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.
For enzymatic synthesis of octyl-β-d-galactopyranoside (octyl-gal) from lactose and n-octanol, Escherichia coli β-galactosidase (β-Gal) was expressed and displayed on the surfaces of Bacillus subtilis spores. The spore-displayed β-Gal was found to be stable when an amphiphilic 1,2-dimethoxyethane (DME) was used as a co-solvent; the transgalactosylation efficiency and octyl-gal conversion were optimal at 50% (v/v) DME. In addition, the product was maximally obtained from 100mM lactose in a phosphate buffer/n-octanol/DME (25/25/50, v/v) mixture. By increasing the agitation speed and the amount of spores displaying β-Gal, a yield of 33.7 mM octyl-gal was obtained over 24h in a batch mode, which is much higher than in other octyl-gal bioconversion processes, such as those involving lipid-coating, reverse micelles, or whole cells. On the other hand, intermittent addition of spore-displayed β-Gal and/or lactose in the reaction medium had no effect on the octyl-gal yield. The synthesized octyl-gal was hydrolyzed by the spore-displayed β-Gal, and a high concentration of octyl-gal competitively inhibited the enzymes (K(i) value of 10.8mM). In summary, we demonstrate that octyl-gal synthesis by spore-displayed β-Gal in non-aqueous medium can be significantly improved with the use of DME as a co-solvent.  相似文献   

2.
3.
The sporulation of Clostridium perfringens NCTC 8798 was studied after exposing vegetative cells to: pH values of 1.5 to 8.0 in fluid thioglycolate broth (for 2h) and then transferring them to Duncan-Strong (DS) sporulation medium; sodium cholate or sodium deoxycholate (0.3 to 6.5 mM) in DS medium; or Rhia-Solberg medium with 0.4% (wt/wt) starch, glucose, or both added at 0 to 55 mM. At pH 1.5, no culturable heat-resistant spores were formed. For cells exposed to pH 3.0, 4.0, 5.0, or 6.0, increases in heat-resistant spores were not seen until after a lag of 12 to 13 h, whereas the lag was only 2 to 3 h for cells exposed to pH 7.0 or 8.0. Maximal spore crops were produced after only 6 to 8 h for cells exposed to pH 7 or 8, but 16 to 18 h was required for production of maximal spore crops by cells exposed to the lower-pH media. The addition of sodium cholate (3.5 to 6.5 mM) to DS medium only slightly reduced the culturable heat-resistant spore count from 1.9 X 10(7) to 3 X 10(6)/ml. The addition of 1.8 mM or more sodium deoxycholate reduced the culturable heat-resistant spore count to less than 10/ ml. When either starch or glucose alone was added to Rhia-Solberg medium there was no production of culturable heat-resistant spores, but a combination of 0.4% (wt/wt) starch and 4.4 mM glucose yielded 6 X 10(5) spores/ml. The spore production remained at this level for glucose concentrations of 6 to 22 mM, but then declined to about 3 X 10(3) spores per ml at higher concentrations.  相似文献   

4.
The sporulation of Clostridium perfringens NCTC 8798 was studied after exposing vegetative cells to: pH values of 1.5 to 8.0 in fluid thioglycolate broth (for 2h) and then transferring them to Duncan-Strong (DS) sporulation medium; sodium cholate or sodium deoxycholate (0.3 to 6.5 mM) in DS medium; or Rhia-Solberg medium with 0.4% (wt/wt) starch, glucose, or both added at 0 to 55 mM. At pH 1.5, no culturable heat-resistant spores were formed. For cells exposed to pH 3.0, 4.0, 5.0, or 6.0, increases in heat-resistant spores were not seen until after a lag of 12 to 13 h, whereas the lag was only 2 to 3 h for cells exposed to pH 7.0 or 8.0. Maximal spore crops were produced after only 6 to 8 h for cells exposed to pH 7 or 8, but 16 to 18 h was required for production of maximal spore crops by cells exposed to the lower-pH media. The addition of sodium cholate (3.5 to 6.5 mM) to DS medium only slightly reduced the culturable heat-resistant spore count from 1.9 X 10(7) to 3 X 10(6)/ml. The addition of 1.8 mM or more sodium deoxycholate reduced the culturable heat-resistant spore count to less than 10/ ml. When either starch or glucose alone was added to Rhia-Solberg medium there was no production of culturable heat-resistant spores, but a combination of 0.4% (wt/wt) starch and 4.4 mM glucose yielded 6 X 10(5) spores/ml. The spore production remained at this level for glucose concentrations of 6 to 22 mM, but then declined to about 3 X 10(3) spores per ml at higher concentrations.  相似文献   

5.
Shao X  Ni H  Lu T  Jiang M  Li H  Huang X  Li L 《New biotechnology》2012,29(3):302-310
An improved surface-immobilisation system was engineered to target heterologous proteins onto vegetative cells and spores of Bacillus thuringiensis plasmid-free recipient strain BMB171. The sporulation-dependent spore cortex-lytic enzyme from B. thuringiensis YBT-1520, SceA, was expressed in vegetative cells and used as the surface anchoring motif. Green fluorescent protein (GFP) and a Bacillus endo-β-1,3-1,4-glucanase (BglS) were used as the fusion partners to test the binding efficiency and the functional activities of immobilised surface proteins. The surface localisation of the SceA-GFP fusion protein on vegetative cells and spores was confirmed by Western blot, immunofluorescence microscopy and flow cytometry. The GFP fluorescence intensity from both vegetative cells and spores was measured and compared to a previously characterised surface display system using a peptidoglycan hydrolase anchor (Mbg). Results demonstrated comparable efficiency of SceA- and Mbg-mediated immobilisation on vegetative cells but a more efficient immobilisation on spores using the SceA anchor, suggesting SceA has greater potential for spore-based applications. The SceA protein was then applied to target BglS onto vegetative cells and spores, and the surface immobilisation was verified by the substantial whole-cell enzymatic activity and enhanced whole-spore enzymatic activity compared to vegetative cells. A dually active B. thuringiensis vegetative cell and spore display system could prove especially valuable for the development of regenerable and heat-stable biocatalysts that function under adverse environmental conditions, for example, an effective feed additive for improved digestion and nutrient absorption by livestock.  相似文献   

6.
A low-cost, low-maintenance system for soilless production of vesicular-arbuscular mycorrhizal (VAM) fungus spores and inoculum was developed and adapted for production of acidophilic and basophilic isolates. Corn (Zea mays) plants were grown with Glomus etunicatum, G. mosseae or Gigaspora margarita in sand automatically irrigated with modified Hoagland's solution. Sand particle size, irrigation frequency, P concentration, and buffer constituents were adjusted to maximize spore production. Modified half-strength Hoagland's solution buffered with 4-morpholine ethane-sulfonic acid (MES) automatically applied 5 times/day resulted in production of 235 G. etunicatum spores/g dry wt. of medium (341000 spores/pot) and 44 G. margarita spores/g dry wt. of medium (64800 spores/pot). For six basophilic isolates of G. mosseae, CaCO3 was incorporated into the sand and pots were supplied with the same nutrient solution as for acidophilic isolates. The increased pH from 6.1±0.2 to 7.2±0.2 resulted in spore production ranging from 70 to 145 spores/g dry wt. (102000–210000 spores/pot). Spore production by all isolates grown in the soilless sand system at Beltsville has exceeded that of traditional soil mixtures by 32–362% in 8–12 weeks.  相似文献   

7.
Berberine, an alkaloid originally extracted from the plant Coptis chinensis and other herb plants, has been used as a pharmacological substance for many years. The therapeutic effect of berberine has been attributed to its interaction with nucleic acids and blocking cell division. However, levels of berberine entering individual microbial cells minimal for growth inhibition and its effects on bacterial spores have not been determined. In this work the kinetics and levels of berberine accumulation by individual dormant and germinated spores were measured by laser tweezers Raman spectroscopy and differential interference and fluorescence microscopy, and effects of berberine on spore germination and outgrowth and spore and growing cell viability were determined. The major conclusions from this work are that: (1) colony formation from B. subtilis spores was blocked ~ 99% by 25 μg/mL berberine plus 20 μg/mL INF55 (a multidrug resistance pump inhibitor); (2) 200 μg/mL berberine had no effect on B. subtilis spore germination with L-valine, but spore outgrowth was completely blocked; (3) berberine levels accumulated in single spores germinating with ≥ 25 μg/mL berberine were > 10 mg/mL; (4) fluorescence microscopy showed that germinated spores accumulated high-levels of berberine primarily in the spore core, while dormant spores accumulated very low berberine levels primarily in spore coats; and (5) during germination, uptake of berberine began at the time of commitment (T1) and reached a maximum after the completion of CaDPA release (Trelease) and spore cortex lysis (Tlysis).  相似文献   

8.
Two strains of Moorella thermoacetica, JW/B-2 and JW/DB-4, isolated as contaminants from autoclaved media for chemolithoautotrophic growth containing 0.1% (wt/vol) yeast extract, formed unusually heat-resistant spores. Spores of the two strains required heat activation at 100 degrees C of more than 2 min and up to 90 min for maximal percentage of germination. Kinetic analysis indicated the presence of two distinct subpopulations of heat-resistant spores. The decimal reduction time (D10-time=time of exposure to reduce viable spore counts by 90%) at 121 degrees C was determined for each strain using spores obtained under different conditions. For strains JW/DB-2 and JW/ DB-4, respectively, spores obtained at approximately 25 degrees C from cells grown chemolithoautotrophically had D10-times of 43 min and 23 min; spores obtained at 60 degrees C from cells grown chemoorganoheterotrophically had D10-times of 44 min and 38 min; spores obtained at 60 degrees C from cells grown chemolithoautotrophically had D10-times of 83 min and 111 min. The thickness of the cortex varied between 0.10 and 0.29 microm and the radius of the cytoplasm from 0.14 to 0.46 microm. These spores are amongst the most heat-resistant noted to date. Electron microscopy revealed structures within the exosporia of spores prior to full maturity that were assumed to be layers of the outer spore coat.  相似文献   

9.
Dolgikh VV  Semenov PB 《Tsitologiia》2003,45(3):324-329
Incubation of Nosema grylli spores in alkaline--saline solution (10 mM KOH, 170 mM KCl) leads to solubilization of the major spore wall protein of 40 kDa (p40). Both the compounds of this solution are crucial for p40 solubilization. After spore incubation in 170 mM KCl no proteins were released in the medium. In contrast, 10 mM KOH causes a release of many spore proteins but only a small amount of p40. A long storage of spores (over a year) in water or 0.02% sodium azide results in a sharp decrease of p40 content. Specific polyclonal antibodies were obtained by immunization of rabbits with isolated p40. The specificity of serum was confirmed by immunoblotting. IFA showed reliable reaction on the envelopes of sporonts and sporoblasts, whereas only part of spores reacted with antibodies. This distinction may be due to changing surface antigens during spore maturation. Solubilization of p40 under alkaline conditions could be associated with spore extrusion, since a subsequent transfer of spores to neutral solution leads to their discharge. Subsequent wash of discharged spores with 1-3% SDS, 9 M urea and treatment by 100% 2-ME result in solubilization of protein of 56 kDa (p56). The maximum concentration of 2-ME is important for isolation of pure p56. Evidence has been provided that p56 is a protein of N. grylli polar tubes. Treatment of discharged spores by 2-ME in the presence of SDS results in solubilization of four additional proteins with molecular weights about 46, 34, 21 and 15 kDa.  相似文献   

10.
The effect of solid substrates, inoculum and incubation time were studied using response surface methodology (RSM) for the production of polygalacturonase enzyme and spores in solid-state fermentation using Aspergillus sojae ATCC 20235. Two-stage optimization procedure was applied using D-optimal and face-centered central composite design (CCD). Crushed maize was chosen as the solid substrate, for maximum polygalacturonase enzyme activity based on D-optimal design. Inoculum and incubation time were determined to have significant effect on enzyme activity and total spore (p<0.01) based on the results of CCD. A second order polynomial regression model was fitted and was found adequate for individual responses. All two models provided an adequate R(2) of 0.9963 (polygalacturonase) and 0.9806 (spores) (p<0.001). The individual optimum values of inoculum and incubation time for maximum production of the two responses were 2 x 10(7) total spores and 5-6 days. The predicted enzyme activity (30.55 U/g solid) and spore count (2.23 x 10(7)spore/ml) were very close to the actual values obtained experimentally (29.093 U/g solid and 2.31 x 10(7)spore/ml, respectively). The overall optimum region considering the two responses together, overlayed with the individual optima. Solid-state fermentation provided 48% more polygalacturonase activity compared to submerged fermentation under individually optimized conditions.  相似文献   

11.
Carbohydrate-protein interaction is one of the most important molecular events deemed critical for numerous biological processes. Therefore, understanding this interaction is essential. In this study, we used bacterial spore display techniques to present multiple copies of streptavidin on the surface of spores to explore carbohydrate-protein interaction in solution and on chips. By applying bacterial spores displaying streptavidin, we developed a new method which allows sensitive, versatile, and passive detection of carbohydrate-protein interactions with a 10-fold increase in sensitivity. The linear relationship of interactions between carbohydrates and labeled concanavalin A (con A) in solution and on functionalized bacterial spore chips has also been confirmed. To the best of our knowledge, this is the first example of development and characterization of binding behavior in carbohydrateprotein interactions using bacterial spore-displayed streptavidin. We believe this strategy may enable new high-throughput screening of carbohydrate interactions as well as establish a basis for monitoring inhibitors of carbohydrate-binding proteins when developing new drugs.  相似文献   

12.
13.
When they were subjected to a range of physical and chemical treatments, spores of Pasteuria penetrans showed properties similar to those of other endospore-forming bacteria. The spores did not take up some stains, were resistant to desiccation and sonication and showed extrusion of spore contents ('spore popping') on prolonged exposure to 0.1% KMnO4 in 0.3 n HNO3. Calcium and dipicolinic acid (DPA) were present at concentrations of 0.28% and 0.96% of the spore dry weight respectively, giving a Ca: DPA molar ratio of 1.2. The infectivity of P. penetrans spores was reduced to a low level after heating at 100°C for 5 min, but spore attachment was not markedly affected by heating at 100°C for 15 min. Evidence for the presence of catalase in P. penetrans spores was equivocal because the low levels of catalase activity observed in spore suspensions may have been due to contamination from catalase-positive nematode tissue. When P. penetrans spores were exposed to a range of substances known to act as germinants for spores of Bacillus spp., germination or loss of refractility was not observed by phase microscopy. In vitro culture of P. penetrans was attempted by inoculating either spores or vegetative mycelial bodies onto a diverse range of simple and complex media and incubating them in aerobic, reduced oxygen, anaerobic and increased CO2 environments. Signs of spore germination or growth of vegetative stages were never observed.  相似文献   

14.
The enzyme CwlJ is involved in the depolymerization of cortex peptidoglycan during germination of spores of Bacillus subtilis. CwlJ with a C-terminal His tag was functional and was extracted from spores by procedures that remove spore coat proteins. However, this CwlJ was not extracted from disrupted spores by dilute buffer, high salt concentrations, Triton X-100, Ca(2+)-dipicolinic acid, dithiothreitol, or peptidoglycan digestion, disappeared during spore germination, and was not present in cotE spores in which the spore coat is aberrant. These findings indicate the following: (i) the reason decoated and cotE spores germinate poorly with dipicolinic acid is the absence of CwlJ from these spores; and (ii) CwlJ is located in the spore coat, presumably tightly associated with one or more other coat proteins.  相似文献   

15.
The exudate of germinated spores of B. cereus IFO 13597 in 0.15 M KCl-50 mM potassium phosphate (pH 7.0) contained a spore-lytic enzyme which has substrate specificity for fragmented spore cortex from wild-type organisms (cortical-fragment-lytic enzyme [CFLE]), in addition to a previously characterized germination-specific hydrolase which acts on intact spore cortex (spore cortex-lytic enzyme [SCLE]) (R. Moriyama, S. Kudoh, S. Miyata, S. Nonobe, A. Hattori, and S. Makino, J. Bacteriol. 178:5330-5332, 1996). CFLE was not capable of degrading isolated cortical fragments from spores of Bacillus subtilis ADD1, which lacks muramic acid delta-lactam. This suggests that CFLE cooperates with SCLE in cortex hydrolysis during germination. CFLE was purified in an active form and identified as a 48-kDa protein which functions as an N-acetylglucosaminidase. Immunochemical studies suggested that the mature enzyme is localized on a rather peripheral region of the dormant spore, probably the exterior of the cortex layer. A gene encoding the enzyme, sleL, was cloned in Escherichia coli, and the nucleotide sequence was determined. The gene encodes a protein of 430 amino acids with a deduced molecular weight of 48,136. The N-terminal region contains a repeated motif common to several peptidoglycan binding proteins. Inspection of the data banks showed no similarity of CFLE with N-acetylglucosaminidases found so far, suggesting that CFLE is a novel type of N-acetylglucosaminidase. The B. subtilis genome sequence contains genes, yaaH and ydhD, which encode putative proteins showing similarity to SleL.  相似文献   

16.
The exudate of fully germinated spores of Clostridium perfringens was found to contain a large amount of a spore lytic enzyme which acted directly on alkali-treated spores of the organism to cause germination. Although no detectable amount of the enzyme was found in dormant spores during germination in a KCl medium, the enzyme was produced rapidly and released into the medium. The optimal conditions for enzyme activity were pH 6.0 and 45 degrees C. Maximum activity occurred in the presence of various univalent cations at a concentration of 50 mM. The enzyme was readily inactivated by several sulfhydryl reagents. A strong reducing condition was generated in the ionic germination of the spores, a minimum Eh level of -350 mV being reached 30 min after initiation of germination. Furthermore, adenosine triphosphate-dependent pyruvate:ferredoxin oxidoreductase (EC 1.2.7.1) was identified in both dorman and germinated spores. The relationship between the release of active enzyme and the generation of reducing conditions during germination is discussed.  相似文献   

17.
Y Chen  S Miyata  S Makino    R Moriyama 《Journal of bacteriology》1997,179(10):3181-3187
The exudate of fully germinated spores of Clostridium perfringens S40 in 0.15 M KCI-50 mM potassium phosphate (pH 7.0) was found to contain another spore-lytic enzyme in addition to the germination-specific amidase previously characterized (S. Miyata, R. Moriyama, N. Miyahara, and S. Makino, Microbiology 141:2643-2650, 1995). The lytic enzyme was purified to homogeneity by anion-exchange chromatography and shown to be a muramidase which requires divalent cations (Ca2+, Mg2+, or Mn2+) for its activity. The enzyme was inactivated by sulfhydryl reagents, and sodium thioglycolate reversed the inactivation by Hg2+. The muramidase hydrolyzed isolated spore cortical fragments from a variety of wild-type organisms but had minimal activity on decoated spores and isolated cell walls. However, the enzyme was not capable of digesting isolated cortical fragments from spores of Bacillus subtilis ADD1, which lacks muramic acid delta-lactam in its cortical peptidoglycan. This indicates that the enzyme recognizes the delta-lactam residue peculiar to spore peptidoglycan, suggesting an involvement of the enzyme in spore germination. Immunochemical studies indicated that the muramidase in its mature form is localized on the exterior of the cortex layer in the dormant spore. A gene encoding the muramidase, sleM, was cloned into Escherichia coli, and the nucleotide sequence was determined. The gene encoded a protein of 321 amino acids with a deduced molecular weight of 36,358. The deduced amino acid sequence of the sleM gene indicated that the enzyme is produced in a mature form. It was suggested that the muramidase belongs to a separate group within the lysozyme family typified by the fungus Chalaropsis lysozyme. A possible mechanism for cortex degradation in C. perfringens S40 spores is discussed.  相似文献   

18.
Simulated raindrops, diameter c. 3 or 4 mm, fell 13 m down a raintower onto suspensions of Septoria nodorum pycnidiospores, depth 0.5 mm, or infected straw pieces. Splash droplets were collected on pieces of fixed photographic film. It was estimated that one drop generated c. 300 spore carrying splash droplets, containing c. 6000 spores, from a concentrated spore suspension (6.5 × 105 spores/ml) and c. 25 spore-carrying droplets, containing c. 30 spores, from infected straw pieces (11 × 106 spores/g dry wt). When the target was a spore suspension in water without surfactant, most spore-carrying droplets were in the 200—400 μm size category and most spores were carried in droplets with diameter >1000 μm. When surfactant was added to spore suspensions, most spore-carrying droplets were in the 0–200 μm category and most spores were carried in droplets with diameter 200–400 μm and none in droplets >1000 μm. Regression analyses showed a significant (p < 0.001) relationship between square root (number of spores per droplet) and droplet diameter; the slope of the regression line was greatest when surfactant was added to the spore suspensions. The distribution of splash droplets with distance travelled from the target was better fitted by an exponential model than by power law or Gaussian models. The distributions of spore-carrying droplets and spores with distance were fitted better by an exponential model than by a power law model. Thus regressions of log, (number collected) against distance were all significant (p < 0.01); the slopes of the regression lines were steepest when surfactant was added to the spore suspension. At a distance of 10 cm from target spore suspensions most splash droplets and spore-carrying droplets were collected at height 10–20 cm, with none above 40 cm; at a distance of 20 cm there were most at heights 0–10 cm and 40–50 cm.  相似文献   

19.
The immobilization of enzymes on edible matrix supports is of great importance for developing stabilized feed enzymes. In this study, probiotic Bacillus spores were explored as a matrix for immobilizing Escherichia coli phytase, a feed enzyme releasing phosphate from phytate. Because Bacillus spore is inherently resistant to heat, solvents and drying, they were expected to be a unique matrix for enzyme immobilization. When mixed with food-grade Bacillus polyfermenticus spores, phytases were adsorbed to their surface and became immobilized. The amount of phytase attached was 28.2 ± 0.7 mg/g spores, corresponding to a calculated activity of 63,960 U/g spores; however, the measured activity was 41,120 ± 990.1 U/g spores, reflecting a loss of activity upon adsorption. Immobilization increased the half life (t1/2) of the enzyme three- to ten-fold at different temperatures ranging from 60 to 90 °C. Phytase was bound to the spore surface to the extent that ultrasonication treatment was not able to detach phytases from spores. Desorption of spore-immobilized phytase was only achieved by treatment with 1 M NaCl, 10% formic acid in 45% acetonitrile, SDS, or urea, suggesting that adsorption of phytase to the spore might be via hydrophobic and electrostatic interactions. We propose here that Bacillus spore is a novel immobilization matrix for enzymes that displays high binding capacity and provides food-grade safety.  相似文献   

20.
Clostridium botulinum spores were sublethally damaged by exposure to 12 or 28 micrograms of available chlorine per ml for 2 min at 25 degrees C and pH 7.0. The damaging dose was 2.7 x 10(-6) to 3.1 x 10(-6) micrograms of available chlorine per spore. Damage was manifested by a consistent 1.6 to 2.4 log difference between the most probable number enumeration of spores (modified peptone colloid medium) and the colony count (modified peptone yeast extract glucose agar); this did not occur with control spores. Damaged spores could be enumerated by the colony count procedure. Germination responses were measured in several defined and nondefined media. Hypochlorite treatment altered the rate and extent of germination in some of the media. Calcium lactate (9 mM) permitted L-alanine (4.5 mM) germination of hypochlorite-treated spores in a medium containing 12 or 55 mM sodium bicarbonate, 0.8 mM sodium thiosulfate, and 100 mM Tris-hydrochloride (pH 7.0) buffer. Tryptose inhibited L-alanine germination of the spores. Treatments with hypochlorite and with hydrogen peroxide (7%, 25 degrees C, 2 min) caused similar enumeration and germination responses, indicating that the effect was due to a general oxidation phenomenon.  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