyacrylamide gel was distinct from the others by showing individual migratory rate, but they were antigenically similar to each other and to the parent enzyme.  相似文献   

9.
The replication terminator protein (RTP) of Bacillus subtilis interacts with its cognate DNA terminators to cause replication fork arrest, thereby ensuring that the forks approaching one another at the conclusion of a round of replication meet within a restricted terminus region. A similar situation exists in Escherichia coli, but it appears that the fork-arrest systems in these two organisms have evolved independently of one another. In the present work, RTP homologs in four species closely related to B. subtilis (B. atrophaeus, B. amyloliquefaciens, B. mojavensis, and B. vallismortis) have been identified and characterized. An RTP homolog could not be identified in another closely related species, B. licheniformis. The nucleotide and amino acid changes from B. subtilis among the four homologs are consistent with the recently established phylogenetic tree for these species. The GC contents of the rtp genes raise the possibility that these organisms arose within this branch of the tree by horizontal transfer into a common ancestor after their divergence from B. licheniformis. Only 5 amino acid residue positions were changed among the four homologs, despite an up to 17.2% change in the nucleotide sequence, a finding that highlights the importance of the precise folded structure to the functioning of RTP. The absence of any significant change in the proposed DNA-binding region of RTP emphasizes the importance of its high affinity for the DNA terminator in its functioning. By coincidence, the single change (E30K) found in the B. mojavensis RTP corresponds exactly to that purposefully introduced by others into B. subtilis RTP to implicate a crucial role for E30 in the fork-arrest mechanism. The natural occurrence of this variant is difficult to reconcile with such an implication, and it was shown directly that RTP.E30K functions normally in fork arrest in B. subtilis in vivo. Additional DNA terminators were identified in the new RTP homolog-containing strains, allowing the definition of a Bacillus terminator consensus and identification of two more terminators in the B. subtilis 168 genome sequence to bring the total to nine.  相似文献   

10.
Nucleotide Sequence and Features of the Bacillus licheniformis gnt Operon   总被引:1,自引:0,他引:1  
Bacillus licheniformis was able to utilize gluconate as thesole carbon source as efficiently as Bacillus subtilis did.Southern analysis indicated that B. licheniformis likely possessesonly one gnt determinant. The nucleotide sequence (6278 bp)of the B. licheniformis DNA containing the gnt operon was determined,revealing the five complete open reading frames (ORF; genes).The putative product of the first gene, oug, did not show anysignificant homology to known proteins, but those of the secondto fifth genes exhibited striking homology to the gntRKPZ genesof B. subtilis, respectively, indicating that they are the correspondinggnt genes of B. licheniformis. Not only is the organizationof the gnt genes of these two Bacilli highly conserved, butso are the cis regulatory elements of their gnt operon. Sequenceanalysis of the upstream regions of these two gnt operons impliedthat a chromosome rearrangement in B. subtilis might have occurredimmediately upstream of the gnt operon during evolution, causingit to diverge from a common ancestor into B. licheniformis andB. subtilis.  相似文献   

11.
We have cloned a DNA fragment containing the gene for a cell wall hydrolase from Bacillus licheniformis FD0120 into Escherichia coli. Sequencing of the fragment showed the presence of an open reading frame (ORF; designated as cwlL), which is different from the B. licheniformis cell wall hydrolase gene cwlM, and encodes a polypeptide of 360 amino acids with a molecular mass of 38 994. The enzyme purified from the E. coli clone is an N-acetylmuramoyl-l-alanine amidase, which has a Mr value of 41 kDa as determined by SDS-polyacrylamide gel electrophoresis, and is able to digest B. licheniformis, B. subtilis and Micrococcus luteus cell walls. The nucleotide and deduced amino acid sequences of cwlL are very similar to those of ORF3 in the putative operon xpaL1-xpaL2-ORF3 in B. licheniformis MC14. Moreover, the amino acid sequence homology of CwlL with the B. subtilis amidase CwlA indicates two evolutionarily distinguishable regions in CwlL. The sequence homology of CwlL with other cell wall hydrolases and the regulation of cwlL are discussed.  相似文献   

12.
A total of 173 bacterial strains were isolated from different sources at different regions such as fermented foods, shrimp guts, sea water, mangrove water, and sediments. These bacteria were screened against white spot syndrome virus (WSSV) infection in Palaemon paucidens. Based on mortality, white spot level, and healthiness, three bacterial strains were selected and identified using 16S rRNA gene sequencing. These bacterial strains were Bacillus subtilis KA1, B. licheniformis KA2, and B. subtilis KA3. WSSV challenge test in pilot scale was conducted using Litopenaeus vannamei with B. subtilis KA1 and B. subtilis KA3. The survival ratio of shrimp was 0% for WSSV control after 17th days, 84% for B. subtilis KA1 plus WSSV after 26th days, and 28% for B. subtilis KA3 with WSSV after 26th days. B. subtilis KA1 showed good growth at 18–37 °C in with and without 3% NaCl, and therefore can be applied to aquaculture at low to high temperatures. B. subtilis KA1 produced protease and lipase which can increase digestion to shrimp; exhibited antibacterial activity against Vibrio parahaemolyticus; and significantly increased the survival of WSSV challenged shrimps.  相似文献   

13.
Bacillus licheniformis PWD-1, the parent strain, and B. subtilis FDB-29, a recombinant strain. In both strains, keratinase was induced by proteinaceous media, and repressed by carbohydrates. A seed culture of B. licheniformis PWD-1 at early age, 6–10 h, is crucial to keratinase production during fermentation, but B. subtilis FDB-29 is insensitive to the seed culture age. During the batch fermentation by both strains, the pH changed from 7.0 to 8.5 while the keratinase activity and productivity stayed at high levels. Control of pH, therefore, is not necessary. The temperature for maximum keratinase production is 37°C for both strains, though B. licheniformis is thermophilic and grows best at 50°C. Optimal levels of dissolved oxygen are 10% and 20% for B. licheniformis and B. subtilis respectively. A scale-up procedure using constant temperature at 37°C was adopted for B. subtilis. On the other hand, a temperature-shift procedure by which an 8-h fermentation at 50°C for growth followed by a shift to 37°C for enzyme production was used for B. licheniformis to shorten the fermentation time and increase enzyme productivity. Production of keratinase by B. licheniformis increased by ten-fold following this new procedure. After respective optimization of fermentation conditions, keratinase production by B. licheniformis PWD-1 is approximately 40% higher than that by B. subtilis FDB-29. Received 16 July 1998/ Accepted in revised form 07 March 1999  相似文献   

14.
The minor extracellular protease (Epr) is secreted into the culture medium during Bacillus licheniformis, strain USC13, stationary phase of growth. Whereas, B. subtilis Epr has been reported to be involved in swarming; the B. licheniformis protease is also involved in milk-clotting as shown by the curd forming ability of culture broths expressing this protein. The objectives of this study are the characterization of recombinant B. licheniformis Epr (minor extracellular protease) and the determination of its calcium-dependent activation process. In this work, we have cloned and expressed B. licheniformis Epr in Escherichia coli. We were also able to construct a tridimensional model for Epr based on its homology to Thermococcus kodakarensis pro-tk-subtilisin 2e1p, fervidolysin from Fervidobacterium pennivorans 1rv6, and B. lentus 1GCI subtilisin. Recombinant Epr was accumulated into inclusion bodies; after protein renaturation, Epr undergoes an in vitro calcium-dependent activation, similar to that described for tk protease. The recombinant Epr is capable of producing milk curds with the same clotting activity previously described for the native B. licheniformis Epr enzyme although further rheological and industrial studies should be carried out to confirm its real applicability. This work represents for the first time that Epr may be successfully expressed in a non-bacilli microorganism.  相似文献   

15.
Summary Potato peel was found to be a superior substrate for solid state fermentation, compared to wheat bran, for the production of α-amylase by two thermophilic isolates of Bacillus licheniformis and Bacillus subtilis. Under optimal conditions, B. licheniformis produced 270 units/ml and 175 units/ml of α-amylase on potato peel and wheat bran, respectively, while the corresponding values for B. subtilis were 600 units/ml and 265 units/ml. The enzyme from B.␣licheniformis was optimally active at 90 °C and pH 9.0, while that from B. subtilis at 60 °C and pH 7.0. The nature of the experimental data permitted excellent polynomial fits, on the basis of which, two master equations, corresponding to the isolated strains, were derived for estimation of enzyme activity for any set of values of temperature, particle size, moisture, and incubation time within the indicated ranges.  相似文献   

16.
Proteins that are produced for commercial purposes in Bacillus subtilis are commonly secreted via the Sec pathway. Despite its high secretion capacity, the secretion of heterologous proteins via the Sec pathway is often unsuccessful. Alternative secretion routes, like the Tat pathway, are therefore of interest. Two parallel Tat pathways with distinct specificities have previously been discovered in B. subtilis. To explore the application potential of these Tat pathways, several commercially relevant or heterologous model proteins were fused to the signal peptides of the known B. subtilis Tat substrates YwbN and PhoD. Remarkably, the YwbN signal peptide directed secretion of active subtilisin, a typical Sec substrate, via the B. subtilis TatAyCy route. In contrast, the same signal peptide directed Tat-independent secretion of the Bacillus licheniformis α-amylase (AmyL). Moreover, the YwbN signal peptide directed secretion of SufI, an Escherichia coli Tat substrate, in a Tat-independent manner, most likely via Sec. Our results suggest that cytoplasmic protein folding prior to translocation is probably a major determinant of Tat-dependent protein secretion in B. subtilis, as is the case with E. coli. We conclude that future applications for the Tat system of B. subtilis will most likely involve commercially interesting proteins that are Sec incompatible.  相似文献   

17.
Competent Bacillus subtilis cells were exposed to radioactive and density labeled donor DNA extracted from B. pumilus and B. licheniformis. The DNA from these strains hybridized with B. subtilis DNA in vitro at a rate of 24% and 11%, respectively. After entry the vast majority of heterologous DNA was found at the single-strand DNA position in CsCl gradients, and was gradually degraded during incubation. Much less donor DNA than expected from the hybridization values participated in the formation of the donorrecipient complex (DRC). By subjecting the heterologous DRC to sonication and alkaline CsCl gradient centrifugation, it was established that the DRC consisted of three components: (1) recipient DNA in which breakdown products of donor DNA were incorporated through DNA synthesis, (2) recipient DNA in which donor DNA was covalently integrated and (3) recipient DNA in which the donor moiety was not covalently integrated.  相似文献   

18.
Bacterial alkaline phosphatases (APases), except those isolated from Bacillus licheniformis, are approximately 45-kDa proteins while eucaryotic alkaline phosphatases are 60 kDa. To answer the question of whether the apparent 60-kDa alkaline phosphatase from Bacillus licheniformis accurately reflected the size of the protein, the entire gene was analyzed. DNA sequence analysis of the alkaline phosphatase I (APaseI) gene of B. licheniformis MC14 indicated that the gene could code for a 60-kDa protein of 553 amino acids. The deduced protein sequence of APaseI showed about 32% identity to those of B. subtilis APase III and IV and had apparent sequence homologies in the core structure and active sites that are conserved among APases of various sources. The extra carboxy-terminal sequence of APaseI, which made the enzyme bigger than other procaryotic APases, was not homologous to those of eucaryotic APases. The amino acid composition of APaseI was most similar to that of salt-dependent APase among the isozymes of B. licheniformis MC14. Another open reading frame of 261 amino acids was present 142 nucleotide upstream of the APaseI gene and its predicted amino acid sequence showed 68% identity to that of glucose dehydrogenase of B. megaterium.  相似文献   

19.
The antifungal activity of bacterial strains Bacillus subtilis EF 617317 and B. licheniformis EF 617325 was demonstrated against sapstaining fungal cultures Ophiostoma flexuosum, O. tetropii, O. polonicum, and O. ips in both in vitro and in vivo conditions. The crude active supernatant fractions of 7 days old B. subtilis and B. licheniformis cultures inhibited the growth of sapstaining fungi in laboratory experiments. Thermostability and pH stability of crude supernatants were determined by series of experiments. FT-IR analysis was performed to confirm the surface structural groups of lipoproteins present in the crude active supernatant. Partial purification of lipopeptides present in the crude supernatant was done by using Cellulose anion exchange chromatography and followed by Sephadex gel filtration chromatography. Partially purified compounds significantly inhibited the sapstaining fungal growth by in vitro analysis. The lipopeptides responsible for antifungal activity were identified by electrospray ionization mass spectrometry after partial purification by ion exchange and gel filtration chromatography. Four major ion peaks were identified as m/z 1023, 1038, 1060, and 1081 in B. licheniformis and 3 major ion peaks were identified as m/z 1036, 1058, and 1090 in B. subtilis. In conclusion, the partially purified lipopeptides may belong to surfactin and iturin family. In vivo analysis for antifungal activity of lipopeptides on wood was conducted in laboratory. In addition, the potential of extracts for fungal inhibition on surface and internal part of wood samples were analyzed by scanning electron microscopy.  相似文献   

20.
The synthesis of polyglutamic acid (PGA) was repressed by exogenous glutamate in strains of Bacillus licheniformis but not in strains of Bacillus subtilis, indicating a clear difference in the regulation of synthesis of capsular slime in these two species. Although extracellular γ-glutamyltranspeptidase (GGT) activity was always present in PGA-producing cultures of B. licheniformis under various growth conditions, there was no correlation between the quantity of PGA and enzyme activity. Moreover, the synthesis of PGA in the absence of detectable GGT activity in B. subtilis S317 indicated that this enzyme was not involved in PGA biosynthesis in this bacterium. Glutamate repression of PGA biosynthesis may offer a simple means of preventing unwanted slime production in industrial fermentations using B. licheniformis.  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