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.
Temperature responses of carbon assimilation processes were studied in four dominant species from mountain grassland ecosystem, i.e. Holcus mollis (L.), Hypericum maculatum (Cr.), Festuca rubra (L.), and Nardus stricta (L.), using the gas exchange technique. Leaf temperature (T L) of all species was adjusted within the range 13–30 °C using the Peltier thermoelectric cooler. The temperature responses of metabolic processes were subsequently modelled using the Arrhenius exponential function involving the temperature coefficient Q 10. The expected increase of global temperature led to a significant increase of dark respiration rate (R D; Q 10 = 2.0±0.5), maximum carboxylation rate (V Cmax; Q 10 = 2.2±0.6), and maximum electron transport rate (J max; Q 10 = 1.6±0.4) in dominant species of mountain grassland ecosystems. Contrariwise, the ratio between J max and V Cmax linearly decreased with T L [y = −0.884 T L + 5.24; r 2 = 0.78]. Hence temperature did not control the ratio between intercellular and ambient CO2 concentration, apparent quantum efficiency, and photon-saturated CO2 assimilation rate (P max). P max primarily correlated with maximum stomatal conductance irrespective of T L. Water use efficiency tended to decrease with T L [y = −0.21 T L + 8.1; r 2 = 0.87].  相似文献   

2.
The vapor pressure deficit (D) of the atmosphere can negatively affect plant growth as plants reduce stomatal conductance to water vapor (gwv) in response to increasing D, limiting the ability of plants to assimilate carbon. The sensitivity of gwv to changes in D varies among species and has been correlated with the hydraulic conductance of leaves (Kleaf), but the hydraulic conductance of other tissues has also been implicated in plant responses to changing D. Among the 19 grass species, we found that Kleaf was correlated with the hydraulic conductance of large longitudinal veins (Klv, r2 = 0.81), but was not related to Kroot (r2 = 0.01). Stomatal sensitivity to D was correlated with Kleaf relative to total leaf area (r2 = 0.50), and did not differ between C3 and C4 species. Transpiration (E) increased in response to D, but 8 of the 19 plants showed a decline in E at high D, indicative of an ‘apparent feedforward’ response. For these individuals, E began to decline at lower values of D in plants with low Kroot (r2 = 0.72). These results show the significance of both leaf and root hydraulic conductance as drivers of plant responses to evaporative demand.  相似文献   

3.
Size-dependent variations in morphological and physiological variables adduced to influence growth rate (e.g., cell surface area and volume, chlorophyll a concentration per cell) were determined by reevaluating published data from unicellular and multicellular plants and animals. With respect to cell volume, reduced major axis regression of the available data indicated that cell surface area decreases roughly as the 0.69-power, the concentration of Chl a decreased roughly as the 0.80-power, and cell mass decreased as the 0.77-power. Computer simulations indicated that the scaling exponent for cell surface area was the consequence of size-dependent variations in cell geometry and aspect ratio (i.e., cell length/width) rather than the result of geometric similitude among cells differing in size. The anisometric relation between cell mass and volume indicated that bulk cell density declines with increasing cell volume. Reanalyses of published data showed that growth rate and weight-specific growth rate scale as the ¾- and negative ¼-power, respectively, with respect to the body mass of unicellular and multicellular plants and animals. It is speculated that the anisometric relation between the growth rate and mass of unicellular plants is attributable to a “dilution” of metabolically active cellular constituents with increasing cell size in combination with the scaling of surface area with respect to volume (and therefore cell mass). It is further speculated that similar biological scalings may account for the ¾-power rule obtained for taxonomically and ecologically diverse multicellular plants and animals.  相似文献   

4.
We postulated that dietary ingestion of vitamin D may be used by some Alaskan Arctic marine mammal species in addition to, or instead of, cutaneous production to meet nutritional requirements. Zooplankton (n=5) sampled near Kaktovik, Alaska, contained no measurable vitamin D2 or D3, but did contain provitamin D (7‐dehydrocholesterol), the cutaneous precursor for previtamin D3 in mammals. Fillets and livers from five fish species were sampled near Barrow, Alaska, and evaluated for vitamin D3 content (no vitamin D2 was detected). Differences in vitamin D3 content appeared significant (P≤0.10) among fish livers (Kruskal‐Wallis [H test]=8.25, df=4, P=0.08) and among fish fillets (H=7.80, df=4, P=0.01). We also found significant differences in several pairwise comparisons (Mann‐Whitney U‐test) of vitamin D3 levels in fillets and livers. Blubber from six species of marine mammals had no detectable vitamin D2. The H test results for blubber vitamin D3 concentration were highly significant: 28.12, df=5, P<0.001. There were also significant differences in vitamin D3 content from blubber in pairwise comparisons of primarily invertebrate feeders (bowhead whale (Balaena mysticetus) [mean=4.20 SD±1.10 ng/g], and Pacific walrus (Odobenus rosmarus divergens) [5.43±2.82 ng/g]) vs. primarily piscivorous feeders (ringed seal (Phoca hispida) [746.57±493.00 ng/g] and beluga whale (Delphinapterus leucas) [426.00±174.92 ng/g]) and a semiaquatic terrestrial carnivore (polar bear (Ursus maritimus) [406.17±311.70 ng/g]). The bearded seal (Erignathus barbatus) had intermediate blubber vitamin D3 concentration (156.83±139.25 ng/g), which may reflect an intermediate‐type feeding strategy or an artifact of the small sample size. Zoo Biol 23:33–43, 2004. © 2004 Wiley‐Liss, Inc.  相似文献   

5.
To examine trophic dynamics over different size classes, an isotopic study of sailfish Istiophorus platypterus life‐history stages was carried out. Samples were collected from eastern Taiwan and the South China Sea during April 2009 and February 2012. A total of 263 samples (111–245 cm, lower jaw fork length, LLJFL) were examined for changes in trophic structure in relation to LLJFL by using stable isotope analysis of carbon (δ13C) and nitrogen (δ15N). The δ15N values for I. platypterus ranged from 7·51 to 14·19‰ (mean ± s.d . = 12·06 ± 1·16‰) and the δ13C values ranged from ?22·04 to ?15·48‰ (mean ± s.d . = ?17·62 ± 1·10‰). The δ15N values were positively dependent on LLJFL (r2 = 0·377), whereas δ13C were negatively dependent on LLJFL (r2 = 0·063). There were significantly different seasonal changes in nitrogen and carbon isotopic concentration, but no significant differences in concentrations between eastern Taiwan and the South China Sea were reported. The trophic level (TL) of each LLJFL class was correlated, starting from 2·84 TL for size class I (LLJFL < 140 cm) and reaching 5·03 TL for size class VI (LLJFL > 221 cm). The mean ± s.d . TL was 4·43 ± 0·19 for all samples. The results reveal that I. platypterus occupies a wide range of trophic levels and different size classes occupy different trophic positions in the pelagic ecosystem.  相似文献   

6.
The allometry of reproductive biomass M was determined for one moss (Polytrichum commune), four pteridophytes (Psilophyton princeps, Psilotum nudum, Lycopodium lucidulum, and L. clavatum), six gymnosperms (Larix decidua, Pseudotsuga douglasii, Tsuga canadensis, Pinus rigida, Picea abies, and Taxus baccata), and two angiosperms (Crataegus sp. and Quercus bicolor). Biomass M was measured for individual and grouped reproductive organs and regressed as a function of stem diameter D measured proximal to M for representative reproductive stems from each species. Published data for Cooksonia pertoni were used to estimate sporangial M. The data from this vascular plant fossil were compared to those from other species. Least squares regression of the entire data set yielded M = 0.12D2,9 (r2 = 0.94, N = 215), indicating that M, on the average, compiled with M Dα3.0. The allometries of the moss, pteridophyte, gymnosperm, and angiosperm species, however, were M = 0.46D4.6 (r2 = 0.99, N = 41), M = 0.07D3.2 (r2 = 0.91, N = 65), M = 0.56D2.2 (r2 = 0.92, N =85), and M = 0.44D1.8 (r2 = 0.97, N = 21), respectively, indicating that the proportionality M Dα3.0. recedes with finer taxonomic resolution. The data for Cooksonia were found to comply with the allometry of Polytrichum when the regression curve of this moss was extrapolated into the size range of the fossil species. Analyses showed that intraspecific allometric scaling factors α were dependent upon the manner in which plant stems taper. Species or portions of branching systems with α >; 4.0 had essentially untapered stems (e.g., Polytrichum commune, Psilotum nudum, twigs of Larix decidua); species with α <; 2.2 had tapered stems resulting from secondary growth in most cases. The evolution of tapered primary stems and secondary growth was interpreted to alter reproductive allometry.  相似文献   

7.
8.
Objective: Obese patients without clinically apparent heart disease may have a high output state and elevated total and central blood volumes. Central circulatory congestion should result in elevated pulmonary diffusing capacity (DLCO) and capillary blood volume (Vc) reflecting pulmonary capillary recruitment; however, the effect on membrane diffusion (Dm) is uncertain. We examined DLCO and its partition into Vc and Dm in 13 severely obese subjects (BMI = 51 ± 14 kg/m2) without manifest cardiopulmonary disease before and after surgically induced weight loss. Research Methods and Procedures: DLCO and its partition into Vc and Dm [referenced to alveolar volume (VA)] as described by Roughton and Forster, total body water by tritiated water, and fat distribution by waist‐to‐hip ratio were performed. Results: Despite normal DLCO (mean 98 ± 16% predicted), Vc/VA was increased (mean 118 ± 30% predicted), and Dm/VA was reduced (mean 77 ± 34% predicted). Nine of 13 subjects were restudied after weight loss (mean 52 ± 43 kg); Vc/VA decreased to 89 ± 18% predicted (p = 0.01), and Dm/VA increased to 139 ± 30% predicted (p < 0.01). Increasing total body water was associated with both increasing Vc (r = 0.74, p = 0.01) and increasing waist‐to‐hip ratio (r = 0.65, p = 0.02), indicating that circulatory congestion increases with increasing central obesity. Discussion: Severely obese subjects without manifest cardiopulmonary disease may have increased Vc indicating central circulatory congestion and reduced Dm suggesting associated alveolar capillary leak, despite normal DLCO. Reversibility with weight loss is in accord with reversibility of the hemodynamic abnormalities of obesity.  相似文献   

9.
Age and growth of Pinna bicolor were examined in the seagrass beds of Merambong shoal (N 1°19′55.62″; E 103°35′57.75″) off the south‐western coast of Johor, Peninsular Malaysia between May 2006 and April 2007. Monthly growth increment data of P. bicolor were analyzed using FiSAT software (FAO‐ICLARM Stock Assessment Tools) to estimate the asymptotic length (L) and growth coefficient (K). Average growth rate of P. bicolor was 1.42 (±0.01) cm per month; the estimated asymptotic length (L) and growth coefficient (K) were 34.66 cm and 0.88 per year, respectively. In their natural habitat, P. bicolor attain shell heights of approximately 17, 25 and 30 cm at the end of their first, second and third years of growth. The length–weight relationship was estimated as Log W = ?5.397 + 3.111Log L, and in exponential form the equation was W = 0.000004L3.111 (r2 = 0.99, P < 0.01). Habitat temperature and salinity ranged between 27.47 and 29.66°C and 28.66–33.00 ppt with a mean of 29.10 (±0.66) m°C and 30.52 (±1.41) ppt, respectively.  相似文献   

10.
The spotback skate Atlantoraja castelnaui (Arhynchobatidae) is a large and threatened skate species subjected to fishing pressure, endemic to the Southwest Atlantic that occurs from Rio de Janeiro, Brazil, to San Jorge Gulf, Argentina. The age, growth, age at maturity and the maximum intrinsic rate of population increase rmax of A. castelnaui were studied using 152 specimens collected from off Uruguay and north Argentina (35°–42° S), between June 2013 and February 2020. Vertebrae from 143 individuals were used for ageing (females: n = 83, size range 404–1300 mm total length, TL; males: n = 60, size range 400–1270 mm TL). Maximum ages determined for females and males were 30 and 28 years, respectively. To fit growth models, non-linear and Bayesian estimation approaches were considered. For the first approach, a set of four candidate growth (size-at-age) models were fitted: three-parameter von Bertalanffy, two-parameter von Bertalanffy with fixed L0, Gompertz and Logistic. In the second approach, von Bertalanffy, Gompertz and Logistic were fitted. For non-linear estimation, model selection indicated that the entire set of candidate growth models were supported by the data. The von Bertalanffy was selected as the best model for Bayesian estimation. There were no differences in growth between sexes. For the sexes combined, the von Bertalanffy growth model by Bayesian method was considered the most adequate to describe the growth of A. castelnaui (growth mean parameters ± S.D. : L = 1210.29 ± 40.68 mm; k = 0.12 ± 0.01 years−1; L0 = 179.20 ± 11.62 mm). The age at maturity was estimated at 16.21 and 14.04 years for females and males, respectively. The maximum intrinsic rate of population increase rmax was estimated as 0.252 years−1. Life-history traits and rmax provided in the present study suggest that this species has a relatively low productivity and may be vulnerable to an intense fishing pressure.  相似文献   

11.
The size, growth and reproductive biology of the giant mudskipper, Periophthalmodon schlosseri, was studied in the inter‐tidal mudflat of a mangrove swamp along a 0.4–0.6 × 5 km stretch of the Sepang coastal waters, Selangor Malaysia. The size ranges of the giant mudskipper specimens caught in the study area were between 10–27.5 cm TL (20.6 ± 3.1 cm) for males and 14.5–28.5 cm TL (21.9 ± 2.3 cm) for females. Male to female sex ratio during the study was 1 : 1.1. The estimated von Bertalanffy growth parameters were L = 29 cm and K = 1.4 year?1, with a growth performance index of Ø′ = 3.1. Otolith microstructure analysis revealed a strong correlation (Female: r = 0.920, n = 35; Male: r = 0.943, n = 41, P < 0.05) between the number of daily rings (number of days old) over the respective size ranges of male (8.5–19.5 cm SL) and female fish (11.8–22.3 cm SL). Four maturity stages were described to follow gonadal development based on external features; these stages were further validated through histological examination of the ovary and oocyte diameter measurements. The gonadal development cycle of P. schlosseri was asynchronous, as marked by the random mixture of early stages oocytes (diameter: 70.7 ± 14.9 μm) to matured stages (356.9 ± 20.2 μm), and had a prolonged spawning season extending from June to October. Monthly fluctuation of the gonadosomatic and hepatosomatic indices for this species was highly correlated (r = 0.126, P < 0.05).  相似文献   

12.
A sample of 346 Uranoscopus scaber (L., 1758) was collected from the southeastern Black Sea between January 2002 and May 2005 in order to provide information on age, growth, length–weight relationship, and stomach contents of this species. Total length and total weight of sampled fish ranged from 5.2 to 21.9 cm and from 2.0 to 182.5 g, respectively. The sex ratio (1 : 1.98) was biased toward females (P < 0.05). Isometric growth was determined in the population. Length–weight relationships for all individuals were described by the parameters: a = 0.0167, b = 3.00, with the r2 = 0.99. The population was composed of five age‐classes (I–V years). The von Bertalanffy growth parameters (±SE) and confidence limit (CL, 95%) for the entire population were: L = 26.31 ± 0.838 cm (CL: 24.66–27.96), k = 0.339 ± 0.023 1/year (CL: 0.294–0.385), and to = 0.087 ± 0.0346 year (CL: 0.019–0.155) (r2 = 0.96). Growth performance index (Φ) was 2.37. The diet was composed of Gobius sp. (2.2%),Trachurus trachurus (2.1%), Merlangius merlangus euxinus (4.3%) and unidentified fish species (17.4%); 39.1% were unidentified remains.  相似文献   

13.
Dynamic and static intensity light scattering techniques were employed to identify conditions allowing preparation of homogeneous solutions of distinct oligomeric states of RecA protein. These hydrodynamically distinguishable oligomer populations of RecA protein were obtained in homogeneous pure quantities sufficient for physical studies. Results indicate two fairly narrow distributions of RecA oligomers comprised on average of 42 ± 3 and 18 ± 1 RecA monomers. These structures, denoted RecA42 and RecA18, respectively, could be obtained reproducibly in milligram quantities and were stable for at least one week. This enabled reliable characterizations of their hydrodynamic properties by dynamic and total intensity light scattering. These measurements revealed RecA42 had an average translational diffusion coefficient, D20(L) = 8 ± 2 × 10−8 cm2/s, molecular weight, Mr = 1.6 ± 0.1 × 106, and radius of gyration, RG = 465 ± 29 Å. The smaller aggregate, RecA18, had D20(S) = 20.5 ± 2.5 × 10−8 cm2/s. Mr = 7.0 ± 0.4 × 105, and RG = 300 ± 20 Å. Heating RecA18 at 37°C overnight resulted in conversion to a species with hydrodynamic properties indistinguishable from RecA42, called RecA18/42. Conversion of RecA42 to RecA18 occurred almost instantaneously by 50% dilution at 38°C or very slowly with incubation at 4°C for at least 39 days. Self-association reactions of the three starting oligomeric states (RecA18, RecA42, and RecA18/42) induced by MgCl2 were monitored at several temperatures by dynamic light scattering. Results of these experiments provided evaluations of kinetic activation parameters of the self-association reactions. The activation parameters found for each starting oligomeric state of the protein were significantly different, revealing the variable influence of MgCl2 on the activation barriers to RecA self-association. Highly aggregated equilibrium solutions that ultimately form in solutions of each starting oligomeric species, incubated in MgCl2 at 38°C for four days, were characterized by total intensity light scattering. Interpretations of these data in terms of characteristic behavior of random polymers suggests the surface morphologies of these highly associated equilibrium states formed from RecA42 and RecA18/42 are similar but contrast with that of RecA18. Calculated values of the translational diffusion coefficient D0 were obtained for oligomeric structures modeled as helical arrays of connected monomer spheres. Best agreement with experimentally determined diffusion coefficients required that constituent monomer spheres of RecA42 have radii 33–40% larger than the monomer spheres of RecA18. Results suggest the hydrodynamically distinct oligomeric forms of RecA may reside in conformational states with different surface exposure of hydrophobic residues, which results in substantial differences in local solvation/hydration. © 1996 John Wiley & Sons, Inc.  相似文献   

14.
Regression curves for the relation between the critical buckling height Hcrit, and the diameter D of columnar support members composed exclusively of different tissues were established based on Greenhill's formula and previously reported mean values for the density-specific stiffness and density-specific strength of parenchyma, primary xylem, sclerenchyma, and wood. These regression curves were used to determine the extent to which the actual heights H of 249 plant species approach or transgress the Hcrit for stems relying principally upon different tissue-types for stiffness. Based on empirically determined H and estimated Hcrit, the safety-factor Hcrit/H (computed on the basis of E/p) against elastic instability resulting from self-loading imposed on stems was determined for dicot and gymnosperm tree species (N = 56), mosses (N = 40), pteridophytes (N = 16), dicot herbs (N = 120), and palms (N = 17). With the exception of tree species, Hcrit/H was size-dependent, decreasing with increasing D. This was a consequence of the scaling exponents (i.e., the slopes of the regression curves) for tree Hcrit, vs. D and H vs. D which were nearly identical, whereas the scaling exponents for H vs. D for “nonwoody” species were in excess of those for Hcrit, vs. D. With the exception of a few very tall specimens of palm species, however, the majority of nonwoody and woody species did not exceed their estimated Hcrit. The upper size-range obtained by the procession of taller plant grades and clades was bounded by the regression curves of Hcrit, vs. D established for progressively stiffer plant tissues: parenchyma Å primary xylem Å sclerenchyma Å wood. This appears to be a consequence of the incorporation of progressively stiffer tissues within the stems of taller nonwoody species and the adjustment in the girth of stems, which developmentally occurs for trees.  相似文献   

15.
An Exact theoretical expression for the apparent diffusion coefficient Dapp(K) of a thin rigid rod with arbitrary anisotropy of its translational diffusion diffusion coefficient is derived from the first cumulant of its dynamic structure factor. Dapp(K) is predicted to reach a limiting plateau value at extermely large values of KL, where K is the scattering vector and L the rod length. Howerver, that limiting plateau value is approached only very slowly along a quasi-plateau with a very gradual slope. Dynamic light-scattering studies have been performed on tobacco mosaic virus from K2 = (0.4–20) × 1010 cm?2 using 632-8-nm laser radiation. The present data yield D0 = (4.19 ± 0.10) × 10?8 cm2/s (corrected to 20,w conditions) and, with literature data to establish L = 2980 Å and the rotational diffusion coefficient DR = 318s?1, yield also Δ ≡ D ? D = (1.79 ± 0.38) × 10?8 cm2/s. The experimental data closely follow the curve of Dapp(K) vs K2 calcuated for these parameters. The present value of D0 substantially exceeds all previous dynamic light-scattering values, but is in good aggreement with previous sedimentation data, which were confirmed for the presemt sample. The anisotropy ratio Δ/D0 = 0.43 ± 0.09 is in accord with theoretical predictions based on the modified Kirkwood algorithm, despite the fact the D0 lies significantly below its corresponding theoretical value. The present data largely predlude the possibility that both D0 and Δ/D0 could simultaneously match their theoretical predictions. We present a detailed comparison of the experimental data with the calculations of Tirado and Garcia de la Torre based on the modified Kirkwood algorithm and with the Broersma formulas.  相似文献   

16.
T. Raj  W. H. Flygare 《Biopolymers》1977,16(3):545-549
The translational diffusion coefficient of a pure sample of α-chymotrypsinogen A is measured by laser light scattering to give a value of D20,w0 = (8.40 ± 0.15) × 10?7 cm2/sec.  相似文献   

17.
Prolonged swimming performances of two as yet unnamed species of three‐spined stickleback, Gasterosteus spp., were compared. The two fishes (not yet formally described, referred to here as benthic and limnetic) inhabit different niches within Paxton Lake, Texada Island, British Columbia, Canada, and are recent, morphologically distinct species. Limnetics had longer endurance during prolonged swimming than did benthics. The mean regression of the log10 of fatigue time (Ft, s) on swimming speed (U, standard length, LS s?1) for limnetics (log10Ft = 7·03 ? 0·46U) had a similar slope, but a significantly higher intercept than that for benthics (log10Ft = 5·55 ? 0·43U). Adult benthics were larger, heavier and deeper‐bodied fish than limnetics. Limnetics, however, had a significantly greater pectoral fin edge:base ratio (mean ± s .e .: limnetics, 4·58 ± 0·43; benthics, 3·63 ± 0·27). In addition, limnetics had significantly lower drag coefficients (CD) than benthics (limnetics, log10CD = ?0·49log10Re + 0·66; benthics, log10CD = ?0·26log10Re ? 0·30) where Re is the Reynolds number [(LSU?1), where U and ν are swimming velocity and the kinematic viscosity of the water, respectively]. Compared to their ancestral form, the anadromous three‐spined stickleback Gasterosteus aculeatus L., limnetics and benthics had significantly longer and shorter endurance times, respectively. In addition, both these fishes had significantly higher fast‐start velocities than their ancestral form. Selection due to differential resource use may have lead to divergence of body form, and, therefore, of steady swimming performance. Therefore predation may be the selective force for the similar high escape performance in these two fishes.  相似文献   

18.
D-Galacturonic acid reductase, a key enzyme in ascorbate biosynthesis, was purified to homogeneity from Euglena gracilis. The enzyme was a monomer with a molecular mass of 38–39 kDa, as judged by SDS–PAGE and gel filtration. Apparently it utilized NADPH with a Km value of 62.5±4.5 μM and uronic acids, such as D-galacturonic acid (Km=3.79±0.5 mM) and D-glucuronic acid (Km=4.67±0.6 mM). It failed to catalyze the reverse reaction with L-galactonic acid and NADP+. The optimal pH for the reduction of D-galacturonic acid was 7.2. The enzyme was activated 45.6% by 0.1 mM H2O2, suggesting that enzyme activity is regulated by cellular redox status. No feedback regulation of the enzyme activity by L-galactono-1,4-lactone or ascorbate was observed. N-terminal amino acid sequence analysis revealed that the enzyme is closely related to the malate dehydrogenase families.  相似文献   

19.
Chlorinated phenols have been intensively investigated from an eco-toxicological point of view, however almost nothing is known about toxicity of tetrachlorophenol (TeCP) to higher terrestrial plants. This article applied the willow tree acute toxicity test to study the toxicity of 2,3,5,6-TeCP to willows Salix viminalis (S. viminalis) at neutral and acidic conditions (roughly pH 7 and 4) with inhibition of transpiration as toxic endpoint. At neutral pH the EC50 was >10 mg L?1 while the EC50 at acidic conditions was 0.32 ± 0.17 mg L?1, clearly indicating that toxicity is exerted by the non-ionic chemical fraction. Standard tests running at neutral pH are therefore not capturing the full toxicity of weak acids and bases.  相似文献   

20.
The present study attempts to assess the potential of artificial substrates to enhance fish production in inland saline groundwater ponds through periphyton production. Grey mullet, Mugil cephalus, was cultured for 100 days in ponds with substrate (treatment ponds) and without substrate (control ponds). To enhance the surface area, bamboo poles were used as substrate. The periphyton population, pigment concentration and hydrobiological characteristics of pond water were monitored. The studies revealed little difference in most of the water quality parameters observed in the two treatments. However, turbidity (27.0 ± 0.1–35.0 ± 0.1 Nephalo Turbidity Unit (NTU)), chlorophyll ‘a’ (6.6 ± 0.6–7.6 ± 0.6 μg L?1), plankton population (phytoplankton 8.4 × 103–9.4 ×103 numbers L?1; zooplankton 4.0 × 103–5.1 × 103 numbers L?1) and NH4–N (2.0 ± 0.2–2.3 ± 0.1 mg L?1) were high in the treatment with no additional substrate; however, in the treatment with substrate the total Kjeldahl nitrogen (9.8 ± 0.8–10.8 ± 0.7 mg L?1) and o‐PO4 (0.1 ± 0.01–0.1 mg L?1) remained significantly (P < 0.05) higher. Highest periphyton biomass in terms of dry matter (DM) (0.8 ± 0.01–1.4 ±0.01 mg cm?2), ash free DM (0.4 ± 0.0–0.6 ± 0.01 mg cm?2), chlorophyll ‘a’ (3.1 ± 0.2–8.1 ± 0.8 μg cm?2) and pheophytin ‘a’ (1.9 ± 0.4–3.9 ± 0.5 μg cm?2) was observed at 50 cm depth in ponds provided with additional substrate. Fifteen plankton genera showing periphytic affinity colonized the bamboo substrates. Fish growth (mean fish weight 524.3 ± 8.7 g and SGR 2.5 ± 0.1) was significantly (P < 0.05) higher in ponds provided with additional substrate compared with control ponds (387.2 ± 6.0). Length–weight relationship (LWR) (W = cLn) also showed that the exponential value (‘n’) of length was high in substrate‐supported ponds (n = 2.36) in comparison with controls (n = 1.09). These studies suggest that a periphyton‐supported aquaculture system can be used successfully for the culture of herbivorous brackishwater fish species like M. cephalus in inland saline groundwaters and thus could contribute to the development of sound and sustainable aquaculture technology.  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