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.
We have isolated Chl a-Chl c-carotenoid binding proteins from the dinoflagellates Prorocentrum minimum and Heterocapsa pygmaea grown under high (500 mol m–2 s–1, HL) and low (35 mol m–2 s–1, LL) light conditions. We compared various isolation procedures of membrane bound light harvesting complexes (LHCs) and assayed the functionality of the solubilized proteins by determining the energy transfer efficiency from the accessory pigments to Chl a by means of fluorescence excitation spectra. The identity of the newly isolated protein-complexes were confirmed by immunological cross-reactions with antibodies raised against the previously described membrane bound Chl a-c proteins (Boczar et al. (1980) FEBS Lett 120: 243–247). Spectroscopic analysis demonstrated the relatedness of these proteins with the recently described Chl-a-c 2-peridinin (ACP) binding protein (Hiller et al. (1993) Photochem Photobiol 57: 125–131; Iglesias Prieto et al. (1993) Phil Trans R Soc London B 338: 381–392). The water-soluble peridinin-Chl a binding-protein (PCP) was not detectable in P. minimum. Two functional forms of ACP with different pigmentation were isolated. A variant of ACP which was isolated from high-light grown cells, that specifically binds increased amounts of diadinoxanthin was compared to the previously described ACPs that bind proportionately more peridinin.Abbreviations ACP Chl a-Chl c-peridinin binding protein - AEBSF 4-(2-aminoethyl)-benzenesulfonyl fluoride hydrochloride - DDM dodecyl -d maltoside - Deriphat 160 N-lauryl-beta-iminopropionic acid - HEPES (N-2-hydroxyethylpiparizine-N-2-ethanesulphonic acid) - HL high light (500 mol m–2 s–1) - LL low light (35 mol m–2 s–1) - 730 fluorescence yield (emission at 730 nm) - PCP peridinin-Chl a-binding protein - PMSF phenyl-methyl-sulfonyl-fluoride - PS I Photosystem I - PS II Photosystem II  相似文献   

2.
Summary Antelope ground squirrels (Ammospermophilus leucurus, 80–100 g) began surface activity 1.0–1.5 h after sunrise and ended it 0.5–1.25 h before sunset throughout the year near Barstow, California. Daily energy expenditure (DEE) of free-living animals measured with doubly labeled water (H3H18O) decreased from 1,340 kJ kg-1 d-1 in April to 970 in October. Resting metabolic rates (RMR) of freshly-captured, fed, ground squirrels varied through the year (22.1 J g-1 h-1 in August, 19.1 in January) but most of the change in DEE could be explained by differences in thermoregulatory costs between seasons. The ground squirrels had lower rates of resting metabolism at night (15.3J g-1 H-1) than during the day.The cost of activity (calculated by subtracting 24 h resting costs from total DEE during August and October, periods when thermoregulatory costs were negligible) was 550 kJ kg-1 d-1 in August and October. Thus, activity accounted for about 50% of the total DEE. The mean rate of energy expenditure during the activity period, calculated as activity cost (kJ d-1) divided by activity time (h d-1), then plus RMR, was about 3xRMR. This multiplication factor may be useful as an estimator of foraging costs or in estimating DEE from time budgets.  相似文献   

3.
The growth and survival of juvenile Haliotis rubra, when fed with the diatom Navicula sp. cultured in f/2 medium containing combined nitrogen at 24.71 mg NO3-N L–1 (high), 12.35 mg NO3-N L–1 (standard) or 2.47 mg NO3-N L–1 (low), were compared in a 33-day trial. The alga in the low nitrogen medium contained 37% less total amino acid than that in the high and standard nitrogen media. There was a slightly greater reduction in essential amino acids (40%) compared to non-essential amino acids (35%). Juvenile abalone feeding on Navicula grown in medium with low nitrate and lower total amino acid content grew more slowly than when fed on the same species grown in standard or higher nitrogen medium with a higher amino acid content. The growth rate of juveniles was highest (43 m d–1) in the high nitrate treatment followed (40 m d–1) by the standard nitrate treatment and lowest (31 m d–1) in the low nitrate treatment. The survival of the juveniles was also effected by the diet. Survival was better in the high and standard nitrogen media (88%) than the low nitrogen medium (75%). The results suggest that in order to achieve uniformity in nutritional quality of diatoms and good growth of abalone juveniles in commercial abalone nurseries, the nitrogen concentration in tanks should be monitored and additional nitrate added to provide an optimum concentration of between 2 and 12 mg NO3-N L–1.  相似文献   

4.
1. In Lymnaea stagnalis L. (Pulmonata, Basommatophora) the neurons in the osphradium were visualized by staining through the inner right parietal nerve by 5,6-carboxyfluorescein (5,6-CF). Three types of neurons were identified: three large ganglionic cells (GC1-3; 80–100 m), the small putative sensory neurons (SC; 20 m) and very small sensory cells (3–5 m).2. The ganglionic and putative sensory neurons were investigated by whole cell patch-clamp method in current-clamp condition. The three giant ganglionic neurons (GC1-3) located closely to the root of osphradial nerve, had a membrane potential (MP) between –30 and –70 mV and showed tonic or bursting activities. The small putative sensory cells (SCs) scattered throughout the osphradial ganglion, possessed a MP between –25 and –55 mV and showed an irregular firing pattern with membrane oscillations. At resting MP the GC1-3 cells were depolarized and increased the frequency of their firing, while the SCs were hyperpolarized and inhibited by NaCl (10–2 M) and L-aspartate (10–5 M) applied to the osphradium.3. 5-Hydroxytryptamine (5HT, 10–6 M), -aminobutyric acid (GABA; 10–6 M) and the GABAB agonist baclofen (10–6 M) depolarized the neurons GC1-3 and increased their firing frequency. In contrast, on the GC1-3 neurons, acetylcholine (Ach; 10–6 M) and FMRFamide (10–6 M) caused hyperpolarization and cessation of the firing activity. The 5HT effect was blocked by mianserin (10–6 M) but picrotoxin (10–5 M) failed to block the GABA-induced effect on the GC1-3 cells.4. The small putative sensory neurons (SCs) were excited by Ach (10–6 M) and 5HT (10–6 M) but were inhibited by GABA (10–6 M). FMRFamide (10–6 M) had a biphasic response. The Ach effect was blocked by hexamethonium (10–6 M) and tetraethylammonium (10–6 M), indicating the involvement of nicotinic cholinergic receptors.5. The distinct responses of the two populations of osphradial neurons to chemical stimuli and neurotransmitters suggest that they can differently perceive signals from environment and hemolymph.  相似文献   

5.
Ulva rigida was cultivated in 7501 tanks at different densities with direct and continuous inflow (at 2, 4, 8 and 12 volumes d–1) of the effluents from a commercial marine fishpond (40 metric tonnes, Tm, of Sparus aurata, water exchange rate of 16 m3 Tm–1) in order to assess the maximum and optimum dissolved inorganic nitrogen (DIN) uptake rate and the annual stability of the Ulva tank biofiltering system. Maximum yields (40 g DW m–2 d–1) were obtained at a density of 2.5 g FW 1–1 and at a DIN inflow rate of 1.7 g DIN m–2 d–1. Maximum DIN uptake rates were obtained during summer (2.2 g DIN M–2 d–1), and minimum in winter (1.1 g DIN m–2 d–1) with a yearly average DIN uptake rate of 1.77 g DIN m–2 d–1 At yearly average DIN removal efficiency (2.0 g DIN m–2 d–1, if winter period is excluded), 153 m2 of Ulva tank surface would be needed to recover 100% of the DIN produced by 1 Tm of fish.Abbreviations DIN= dissolved inorganic nitrogen (NH inf4 sup+ + NO inf3 sup– + NO inf2 sup– ); - FW= fresh weight; - DW= dry weight; - PFD= photon flux density; - V= DIN uptake rate  相似文献   

6.
The field metabolic rates (FMR) and rates of water flux were measured in two species of varanid lizards over five periods of the year in tropical Australia. The energetics of these species were further investigated by directly measuring activity (locomotion) and body temperatures of free-ranging animals by radiotelemetry, and by measuring standard metabolic rate (over a range of body temperatures) and activity metabolism in the laboratory. Seasonal differences in the activity and energetics were found in these goannas despite similar, high daytime temperatures throughout the year in tropical Australia. Periods of inactivity were associated with the dry times of the year, but the onset of this period of inactivity differed with respect to habitat even within the same species. Varanus gouldii, which inhabit woodlands only, were inactive during the dry and late dry seasons. V. panoptes that live in the woodland had a similar seasonal pattern of activity, but V. panoptes living near the floodplain of the South Alligator River had their highest levels of activity during the dry season when they walked long distances to forage at the receding edge of the floodplain. However, during the late dry season, after the floodplain had dried completely, they too became inactive. For V. gouldii, the rates of energy expenditure were 196 kJ kg–1 day–1 for active animals and 66 kJ kg–1 day–1 for inactive animals. The rates of water influx for these groups were respectively 50.7 and 19.5 ml kg–1 day–1. For V. panoptes, the rates of energy expenditure were 143 kJ kg–1 day–1 for active animals and 56 kJ kg–1 day–1 for inactive animals. The rates of water influx for these two groups were respectively 41.4 and 21.0 ml kg–1 day–1. We divided the daily energy expenditure into the proportion of energy that lizards used when in burrows, out of burrows but inactive, and in locomotion for the two species during the different seasons. The time spent in locomotion by V. panoptes during the dry season is extremely high for a reptile (mean of 3.5 h/day spent walking), and these results provide an ecological correlate to the high aerobic capacity found in laboratory measurements of some species of varanids.  相似文献   

7.
The seasonal variation in primary production, individual numbers, and biomass of phyto- and zooplankton was studied in the River Danube in 1981. The secondary production of two dominant zooplankton species (Bosmina longirostris and Acanthocyclops robustus) was also estimated. In the growing season (April–Sept.) individual numbers dry weights and chlorophyll a contents of phytoplankton ranged between 30–90 × 106 individuals, l–1, 3–12 mg l–1, and 50–170 µg l–1, respectively. Species of Thalassiosiraceae (Bacillariophyta) dominated in the phytoplankton with a subdominance of Chlorococcales in summer. Individual numbers and dry weights of crustacean zooplankton ranged between 1400–6500 individuals m–3, and 1.2–12 mg m–3, respectively. The daily mean gross primary production was 970 mg C m–3 d–1, and the net production was 660 mg C m–3 d–1. Acanthocyclops robustus populations produced 0.2 mg C m–3 d–1 as an average, and Bosmina longirostris populations 0.07 mg C m–3 d–1. The ecological efficiency between phytoplankton and crustacean zooplankton was 0.03%.  相似文献   

8.
The focus of this microcosm study was to monitor the performances of 17 butane-utilizing microcosms during a long-term (100–250 days) aerobic cometabolic depletion of chloroform (CF). The depletion of the contaminant began after a lag-time variable between 0 and 23 days. All microcosms quickly reached a pseudo steady-state condition, in terms of biomass concentration (with an average of 9.3 × 106 CFU ml–1), chloroform depletion rate (5 mol l–1 d–1) and butane utilization rate (730 mol l–1 d–1). After about 100 days of CF depletion, a sudden 5- to 7-fold increase of the chloroform rate was observed in two microcosms, where the highest amount of contaminant had been depleted. In one of these high-performing microcosms, an experiment of chloroform depletion in the absence of butane resulted in the depletion of a surprisingly high amount of contaminant (765 molCF kgdry soil–1 in 2 months) and in a marked selection of a single bacterial strain. Bioaugmentation assays conducted with the biomass selected in this microcosm and with a pure culture of the selected strain immediately resulted in very high chloroform depletion rates. Preliminary results of a study conducted with resting cells of the selected strain indicated that it can degrade chloroform concentrations up to 119 M (14.2 mg l–1) without any sign of substrate toxicity, and that it is able to transform vinyl chloride and 1,1,2-trichloroethane.  相似文献   

9.
Summary The combined effect of various temperatures and light intensities on the growth of seven species of antarctic diatoms in culture has been studied. With the exception of Chaetoceros deflandrei whose thermal tolerance is fairly good, these obligatory psychrophils cannot survive in temperatures above 6° to 9° C. Their mean growth rate is relatively low, between 0.24 div d–1 for Corethron criophilum and 0.63 div d–1 for C. deflandrei. Regardless of light intensity, growth rate increased with the temperature to reach a maximum between 3° and 5° C. The highest rates were obtained between 115 and 220 mol m–2 s–1 with 0.38 div d–1 for C. criophilum, 0.56 div d–1 for Synedra sp. and between 0.71 and 0.88 div d–1 for the other 5 species. A reduction in light intensity from 220 to 46 mol m–2 s–1 slowed growth by nearly 50%. These results suggest that the combined effect of temperature and light is one of the factors involved in the limitation of antarctic phytoplankton growth. The low temperatures of the environment do not permit rapid growth, which, even under optimal light conditions remains low. In addition, in the euphotic layer, the overall light energy available for algae is considerably reduced due to turbulence, a factor which exacerbates the reduced growth rate.  相似文献   

10.
Summary The function of the caecal bulb, and its adaptation to chronic high- or low-Na+ intake, was investigated by in vivo perfusion of anaesthetised birds. Effects of acute aldosterone injection (125 g·kg–1 body mass) were also measured.Evidence was found for primary active net absorption of Na+, inducing parallel Na-linked absorption of water and Cl and secretion of K+. Around 20–35% of total Cl absorption and K+ secretion were independent of Na+ fluxes, and these components appear to be driven by passive processes with apparent conductances of 6.3×10–3 (G Cl) and 1.1×10–3 (G K) S·cm–2.Acetate (40mM) stimulated Na+ fluxes (8.5–9.9 Eq·cm–2·h–1) and Na-linked water fluxes (27–44 l·cm–2·h–1). Increased coupling ratios (2.9–4.6 l·Eq–1) and other data indicate that these effects may be due to increased osmotic permeabilities of barriers involved in the Na-linked water transfer pathway.Low-Na+ maintenance enhanced EPD (49–69 mV, serosa positive) and all net fluxes:J Na (6.8–11.6);J K (–3.2––4.3);J Cl (4.3–5.6 Eq·cm serosal area–2·h–1);J v (28–43 l·cm–2·h–1) (mucosal-serosal fluxes positive).Acute aldosterone enhancedJ Na (10.8–14.0 Eq·cm–2·h–1) and EPD (54–66 mV) by 3 h after injection, but had no effect on the Na-linked components ofJ K orJ Cl.Abbreviations ECPD, EPD Electrochemical or electrical potential difference - G Cl ,G K ionic conductances (Cl, K+) - J v ,J ion net volume or ion flux rate, mucosa-serosa positive;P d (Cl) diffusive permeability coefficient (of Cl) - SEDM standard error of difference between means  相似文献   

11.
Three Bromeliaceae species of the medium Orinoco basin, Venezuela, were compared in their light-use characteristics. The bromeliads studied were two species of pineapple, i.e. the wild species Ananas ananassoides originating from the floor of covered moist forest, and the primitive cultivar Panare of Ananas comosus mostly cultivated in semi-shaded palm swamps, and Pitcairnia pruinosa, a species abundant in highly sun exposed sites on rock outcrops. Ananas species are Crassulacean acid metabolism (CAM) plants, P. pruinosa is C3 plant. Plants were grown at low daily irradiance (LL = 1.3 mol m–2 d–1 corresponding to an incident irradiance of 30 mol m–2 s–1) and at high irradiance (HL = 14.7 mol m–2 d–1 or 340 mol m–2 s–1), and CO2 and H2O-vapour gas exchange and photochemical (qP) and non-photochemical quenching (qNP) of chlorophyll a fluorescence of photosystem 2 (PS2) were measured after transfer to LL, medium irradiance (ML = 4.1 mol m–2 d–1 or 95 mol m–2 s–1) and HL. All plants showed flexible light-use, and qP was kept high under all conditions. LL-grown plants of Ananas showed particularly high rates of CAM-photosynthesis when transferred to HL and were not photoinhibited.  相似文献   

12.
A modified Rotating Biological Contactor (RBC) was used for the treatability studies of synthetic tapioca wastewaters. The RBC used was a four stage laboratory model and the discs were modified by attaching porous nechlon sheets to enhance biofilm area. Synthetic tapioca wastewaters were prepared with influent concentrations from 927 to 3600 mg/l of COD. Three hydraulic loads were used in the range of 0.03 to 0.09 m3·m–2·d–1 and the organic loads used were in the range of 28 to 306 g COD· m–2·d–1. The percentage COD removal were in the range from 97.4 to 68. RBC was operated at a rotating speed of 18 rpm which was found to be the optimal rotating speed. Biokinetic coefficients based on Kornegay and Hudson models were obtained using linear analysis. Also, a mathematical model was proposed using regression analysis.List of Symbols A m2 total surface area of discs - d m active depth of microbial film onany rotating disc - K s mg ·l–1 saturation constant - P mg·m–2·–1 area capacity - Q l·d–1 hydraulic flow rate - q m3·m–2·d–1 hydraulic loading rate - S 0 mg·l–1 influent substrate concentration - S e mg·l–1 effluent substrate concentration - w rpm rotational speed - V m3 volume of the reactor - X f mg·l–1 active biomass per unit volume ofattached growth - X s mg·l–1 active biomass per unit volume ofsuspended growth - X mg·l–1 active biomass per unit volume - Y s yield coefficient for attachedgrowth - Y A yield coefficient for suspendedgrowth - Y yield coefficient, mass of biomass/mass of substrate removed Greek Symbols hr mean hydraulic detention time - (max)A d–1 maximum specific growth rate forattached growth - (max)s d–1 maximum specific growth rate forsuspended growth - max d–1 maximum specific growth rate - d–1 specific growth rate - v mg·l–1·hr–1 maximum volumetric substrateutilization rate coefficient  相似文献   

13.
James  Charles M.  Rezeq  T. Abu 《Hydrobiologia》1989,186(1):423-430
Continuous production of the rotifer Brachionus plicatilis rotundiformis (S-type) in an intensive chemostat culture system has been investigated. The production dynamics of rotifers in relation to different flow rates and feed regimes show that the growth rate and production depends on the type of algal feed and flow rate utilized in the culture system. It was possible to achieve a mean production of up to 318.84 × 106 rotifers m–3 d–1 at a flow rate of 6 1 h–1 in 100 1 chemostats and up to 261.21 × 106 rotifers m–3 d–1 at a flow rate of 40 1 h –1 while using 1 m3 capacity rotifer chemostats as production units. The 3 fatty acid composition of rotifers while using Chlorella and Nannochloropsis in the culture system has been described. The results of this investigation show that the rotifer productivity in the continuous culture system is considerably higher than in any of the conventional culture systems described to date for aquacultural purposes.This research was financed by the Kuwait Foundation for the Advancement of Sciences (KFAS), Kuwait, under a contract research project code 86-04-02.  相似文献   

14.
Summary NaCl was added to the nutrient solution of 4–6-week old Mesembryanthemum crystallinum plants so that the concentration rose by 50 mM per day. Ten to fifteen days after a concentration of 400 mM was reached, pronounced diurnal oscillations of malate levels indicated that plants had changed from C3-photosynthesis to crassulacean acid metabolism (CAM). Due to the NaCl-treatment the solute potential (s) decreased from about -6 bar to -25 bar, and the water potential () changed from about -5 bar to -23 bar on average. showed small diurnal oscillations both in controls and NaCl-treated plants, with an amplitude of 1 to 3 bar, the value at the end of the dark phase being less negative than that at the end of the light phase. Changes of ion levels due to the NaCl-treatment were average increases in Na+ and Cl- from 10–20 to 370–470 mmol kg-1 FW and from below 10 to 280–325 mmol kg-1 FW, respectively, and a decrease in K+ from 70–80 to 25 mmol kg-1 FW. These changes of ion levels corresponded very closely to an increase of dry weight in per cent of fresh weight observed during the NaCl-treatment (e.g. a change of 2% in one experiment), and osmotically they matched the measured change in s (e.g. about 18–20 bar in one experiment).Most of the organic solutes analysed did not show any significant changes as a result of the NaCl-treatment. The following compounds were identified within the respective ranges of concentrations: mannitol (0.2 to 0.5 mmol kg-1 FW), sum of quaternary ammonium compounds (60 to 140 mg kg-1 FW), choline (0.1 to 0.4 mmol kg-1 FW), betaine (0.3 to 0.7 mmol kg-1 FW), hexoses (2–9 mmol kg-1 FW), pentoses (1–5 mmol kg-1 FW) and sucrose (2–4 mmol kg-1 FW). The levels of proline and of total amino acids minus proline rose during the NaCl-treatment from 0.1–1 mmol kg-1 FW to 2.5–5 mmol kg-1 FW and from 2.5–4 mmol kg-1 FW to 6–8 mmol kg-1 FW, respectively.The changes of s and , and of Na+- and Cl--levels were complete, and new steady levels were attained by the time 400 mM NaCl was reached in the nutrient solution, i.e. many days before pronounced diurnal malate oscillations indicated that the change from C3-photosynthesis to CAM had occurred. The attainment of new steady levels of proline and K+, however, was much slower and coincided with the onset of CAM.Dedicated to Professor Dr.Dr.h.c. Michael Evenari on the occasion of his 75th birthday in appreciation of his great achievements in promoting experimental ecology and with gratitude for encouragements and help with the investigation of the adaptation of Mesembryanthemum crystallinum  相似文献   

15.
The chemical and biological conditions, and the bacteria-heterotrophic nanoflagellate (HNF) relationship were investigated in the vicinity of Funka Bay, southwest of Hokkaido, Japan during early spring 1999. At the time of sampling, chlorophyll a concentration, bacteria, phycoerythrin rich-cyanobacteria, and HNF abundance were in the following ranges: 0.3–3.6 g l–1, 2.5–5.6 × 105 cells ml–1, 0.6–1.2 × 103 cells ml–1, and 2.2–4.2 × 103 cells ml–1, respectively. Dissolved inorganic nitrogen, phosphate and silicate concentrations were in the ranges: 8.7–12.2 M, 0.9–2.0 M, and 21.6–25.5 M, respectively. Primary production ranged from 6.4 to 76.3 mg C m–3 d–1. Using water samples from regions of different productivity levels (in and outside bay), the bacteria - HNF relationship was uncoupled experimentally by the size-fractionation technique. Higher primary production (19.9 mg C m–3 d–1) in the bay supported higher bacterial growth rate (0.029 h–1). However, outside the bay both primary production (6.4 mg C m–3 d–1) and bacterial growth rate (0.007 h–1) were lower. The HNF growth rates and grazing rates were similar for both but by comparing both HNF grazing capacity and bacterial production, there was net decrease in bacterial abundance outside the bay and net increase inside the bay. The microbial parameters (rates and abundance) and the amount of carbon flow estimated through the phytoplankton – dissolved organic matter (DOM) – bacteria loop were different between the coastal station and the open ocean station. However HNF grazing and growth rates was similar for both stations.  相似文献   

16.
The surface elevation of Mono Lake, California, rose 2 m and mixed-layer salinities declined about 5 g kg–1 during the 3 years (1995–1997) following the decision to restrict water diversions out of the Mono Basin. Abundant (18000 m–2) Hexarthra jenkinae de Beauchamp were noted in pelagic samples in October 1997 after three decades of absence or very low abundance. Abundance subsequently increased to 100000 m–2 in December 1997 before declining to low numbers through 1998 and 1999. The re-appearance of Branchionus plicatilis Müller in pelagic samples occurred in September 1998. B. plicatilis areal abundance increased to 15000 m–2 in October–December of both 1998 and 1999 but was low throughout the rest of the year. Both rotifers were noted in nearshore ponds, but were only abundant in those with salinities below 53 g kg–1. During 1998–1999 when the salinities of the upper water column were 73–75 g kg–1, less saline shoreline habitats may have been seeding the offshore rotifer populations.  相似文献   

17.
The potential of three estuarine macroalgae (Ulvarotundata, Enteromorpa intestinalis andGracilaria gracilis) as biofilters for phosphate ineffluents of a sea bass (Dicentrarchus labrax) cultivationtank was studied. These seaweeds thrive in Cádiz Bay and were alsoselected because of their economic potential, so that environmental andeconomicadvantages may be achieved by future integrated aquaculture practices in thelocal fish farms. The study was designed to investigate the functioning of Pnutrition of the selected species. Maximum velocity of phosphate uptake (2.86mol PO4 g–1 dry wth–1) was found in U. rotundata.This species also showed the highest affinity for this nutrient. At low flowrates (< 2 volumes d–1), the three species efficientlyfiltered the phosphate dissolved in the waste water, with a minimum efficiencyof 60.7% in U. rotundata. Net phosphate uptake rate wassignificantly affected by the water flow, being greatest at the highest rateassayed (2 volumes d–1). The marked decrease in tissue P shownby the three species during a flow-through experiment suggested that growth wasP limited. However, due to the increase in biomass, total P biomass increasedinthe cultures. A significant correlation was found between growth rates and thenet P biomass gained in the cultures. A three-stage design under low water flow(0.5 volumes d–1) showed that the highest growth rates (up to0.14 d–1) and integrated phosphate uptake rates(up to 5.8 mol PO4 3– g–1dry wt d–1) were found in E.intestinalis in the first stage, with decreasing rates in thefollowing ones. As a result, phosphate become limiting and low increments oreven losses of total P biomass in these stages were found suggesting thatphosphate was excreted from the algae. The results show the potential abilityofthe three species to reduce substantially, at low water flow, the phosphateconcentration in waste waters from a D. labrax cultivationtank, and thus the quality of effluents from intensive aquaculture practices.  相似文献   

18.
Eurytemora affinis, a calanoid copepod, has been encountered in Volkerak-Zoommeer (Rhine delta region, S.W. Netherlands) both before this lake system was isolated in 1987 from the estuarine influence, and after. It was the main particle-feeding crustacean at all the 3 sampling stations in March–April 1990 when it reached densities of up to 215 ind.l–1. Its decline from mid April onwards, and low densities through summer, coincided with increase in cladocerans, especiallyDaphnia spp. (D. pulex andD. galeata), a decrease in seston (<33m) and chlorophyll concentrations and in primary production rates. The clearance rates (CR) ofEurytemora measured in the spring period varied enormously (0.6–24 ml.ind–1.d–1) depending mainly on size (0.44–1.06 mm), food concentration (0.8–2.2 mg C.l–1), and the water temperature which varied only narrowly (8.0–9.0°C). Mean ingestion rates of the animals measuring 0.68±0.02 mm during the study was 6.7±3.2 gC.ind–1.d–1; and assimilation efficiency varied between 27 and 53% (mean: 41±9%). The weight specific CR (SCR) varied between 0.96 and 6.4 litre.mg–1 body C.d–1. Pooled regression of SCR on the animal's body weight at the 3 study stations revealed a significant inverse relationship. Also daily ration and specific assimilation ofE. affinis varied greatly and inversely with the body weight. This calanoid contributed from about 50 to 100% to the zooplankton community grazing rates and assimilation rates, the former often exceeding the phytoplankton primary production.  相似文献   

19.
Summary The digestion and metabolism ofEucalyptus radiata foliage was studied in a small (1–1.5 kg) arboreal marsupial, the greater glider (Petauroides volans). Mean dry matter intake was 44 g·kg–0.75·d–1 and mean cell wall digestibility was 34%; these values fall within the range of other marsupials fedEucalyptus foliage. Digestible energy content ofE. radiata was high compared to other eucalypts because of the high content and digestibility of essential oils. However, excretion of essential oils and their metabolites in the urine meant that greater gliders retained only 55% of their digestible energy intakes (0.61 MJ · kg–0.75· d–1) as metabolizable energy (ME). Low ME intakes were not offset by low standard metabolic rates (2.39 W · kg–0.75), but the efficiency with which ME substituted for tissue energy was high (94%), so that greater gliders were able to maintain energy balance and body mass onE. radiata foliage.Abbreviations ME metabolizable energy - DE digestible energy - RQ respiratory quotient - FHP fasting heat production  相似文献   

20.
Behavioral and physiological responses to hypoxia were examined in three sympatric species of sharks: bonnethead shark Sphyrna tiburo, blacknose shark, Carcharhinus acronotus, and Florida smoothhound shark, Mustelus norrisi, using closed system respirometry. Sharks were exposed to normoxic and three levels of hypoxic conditions. Under normoxic conditions (5.5–6.4mg l–1), shark routine swimming speed averaged 25.5 and 31.0cm s–1 for obligate ram-ventilating S. tiburo and C. acronotus respectively, and 25.0cm s–1 for buccal-ventilating M. norrisi. Routine oxygen consumption averaged about 234.6 mg O2kg–1h–1 for S. tiburo, 437.2mg O2kg–1h–1 for C. acronotus, and 161.4mg O2 kg–1 h–1 for M. norrisi. For ram-ventilating sharks, mouth gape averaged 1.0cm whereas M. norrisi gillbeats averaged 56.0 beats min–1. Swimming speeds, mouth gape, and oxygen consumption rate of S. tiburo and C. acronotus increased to a maximum of 37–39cm s–1, 2.5–3.0cm and 496 and 599mg O2 kg–1 h–1 under hypoxic conditions (2.5–3.4mg l–1), respectively. M. norrisi decreased swimming speeds to 16cm s–1 and oxygen consumption rate remained similar. Results support the hypothesis that obligate ram-ventilating sharks respond to hypoxia by increasing swimming speed and mouth gape while buccal-ventilating smoothhound sharks reduce activity.  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