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 531 毫秒
1.
The bacterium Serratia entomophila (Enterobacteriaceae) has been developed as a commercially available biopesticide for control of the pasture pest Costelytra zealandica. The influence of culture medium composition, dissolved oxygen (DO) concentration and harvesting time were investigated in order to optimise the production of S. entomophila. In batch fermentations, highest yields were achieved using sucrose (40 g L-1) as the carbon source, followed closely by fructose and molasses. The effect of yeast extract (YE), marmite and bakery yeast as cell growth enhancers was also examined in both batch and fed-batch mode. Culture medium containing 20 g L-1 of YE (fed-batch) produced the highest cell density. No significant effect on cell yield was detected when cultures were supplemented with bakery yeast or marmite. The DO concentration influenced biomass production: a 5-fold increase in cell density was achieved when the concentration of DO was maintained in the range of 20-50% (5.7×1010 CFUs mL-1) in comparison with 1% (1.2×1010 CFUs mL-1). In cultures maintained at 1 and 20% DO concentration, cells harvested from the exponential growth phase survived for less than 2 weeks when stored at 4°C. In contrast, high cell survival (85-100%) was achieved when cells were harvested after they had entered the stationary growth phase. Recommendations are provided for the production of robust, high cell density cultures of S. entomophila.  相似文献   

2.
The interaction between sialosyl cholesterol (- or neuraminyl cholesterol, - or β-SC) and the plasma membrane of astrocytes was investigated by the use of 14C-labeled - or β-SC. Both - and β-SC were dose-dependently and time-dependently bound to rat astrocytes. The Scatchard plot analyses showed that rat astrocytes bound apparently 9.69 × 109 molecules of both -SC/cell (apparent Kd = 2.29 × 10−5 M) and β-SC/cell (apparent Kd = 5.39 × 10−5 M) at 37°C. Both the binding of -SC to astrocytes and the subsequent inhibition of DNA synthesis were decreased at the low temperature (4°C), and also suppressed by serum proteins including albumin. One molecule of bovine serum albumin (BSA) bound 2.3 molecules of -SC with the slightly lower Kd-value (8.03 × 10−6 M) than that for the binding site on astrocytes. BSA not only suppressed the -SC-binding to astrocytes but also increased its release from the cells to the culture media. Gangliosides such as GM1 and GM3 unaffected the -SC-binding, promoted the small release of -SC from the cell surface, and inhibited the morphological changes of astrocytes induced by -SC. The mechanism of -SC-binding to cultured astrocytes with reference to the effects of serum or gangliosides is discussed.  相似文献   

3.
A rapid method for microorganism detection using a piezoelectric quartz crystal sensor (PQC) coated with a thin liquid culture medium film was developed and applied to detect the cell number of Proteus vulgaris. This method employed the viscosity and density response of PQC and utilized the coagulation of gelatine medium solution in which the microorganisms had grown to determine the microorganism indirectly. Three time points (TT1, DT, TT2) were obtained from the coagulation curve and were found to be in good linear relationship with the logarithm of the initial number of P. vulgaris in the range 1·3 × 102−1·3 × 105 cells/ml. The detection was rapid and accurate because the coagulation of the thin liquid culture medium film was quick and the time points in the response curve were sharp and so were easy to determine accurately. The detection time was less than 4 h and only a micro sample was needed. A 5 h preincubation was needed before detection. Some experimental conditions are discussed in detail.  相似文献   

4.
Inundative mycoherbicidal biocontrol agents are typically insufficiently virulent to be commercially competitive with herbicides in row crop agriculture, and require enhancement. Pectinase and cellulase are typically used by pathogens during infection. Thus, it was hypothesized that adding exogenous cell wall degrading enzymes might enhance fungal infection. Pectinase or cellulase was added to inocula of aqueous chopped mycelial suspensions of a strain of Colletotrichum coccodes for control of Abutilon theophrasti. Plants treated with 5.3×106 C. coccodes propagules mL-1 and 1.65 U mL-1 pectinase had more rapid and complete disease development. Similar trend was achieved when 10 U mL-1 of cellulase were added to 2.2×106 C. coccodes propagules mL-1. Adding pectinase or cellulase did not increase the host range of the wild-type fungus. The results suggest that there might be value to transforming biocontrol agents to overproduce these enzymes.  相似文献   

5.
Hypochlorous acid (HOCl) is an oxygen-derived species involved in physiological processes related to the defence of the organism that may cause adverse effects when its production is insufficiently controlled. In order to examine its reactivity with potential scavenging molecules from the non steroidal anti-inflammatory drugs (NSAIDs) family, a competition assay based on para-aminobenzoic acid (PABA) chlorination was developed. The original optimised in vitro fluorimetric procedure offered the possibility to determine rate constants (ks) for the reaction with HOCl in physiologically relevant conditions. The specificity of the system was improved by a liquid chromatography (LC) which allows the separation of the drugs and their oxidation products. After determination of the rate constant for PABA chlorination by HOCl (mean±SD in M-1 s-1: 4.3±0.3×103), the applied mathematical model for a chemical competition permits to obtain linear curves from competition studies between several NSAIDs and PABA. Their slopes provided the following rate constants for the different studied drugs: tenoxicam: 4.0±0.7×103, piroxicam: 3.6±0.7×103, lornoxicam: 4.3±0.7×103, meloxicam: 1.7±0.3×104, nimesulide: 2.3±0.6×102. Meloxicam therefore reacted significantly faster than the other oxicams and nimesulide, which is the weakest scavenger of the studied series. The identification of some of the oxidation products by NMR or MS permitted to explore the reaction mechanism and to examine some aspects of the structure/activity relationships for the molecules of the same chemical family.  相似文献   

6.
Microbial characterization during composting of municipal solid waste   总被引:29,自引:0,他引:29  
This study investigates the prevailing physico-chemical conditions and microbial community; mesophilic bacteria, yeasts and filamentous fungi, bacterial spores, Salmonella and Shigella as well as faecal indicator bacteria: total coliforms, faecal coliforms and faecal Streptococci, present in a compost of municipal solid waste. Investigations were conducted in a semi-industrial pilot plant using a moderate aeration during the composting process. Our results showed that: (i) auto-sterilization induced by relatively high temperatures (60–55°C) caused a significant change in bacterial communities. For instance, Escherichia coli and faecal Streptococci populations decreased, respectively, from 2×107 to 3.1×103 and 107 to 1.5×103 cells/g waste dry weight (WDW); yeasts and filamentous fungi decreased from 4.5×106 to 2.6×103 cells/g WDW and mesophilic bacteria were reduced from 5.8×109 to 1.8×107 bacteria/g WDW. On the other hand, the number of bacterial spores increased at the beginning of the composting process, but after the third week their number decreased notably; (ii) Salmonella disappeared completely from compost by the 25th day as soon as the temperature reached 60°C; and (iii) the bacterial population increased gradually during the cooling phase. While Staphylococci seemed to be the dominant bacteria during the mesophilic phase and at the beginning of the thermophilic phase, bacilli predominated during the remainder of the composting cycle. The appearance of gram-negative rods (opportunistic pathogens) during the cooling phase may represent a serious risk for the sanitary quality of the finished product intended for agronomic reuse. Compost sonication for about 3 min induced the inactivation of delicate bacteria, in particular gram-negatives. By contrast, gram-positive bacteria, especially micrococcus, spores of bacilli, and fungal propagules survived, and reached high concentrations in the compost.  相似文献   

7.
Electron self-exchange in solutions of the ‘blue’ copper protein plastocyanin is catalysed by the redox-inert multivalent cations Mg2+ or Co(NH3)3+6. Measurements of specific 1H-NMR line broadening with 50% reduced solutions in the presence of these cations show that electron exchange proceeds through encounters of cation-protein complexes which dissociate at high ionic strength. In the presence of 8mM (5 equivalents/total protein) Co(NH3)3+6, with 10 mM cacodylate (pH*6.0) as background electrolyte, the bimolecular rate constant at 25°C is 7 × 104 M−1·s−1. For comparison, the ‘electrostatically screened’ rate constant measured in 0.1 M KCl in the absence of added multivalent cations is ˜ 4 × 103 M1·s−1.

Plastocyanin Electron self-exchange NMR Protein-protein interaction Multivalent cation Blue copper protein  相似文献   


8.
Palythoa psammophilia Walsh & Bowers has a well coordinated, stereotyped feeding response, the culminating step of which is ingestion; this may be elicited by the synergistic effect of the tripeptide glutathione and the -imino acid, proline. Either activator acting separately causes responses only at high concentrations (above 10−5 M for glutathione; above 10−4 M for proline) in a reduced number of animals and at a low rate (5.00 ± 1.73 min in 5 × 10−3 M solutions of glutathione; 11.10±3.74 min in 5 × 10−3 M solutions of proline). Highest percentages of response were obtained in combinations where glutathione was at a concentration of 5 × 10−3 M and proline at 5 × 10−4 M or in combinations of glutathione at concentrations 5 × 10−6 M and proline at 5 × 10−5 M. The speed of ingestion is considerably enhanced when these activators are combined (1.17±1.18 min).  相似文献   

9.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


10.
Laboratory experiments were conducted to assess effects of nutrients on germination of Verticillium lecanii (=Lecanicillium sp.) conidia and infection of the greenhouse whitefly, Trialeurodes vaporariorum. Suspensions of V. lecanii conidia were prepared in four nutrient solutions: 2% glucose, 2% sucrose, 2% maltose, and 2% peptone. Suspensions in de-mineralized water served as the control. At 23°C the germination rate was highest in the 2% glucose solution, followed by sucrose, maltose, demineralized water, and peptone, respectively. Germ tube growth was greatest at 23°C in the 2% glucose solution after 10 h incubation. Results of the bioassays indicated that the nutrients influenced whitefly infection. Infection levels were highest for conidial suspensions (1×106 conidia/mL) prepared in 2% glucose, and were significantly greater than for peptone, demineralized water and maltose. Infection levels at 1×108 conidia/mL were not significantly different from each other for all materials tested. The potential use of nutrients in a spray formulation as a means of enhancing field efficacy are discussed.  相似文献   

11.
Low-molecular-weight chitosan were prepared using 85% phosphoric acid at different reaction temperatures and reaction time. At room temperature, the viscosity average-molecular weights (Mv) of chitosan decreased to 7.1×104 from 21.4×104 after 35 days treatment. The degradation rate decreased with increasing hydrolysis time. The yields of chitosan also continuously decreased from 68.4 to 40.2% after 35 days. At 40, 60 and 80 °C, the molecular weight decreased to 3.70×104, 3.50×104 and 2.00×104 on 8 h hydrolysis, respectively. The yields of chitosan remain at a high level compared with that at room temperature and were 86.5, 71.4 and 61.3% at 40, 60 and 80 °C treatment, respectively. The different reaction time gave chitosan with different molecular weights. At 60 °C, the molecular weight of products decreased to 7.40×104 from 21.4×104 within 4 h, then decreased slowly to 1.90×104 in 15 h. It was also found that the water-solubility of chitosan increased as the molecular weight decreased. Results show the changes in yields and molecular weight of chitooligomers were strongly dependent on the reaction temperature and reaction time.  相似文献   

12.
Under physiological pH conditions (pH 7.2-7.4) the rate constant of the reaction NO + O2 yielding peroxonitrite (ONOO) was determined as k = (3.7 ± 1.1) × 107 M 1 s 1. The decay of peroxonitrite at this pH follows first order kinetics with a rate constant of 1.4 s 1. At alkaline pH peroxonitrite is practically stable.

Possible consequences of these reactions for the biological lifetime of EDRF will be discussed.  相似文献   

13.
The oxidation of TEMPO (2,2,6,6-tetramethyl-piperidine-1-oxyl radical) has been studied in the presence of recombinant laccases (benzenediol:oxygen oxidoreductase, EC 1.10.3.2) from Polyporus pinsitus (rPpL), Myceliophthora thermophila (rMtL), Coprinus cinereus (rCcL) and Rhizoctonia solani (rRsL) in buffer solution pH 4.5–7.3 and at 25 °C. At pH 5.5 the oxidation constant calculated from the initial rate of TEMPO oxidation was 1.7 × 104, 1.4 × 103, 7.8 × 102 and 5.2 × 102 M−1 s−1 for rPpL, rRsL, rCcL and rMtL, respectively. The maximal activity of rPpL-catalysed TEMPO oxidation was at pH 5.0. The pKa obtained in neutral pH range was 6.2. The reactivity of laccases is in a good agreement with laccases copper type I redox potential.

TEMPO oxidation rate increased 541 times in the presence of 10-(3-propylsulfonate) phenoxazine (PSPX). The model of synergistic TEMPO and PSPX oxidation was proposed. Experimentally obtained rate constants for rPpL-catalysed PSPX oxidation were in a good agreement with those calculated from the synergistic model, therefore confirming the feasibility of the model. The acceleration of TEMPO oxidation with high reactive laccase substrates opens new possibilities for TEMPO application as a mediator.  相似文献   


14.
The effects of ambient temperature and humidity, month, age and genotype on sperm production and semen quality in AI bulls in Brazil were evaluated. Data from two consecutive years were analyzed separately. Seven Bos indicus and 11 Bos taurus bulls from one artificial insemination (AI) center were evaluated in Year 1 and 24 B. indicus and 16 B. taurus bulls from three AI centers were evaluated in Year 2. Ambient temperature and humidity did not significantly affect sperm production and semen quality, probably because there was little variation in these variables. Month accounted for less than 2% of the variation in sperm production and semen quality. Increased bull age was associated with decreased sperm motility (P<0.10) and increased minor sperm defects (P<0.001) in Year 1. B. indicus bulls had greater (P<0.005) sperm concentration than B. taurus bulls in both years (1.7×109/ml versus 1.2×109/ml in Year 1 and 1.6×109/ml versus 1.2×109/ml in Year 2, respectively). Ejaculate volume was not significantly affected by genotype in Year 1 (6.6 ml versus 6.9 ml in B. indicus and B. taurus bulls, respectively), but B. indicus bulls had greater (P<0.05) total (11.4×109 versus 8.2×109) and viable (6.7×109 versus 4.9×109) numbers of spermatozoa in the ejaculate than B. taurus bulls. In Year 2, B. taurus bulls had greater (P<0.05) ejaculate volume than B. indicus bulls (8.2 ml versus 6.7 ml, respectively) and total and viable number of spermatozoa in the ejaculate were not significantly different between genotypes (10.3×109 versus 9.1×109 and 6.1×109 versus 5.4×109 in B. indicus and B. taurus bulls, respectively). Sperm motility was not significantly affected by genotype (mean, 59%). In Year 1, B. indicus bulls tended (P<0.10) to have more major sperm defects and had more (P<0.05) total sperm defects than B. taurus bulls (11.8% versus 8.7% and 13.6% versus 10.0%, respectively). In Year 2, B. indicus bulls tended (P<0.10) to have more total sperm defects than B. taurus bulls (16.2% versus 13.3%, respectively). In conclusion, neither ambient temperature and humidity nor month (season) significantly affected sperm production and semen quality. B. indicus bulls had significantly greater sperm concentration and B. taurus bulls had significantly fewer morphologically defective spermatozoa.  相似文献   

15.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


16.
PDC-109 (13 kDa) is the most abundant component, and the major heparin-binding protein, of bovine (Bos taurus) seminal plasma. Here, we show that PDC-109 contains a single O-linked oligosaccharide (NeuNAc(2–6)-Galβ(1–3)-GalNAc-) attached to Thr11. Immunoquantitation of PDC-109 indicates that its concentration in seminal plasma is 15–20 mg/ml. Though PDC-109 is not present on epididymal sperm, ejaculated spermatozoa on average are coated with (9.5 ± 0.3) × 106 molecules of PDC-109/cell. This value remained constant in swim-up sperm and decreased to (7.7 ± 0.4) × 106/spermatozoon after incubation for 24 h in capacitation medium at 39°C. These data substantiate the hypothesis that PDC-109 may be one of the seminal plasma components that enhance the fertilizing capacity of bull spermatozoa upon interaction with heparin-like glycosaminoglycans present in the female genital tract.  相似文献   

17.
The neogregarine, Mattesia oryzaephili (Neogregarinorida: Lipotrophidae) has only been reported from the sawtoothed grain beetle, Oryzaephilus surinamensis. The pathogen's presence in cadavers of the rusty grain beetle, Cryptolestes ferrugineus, in collapsed colonies prompted studies of its potential to control stored-product insects. Respective mortality rates in fourth instar C. ferrugineus and C. pusillus were 15.3 and 17.7% at 102 oocysts/g of diet and 89.4 and 80.5% at 105 oocysts/g. The mortality of fourth instar O. surinamensis exposed to 105 oocysts/g was only 12%. For C. ferrugineus larvae, there were no significant differences in mortality and infection between exposure to Mattesia dispora and exposure to M. oryzaephili (P>0.05), but for C. pusillus larvae, both responses were significantly higher for M. oryzaephili than M. dispora. Adult C. ferrugineus and O. surinamensis were similar in their responses to M. oryzaephili, with mortality not exceeding 20%, but differed in their responses to M. dispora, with O. surinamensis being more susceptible. The median lethal doses for larval Mediterranean flour moths, Ephestia kuehniella, were 7.9×107M. oryzaephili oocysts/g of diet and 2.7×103M. dispora oocysts/g of diet. In single dose assays of M. oryzaephili physiological host range, greater than 75% infection was achieved for Rhyzopertha dominica and Plodia interpunctella. More than half of oocysts germinated during passage through the guts of susceptible and resistant insects. Second and third instar Galleria mellonella were highly susceptible to M. oryzaephili infection, but fifth instars were not. Infection percentages in fifth instars exposed to 106 oocysts/g were significant only when boric acid or the stilbene, Blankophor®RHK were incorporated into the diet. Host range and general morphology confirm the identity of Mattesia oryzaephili.  相似文献   

18.
We report spectrophotometric equilibrium studies of both the self-association of the new antibiotic iremycin and of its binding to calf thymus DNA in solution (ionic strength 0.2 M; pH 6.0). Iremycin forms dimers in this solution with a dimerization constant K4=(1.19 ± 0.10) × 103 M−1. This equilibrium is taken into account in the evaluation of the interaction of iremycin with DNA. The binding behaviour can be completely described by a single binding mechanism of monomeric iremycin to DNA with allowance both for neighbour exclusion and for cooperativity of interaction. The three intrinsic binding parameters for the homogeneous model were determined simultaneously by a least squares fit of the original titration data: equilibrium constant of cooperative binding K = (2.72 ± 0.66) × 105 M−1 cooperativity parameter σ=0.38±3.27 ± 0.32. The binding parameters of iremycin and adriamycin and their microbial activities are compared.  相似文献   

19.
Common cocklebur has several biotypes including multiple seeded cocklebur (MSC), NCC-TX, and NCC-MS. Alternaria helianthi applied at 2.5×104 conidia mL-1 in a 50% micro-emulsion of unrefined corn oil (MESUCO) or 0.2% Silwet L 77 caused 60-75% mortality on NCC-TX and MSC. Increasing the conidial concentration to 5×104 mL-1 increased mortality to 100% on MSC and NCC-TX, and 75% on NCC-MS. At 10×104 conidia mL-1, A. helianthi caused 100% mortality in all three biotypes. No mortality occurred in any biotype at inoculation rates of 2.5 and 5×104 conidia mL-1 when applied in water. Increasing the dew period from 0 to 12 h increased mortality from 0 to 100% on all three biotypes at a rate of 2.5×104 conidia mL-1 in Silwet and MESUCO. MSC appears to be the most sensitive biotype.  相似文献   

20.
Roger N.F. Thorneley 《BBA》1974,333(3):487-496
1. Single reduced methyl viologen (MV.+) acts as an electron donor in a number of enzyme systems. The large changes in extinction coefficient upon oxidation (λmax 600 nm; MV.+, = 1.3 · 104 M−1 · cm−1; oxidised form of methyl viologen (MV2+), = 0.0) make it ideally suited to kinetic studies of electron transfer reactions using stopped-flow and standard spectrophotometric techniques.

2. A convenient electrochemical preparation of large amounts of MV.+ has been developed.

3. A commercial stopped-flow apparatus was modified in order to obtain a high degree of anaerobicity.

4. The reaction of MV.+ with O2 produced H2O2 (k > 5 · 106 M−1 · s−1, pH 7.5, 25 °C). H2O2 subsequently reacted with excess MV.+ (k = 2.3 · 103 M−1 · s−1, pH 7.5, 25 °C) to produce water. The kinetics of this reaction were complex and have only been interpreted over a limited range of concentrations.

5. The results support the theory that the herbicidal action of methyl viologen (Paraquat, Gramoxone) is due to H2O2 (or radicals derived from H2O2) induced damage of plant cell membrane.  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