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.
Aeration and agitation are important variables to ensure effective oxygen transfer rate during aerobic bioprocesses; therefore, the knowledge of the volumetric mass transfer coefficient (kLa) is required. In view of selecting the optimum oxygen requirements for extractive fermentation in aqueous two-phase system (ATPS), the kLa values in a typical ATPS medium were compared in this work with those in distilled water and in a simple fermentation medium, in the absence of biomass. Aeration and agitation were selected as the independent variables using a 22 full factorial design. Both variables showed statistically significant effects on kLa, and the highest values of this parameter in both media for simple fermentation (241 s−1) and extractive fermentation with ATPS (70.3 s−1) were observed at the highest levels of aeration (5 vvm) and agitation (1200 rpm). The kLa values were then used to establish mathematical correlations of this response as a function of the process variables. The exponents of the power number (N3D2) and superficial gas velocity (Vs) determined in distilled water (α = 0.39 and β = 0.47, respectively) were in reasonable agreement with the ones reported in the literature for several aqueous systems and close to those determined for a simple fermentation medium (α = 0.38 and β = 0.41). On the other hand, as expected by the increased viscosity in the presence of polyethylene glycol, their values were remarkably higher in a typical medium for extractive fermentation (α = 0.50 and β = 1.0). A reasonable agreement was found between the experimental data of kLa for the three selected systems and the values predicted by the theoretical models, under a wide range of operational conditions.  相似文献   

2.
Nine microalgal species from the classes Bacillariophyceae, Cryptophyceae, Prymnesiophyceae and Dinophyceae were isolated from Australian waters, cultured to stationary phase and analyzed for their lipid and fatty acid composition and yield. Five species (Pavlova pinguis, Heterocapsa niei, Proteomonas sulcata, Navicula jeffreyi and Thalassiosira pseudonana) produced high proportions of triacylglycerol (TAG: 22–57% total lipid). An unidentified Navicula-like diatom (CS-786), despite having a low TAG content, had the highest EPA yield (5.8 mg L−1), due to high biomass and a high relative proportion of EPA. Heterocapsa niei had the highest DHA yield (2.9 mg L−1), due to a high cellular lipid and DHA content (171 pg cell−1 and 13.7 pg cell−1, respectively) despite its relatively low biomass. The desirable PUFA composition and yield of both diatom CS-786 and H. niei make them potential candidates for optimization of biomass and PUFA production for use as live-feeds in aquaculture. In addition, H. niei may have potential as a source of DHA for other uses. Low proportions (< 1.2%) of 24:6(n−3) accompanied by trace proportions of 24:5(n−6) were detected in most strains, while 28:8(n−3) was found in dinoflagellates and also in the prymnesiophyte P. pinguis. All non-diatomaceous species contained 26:7(n−3) in minor quantities. This is the first time these unusual C24 and C26 PUFA have been reported in microalgae and the first report of C28 PUFA in a microalga other than dinoflagellates. Possible biosynthetic reasons why these might occur in stationary phase cultures are considered and the likely dietary transfer of these PUFA to higher aquatic life is discussed.  相似文献   

3.
The goal of the present study was to assess how genetic loss of microsomal prostaglandin E2 synthase-1 (mPGES-1) affects acute cardiac ischemic damage after coronary occlusion in mice. Wild type (WT), heterozygous (mPGES-1+/−), and homozygous (mPGES-1−/−) knockout mice were subjected to left coronary artery occlusion. At 24 h, myocardial infarct (MI) volume was measured histologically. Post-MI survival, plasma levels of creatine phosphokinase (CPK) and cardiac troponin-I, together with MI size, were similar in WT, mPGES-1+/− and mPGES-1−/− mice. In contrast, post-MI survival was reduced in mPGES-1−/− mice pretreated with I prostanoid receptor (IP) antagonist (12/16) compared with vehicle-treated controls (13/13 mPGES-1−/−) together with increased CPK and cardiac troponin-I release. The deletion of mPGES-1 in mice results in increased prostacyclin I2 (PGI2) formation and marginal effects on the circulatory prostaglandin E2 (PGE2) level. We conclude that loss of mPGES-1 results in increased PGI2 formation, and in contrast to inhibition of PGI2, without worsening acute cardiac ischemic injury.  相似文献   

4.
Net ecosystem exchange of CO2 (NEE) was measured during 2005 using the eddy covariance (EC) technique over a reed (Phragmites australis (Cav.) Trin. ex Steud.) wetland in Northeast China (121°54′E, 41°08′N). Diurnal NEE patterns varied markedly among months. Outside the growing season, NEE lacked a diurnal pattern and it fluctuated above zero with an average value of 0.07 mg CO2 m−2 s−1 resulting from soil microbial activity. During the growing season, NEE showed a distinct V-like diel course, and the mean daily NEE was −7.48 ± 2.74 g CO2 m−2 day−1, ranging from −13.58 g CO2 m−2 day−1 (July) to −0.10 g CO2 m−2 day−1 (October). An annual cycle was also apparent, with CO2 uptake increasing rapidly in May, peaking in July, and decreasing from August. Monthly cumulative NEE ranged from −115 ± 24 g C m−2 month−1 (the reed wetland was a CO2 sink) in July to 75 ± 16 g C m−2 month−1 (CO2 source) in November. The annual CO2 balance suggests a net uptake of −65 ± 14 g C m−2 year−1, mainly due to the gains in June and July. Cumulative CO2 emission during the non-growing season was 327 g C m−2, much greater than the absolute value of the annual CO2 balance, which proves the importance of the wintertime CO2 efflux at the study site. The ratio of ecosystem respiration (Reco) to gross primary productivity (GPP) for this reed ecosystem was 0.95, indicating that 95% of plant assimilation was consumed by the reed plant or supported the activities of heterotrophs in the soil. Daytime NEE values during the growing season were closely related to photosynthetically active radiation (PAR) (r2 > 0.63, p < 0.01). Both maximum ecosystem photosynthesis rate (Amax) and apparent quantum yield (α) were season-dependent, and reached their peak values in July (1.28 ± 0.11 mg CO2 m−2 s−1, 0.098 ± 0.027 μmol CO2 μmol−1 photon, respectively), corresponding to the observed maximum NEE in July. Ecosystem respiration (Reco) relied on temperature and soil water content, and the mean value of Q10 was about 2.4 with monthly variation ranging from 1.8 to 4.1 during 2005. Annual methane emission from this reed ecosystem was estimated to be about 3 g C m−2 year−1, and about 5% of the net carbon fixed by the reed wetland was released to the atmosphere as CH4.  相似文献   

5.
The seasonal variability of specific growth rate and the carbon stable isotope ratio (δ13C) of leaf blades (δ13Cleaf) of a temperate seagrass, Zostera marina (within 10 days old) were measured simultaneously, together with the δ13C of dissolved inorganic carbon (δ13CDIC) at three sites in the semi-closed Akkeshi estuary system, northeastern Japan, in June, September, and November 2004. The δ13Cleaf ranged from −16.2 to −6.3‰ and decreased from summer to winter. The simultaneous measurement of the δ13Cleaf, growth rate, and morphological parameters (mean leaf length and width, mean number of leaves per shoot, and sheath length) of the seagrass and δ13CDIC in the surrounding water allowed us to compare directly the δ13Cleaf and specific growth rate of seagrass. The difference in the δ13C of seagrass leaves relative to the source DIC (Δδ13Cleaf − DIC) was the least negative (−11 to −7‰) in June at all three sites and became more negative (−17 to −8‰) as the specific growth rate decreased. This positive correlation between Δδ13Cleaf − DIC and specific growth rate can be used to diagnose the growth of seagrasses. Δδ13Cleaf − DIC changed by −1.7 ± 0.2‰ when the leaf specific growth rate decreased by 1% d−1.  相似文献   

6.
In this paper the utilization of the cyanobacteria Anabaena sp. in carbon dioxide removal processes is evaluated. For this, continuous cultures of this strain were performed at different dilution rates; alternatives for the recovery of the organic matter produced being also studied. A maximum CO2 fixation rate of 1.45 g CO2 L−1 day−1 was measured experimentally, but it can be increased up to 3.0 g CO2 L−1 day−1 outdoors. The CO2 is mainly transformed into exopolysaccharides, biomass representing one third of the total organic matter produced. Organic matter can be recovered by sedimentation with efficiencies higher than 90%, the velocity of sedimentation being 2 · 10−4 s−1. The major compounds were carbohydrates and proteins with productivities of 0.70 and 0.12 g L−1 day−1, respectively. The behaviour of the cultures of Anabaena sp. has been modelized, also the characteristics parameters requested to design separation units being reported. Finally, to valorizate the organic matter as biofertilizers and biofuels is proposed.  相似文献   

7.
The purpose of this study was to localize vasopressin (VP) V1a receptor in stomach and to characterize the role of VP in the regulation of gastric motility in rats. Double staining was used to locate the V1a receptor in the gastric body of the rat. The contraction of the circular muscle strips of gastric body was monitored by a polygraph. V1a receptor was expressed on the neurons of myenteric plexus of the gastric body. VP (10− 10–10− 6 M) caused a concentration-dependent contractile effect on the circular muscle strips of gastric body in vitro. V-1880 ([deamino-Pen1, O-Me-Tyr2, Arg8]-Vasopressin, 10− 7 M), a V1 receptor antagonist, inhibited the spontaneous contraction of the strips. Tetradotoxin (TTX, 10− 6 M) and V-1880 (10− 7 M) abolished the excitatory effect of VP. Atropine (10− 6 M) partially inhibited VP-induced excitatory effect on the muscle strips but hexamethonium (10− 4 M) did not influence it. These results suggest that V1a receptor was expressed on the neurons of myenteric nerves. The cholinergic nerve was involved in the excitatory effect of VP on the contraction of gastric body.  相似文献   

8.
One and a half year-old Ginkgo saplings were grown for 2 years in 7 litre pots with medium fertile soil at ambient air CO2 concentration and at 700 μmol mol−1 CO2 in temperature and humidity-controlled cabinets standing in the field. In the middle of the 2nd season of CO2 enrichment, CO2 exchange and transpiration in response to CO2 concentration was measured with a mini-cuvette system. In addition, the same measurements were conducted in the crown of one 60-year-old tree in the field. Number of leaves/tree was enhanced by elevated CO2 and specific leaf area decreased significantly.CO2 compensation points were reached at 75–84 μmol mol−1 CO2. Gas exchange of Ginkgo saplings reacted more intensively upon CO2 than those of the adult Ginkgo. On an average, stomatal conductance decreased by 30% as CO2 concentration increased from 30 to 1000 μmol mol−1 CO2. Water use efficiency of net photosynthesis was positively correlated with CO2 concentration levels. Saturation of net photosynthesis and lowest level of stomatal conductance was reached by the leaves of Ginkgo saplings at >1000 μmol mol−1 CO2. Acclimation of leaf net CO2 assimilation to the elevated CO2 concentration at growth occurred after 2 years of exposure. Maximum of net CO2 assimilation was 56% higher at ambient air CO2 concentration than at 700 μmol mol−1 CO2.  相似文献   

9.
The direct immobilization of glucose oxidase (GOD) on TiO2/SiO2 nanocomposite and its application as glucose biosensor were investigated. The room-temperature phosphorescence of TiO2/SiO2 nanocomposite can be quenched by hydrogen peroxide (H2O2). The detection of glucose may be accomplished by monitoring the formation of hydrogen peroxide which generated in the oxidation process of glucose with the catalysis of GOD. To our surprise, by using a 96-hole polyporous plate accessory of fluorescence spectrophotometer, the biosensor exhibits excellent linear response to glucose concentrations ranging from 1.0 × 10−9 to 1.0 × 10−2 M with a detection limit of 1.2 × 10−10 M. The TiO2/SiO2 nanocomposite can be used as both supporting material and signal transducer. The phosphorescence intensity and color of the biosensor change obviously and even could be observed with naked eyes by continuous addition of glucose. Based on the room-temperature phosphorescence of TiO2/SiO2 nanocomposite, a new method of solid substrate-room-temperature phosphorimetry (SS-RTP) for glucose determination is proposed. A glucose biosensor was fabricated with wide determination concentration range, low detection limit, high sensitivity, and fast response time. And the biosensor has been successfully applied to the determination of glucose in human blood serum. The coacervation of GOD enzyme and its interaction with TiO2/SiO2 nanocomposite enlarge the surface area and enhance the chemical stability of GOD. The nice biocompatibility, large surface area, good chemical stability and nontoxicity of the TiO2/SiO2 nanocomposite have made this material suitable for functioning as biosensor.  相似文献   

10.
The nitrogen uptake and growth capabilities of the potentially harmful, raphidophycean flagellate Heterosigma akashiwo (Hada) Sournia were examined in unialgal batch cultures (strain CCMP 1912). Growth rates as a function of three nitrogen substrates (ammonium, nitrate and urea) were determined at saturating and sub-saturating photosynthetic photon flux densities (PPFDs). At saturating PPFD (110 μE m−2 s−1), the growth rate of H. akashiwo was slightly greater for cells grown on NH4+ (0.89 d−1) compared to cells grown on NO3 or urea, which had identical growth rates (0.82 d−1). At sub-saturating PPFD (40 μE m−2 s−1), both urea- and NH4+-grown cells grew faster than NO3-grown cells (0.61, 0.57 and 0.46 d−1, respectively). The N uptake kinetic parameters were investigated using exponentially growing batch cultures of H. akashiwo and the 15N-tracer technique. Maximum specific uptake rates (Vmax) for unialgal cultures grown at 15 °C and saturating PPFD (110 μE m−2 s−1) were 28.0, 18.0 and 2.89 × 10−3 h−1 for NH4+, NO3 and urea, respectively. The traditional measure of nutrient affinity—the half saturation constants (Ks) were similar for NH4+ and NO3 (1.44 and 1.47 μg-at N L−1), but substantially lower for urea (0.42 μg-at N L−1). Whereas the α parameter (α = Vmax/Ks), which is considered a more robust indicator for substrate affinity when substrate concentrations are low (<Ks), were 19.4, 12.2 and 6.88 × 10−3 h−1/(μg-at N L−1) for NH4+, NO3 and urea, respectively. These laboratory results demonstrate that at both saturating and sub-saturating N concentrations, N uptake preference follows the order: NH4+ > NO3 > urea, and suggests that natural blooms of H. akashiwo may be initiated or maintained by any of the three nitrogen substrates examined.  相似文献   

11.
Prostaglandin I2 potentiated the paw swelling induced by carrageenin in rats. Prostaglandin I2 (0.1 μg) showed similar activity to PGE1 (0.01 μg). This potentiating property disappeared in 60 minutes and was completely abolished by diphenhydramine (25 mg kg−1, i.p.). In vascular permeability tests, PGI2 itself (2.5 × 10−10 mol, 88 ng) caused no dye leakage reaction, but PGE1 (2.5 × 10−10 mol, 88.5 ng) caused a significant dye leakage. This effect of PGE1 was statistically significant compared with vehicle- or PGI2-treated group (p<0.05). Prostaglandin I2 potentiated the increased vascular permeability induced by 5-hydroxytriptamine (2.5 × 10−10 mol), bradykinin (5 × 10−10 mol) and histamine (2 × 10−10 to 2 × 10−8 mol). The potentiation was the most evidence in the case of histamine.  相似文献   

12.
Effects of salinity and nitrate nitrogen (NO3-N) on ion accumulation and chlorophyll fluorescence were monitored for two populations of Suaeda salsa grown from seeds in a greenhouse experiment. One population inhabits the intertidal zone and the other occurs on inland saline soils. Ion contents in soils and in leaves of the two populations were also investigated in field. In the greenhouse, seedlings were exposed to a NaCl concentration of 0.6 and 35.1 ppt, with 0.1 or 5 mM NO3-N treatments for 20 days. The contents of Na+ and Cl were higher, but NO3 was lower in soils of the intertidal zone than at the inland site. In the field, ion concentrations and the estimated contribution of these ions to osmotic potential in leaves showed no difference between the two populations, except that the estimated contribution of Na+ to osmotic potential in leaves of the intertidal population was lower than that in the inland population. In the greenhouse, in contrast, the concentration of Cl was lower, but NO3 concentration and the estimated contribution of NO3 to osmotic potential were higher, in the leaves of plants from the intertidal zone. Salinity had no effect on the maximal efficiency of PSII photochemistry (Fv/Fm) and the actual PSII efficiency (ΦPSII). The results indicated that S. salsa from the intertidal zone was better able to regulate Cl to a lower level, and accumulate NO3 even with low soil NO3 concentrations. Tolerance of the PSII machinery to high salinity stress may be an important characteristic for the studied species supporting growth in highly saline environments.  相似文献   

13.
Can elevated CO(2) improve salt tolerance in olive trees?   总被引:2,自引:0,他引:2  
We compared growth, leaf gas exchange characteristics, water relations, chlorophyll fluorescence, and Na+ and Cl concentration of two cultivars (‘Koroneiki’ and ‘Picual’) of olive (Olea europaea L.) trees in response to high salinity (NaCl 100 mM) and elevated CO2 (eCO2) concentration (700 μL L−1). The cultivar ‘Koroneiki’ is considered to be more salt sensitive than the relatively salt-tolerant ‘Picual’. After 3 months of treatment, the 9-month-old cuttings of ‘Koroneiki’ had significantly greater shoot growth, and net CO2 assimilation (ACO2) at eCO2 than at ambient CO2, but this difference disappeared under salt stress. Growth and ACO2 of ‘Picual’ did not respond to eCO2 regardless of salinity treatment. Stomatal conductance (gs) and leaf transpiration were decreased at eCO2 such that leaf water use efficiency (WUE) increased in both cultivars regardless of saline treatment. Salt stress increased leaf Na+ and Cl concentration, reduced growth and leaf osmotic potential, but increased leaf turgor compared with non-salinized control plants of both cultivars. Salinity decreased ACO2, gs, and WUE, but internal CO2 concentrations in the mesophyll were not affected. eCO2 increased the sensitivity of PSII and chlorophyll concentration to salinity. eCO2 did not affect leaf or root Na+ or Cl concentrations in salt-tolerant ‘Picual’, but eCO2 decreased leaf and root Na+ concentration and root Cl concentration in the more salt-sensitive ‘Koroneiki’. Na+ and Cl accumulation was associated with the lower water use in ‘Koroneiki’ but not in ‘Picual’. Although eCO2 increased WUE in salinized leaves and decreased salt ion uptake in the relatively salt-tolerant ‘Koroneiki’, growth of these young olive trees was not affected by eCO2.  相似文献   

14.
The spatial upscaling of soil respiration from field measurements to ecosystem levels will be biased without studying its spatial variation. We took advantage of the unique spatial gradients of an oak–grass savanna ecosystem in California, with widely spaced oak trees overlying a grass layer, to study the spatial variation in soil respiration and to use these natural gradients to partition soil respiration according to its autotrophic and heterotrophic components. We measured soil respiration along a 42.5 m transect between two oak trees in 2001 and 2002, and found that soil respiration under tree canopies decreased with distance from its base. In the open area, tree roots have no influence on soil respiration. Seasonally, soil respiration increased in spring until late April, and decreased in summer following the decrease in soil moisture content, despite the further increase in soil temperature. Soil respiration significantly increased following the rain events in autumn. During the grass growing season between November and mid-May, the average of CO2 efflux under trees was 2.29 μmol m−2 s−1, while CO2 efflux from the open area was 1.40 μmol m−2 s−1. We deduced that oak root respiration averaged as 0.89 μmol m−2 s−1, accounting for 39% of total soil respiration (oak root + grass root + microbes). During the dry season between mid-May and October, the average of CO2 efflux under trees was 0.87 μmol m−2 s−1, while CO2 efflux from the open areas was 0.51 μmol m−2 s−1. Oak root respiration was 0.36 μmol m−2 s−1, accounting for 41% of total soil respiration (oak root + microbes). The seasonal pattern of soil CO2 efflux under trees and in open areas was simulated by a bi-variable model driven by soil temperature and moisture. The diurnal pattern was influenced by tree physiology as well. Based on the spatial gradient of soil respiration, spatial analysis of crown closure and the simulation model, we spatially and temporally upscaled chamber measurements to the ecosystem scale. We estimated that the cumulative soil respiration in 2002 was 394 gC m−2 year−1 in the open area and 616 gC m−2 year−1 under trees with a site-average of 488 gC m−2 year−1.  相似文献   

15.
The effects of NaCl salinity on growth, morphology and photosynthesis of Salvinia natans (L.) All. were investigated by growing plants in a growth chamber at NaCl concentrations of 0, 50, 100 and 150 mM. The relative growth rates were high (ca. 0.3 d−1) at salinities up to 50 mM and decreased to less than 0.2 d−1 at higher salinities, but plants produced smaller and thicker leaves and had shorter stems and roots, probably imposed by the osmotic stress and lowered turgor pressure restricting cell expansion. Na+ concentrations in the plant tissue only increased three-fold, but uptake of K+ was reduced, resulting in very high Na+/K+ ratios at high salinities, indicating that S. natans lacks mechanisms to maintain ionic homeostasis in the cells. The contents of proline in the plant tissue increased at high salinity, but concentrations were very low (<0.1 μmol g−1 FW), indicating a limited capacity of S. natans to synthesize proline as a compatible compound. The potential photochemical efficiency of PSII (Fv/Fm) of S. natans remained unchanged at 50 mM NaCl but was reduced at higher salinities, and the photosynthetic capacity (ETRmax) was significantly reduced at 50 mM NaCl and higher. It is concluded that S. natans is a salt-sensitive species lacking physiological measures to cope with exposure to high NaCl salinity. At low salinities salts are taken up and accumulate in old leaves, and high growth rates are maintained because new leaves are produced at a higher rate than for plants not exposed to salt.  相似文献   

16.
Red tide blooms of Cochlodinium polykrikoides in a coastal cove   总被引:1,自引:0,他引:1  
Successive blooms of the dinoflagellate Cochlodinium polykrikoides occurred in Pettaquamscutt Cove, RI, persisting from September through December 1980 and again from April through October 1981. Cell densities varied from <100 cells L−1 at the onset of the bloom and reached a maximum density exceeding 3.4 × 106 cells L−1 during the summer of 1981. The bloom was mainly restricted to the mid to inner region of this shallow cove with greatest concentrations localized in surface waters of the southwestern region during summer/fall periods of both years. Highly motile cells consisting of single, double and multiple cell zooids were found as chains of 4 and 8 cells restricted to the late August/September periods. The highest cell densities occurred during periods when annual temperatures were between 19 and 28 °C and salinities between 25 and 30. A major nutrient source for the cove was Crying Brook, located at the innermost region at the head of the cove. Inorganic nitrogen (NH3 and NO2 + NO3) from the brook was continually detectable throughout the study with maximum values of 57.5 and 82.5 μmol L−1, respectively. Phosphate (PO4-P) was always present in the source waters and rarely <0.5 μmol L−1; silicate always exceeded 30 μmol L−1 with maximum concentrations reaching 226 μmol L−1. Chlorophyll a and ATP concentrations during the blooms varied directly with cell densities. Maximum Chl a levels were 218 mg m−3 and ATP-carbon was >20 g C m−3. Primary production by the dinoflagellate-dominated community during the bloom varied between 4.3 and 0.07 g C m−3 d−1. Percent carbon turnover calculated from primary production values and ATP-carbon varied from 6 to 129% d−1. The dinoflagellates dominated the entire summer period; other flagellates and diatoms were present in lesser amounts. A combination of low washout rate due to the cove dynamics, active growth, and life cycles involving cysts allowed C. polykrikoides to maintain recurrent bloom populations in this area.  相似文献   

17.
A novel β-glucosidase from Fusarium proliferatum ECU2042 (FPG) was successfully purified to homogeneity with a 506-fold increase in specific activity. The molecular mass of the native purified enzyme (FPG) was estimated to be approximately 78.7 kDa, with two homogeneous subunits of 39.1 kDa, and the pI of this enzyme was 4.4, as measured by two-dimensional electrophoresis. The optimal activities of FPG occurred at pH 5.0 and 50 °C, respectively. The enzyme was stable at pH 4.0–6.5 and temperatures below 60 °C, and the deactivation energy (Ed) for FPG was 88.6 kJ mo1−1. Moreover, it was interesting to find that although the purified enzyme exhibited a very low activity towards p-nitrophenyl β-d-glucoside (pNPG), and almost no activity towards cellobiose, a relatively high activity was observed on ginsenoside Rg3. The enzyme hydrolyzed the 3-C, β-(1 → 2)-glucoside of ginsenoside Rg3 to produce ginsenoside Rh2, but did not sequentially hydrolyze the β-d-glucosidic bond of Rh2. The Km and Vmax values of FPG for ginsenoside Rg3 were 2.37 mM and 0.568 μmol (h mg protein)−1, respectively. In addition, this enzyme also exhibited significant activities towards various alkyl glucosides, aryl glucosides and several natural glycosides.  相似文献   

18.
Preference for NH4+ or NO3 nutrition by the perennial legume Sesbania sesban (L.) Merr. was assessed by supplying plants with NH4+ and NO3 alone or mixed at equal concentrations (0.5 mM) in hydroponic culture. In addition, growth responses of S. sesban to NH4+ and NO3 nutrition and the effects on root nodulation and nutrient and mineral composition of the plant tissues were evaluated in a hydroponic setup at a range of external concentration of NH4+ and NO3 (0, 0.1, 0.2, 0.5, 2 and 5 mM). Seedlings of S. sesban grew equally well when supplied with either NH4+ or NO3 alone or mixed and had high relative growth rates (RGRs) ranging between 0.19 and 0.21 d−1. When larger plants of S. sesban were supplied with NH4+ or NO3 alone, the RGRs and shoot elongation rates were not affected by the external concentration of inorganic N. At external N concentrations up to 0.5 mM nodulation occurred and contributed to the N nutrition through fixation of gaseous N2 from the atmosphere. For both NH4+ and NO3-fed plants the N concentration in the plant tissues, particularly water-extractable NO3, increased at high supply concentrations, and concentrations of mineral cations generally decreased. It is concluded that S. sesban can grow without an external inorganic N supply by fixing atmospheric N2 gas via root nodules. Also, S. sesban grows well on both NH4+ and NO3 as the external N source and the plant can tolerate relatively high concentrations of NH4+. This wide ecological amplitude concerning N nutrition makes S. sesban very useful as a N2-fixing fallow crop in N deficient areas and also a candidate species for use in constructed wetland systems for the treatment of NH4+ rich waters.  相似文献   

19.
The dimorphic fungus Mucor indicus was grown in different forms classified as purely filamentous, mostly filamentous, mostly yeast-like and purely yeast-like, and the relationship between morphology and metabolite production, inhibitor tolerance and the cell wall composition was investigated. Low concentrations of spores in the inoculum with subsequent aeration promoted filamentous growth, whereas higher spore concentrations and anaerobic conditions promoted yeast-like growth. Ethanol was the main metabolite with glycerol next under all conditions tested. The yields of ethanol from glucose were between 0.39 and 0.42 g g−1 with productivities of 3.2–5.0 g l−1 h−1. The ethanol productivity of mostly filamentous cells was increased from 3.9 to 5.0 g l−1 h−1 by the presence of oxygen, whereas aeration of purely yeast-like cells showed no such effect. All growth forms were able to tolerate 4.6 g l−1 furfural and 10 g l−1 acetic acid and assimilate the sugars, although with different consumption rates. The cell wall content of the fungus measured as alkali insoluble materials (AIM) of the purely yeast-like cells was 26% of the biomass, compared to 8% of the pure filaments. However, the chitosan concentration of the filaments was 29% of the AIM, compared to 6% of the yeast-like cells.  相似文献   

20.
Three columns were differentiated with feeding mixture of H2S and NH3 (MFC), feeding NH3 followed by H2S (NFC), and feeding H2S followed by NH3 (SFC). Removal performance, biodegradation capacity and microbial community structures in the three columns were compared. The results show that NFC has a shorter acclimation period for the removal of NH3 gas and nitrification than MFC. Under the high loading of H2S and NH3 at 164 and 82 g m−3 h−1, respectively, NFC exhibited high removal efficiency of NH3 (>95%) while the removal efficiencies were obtained at 63 and 75% in MFC and SFC, respectively. The removal of NH3 gas in NFC was significantly attributed to nitrification (over 50%), while adsorption and chemical reaction contributed to the removal of NH3 in MFC and SFC. The different biodegradation capacities of NH3 could be due to the dissimilarity in the microbial population presented in each column.  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