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 112 毫秒
1.
Metal concentrations and population parameters of the seagrass Halodule wrightii were determined at three locations at Rio de Janeiro State, Brazil. The possible increase of metal availability in one of these areas, Sepetiba Bay, as a result of dredging of contaminated bottom sediments which ocurred, was evaluated by analyses of Al, Cd, Cr, Cu, Fe, Ni, Pb and Zn in root, rhizome and shoots. In addition, analyses were carried out in H. wrightii populations from non-contaminated areas located at northwestern (Cabo Frio) and southeastern (Angra do Reis) regions of Rio de Janeiro State. Concurrently, abundance and density data of the seagrass populations were obtained. It was found that concentration from Sepetiba Bay samples up to 1.6 ± 0.4 μg g−1 of Cd, 12 ± 1.0 μg g−1 of Cr, 27 ± 2.4 μg g−1 of Pb, 291 ± 47 μg g−1 of Mn, 128 ± 23 μg g−1 of Zn were significantly higher than that from two other collection sites. An increase in Cd and Zn concentration was observed in H. wrightii from Sepetiba Bay indicating that metal mobilization from contaminated sediments through dredging activities were, at least in part, transferred to the biotic compartment via accumulation by the seagrass. The populations of seagrass within the region demonstrated quite substantial changes in biomass data but not in shoot or rhizome density during the study. Such changes in biomass are to be expected, as these dynamics are typical of the small, isolated monospecific populations of H. wrightii along the Rio de Janeiro coast.  相似文献   

2.
Resistance in Cicer bijugum Rech. f. a wild relative of chickpea, to Botrytis grey mould (BGM), caused by Botrytis cinerea Pers., was shown to be associated with high concentrations of maackiain when compared to three susceptible species. The two BGM resistant accessions of C. bijugum contained between 200 and 300 μg maackiain g−1 of foliage whereas the BGM susceptible species C. arietinum, C. echinospermum and C. reticulatum contained less than 70 μg g−1. Furthermore, the concentration of maackiain increased to more than 400 μg g−1 in the resistant wild species after being inoculated with the pathogen whereas no significant increase was recorded in the susceptible species. The germination of spores of Botrytis cinerea, treated with maackiain, was inhibited in a dose dependent manner; less than 10% of spores germinated when treated with 500 μg ml−1. The data indicate that maackiain may be an important component in BGM resistance in the wild chickpea C. bijugum and that the resistance is enhanced in the presence of the pathogen.  相似文献   

3.
Conidiation and lytic enzyme production by Trichoderma viride at different solids concentration of pre-treated municipal wastewater sludge was examined in a 15-L fermenter. The maximum conidia concentration (5.94 × 107 CFU mL−1 at 96 h) was obtained at 30 g L−1 suspended solids. The maximum lytic enzyme activities were achieved around 12–30 h of fermentation. Bioassay against a fungal phytopathogen, Fusarium sp. showed maximum activity in the sample drawn around 96 h of fermentation at 30 g L−1 suspended solids concentration. Entomotoxicity against spruce budworm larvae showed maximum value ≈17290 SBU μL−1 at 30 g L−1 suspended solids concentration at the end of fermentation (96 h). Plant bioassay showed dual action of T. viride, i.e., disease prevention and growth promotion. The rheological analyses of fermentation sludges showed the pseudoplastic behaviour. In order to maintain required dissolved oxygen concentration ≥30%, the agitation and aeration requirements significantly increased at 35 g L−1 compared to 30 and 25 g L−1. The oxygen uptake rate and volumetric oxygen mass transfer coefficient, kLa at 35 g L−1 did not increase in comparison to 30 g L−1 due to rheological complexity of the broth during fermentation. Thus, the successful fermentation operation of the biocontrol fungus T. viride is a rational indication of its potential for mass-scale production for agriculture and forest sector as a biocontrol agent.  相似文献   

4.
Andreas Hussner  Rainer Lsch 《Flora》2007,202(8):653-660
Floating Pennywort (Hydrocotyle ranunculoides L. fil.) is a worldwide distributed aquatic plant. The species is native to North America and quite common also in Central and South America. In Europe, Japan and Australia it is known as an alien plant, sometimes causing serious problems for affected ecosystems and human use of water bodies. Starting from Western Europe with an eastwards directed spread, Floating Pennywort was recorded in Germany in 2004 for the first time. Since then, the species spread out and got established in western parts of Central Europe. For a definite prediction of the potential of a further spread, data about biology, in particular growth and photosynthesis are needed. Here, regeneration capacity, growth at different nutrient availabilities and photosynthesis of H. ranunculoides were investigated. In addition biomass samples were taken in the field. Results show an enormous regeneration capacity (e.g., by forming new shoots from small shoot fragments), increasing growth rates under increasing nutrient availability and a maximum increase of biomass reaching 0.132±0.008 g g−1 dw d−1. Dense populations of H. ranunculoides growing in ponds and oxbows were found at high nutrient content of the substrate, the biomass reaching there up to 532.4±14.2 g dw m−2. Gas exchange analysis showed a physiological optimum of H. ranunculoides CO2 uptake at temperatures between 25 and 35 °C and high photon flux densities (PPFD) above 800 μmol photons m−2 s−1. In comparison, native Hydrocotyle vulgaris showed an optimum of net photosynthesis at 20–30 °C and a light saturation of CO2 gas exchange at 350 μmol photons m−2 s−1.  相似文献   

5.
Because of its novel bioactive properties the production of gymnodimine for use as a pharmaceutical precursor has aroused interest. The dinoflagellate, Karenia selliformis produces gymnodimine when grown in bulk culture using GP + selenium medium but the growth rates (μ) and levels of gymnodimine are low (μ, 0.05 days−1; gymnodimine 250 μg L−1 max). We describe the effects of organic acid additions (acetate, glycolate, alanine and glutamate additions and combinations of these) in enhancing growth and gymnodimine production in axenic cultures. The most effective organic acid combinations in decreasing order were: glycolate/alanine > acetate > glycolate. Glycolate/alanine optimised gymnodimine production by prolonging growth (maximum cell yield, 1.76 × 105 cells mL−1; gymnodimine, 1260 μg L−1; growth rate (μ), 0.2 days−1) compared to the control (growth maximum cell yield, 7.8 × 104 cells mL−1; gymnodimine, 780 μg L−1; μ, 0.17 days−1). Acetate enhanced gymnodimine by stimulating growth rate (μ, 0.23 days−1) and the large concentration of gymnodimine per cell (16 pg cell−1 cf. 9.8 pg cell−1 for the control) suggests a role for this compound in gymnodimine biosynthesis. Amending culture media with Mn2+ additions resulted in slightly decreased growth in control cultures and increased the gymnodimine while in glycolate/alanine cultures growth was stimulated but gymnodimine production decreased. The results suggest that the organic acid can enhance gymnodimine production by either enhancing growth maximum or the biosynthetic pathway.  相似文献   

6.
The potential for nutrient load (30, 100 and 350 g N m−2 per year) to alter plant performance under saline conditions (control, 4.5, 9 and 13 dS m−1) was examined in the sedge Bolboschoenus medianus. Relative growth rates (RGR) across nutrient loadings ranged from 30.2 to 41.8 mg g−1 per day in controls and were reduced to 20.9–28.5 mg g−1 per day by salinities of 13 dS m−1. Whilst higher nutrient loads generally increased RGR, the response was smaller at higher salinities. Responses to salinity and nutrient load were specific. Nutrient load increased the RGR via increases in the leaf area ratio (LAR). The LAR ranged from 1.9 to 2.1 m2 kg−1 across salinity treatments at 30 g N m−2 per year, and increased to 2.5–2.8 m2 kg−1 at 350 g N m−2 per year. Salinity reduced the RGR via a reduction in the net assimilation rate (NAR). The NAR in control plants ranged from 14.7 to 16 g m−2 per day across nutrient loadings and decreased to 11–12 g m−2 per day at 13 dS m−1. Carbon isotope discrimination of leaves decreased by 2–3‰ in response to 13 dS m−1 at the lower nutrient loadings. A prominent response of B. medianus to salinity was a change in biomass allocation from culms to tubers. In contrast, the response to nutrient load was characterised by a shift in biomass allocation from roots to leaves.  相似文献   

7.
The effect of changing dilution rate (D) on Bacillus sp. CCMI 1051 at dilution rates between 0.1 and 0.55 h−1 in a glucose-limited medium was studied. Biomass values varied between 0.88 and 1.1 g L−1 at D values of 0.15–0.35 h−1. Maximal biomass productivity was found to be 0.39 g L−1 h−1, obtained at D = 0.35 h−1 and corresponding to a 54.4% conversion of the carbon into cell mass. The highest rate of glucose consumption was 4.45 mmol g−1 h−1 occurring at D = 0.4 h−1. The glucose concentration inside the chemostat was below the detection level starting to accumulate around 0.4 h−1. Growth inhibition of fifteen strains of fungi by the broth of the steady-state cell-free supernatants was assessed. Results showed that the relative inhibition differ among the target species but was not influenced by the dilution rate changing.  相似文献   

8.
In this study the effect of ontogenetic drift on crassulacean acid metabolism (CAM) was investigated in the aquatic CAM-isoetid Littorella uniflora. The results of this study strengthen the general hypothesis of CAM being a carbon-conserving mechanism in aquatic plants, because high-CAM capacity (45–183 μequiv. g−1 FW) was present in all leaves of L. uniflora irrespective of age. Since possession of CAM in aquatic plants allows CO2 uptake throughout the light/dark cycle, presence of CAM in all leaves influences the carbon balance of L. uniflora positively. On average for all lakes, different leaf classes accounted for 11–36% of the total dark CO2 uptake by the individual plant.

The capacity for both CAM and photosynthesis declined with increasing leaf age, and was in the oldest leaves only 25–53% of the capacity in the youngest. The photosynthetic capacity was estimated to be sufficiently high to ensure refixation of the CO2 released from malate during decarboxylation in the daytime. In line with this, a linear coupling between CAM capacity and photosynthetic capacity was found. Parallel to the change in photosynthetic capacity, an age-related change in total ribulose-bisphosphate carboxylase/oxygenase (rubisco) activity from 732 μmol C g−1 DW h−1 in the youngest leaves to 346 μmol C g−1 DW h−1 in the oldest was observed. In contrast, no significant change in phosphoenolpyruvate carboxylase (PEPcase) activity with leaf age was observed (means ranged between 46 and 156 μmol C g−1 DW h−1).  相似文献   


9.
The photosynthetic capacity of Myriophyllum salsugineum A.E. Orchard was measured, using plants collected from Lake Wendouree, Ballarat, Victoria and grown subsequently in a glasshouse pond at Griffith, New South Wales. At pH 7.00, under conditions of constant total alkalinity of 1.0 meq dm−3 and saturating photon irradiance, the temperature optimum was found to be 30–35°C with rates of 140 μmol mg−1 chlorophyll a h−1 for oxygen production and 149 μmol mg−1 chlorophyll a h−1 for consumption of CO2. These rates are generally higher than those measured by other workers for the noxious Eurasian water milfoil, Myriophyllum spicatum L., of which Myriophyllum salsugineum is a close relative. The light-compensation point and the photon irradiance required to saturate photosynthetic oxygen production were exponentially dependent on water temperature. Over the temperature range 15–35°C the light-compensation point increased from 2.4 to 16.9 μmol (PAR) m−2 s−1 for oxygen production while saturation photon irradiance increased from 41.5 to 138 μmol (PAR) m−2 s−1 for oxygen production and from 42.0 to 174 μmol (PAR) m−2 s−1 for CO2 consumption. Respiration rates increased from 27.1 to 112.3 μmol (oxygen consumed) g−1 dry weight h−1 as temperature was increased from 15 to 35°C. The optimum temperature for productivity is 30°C.  相似文献   

10.
The aim of this work was to assess the potential for bacterial oxidation of hydrogen sulphide as a purification method of sour gas. Using a continuous culture of Chlorobium limicola, high efficiencies of oxidation of both soluble and gaseous sulphide were achieved, with efficiencies for the latter exceeding 95%. Sulphide added as aqueous sodium sulphide was converted to sulphur and sulphate with almost total removal of the initial 100 mg S l−1 within 24 h. Gaseous sulphide was oxidized at an efficiency of 95% (approximately 3 mmol S h−1 (unit biomass Abs)−1) over 1 h runs at a gas flow rate of 60 ml min−1. With a sulphur recovery system to prevent sulphur accumulation, an efficiency of 70% was maintained. Biological removal of sulphide represents a potentially important biotechnological process, with high potential for viable scale up.  相似文献   

11.
Sulfur cycling was examined in sediments inhabited with the isoetids Littorella uniflora and Isoetes lacustris in the oligotrophic soft-water Lake Kalgaard, Denmark. Based on short-term tracer incubations sulfate reduction was measured along a transect from the shore (0.6 m) to profundal sediments (4.6 m). The sulfate reduction rates were low (0.008–0.8 mmol m−2 d−1) in the sandy shallow sediments with low organic content (<1.3 mmol C g−1 sed DW) and high redox potentials (>100 mV), whereas sulfate reduction was higher at the deeper sites (2.7–4.6 mmol m−2 d−1) with high organic content (max. 11.5 mmol C g−1 sed DW) and lower redox potentials (<100 mV). High concentrations of dissolved organic carbon (DOC) were found in the low particulate organic sediments (up to 18.4 mM), and most of the DOC pool consisted of acetate (40–77%). Reoxidation of sulfides due to root oxygen release was probably important at all sites and a positive efflux of sulfate across the sediment–water interface was measured, attaining rates (up to 4.8 mmol m−2 d−1) similar to the sulfate reduction rates. Reoxidation of sulfides was also manifested by high fraction (>80%) of reduced sulfides being accumulated as elemental sulfur or pyrite (chromium reducible sulfur, CRS). The largest pools of CRS were found in high organic sediment with vertical distributions resembling those of the sulfate reduction rates. The overall effect of isoetid growth on sulfur cycling in the rhizosphere is a suppression of sulfate reduction in low organic sediments and the governing of sulfide reoxidation in sediments with higher organic content.  相似文献   

12.
Optimal utilisation of tannin-rich browse tree fodders including Acacia spp. foliages as crude protein (CP) supplements to ruminants in the tropics is limited by less available information on their feed nutritive potential. Two studies were conducted to: (1) determine rate and extent of ruminal dry matter (DM) degradability (DMD) and (2) investigate effect of sun-dried Acacia nilotica (NLM), A. polyacantha (PLM) and Leucaena leucocephala leaf meal (LLM) supplementation on growth performance of 20 growing (7–9 months old) Small East African male goats (14.6 ± 0.68 kg) fed on native pasture hay (NPH) basal diet for 84 days in a completely randomised design experiment in north-western Tanzania. The goats were randomised into four treatment groups consisting of five animals each. Three supplement diets: 115.3 g NLM (T2), 125.9 g PLM (T3) and 124.1 g LLM (T4), which was used as a positive control, were supplemented at 20% of the expected DM intake (DMI; i.e., 3% body weight) to the three animal groups fed on NPH (basal diet) compared to the animals in a control group that were fed on NPH without browse supplementation (T1).

NPH had significantly the lowest (P < 0.05) CP of 45.5 g kg−1 DM compared to NLM, PLM and LLM (159, 195 and 187 g kg−1 DM, respectively). NPH had higher (P < 0.05) fibre fractions; lower ruminal DM degradability characteristics and ME than NLM, PLM and LLM. Supplementation of the animals with browse resulted to (P < 0.05) higher average daily weight gains (ADG) of 157.1 g day−1 in T4 than the animals fed on T2 (114.3 g day−1) and T3 (42.9 g day−1), and even to those fed on T1 (control), which lost weight (−71.4 g day−1). Improved weight gains were mainly due to corrected feed nitrogen (N) or CP due to supplementation of the animals with browse fodder. Too low CP of the NPH would not meet the normal requirements of CP (80 g CP kg−1 DM) for optimal rumen microbial function in ruminants. Higher ADG due to LLM (T4) and NLM (T2) supplementation suggest optimised weight gains due to browse supplementation (20% of expected DMI); while lower weight gains from supplementation with PLM (T3) indicate the possible utilisation of A. polyacantha leaves to overcome weight losses especially during dry seasons.  相似文献   


13.
The therapeutic enzyme asparaginase, which is used for the treatment of acute lymphoblastic leukaemia, is industrially produced by the bacteria Escherichia coli or Erwinia crysanthemi. In spite of its effectiveness as a therapeutic agent, the drug causes severe immunological reactions. As asparaginase is also produced by the yeast Saccharomyces cerevisiae, this microorganism could be considered for the production of the enzyme, providing an alternative antitumoral agent. In this study the ASP3 gene, that codes for the periplasmic, nitrogen regulated, asparaginase II from S. cerevisiae, was cloned and expressed in the methylotrophic yeast Pichia pastoris, under the control of the AOX1 gene promoter. Similarly to S. cerevisiae the heterologous enzyme was addressed to the P. pastoris cell periplasmic space. Enzyme yield per dry cell mass reached 800 U g−1, which was seven fold higher than that obtained using a nitrogen de-repressed ure2 dal80 S. cerevisiae strain. High cell density cultures performed with P. pastoris harbouring the ASP3 gene using a 2 l instrumented bioreactor, where biomass concentration reached 107 g l−1, resulted in a dramatic increase in volumetric yield (85,600 U l−1) and global volumetric productivity (1083 U l−1 h−1).  相似文献   

14.
The production of hGM-CSF was investigated in both a flask and a 5-l bioreactor, using transgenic Nicotiana tabacum suspension cells. While the maximum cell density and secreted hGM-CSF in the flask were 15.4 g l−1 and 6.5 μg l−1, respectively, those in the bioreactor were 15.6 g l−1 and 7.6 μg l−1. No detectable growth inhibition, shorter production of hGM-CSF and reduced cell viability in the batch bioreactor were observed under the specific conditions used compared with the flask culture. To improve the productivity, a perfusion culture was carried out in the bioreactor, with three different perfusion rates (0.5, 1.0 and 2.0 day−1). In all cases, the hGM-CSF in the medium was significantly increased during the overall culture period (16 days), with maximum values 3.0-, 9.4- and 6.0-fold higher than those obtained in the batch cultures, respectively, even though the intracellular hGM-CSF content was not significantly varied by the perfusion rate. In terms of the total amount of hGM-CSF secreted, 205.5, 1073.2 and 1246.3 μg accumulated in the perfusate within 16 days at the perfusion rates of 0.5, 1.0 and 2.0 day−1, respectively. It was concluded that the beneficial effect of perfusion on the production of hGM-CSF originated from the reduced proteolytic degradation due to the lower protease activity caused by the perfusion. Additionally, the cell growth and physiology in the perfusion culture were somewhat negatively affected by the increased perfusion rate, although the dry cell density steadily increased, and as a result, 19.4, 22.4 and 22.9 g l−1 of maximum cells were obtained with perfusion rates of 0.5, 1.0 and 2.0 day−1, respectively. This work highlighted the importance of proteolytic degradation in plant cell cultures for the production of secretory proteins and the feasibility of perfusion strategies for the continuous production of foreign proteins by the prevention of protein loss due to proteolytic enzymes.  相似文献   

15.
The effects of oxygen supply within the range 20.8–50% (using pure oxygen and air), on cell cultures of Panax ginseng were investigated in a balloon-type bubble bioreactor (5 L capacity, containing 4 L Murashige and Skoog medium, supplemented with 7.0 mg L−1 indolebutyric acid, 0.5 mg L−1 kinetin and 30 g L−1 sucrose). A 40% oxygen supply was found to be optimal for the production of both cell mass and saponin yielding values of 12.8 g (DW) L−1, 4.5 mg (g DW)−1 on day 25, respectively. Low (20.8%, 30%) and high (50%) oxygen concentration supplies were unfavorable to cell growth and saponin accumulation. The results indicate that oxygen supplementation to bioreactor-based ginseng cultures was beneficial for biomass accumulation and saponin production.  相似文献   

16.
The bioconversion of propionitrile to propionamide was catalysed by nitrile hydratase (NHase) using resting cells of Microbacterium imperiale CBS 498-74 (formerly, Brevibacterium imperiale). This microorganism, cultivated in a shake flask, at 28 °C, presented a specific NHase activity of 34.4 U mgDCW−1 (dry cell weight). The kinetic parameters, Km and Vmax, tested in 50 mM sodium phosphate buffer, pH 7.0, in the propionitrile bioconversion was evaluated in batch reactor at 10 °C and resulted 21.6 mM and 11.04 μmol min−1 mgDCW−1, respectively. The measured apparent activation energy, 25.54 kJ mol−1, indicated a partial control by mass transport, more likely through the cell wall.

UF-membrane reactors were used for kinetic characterisation of the NHase catalysed reaction. The time dependence of enzyme deactivation on reaction temperature (from 5 to 25 °C), on substrate concentrations (from 100 to 800 mM), and on resting cell loading (from 1.5 to 200 μg  ml−1) indicated: lower diffusional control (Ea=37.73 kJ mol−1); and NHase irreversible damage caused by high substrate concentration. Finally, it is noteworthy that in an integral reactor continuously operating for 30 h, at 10 °C, 100% conversion of propionitrile (200 mM) was attained using 200 μg  ml−1 of resting cells, with a maximum volumetric productivity of 0.5 g l−1 h−1.  相似文献   


17.
Hairy root cultures of red beet (Beta vulgaris) were grown in 3 l bubble column reactor for studying growth and pigment production under the influence of polyamines (PA) and elicitor treatment. Earlier studies with shake flask cultures had shown that combined feeding of spermidine (spd) and putrescine (put) (each 0.75 mM) significantly enhanced betalaine productivity in hairy root cultures of red beet. The present study has been focused on betalaine production in 3 l bubble column bioreactor where the growth pattern and betalaine synthesis under the influence of similar levels of polyamines were followed. A combination of spermidine and putrescine fed to the roots each at levels of 0.75 mM efficiently increased growth and pigment production resulting in 1.23-fold higher biomass (39.2 g FW l−1) and 1.27-fold higher betalaine content (32.9 mg g−1 DW) than control. Treatments with various levels of elicitor-methyl jasmonate (MJ), though progressively retarded biomass, at 40 μM level resulted in a significant increase in betalaine content resulting in 36.13 mg g−1 DW which was 1.4-fold higher than the control. Further higher concentrations of methyl jasmonate treatments supported high as well as rapid accumulation of betalaines, the overall betalaine productivity was hampered mainly because of the inhibitory action on biomass. Pigment release studies with cetyl trimethyl ammonium bromide (CTAB) resulted in optimization of concentration for better efflux of betalaines without showing any inhibitory effect on hairy root viability. These studies on product enhancement and on-line extraction of pigment are useful for developing a bioreactor system for betalaine production using B. vulgaris hairy root cultures. In particular the use of elicitors and efflux studies provide an insight for integrating unit operations and developing a process for continuous operation and higher production of phytochemicals.  相似文献   

18.
Combined effects of UVB radiation and CO2 concentration on plant reproductive parts have received little attention. We studied morphological and physiological responses of siliquas and seeds of canola (Brassica napus L. cv. 46A65) to UVB and CO2 under four controlled experimental conditions: UVB radiation (4.2 kJ m−2 d−1) with ambient level of CO2 (370 μmol mol−1) (control); UVB radiation (4.2 kJ m−2 d−1) with elevated level of CO2 (740 μmol mol−1); no UVB radiation (0 kJ m−2 d−1) with ambient level of CO2 (370 μmol mol−1); and no UVB radiation (0 kJ m−2 d−1) with elevated level of CO2 (740 μmol mol−1). UVB radiation affected the outer appearance of siliquas, such as colour, as well as their anatomical structures. At both CO2 levels, the UVB radiation of 4.2 kJ m−2 d−1 reduced the size of seeds, which had different surface patterns than those from no UVB radiation. At both CO2 levels, 4.2 kJ m−2 d−1 of UVB decreased net CO2 assimilation (AN) and water use efficiency (WUE), but had no effect on transpiration (E). Elevated CO2 increased AN and WUE, but decreased E, under both UVB conditions. At both CO2 levels, the UVB radiation of 4.2 kJ m−2 d−1 decreased chlorophyll fluorescence, total chlorophyll (Chl), Chl a and Chl b, but had no effect on the ratio of Chl a/b and the concentration of UV-screening pigments. Elevated CO2 increased total Chl and the concentration of UV-screening pigments under 4.2 kJ m−2 d−1 of UVB radiation. Neither UVB nor CO2 affected wax content of siliqua surface. Many significant relationships were found between the above-mentioned parameters. This study revealed that UVB radiation exerts an adverse effect on canola siliquas and seeds, and some of the detrimental effects of UVB on these reproductive parts can partially be mitigated by CO2.  相似文献   

19.
The soluble β-glucan contents in the cell wall of yeasts were estimated by treating cells with Glucanex® 200G that contained mainly β1,3-glucanase and some β1,6-glucanase. The sensitivity of cell walls of 11 yeasts to various concentrations of β-glucanase was compared. The yeasts that are resistant to β-glucanase treatment are expected to contain higher β-glucan content and those that are sensitive to the β-glucanase treatment are expected to contain lower β-glucan content. Two yeast strains were selected for further study by comparing the sensitivity of cell wall to β-glucanase; Candida bombicola and Candida albicans. Candida bombicola was more resistant and C. albicans was more sensitive to the Glucanex® 200G treatment. The results of enzyme sensitivity tests were verified by quantification of soluble β-glucan content purified from the yeasts. Much larger amount of soluble β-glucan was obtained from the cell walls of C. bombicola (0.08 g g−1 dried cell) than C. albicans (0.025 g g−1 dried cell).  相似文献   

20.
Investment by bladderwort (Utricularia foliosa) in carnivory, in terms of total C and N of bladders per leaf, was estimated in places with different nutrient concentrations from the Yahuarcaca Creek in the Colombian Amazon. The aims were to determine whether nutrient limiting conditions stimulate the investment in carnivory, and the relative balance between C and N invested in carnivory versus C and N obtained from prey. There were no significant differences either for phosphate (PO43−) concentration or for ammonia (NH4+) concentration among five sampling areas, along approximately 5 km long stretch of the creek, with a pooled mean ± S.D. of 0.19 ± 0.06 and 8.6 ± 3.0 μM, respectively. However, there were significant differences in the nitrate (NO3) concentration ranging from 0.6 to 2.5 μM. Total C and N of bladders per leaf increased with decreasing NO3. This corroborates the hypotheses that the carnivorous plant U. foliosa optimises its investment in carnivory according to nutrient availability in the water, and that N is a limiting factor that stimulates the investment in carnivory. The numbers of prey per bladder were also higher under NO3 limitation, thus enhancing the input of nutrients toward the plant through the bladders. The ratio of total C of prey captured/total C invested in bladders was always lower than 1. However, the efficiency of N was higher since when NO3 concentration was lower than 1 μM, the ratio of total N of prey captured/total N invested in bladders ranged between 0.97 and 1.67.  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