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.
SUMMARY. Increase in body wet weight of Gammarus pulex fed on decaying elm leaves was followed to senescence and death. Growth in juveniles was approximately exponential; from birth to death it conformed to a logistic growth curve, with maximum absolute increments in weight about half-way through a life span of 350–450 days at 15°C. Some individuals lived longer, for up to 640–700 days. The instantaneous or specific growth rate was maximal near birth, at c. 5–6% wet wt day?1, and declined exponentially with increasing size and age. Over the range 4.7–14.8°C there was a log-log relationship between temperature and specific growth rate. Growth was maximal at 20°C in newborn animals and at 15°C in 6–9-mg animals. The specific growth rate of young individuals was fastest on decaying leaves of elm with a well developed flora of fungi and other microorganisms. Leached elm leaves without this flora supported growth at a lower rate. The latter diet was sufficient for survival and growth of newborn individuals; detritus, faeces or other food items were not needed. Isolated specimens grew as fast as those kept in groups. Growth was generally slower on leached leaves of oak and sycamore. In newborn animals fed on the fine roots of aquatic plants (Veronica, Rorippa and Glyceria), growth was as fast as on decaying elm leaves; growth on the green living leaves of the plants was slower, as on detritus from two streams and on a pure culture of an aquatic fungus. Consumption of leached elm leaves was related to leaf thickness. In a full gut the wet weight (1.34–1.37 mg) and volume (3.8–4.1 mm3) (for 20-mg animals) was independent of leaf thickness but dependent on animal size, increasing 4-fold over the range 2–50 mg body wt. Daily consumption (dry wt) was approximately equivalent to 50% body dry wt at 5 mg and 20% at 50 mg body wet wt. Individuals fed on thick leaves ingested 50% more dry weight per day and absorbed more in the gut than when fed on thin leaves, but the relative efficiency of absorption was the same at 36–59% for 10–20-mg animals. Weight-specific absorption in the gut was highest in juveniles and decreased with increasing body weight; relative efficiency of absorption was generally lower in the larger individuals. Assuming an energy value of 5 cal mg?1 dry wt for elm leaves, daily mean energy intake by absorption in thegutof G. pu/ex was2.2 cal mg?1 animaldry wt (9.2 J mg?1) in individuals of 0.4 mgdry wt (2 mg wet wt), decreasing to 0.3 cal mg?1 (1.3 J mg?1) at 10 mg dry wt (50 mg wet wt). Growth in Gammarus is briefly reviewed in the hght of work on other animals and it is emphasized that all aspects of feeding, growth and metabol-ism should be specifically related to size and age of the individuals, using well defined diets.  相似文献   

2.
The survival time of zooplankton under conditions of total starvation is expressed as a function of weight-specific respiration rate and the fraction of initial (pre-starvation) body weight which may be lost prior to death. Data from the literature on these two components of survival time are used to formulate a general expression of survivorship of zooplankton at 20°C as a function of body weight: t=2·95 w0·25, where t is in days and w is μg dry weight. Survival time data from the literature on 25 marine and freshwater species are compared to this prediction as are new data on Daphnia pulex Leydig, D. magna Straus and Simocephalus serrulatus (Koch). The effects of age-specific (non-starvation) mortality are considered; in particular, older individuals of each of these species survive for shorter periods of time than predicted and an interaction between age-specific and starvation-induced mortality is proposed. The effects of total and partial starvation on the size structure of zooplankton communities are discussed.  相似文献   

3.
The two mayflies (Insecta, Ephemeroptera) Baetis muticus and Baetis rhodani are absent from the acid water (pH 4.8–5.2) streams of the Upper Duddon drainage basin. Field observations and laboratory experiments were used in an attempt to explain this absence. Baetis muticus could not tolerate acid water; in addition it was deduced that available foods in the Upper Duddon streams would not sustain it there, even if the water could be tolerated. Baetis muticus feeds by browsing on decaying, allochthonous leaves, implying an input of micro-organisms, especially fungi, to the diet. Well-decayed, allochthonous leaves do not occur in the acid water, Upper Duddon streams. Baetis rhodani proved to be acid water tolerant, but only if the background waters had high ionic loadings. Since Upper Duddon waters have only low ionic loadings, Baetis rhodani is apparently excluded from them on this account. The evidence was that if the Upper Duddon waters were chemically suitable, then Baetis rhodani could survive in them, because a suitable food, the alga Hormidium subtile is present.  相似文献   

4.
Quantitative analyses have been made of the dietary cholesterol requirement for the growth of the larvae of Musca domestica. The larvae will not grow on diets to which no cholesterol is added, a few pupae and adults are obtained when the concentration of cholesterol is 0·05 μmol/g of diet, but the concentration has to be raised to 0·36 μmol/g of diet before the maximum numbers of pupae and adults are obtained. Further addition of cholesterol above 0·36 μmol/g diet did not have any significant effect on the weight and growth of the larvae. However, the ratios of the cholesterol to phospholipid fractions recovered from the larvae increased rapidly when the concentration of cholesterol in the diet was raised from 0·05 to 0·56 μmol/g of diet. Above this concentration only a slight increase in the ratios was observed. Larvae reared on diets containing 0·05 μmol cholesterol/g of diet contain only 25 per cent of the cholesterol content of the larvae reared on the diets containing more than 0·28 μmol of cholesterol/g of diet, the cholesterol content being expressed relative to the weight of the larvae,The absence of cholesterol synthesis has been demonstrated in the larvae by feeding [4-14C] cholesterol. The specific activity of the cholesterol recovered from the larvae is the same as that of cholesterol added to the diet. Irrespective of the cholesterol concentration of the larval diet, approximately 97 per cent of the radioactivity recovered from the larvae behaved as free cholesterol, less than 1 per cent as cholesterol esters and the rest as unidentified ‘polar sterols’. The results are compared with those from similar studies on other insects.  相似文献   

5.
The aim of the study is to investigate the influence of diet treatment on bone marrow cells. Normal male Wistar rats were divided into six groups (n?=?6 per group): control with normal diet (C), increased fructose (31 % w/w in fodder) (Fr) and high fatty (30 % w/w of animal fat in fodder) diet (Fa), and the same diets with vanadium complex ([VO(4,4′ Me2-2,2′ Bpy)2]SO4)?·?H2O (CV, FrV and FaV). During 5 weeks, the animals had unlimited access to food and water. Immediately after anaesthetizing and sacrificing the animals, bone marrow smears were prepared from the femurs. Different types of cell lines in the animal smears were examined under the microscope: erythroid line, myeloid line, monocytic line, megakariocytic line and lymphoid line. Addition of fructose or animal fat had evident influence on the proportional composition of the bone marrow cells. In erythroid precursors, addition of both investigated products resulted in a statistically significant increase of percentage of this type of cells. A reverse effect was observed for the lymphoid cell line where addition of both tested diets decreased quantity of these cells in comparison to the control diet. In the same lines, addition of vanadium intensified the observed changes. In the case of other types of cell lines, statistically significant changes were not observed.  相似文献   

6.
Factors determining the microflora of stored barley grain   总被引:3,自引:0,他引:3  
Colonisation of barley grain has been studied during storage at different water contents and with and without restriction of the air supply to simulate conditions in unsealed silos. Grain stored with a water activity >0·9 aw (20% water content) heated spontaneously when aeration was unrestricted, the maximum temperature attained increasing with aw to 58 °C at 1·0 aw (39% water content). The presence of many unripe grains increased the tendency to heat at a given mean water content. Although heating was prevented by sheeting to restrict the air supply, it could occur subsequently when the sheeting was removed. Both heating and restriction of the air supply were associated with increased carbon dioxide (to >25%) and decreased oxygen concentrations (to <5%). Germination of grain after 6·9 months storage was correlated with aw; germination levels approaching 100% were retained only at about 0·7 aw (13·5% water content). Colonisation by Aspergillus species was correlated with aw and temperature and similar correlations with Penicillium species were also found, with P. verrucosum var. cyclopium abundant at 0·85-0·90 aw (17·20% water content) and P. piceum, P. funiculosum and P. capsulatum at 0·90-0·95 aw (20–25% water content). In sealed containers P. roquefortii occurred at 1·00 aw (39% water content) and P. hordei at 0·90-0·92 aw (20–22% water content). Spores of fungi and actinomycetes formed during spontaneous heating of grain survived 6 months sealed storage although Absidia corymbifera and Micropolyspora faeni may have declined in numbers. Fungicides applied to the ripening grain had only limited effect on the colonisation of the grain during storage.  相似文献   

7.
The kinetic parameters of the inhibition of pigeon brain acetylchlolinesterase (AChE) by procaine hydrochloride were investigated. Procaine (0·083–1·67 mM) reversibly inhibited AChE activity (15–83 percent) in a concentration dependent manner, the IC50 being about 0·38 mM. The Michaelis-Menten constant (Km) for the hydrolysis of acetylthiocholine iodide was found to be 1·53 × 10?4 M and the Vmax was 1·06 μmol min?1 mg?1 protein. Dixon as well as Lineweaver-Burk plots and their secondary replots indicated that the nature of the inhibition is of the linear mixed type which is considered to be a mixture of partial competitive and pure non-competitive. The values of Ki(slope) and Ki (intercepts) were estimated as 0·14 mM and 0·22 mM respectively by the primary Dixon and by the secondary replots of the Lineweaver-Burk plot. The Ki′/Ki ratio shows that procaine has a greater affinity of binding for the peripheral than for the active site.  相似文献   

8.
The effects of supplementing diets with acetone extract (1% w/w) from four medicinal plants (Bermuda grass Cynodon dactylon, H1, beal Aegle marmelos, H2, winter cherry Withania somnifera, H3 and ginger Zingiber officinale, H4) on growth, the non‐specific immune response and ability to resist pathogen infection in tilapia Oreochromis mossambicus were assessed. In addition, the antimicrobial properties of the extract were assessed against Vibrio alginolyticus, Vibrioparahaemolyticus, Vibrio mimicus, Vibrio campbelli, Vibrio vulnificus, Vibrio harveyi and Photobacterium damselae. Oreochromis mossambicus were fed 5% of their body mass per day for 45 days, and those fed the experimental diets showed a greater increase in mass (111–139%) over the 45 days compared to those that received the control diet (98%). The specific growth rate of O. mossambicus fed the four diets was also significantly greater (1·66–1·93%) than control (1·52%) diet‐fed fish. The blood plasma chemistry analysis revealed that protein, albumin, globulin, cholesterol, glucose and triglyceride levels of experimental fish were significantly higher than that of control fish. Packed cell volume of the blood samples of experimental diet‐fed fish was also significantly higher (34·16–37·95%) than control fish (33·0%). Leucocrit value, phagocytic index and lysozyme activity were enhanced in fish fed the plant extract‐supplemented diets. The acetone extract of the plants inhibited growth of Vibrio spp. and P. damselae with extracts from W. somnifera showing maximum growth inhibition. A challenge test with V. vulnificus showed 100% mortality in O. mossambicus fed the control diet by day 15, whereas the fish fed the experimental diets registered only 63–80% mortality at the end of challenge experiment (30 days). The cumulative mortality index for the control group was 12 000, which was equated to 1·0% mortality, and accordingly, the lowest mortality of 0·35% was registered in H4‐diet‐fed group.  相似文献   

9.
Acclimation to rapidly fluctuating light, simulating shallow aquatic habitats, is altered depending on inorganic carbon (Ci) availability. Under steady light of 50 μmol photons·m?2·s?1, the growth rate of Synechococcus elongatus PCC7942 was similar in cells grown in high Ci (4 mM) and low Ci (0.02 mM), with induced carbon concentrating mechanisms compensating for low Ci. Growth under fluctuating light of a 1‐s period averaging 50 μmol photons·m?2·s?1 caused a drop in growth rate of 28%±6% in high Ci cells and 38%±8% in low Ci cells. In high Ci cells under fluctuating light, the PSI/PSII ratio increased, the PSII absorption cross‐section decreased, and the PSII turnover rate increased in a pattern similar to high‐light acclimation. In low Ci cells under fluctuating light, the PSI/PSII ratio decreased, the PSII absorption cross‐section decreased, and the PSII turnover remained slow. Electron transport rate was similar in high and low Ci cells but in both was lower under fluctuating than under steady light. After acclimation to a 1‐s period fluctuating light, electron transport rate decreased under steady or long‐period fluctuating light. We hypothesize that high Ci cells acclimated to exploit the bright phases of the fluctuating light, whereas low Ci cells enlarged their PSII pool to integrate the fluctuating light and dampen the variation of the electron flux into a rate‐restricted Ci pool. Light response curves measured under steady light, widely used to predict photosynthetic rates, do not properly predict photosynthetic rates achieved under fluctuating light, and exploitation of fluctuating light is altered by Ci status.  相似文献   

10.
Three photosynthetic parameters of 7 species of marine diatoms were studied using Na214CO3 at 5–8 C using log phase axenic cultures. The cell volumes of the different species varied from 70 μm3 to 40 × 105μm3. The present experiment is consistent with the interpretation that the initial slope α (mg C · [mg chl a]?1· h?1· w?1· m2) of photosynthesis vs. light curves is controlled by self-shading of chlorophyll a in the cell. Pm, the rate of photosynthesis at light saturation (mg C · [mg cell, C]?1· h?1) and R, the intercept at zero light intensity (mg C · [mg cell C]?1· H?1) are both dependent on the ratio of surface area to volume of cell.  相似文献   

11.
The desmid Staurastrum luetkemuellerii Donat et Ruttner and the cyanobacterium Microcystis aeruginosa Kütz. showed pronounced differences in chemical composition and ability to maintain P fluxes. The cellular P:C ratio (Qp) and the surplus P:C ratio (Qsp) were higher in M. aeruginosa, indicating a lower yield of biomass C per unit of P. The subsistence quota (Qp) was 1.85 μg P·mg C?1in S. luetkemuellerii and 6.09 μg P·mg C?1in M. aeruginosa, whereas the respective Qp of P saturnted organisms (Qs) were 43 and 63 μg P·mg C?1. These stores could support four divisions in S. luetkemuellerii and three divisions in M. aeruginosa, which suggests that the former exhibited highest storage capacity (Qs/Q0). M. aeruginosa showed a tenfold higher activity of alkaline phosphatase than S. luetkemuellerii when P starved. The optimum N:P ratio (by weight) was 5 in S. luetkemuellerii and 7 in M. aeruginosa. The initial uptake of Pi pulses in the organisms was not inhibited by rapid (<1 h) internal feedback mechanisms and the short term uptake rote could be expressed solely as a function of ambient Pi. The maximum cellular C-based uptake rate (Vm) in P starved M. aeruginosa was up to 50 times higher than that of S. luetkemuellerii. It decreased with increasing growth rate (P status) in the former species and remained fairly constant in the latter. The corresponding cellular P-based value (Um= Vm/Qp) decreased with growth rate in both species and was about 10 times higher in P started M. aeruginosa than in S. luetkemuellerii. The average half saturation constant for uptake (Km) was equal for both species (22 μg P·L?1) and varied with the P status. S. luetkemuellerii exhibited shifts in the uptake rate of Pi that were characterized by increased affinity (Um/Km) at low Pi, concentrations (<4 μg P·L?1) compared to that at higher concentrations. The species thus was well adapted to uptake at low ambient Pi, but M. aeruginosa was superior in Pi uptake under steady state and transient conditions when the growth rate was lower than 0.75 d?1. Moreover, M. aeruginosa was favored by pulsed addition of Pi. M. aeruginosa relpased Pi at a higher rate than S. luetkemuellerii. Leakage of Pi from the cells caused C-shaped μ vs. Pi curves. Therefore, no unique Ks for growth could be estimated. The maximum growth rate (μm) (23° C) was 0.94 d?1for S. luetkemuellerii and 0.81 d?1for M. aeruginosa. The steady state concentration of Pi (P*) was lower in M. aeruginosa than in S. luetkemuellerii at medium growth rates. The concentration of Pi at which the uptake and release of Pi was equal (Pc was, however, lower in S. luetkemuellerii.  相似文献   

12.
NADP-dependent non-phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.9), previously described in higher plants, has been now found to be present in eukaryotic green algae, but in neither cyanobacteria nor non-photosynthetic microorganisms. The enzyme from the unicellular green alga Chlamydomonas reinhardtii, strain 6145c, has been purified to apparent electrophoretic homogeneity. The non-phosphorylating enzyme was effectively separated from the NADP-dependent phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.13) dye-ligand chromatography on Reactive Red-120 agarose. The purified enzyme exhibited an optimum pH in the 8.5–9.0 range and a specific activity of approx. 8 μmol·(mg protein)−1·min−1. The native protein was characterized as a homotetramer with a molecular weight of 190 000, a Stokes radius of 5.2 mn, and an isoelectric point of 6.9. From kinetic studies, Km-values of 9.8 and 51 μM were calculated for NADP and D-glyceraldehyde 3-phosphate, respectively, an absolute specificity for both substrates being observed. L-Glyceraldehyde 3-phosphate was a potent non-competitive inhibior (Ki, 48 μM). The reaction products NADPH and D-3-phosphoglycerate inhibited enzyme activity in a competitive manner with respect to NADP (Ki, 78 μM) and D-glyceraldehyde 3-phosphate (Ki, 1.2 mM), respectively. Thermal inactivation occurred above 45°C and was effectively prevented by either substrate. The presence of essential vicinal thiol groups is suggested by the inactivation produced by diamide, with D-glyceraldehyde 3-phosphate, but not NADP, behaving as a protective agent. The enzyme's possible physiological role in photosynthetic metabolism is discussed briefly.  相似文献   

13.
Sea ice microalgae in McMurdo Sound, Antarctica were examined for photosynthesis-irradiance relationships and for the extent and time course of their photoadaptation to a reduction in in situ irradiance. Algae were collected from the bottom centimeter of coarse-grained congelation ice in an area free of natural snow cover. Photosynthetic rate was determined in short term (1 h) incubations at ?2° C over a range of irradiance from 0 to 286 μE·m?2·s?1. Assimilation numbers were consistently below 0.1 mg C·mg chl a?1·h?1. The Ik's3 averaged only 7 μE·m?2·s?1, and photosynthesis was inhibited at irradiances above 25 μE·m?2·s?1. Photosynthetic parameters of the ice algal community were examined over a nine day period following the addition of 4 cm of surface snow while a control area remained snow-free. A reduction of 40% in PmB relative to the control occurred after two days of snow cover; α, β, Ik, and Im were not significantly altered. Low assimilation numbers and constant standing crop size, however, suggested that the algal bloom may have already reached stationary growth phase, possibly minimizing their photoadaptive response.  相似文献   

14.
Meridic and oligidic diets suitable for the continuous culture of the aphid Myzus persicae were developed through modifications of a holidic diet. These included the addition of various amounts of crude nutrients and the replacement of defined nutrients by crude nutrients. The highest level of aphid growth (mean weights of 600 to 800 μg) was maintained (for 45 successive generations) on a meridic diet made by supplementing a holidic diet with 2.0% yeast extract (NBC).Continuous culture, at a level of growth (400 to 600 μg) comparable to that on the unsupplemented holidic diet (formulated with 34 defined nutrients), was supported by an oligidic diet containing only 10 weighed ingredients: 15 g sucrose, 2.5 g enzymatic casein hydrolysate (NBC), 2 g yeast extract, 123 mg MgSO4·7H2O, 100 mg ascorbic acid, 10 mg niacin, 5 mg Ca pantothenate, 2.5 mg pyridoxine, 2.5 mg thiamine, and appx. 1.5 gm K2HPO4·3H2O (pH 6.8) per 100 ml of diet.Yeast extract at 2.0% provided adequate amounts of choline chloride, biotin, folic acid and inositol, but not of niacin, pantothenate, pyridoxine, and thiamine. Enzymatic casein hydrolyzate at 2.5% could replace the 20 defined amino acids of the holidic diet. Diets with both yeast extract and casein hydrolysate did not have to be supplemented with trace minerals (Cu, Fe, Mn and Zn). Yeast extract at 2.5% provided sufficient amounts of trace minerals, amino acids, and B-vitamins to sustain numerous successive generations, albeit at a low level of growth (200 to 300 μg). The simplicity and low cost of oligidic diets makes them attractive for aphid studies in which nutrition is not the primary consideration.  相似文献   

15.
The kinetics of the light-driven Cl? uptake pump of Synechococcus R-2 (PCC 7942) were investigated. The kinetics of Cl? uptake were measured in BG-11 medium (pHo, 7·5; [K+]o, 0·35 mol m?3; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 mol m?3) or modified media based on the above. Net36Cl? fluxes (?Cl?o,i) followed Michaelis-Menten kinetics and were stimulated by Na+ [18 mol m?3 Na+ BG-11 ?Cl?max= 3·29±0·60 (49) nmol m?2 s?1 versus Na+-free BG-11 ?Cl?max= 1·02±0·13 (54) nmol m?2 s?1] but the Km was not significantly different in the presence or absence of Na+ at pHo 10; the Km was lower, but not affected by the presence or absence of Na+ [Km = 22·3±3·54 (20) mmol m?3]. Na+ is a non-competitive activator of net ?Cl?o,i. High [K+]o (18 mol m?3) did not stimulate net ?Cl?o,i or change the Km in Na+-free medium. High [K+]o (18 mol m?3) added to Na+ BG-11 medium decreased net ?Cl?o,i [18 mol m?3K+ BG-11; ?Cl?max= 2·50±0·32 (20) nmol m?2 s?1 versus BG-11 medium; ?Cl?max= 3·35±0·56 (20) nmol m?2 s?1] but did not affect the Km 55·8±8·100 (40) mmol m?3]. Na+-stimulation of net ?Cl?o,i followed Michaelis-Menten kinetics up to 2–5 mol m?3 [Na+]o but higher concentrations were inhibitory. The Km for Na+-stimulation of net ?Cl?o,i [K1/2(Na+)] was different at 47 mmol m?3 [Cl?]o (K1/2[Na+] = 123±27 (37) mmol m?3]. Li+ was only about one-third as effective as Na+ in stimulating Cl? uptake but the activation constant was similar [K1/2(Li+) = 88±46 (16) mmol m?3]. Br? was a competitive inhibitor of Cl? uptake. The inhibition constant (Ki) was not significantly different in the presence and absence of Na+. The overall Ki was 297±23 (45) mmol m?3. The discrimination ratio of Cl? over Br? (δCl?/δBr?) was 6·38±0·92 (df = 147). Synechococcus has a single Na+-stimulated Cl? pump because the Km of the Cl? transporter and its discrimination between Cl? and Br? are not significantly different in the presence and absence of Na+. The Cl? pump is probably driven by ATP.  相似文献   

16.
Kinetics of net phosphate (Pi) uptake was measured on intact ectomycorrhizal and non‐mycorrhizal Pinus sylvestris seedlings using a semihydroponic cultivation method. The depletion of Pi in a nutrient solution was assessed over a 160–0.2 μM Pi gradient. Growth of the pine seedlings was P limited and measurements were performed 7 and 9 weeks after inoculation. Three ectomycorrhizal fungi were studied: Paxillus involutus, Suillus bovinus and Thelephoraterrestris. Pi uptake was extremely fast in plants colonised by P. involutus. The Pi concentration dropped below 0.2 μM within 4–5 h. In plants colonised with S. bovinus this occurred in 5–6 h and in plants associated with T. terrestris 8 h were needed to run through the whole concentration range. Non‐mycorrhizal plants of similar size and nutrient status decreased Pi to a concentration between 1 and 2 μM in 18 h. Data were curve fitted to a two‐phase Michaelis‐Menten equation. The apparent kinetic constants, Km and Vmax, for the high affinity Pi uptake system of the pine roots could be estimated accurately. Vmax of this system was up to 7 times higher in pines associated with P. involutus than in non‐mycorrhizal seedlings. The intact extraradical mycelium greatly increased the absorption surface area of the roots (Vmax). Non‐mycorrhizal plants had a Km between 7.8 and 16.4 μM Pi. Plants mycorrhizal with P. involutus had Km values between 2.4 and 7.2, plants colonised with S. bovinus had a Km between 5.1 and 12.3, and seedlings associated with T. terrestris had a Km from 4.6 to 10.1 μM Pi. All 3 ectomycorrhizal fungi had a strong impact on the Pi absorption capacity of the pine seedlings. The results also demonstrated that there is substantial heterogeneity in kinetic parameters among the different mycorrhizal root systems.  相似文献   

17.
A further quantitative analysis of the localization of the centromere on chromosomes was made using 16,817 individual chromosomes obtained from 723 mammalian species. Centromeric position was expressed quantitatively by the size of short arms as per cent weight (Sw) relative to the X-containing haploid set. When the class interval of Sw was 0·1 instead of the previous 0·2 (Imai, 1975), the frequency distribution of Sw showed an uneven (W-shaped) pattern with two distinct antimodes lying at Sw 0·6 (reconfirmation) and 0·1 (new finding). Two hypotheses, that are not mutually exclusive, are proposed to explain the non-random distribution of centromere position. One is that there are three structurally different short arms consisting of (1) centromere, (2) constitutive heterochromatin (as determined by C-banding), and (3) euchromatin, each arm-type being approximately characterized by the size of short arms (Sw) as Sw < 0·1, 0·1 ? Sw ? 0·6 and Sw > 0·6. The other possibility is concerned with an “orthogenetical” change of chromosome morphologies. When the chromosomes with Sw < 0·1, 0·1 ? Sw ? 0·6 and Sw > 0·6 are denoted as telocentric (T), acrocentric (A), and meta-, submeta- and subtelocentric (M, SM & ST), it was suggested that the chromosome morphologies tend to change orthogenetically (in a statistical sense) from T to M, SM & ST via A-chromosomes by rearrangements such as tandem growth of constitutive heterochromatic, pericentric inversions, and centric fusions.  相似文献   

18.
The metabolic fate of photosynthetically-fixed CO2 was determined by labeling samples of Merismopedia tenuissima Lemmerman for 30 min with NaH14CO3 and analyzing its incorporation into low molecular weight compounds, polysaccharide and protein. In N- and P-sufficient cultures, relative incorporation into protein increased as the irradiance used during the labeling period was decreased to 20 μE · m-2 s-1. This pattern was found for cells grown at irradiances of either 20 or 180 μE · m-2· s-1, although incorporation into protein was greater in cultures grown at the higher irradiance. In N-limited continuous cultures, relative incorporation into protein was low, independent of growth rate, and the same for samples tested at 20 or 180 μE · m-2· s-1 irradiance. In contrast, 14C incorporation into protein by P-limited cultures increased as growth rate increased, and at relative growth rates greater than 0.25, the incorporation was greater at 20 than at 180 μE · m-2· s-1. However, the total RNA content and maximum photosynthetic rate of the cultures was the same at all growth rates tested. The interaction between nutrient concentration and light intensity was studied by growing-limited continuous cultures at the same dilution rate, but different irradiances. Relative incorporation into protein was highest in cultures grown at 20 μE · m-2· s-1, in which the relative growth rate was 0.4. These results suggest that photosynthetic carbon metabolism may respond to relative growth rate μ/μmax rather than to growth rate directly.  相似文献   

19.
Aims: This study was conducted to characterize the growth of and aflatoxin production by Aspergillus flavus on paddy and to develop kinetic models describing the growth rate as a function of water activity (aw) and temperature. Methods and Results: The growth of A. flavus on paddy and aflatoxin production were studied following a full factorial design with seven aw levels within the range of 0·82–0·99 and seven temperatures between 10 and 43°C. The growth of the fungi, expressed as colony diameter (mm), was measured daily, and the aflatoxins were analysed using HPLC with a fluorescence detector. The maximum colony growth rates of both isolates were estimated by fitting the primary model of Baranyi to growth data. Three potentially suitable secondary models, Rosso, polynomial and Davey, were assessed for their ability to describe the radial growth rate as a function of temperature and aw. Both strains failed to grow at the marginal temperatures (10 and 43°C), regardless of the aw studied, and at the aw level of 0·82, regardless of temperature. Despite that the predictions of all studied models showed good agreement with the observed growth rates, Davey model proved to be the best predictor of the experimental data. The cardinal parameters as estimated by Rosso model were comparable to those reported in previous studies. Toxins were detected in the range of 0·86–0·99 aw with optimal aw of 0·98 and optimal temperature in the range of 25–30°C. Conclusions: The influences of aw and temperature on the growth of A. flavus and aflatoxin production were successfully characterized, and the models developed were found to be capable of providing good, related estimates of the growth rates. Significance and Impact of the Study: The results of this study could be effectively implemented in minimizing the risk of aflatoxin contamination of the paddy at postharvest.  相似文献   

20.
Abstract—The effect of pentylenetetrazol (PTZ) on acetylcholinesterase (E.C.3.1.1.7) was studied in vitro. The kinetics of the reaction were studied on AChE in crude homogenates of rat brain and in purified preparations from Electrophorus electricus. The Km for rat brain AChE was 1·22 × 10-4m, with a Vmax of 1·37 μmol/g/min whereas the K4 for competitive inhibition of the enzyme by PTZ was 4·7 × 10-3m. The commercially purified enzyme exhibited a Km of 1·73 × 10-4m and a Ki of 1·00 × 10-3m.  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