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 15 毫秒
1.
Summary 3H or14C labeled tracers were used to investigate the metabolism of trimethylamine (TMA), trimethylamine oxide (TMAO), choline, and betaine in free swimming kelp bass (Paralabrax clathratus). An indwelling cannula in the ventral aorta was used to administer tracer and withdraw blood samples. The concentrations of TMA and TMAO were determined in liver, muscle, and plasma. The TMA liver content is higher than that of muscle (0.85 vs 0.01 moles/g wet tissue) while the amount of TMAO in muscle greatly exceeds its liver concentration (60 vs 0.04 moles/g wet tissue). Prolonged fasting (21 and 75 days) or feeding the fish a squid diet containing high levels of TMAO did not alter the tissue concentrations of TMA or TMAO, suggesting that these compounds are endogenous in origin and that their tissue concentrations are subject to regulation. Comparison of the radiospecific activities of TMA and TMAO, and the administered TMA tracer suggest that TMA is channled directly to TMAO in the liver without equilibration in the hepatic TMA pool. The conversion kinetics of TMA to TMAO and the distribution of these amines in liver and muscle with time suggest that labeled TMA is rapidly taken up into a sequestered pool from which it is slowly released, oxidized to TMAO in the liver, and then transported via the circulation to the muscle mass. The location of this proposed sequestered TMA pool was not determined. Experiments with labeled choline and betaine suggest that these compounds are interconverted in the liver and that enzymes are present for conversion of choline betaine TMA TMAO. Labeled dimethylamine (DMA) was not metabolized and is, therefore, probably not a precursor of TMA and TMAO. [14C]Trimethylamine (TMA) was also used to investigate the possible role of trimethylamine oxide (TMAO) as an osmoregulatory compound in migrating prespawning cannulated Pacific pink salmon (Oncorhynchus gorbuscha) taken from marine or fresh water environments. Marine and fresh water salmon oxidized administered [14C]TMA to TMAO; labeled metabolites other than TMA and TMAO were not detected. Four hours after [14C]TMA injection about 10% of the administered dose was present in muscle as labeled TMAO and about 33% as TMA. Unlike our finding in kelp bass, [14C]TMAO was not recovered in liver, although low amounts of labeled TMA were found (0.4% of administered dose). Labeled TMA and TMAO, however, were detected in liver after [14C]betaine adminstration to a marine salmon, indicating that TMA-mono-oxygenase is present in salmon liver. The presence of labeled choline indicates that choline and betaine are interconverted as in kelp bass. The amount of [14C]TMA oxidized to [14C]TMAO and then accumulated in the muscle mass is the same in marine and fresh water salmon, as is the amount of chemical TMAO present (4.6 moles/g muscle).  相似文献   

2.
G. Paulson  C. Struble 《Life sciences》1980,27(20):1811-1817
14C-Sulfamethazine [4-amino-N-(4, 6-dimethyl-2-pyrimidinyl)- 14C-benzenesulfonamide] was administered orally to young, 21-kg male swine. A unique metabolite, N-(4, 6-dimethyl-2-pyrimidinyl)- 14C-benzenesulfonamide (14C-desaminosulfamethazine), was identified in skeletal muscle collected 24 hr after dosing. The structure of the metabolite was confirmed by synthesis, comparative electron impact mass spectrometry, chemical ionization mass spectrometry, and infrared spectroscopy.  相似文献   

3.
1. The effects of spawning coho (Oncorhynchus kisutch) and chum salmon (Oncorhynchus keta) on the limnephilid caddisfly Ecclisomyia conspersa were evaluated by experimentally excluding salmon from the upper 14‐m stretch of a spawning channel by a wire‐meshed fence. Density, and development and growth rates, of larvae upstream of the fence (without salmon) were compared with those downstream (with salmon). 2. Larval density in the stretch with salmon declined during spawning, but increased again after spawning subsided and the carcasses of dead fish became available. In the stretch with salmon, larval density on salmon carcasses was seven to 37 times greater than on the adjacent channel substratum. The rate of larval development in the stretch with salmon was greater than that in the stretch without salmon. Two months after carcasses became available, 98% of larvae sampled from the stretch with salmon were in the fifth instar, compared to only 23% from the stretch without salmon. Body weight of E. conspersa in the stretches with and without salmon increased by an average of 3.04 and 2.38 mg, respectively, over a 6‐month period. 3. 15N values of larvae from the stretch with salmon increased following the arrival of the fish, suggesting that the larvae were feeding on salmon‐derived material, such as eggs and carcasses, which contain a high proportion of the heavier stable isotope. In contrast, 15N values of larvae from the stretch without salmon remained relatively constant throughout the experiment. The availability of salmon carcasses as a high‐quality food source late in larval development may increase survival and fecundity of E. conspersa. 4. These substantial differences were consistent with the view that they were due to the experimental exclusion of salmon and salmon carcasses from the upstream stretch, though the study was un‐replicated and thus precludes ascribing causation more definitely.  相似文献   

4.
The catabolic fate of circulating collagen (Col) in the Atlantic salmon (Salmo salar) was studied. Serum t1/2 and organ distribution of circulating Col in salmon were determined using Col conjugated with l25I-tyramine cellobiose (125I-TC), a low molecular weight adduct which is trapped intralysosomally at the site of uptake. Intravenously administered l25I-tyramine cellobioselabelled Col type I was prepared either from salmon skin (sCol) or rat skin (rCol). Biphasic clearance kinetics of l25I-TC-sCol in salmon were apparent, with 78% being removed from the circulation in an initial rapid α-phase (t1/2(α) = 2.4 min), and 22% being removed more slowly in a terminalβ-phase (t]2(β) = 25.8 min). Serum half life of 125I-TC-rCol was found to be 5.4 min (in this type of experiment the number of data points allow the determination of only a monophasic decay slope). Approximately 90% of recovered radioactivity was found in the kidney of the fish. In comparative experiments, 74% of administered 125I-TC-sCol was cleared from the circulation of rats during an initial rapid α-phase with tl/2(α) = 0.8 min, and 26% was eliminated in the terminal β-phase with t1/2(β) = 7.2 min l25I-sCol was endocytosed and degraded in pure cultures of ral liver endothelial cells, which are the main site of clearance of circulating Col in the rat. Moreover, Col from the two species competed for the same receptor on cultured rat liver endothelial cells, Intravenous administration of tetramethyl rhodamine isothiocyanate-labelled sCol (TRITC-sCol) in salmon, and subsequent examination of sections of kidney in the fluorescence microscope, revealed that the fluorochrome was accumulated exclusively in discrete vesicles of sinusoidal lining cells. Analyses of kidney tissue 24h after intravenous administration of a mixture of fluorescein isothiocyanate-labelled latex beads and TRITC-sCol revealed no codistribution of the two fluorochromes, suggesting that the injected Col was taken up in cells different from macrophages. Purified pronephros macrophages prepared after simultaneous injections of stained beads and Col contained only fluorescein-labelled latex particles. Interestingly, the cells which had accumulated TRITC-sCol appeared to be equally distributed in both pronephros and the part of the kidney containing tubuli. We conclude that Col which gains access to the circulation of the Atlantic salmon is cleared mainly by uptake into sinusoidal lining cells of the kidney. These cells are distinct from phagocytosing macrophages, and morphologically similar to the highly specialized scavenging endothelial cells of mammalian liver sinusoids.  相似文献   

5.
A study was conducted to compare astaxanthin binding ability of solubilized muscle proteins of Atlantic salmon (Salmo salar L.), haddock (Melanogrammus aeglefinus L.) and Atlantic halibut (Hippoglossus hippoglossus L.). Muscle proteins of juvenile Atlantic salmon, haddock and halibut were solubilized by sequential extraction of muscle tissue using low ionic strength solutions. Electrophoretic protein profiles of the six solubilized fractions from these species were similar. Each solubilized fraction from the three species was examined for its relative astaxanthin binding capacity. The amount of bound astaxanthin was significantly different (P < 0.05) among the six fractions of each species. Significant differences in astaxanthin binding were only found for fractions A and E among the species. The amount of bound astaxanthin in various fractions of each species showed a good correlation (R2 = 0.80–0.92) with the ANS (8-anilino-1-naphthalenesulfonate) fluorescence intensity of those fractions. The pattern and extent of astaxanthin binding to the muscle proteins of juvenile salmon, haddock and halibut is comparable to that reported previously for adult Atlantic salmon [Saha, M.R., Ross, N.W., Gill, T.A., Olsen, R.E., Lall, S.P., 2005. Development of a method to assess binding of astaxanthin to Atlantic salmon S. salar L. muscle proteins. Aquacult. Res. 36, 336–343.]. These combined observations suggest that the carotenoid binding capacity of the muscle proteins of salmon is not the limiting factor in the deposition of carotenoid in their flesh.  相似文献   

6.
Salmon eggs and unfed fry were planted in reaches (total length 2.8 km, mean width 4 m) of a Scottish stream between 1971 and 1977 and their subsequent progress was studied by sampling 16 sections (areas 38–126 m2) of the stream. Brown trout are the only fish which spawn in the stream, waterfalls and a dam near its mouth preventing adult salmon and sea-trout passing upstream. There were no restraints on the downstream movement of fish except in 1977, when a fry trap was operated. In 1971 and 1974 boxes each containing 300 eggs were buried in groups of 3–6. In other years fry were evenly distributed at 3.6–29.3 m?2. At the end of the first growing season, salmon occurred at decreasing population densities for a distance of 600 m below the plantings, but after two growing seasons there was little remaining indication of their pattern of dispersion when planted. Rates of survival between planting and the end of the growing season were 9.4–31%. Survival when eggs were planted (11.1–14.8%) was not affected by the numbers planted together at one point (900–1800) or the distance apart of groups of boxes (10–85 m). When fry were planted the instantaneous mortality rate (M) of the 0+ salmon during their first growing season was related to the initial stocking density (Dp) by the formula M= 0.00637 + 0.00444 log10Dp. Twenty-two to 88% of 0+ salmon present at the end of the growing season were still surviving in the stream as 1+ fish one year later. In 1973–1976 only a small number of 2+ salmon occurred, the majority having migrated between the end of the second growing season and the following spring. There were more 2+ salmon in 1977 and 1978 resulting from higher stocking densities in 1975 and 1976 and slower growth. Trout of several age classes were present but their population densities were never high (<0.6 m?2). Salmon reached a greater size than trout by the end of the first growing season. Their mean weight (Wo, g) at this time was inversely related to their population density (Do No. m?2) and the biomass (B1, g m?2) of 1+ salmon present, giving the relationship log10wo= 0.6584–0.0558 D0-0.0352B1. The mean weight of 1+ salmon tended to be highest in sections where the 0+ salmon had reached a relatively large size the previous year. When a reach of the stream was planted twice (11 and 30 May 1977) with salmon fry (total 13.9 m?2) at the same stage of development, M during the first growing season was 0.0099 per day. This was less than that of fry in a control (M= 0.0107) where the stocking density was lower (6.8 m?2) and also less than in previous years when single planting rates of approximately 14 m?2 were used (M=0.0115). The double planting resulted in a wide range of lengths of 0+ salmon in September and the highest biomass values encountered during all experiments.  相似文献   

7.
Landscape genomics is a rapidly growing field with recent advances in both genotyping efficiency and statistical analyses that provide insight towards local adaptation of populations under varying environmental and selective pressure. Chinook salmon (Oncorhynchus tshawytscha) are a broadly distributed Pacific salmon species, occupying a diversity of habitats throughout the northeastern Pacific with pronounced variation in environmental and climate features but little is understood regarding local adaptation in this species. We used a multivariate method, redundancy analysis (RDA), to identify polygenic correlations between 19 703 SNP loci and a suite of environmental variables in 46 collections of Chinook salmon (1956 total individuals) distributed throughout much of its North American range. Models in RDA were conducted on both rangewide and regional scales by hierarchical partitioning of the populations into three distinct genetic lineages. Our results indicate that between 5.8 and 21.8% of genomic variation can be accounted for by environmental features, and 566 putatively adaptive loci were identified as targets of environmental adaptation. The most influential drivers of adaptive divergence included precipitation in the driest quarter of the year (Rangewide and North Coastal Lineage, anova = 0.002 and 0.01, respectively), precipitation in the wettest quarter of the year (Interior Columbia River Stream‐Type Lineage, anova = 0.03), variation in mean diurnal range in temperature (South Coastal Lineage, anova = 0.005), and migration distance (Rangewide, anova = 0.001). Our results indicate that environmental features are strong drivers of adaptive genomic divergence in this species, and provide a foundation to investigate how Chinook salmon might respond to global environmental change.  相似文献   

8.
Abstract

The effects on uptake and biodistribution of radiolabelled lipopolysaccharide (LPS) due to changing routes of administration, encapsulation of LPS within liposomes and altering liposomal surface charge were examined in rainbow trout (Oncorhynchus mykiss). 3H-labelled LPS, positively- and negatively-charged (14C-labelled) liposomes or 14C-labelled liposomes containing 3H-LPS were administered to trout via intravenous, intraperitoneal, intramuscular, or oral routes. Twenty-four hours following administration, relative uptake of LPS and multilamellar vesicles (MLV) based on detection of 3H and 1AC, respectively, was determined in samples taken from the kidney, spleen, liver, plasma, blood cells and skeletal muscle. In general, regardless of the route of administration, 3H-LPS, 1AC-MLV and liposomally encapsulated LPS were recovered primarily in the kidney and spleen. Intravenous administration resulted in the greatest uptake of radiolabel by the kidney and spleen, followed by the intraperitoneal and intramuscular routes. Although oral administration yielded the lowest overall uptake of labelled material, detection of 3H and 14C in the liver was enhanced when compared with the other routes. Negatively-charged MLV were delivered more efficiently to the kidney and spleen than positively-charged MLV; but negatively- and positively-charged MLV containing LPS demonstrated the opposite relationship between charge and distribution among the kidney and spleen. These results suggest that liposomal encapsulation (particularly within positively-charged MLV) enhances delivery of LPS to the primary hemopoietic organs in rainbow trout.  相似文献   

9.
Little is known about the behaviour patterns and swimming speed strategies of anadromous upriver migrating fish. We used electromyogram telemetry to estimate instantaneous swimming speeds for individual sockeye (Oncorhynchus nerka) and pink salmon (O. gorbuscha) during their spawning migrations through reaches which spanned a gradient in river hydraulic features in the Fraser River, British Columbia. Our main objectives were to describe patterns of individual-specific swim speeds and behaviours, identify swimming speed strategies and contrast these between sexes, species and reaches. Although mean swimming speeds did not differ between pink salmon (2.21 BL s–1) and sockeye salmon (1.60 BL s–1), sockeye salmon were over twice as variable (mean CV; 54.78) in swimming speeds as pink salmon (mean CV; 22.54). Using laboratory-derived criteria, we classified swimming speeds as sustained (<2.5 BL s–1), prolonged (2.5–3.2 BL s–1), or burst (>3.2 BL s–1). We found no differences between sexes or species in the proportion of total time swimming in these categories – sustained (0.76), prolonged (0.18), burst (0.06); numbers are based on species and sexes combined. Reaches with relatively complex hydraulics and fast surface currents had migrants with relatively high levels of swimming speed variation (e.g., high swimming speed CV, reduced proportions of sustained speeds, elevated proportions of burst speeds, and high rates of bursts) and high frequency of river crossings. We speculate that complex current patterns generated by river constrictions created confusing migration cues, which impeded a salmon's ability to locate appropriate pathways.  相似文献   

10.
We quantified the amount, spatial distribution, and importance of salmon (Oncorhynchus spp.)-derived nitrogen (N) by brown bears (Ursus arctos) on the Kenai Peninsula, Alaska. We tested and confirmed the hypothesis that the stable isotope signature (δ15N) of N in foliage of white spruce (Picea glauca) was inversely proportional to the distance from salmon-spawning streams (r=–0.99 and P<0.05 in two separate watersheds). Locations of radio-collared brown bears, relative to their distance from a stream, were highly correlated with δ15N depletion of foliage across the same gradient (r=–0.98 and –0.96 and P<0.05 in the same two separate watersheds). Mean rates of redistribution of salmon-derived N by adult female brown bears were 37.2±2.9 kg/year per bear (range 23.1–56.3), of which 96% (35.7±2.7 kg/year per bear) was excreted in urine, 3% (1.1±0.1 kg/year per bear) was excreted in feces, and <1% (0.3± 0.1 kg/year per bear) was retained in the body. On an area basis, salmon-N redistribution rates were as high as 5.1±0.7 mg/m2 per year per bear within 500 m of the stream but dropped off greatly with increasing distance. We estimated that 15.5–17.8% of the total N in spruce foliage within 500 m of the stream was derived from salmon. Of that, bears had distributed 83–84%. Thus, brown bears can be an important vector of salmon-derived N into riparian ecosystems, but their effects are highly variable spatially and a function of bear density. Received: 11 February 1999 / Accepted: 7 July 1999  相似文献   

11.
Newly emerged Atlantic salmon (Salmo salar) were observed from May to August 1981 at six isolated redds in Washington County, Maine, USA. Territorial size and distribution were measured. At the end of the emergence period (12 to 28 May), fish maintained positions (stations) at redds where water velocity did not exceed 52 cm s–1 By 12 June, most salmon (80–96%) had moved off the redds of origin and had established territories 1 to 5 m from the redd. The area defended increased substantially after mid-June, but territorial aggression diminished by 15 July, and the fry dispersed downstream. All fish observed were territorial, and the percentage of time during which stations were held decreased from 89 in mid-May to as low as 40% in mid-June.Cooperators are the Maine Department of Inland Fisheries and Wildlife, University of Maine, U.S. Fish and Wildlife Service, and the Wildlife Management Institute  相似文献   

12.
The following organophosphates were tested for their ability to induce DNA damage in a rec-type repair test with Proteus mirabilis strains PG713 (rec? hcr?) and PG273 (wild type) and point mutations in his? strain TA100 of Salmonella typhimurium — butonate: O,O-dimethyl-(1-n-butyryloxy-2,2,2-trichloroethyl)-phosphonate; vinylbutonate: O,O-dimethyl-(n-butyryloxy-2,2-dichlorovinyl)-phosphonate; trichlorfon: O,O-dimethyl-(1-hydroxy-2,2,2-trichloroethyl)-phosphonate; dichlorvos: O,O-dimethyl-O-(2,2-dichlorovinyl)-phosphate; the demethylated derivatives — demethyldichlorvos: O-methyl-O-(2,2-dichlorovinyl)-phosphoric acid; demethyl vinylbutonate: O-methyl-(1-n-butyryloxy-2,2-dichlorovinyl)phosphonic acid. Of the six compounds tested, dichlorvos and trichlorfon induced base pair substitutions and DNA damage. No mutagenicity and DNA damage were found in experiments with butonate, vinylbutonate, demethyl vinylbutonate and demethyl dichlorvos. Genotoxic activity for dichlorvos and the absence of both mutagenic and DNA damaging properties for its non-alkylating demethyl derivative favors the hypothesis that alkylation of DNA is the essential step for mutation induction by this organophosphate. Furthermore, the absence of genetic effects after treatment with vinylbutonate and demethyl dichlorvos does not support a crucial role of vinyl or allyl groups in side chains of organophosphates for genetic activity. Microsomal enzymes decreased genetic activity of dichlorvos and trichlorfon in vitro. No evidence for a role of metabolic activation in the mutagenic activity of any of these compounds was found.  相似文献   

13.
We isolated cDNA clones encoding a transglutaminase (TGase: EC 2.3.2.13) from a salmon (Onchorhynchus keta) cDNA library prepared from the liver. In the cDNA sequence combined, an open reading frame coding for a protein of 680 aa was found. The deduced sequence showed a considerable similarity (62.4%) to that of red sea bream TGase. By comparison of sequence similarity to other TGases, the structure of salmon TGase was like tissue type TGases, rather than membrane-associated type or plasma type TGases. As a structural feature of salmon TGase, 3 aa residues were substituted in the 25 aa sequence around the active site Cys residue, which is conserved among several tissue type TGases. The critical residues thought to form the catalytic-center triad (Cys272, His331, and Asp301) were found in the highly conserved region, but the region surrounding Tyr511, which corresponds to the residue participates in hydrogen-bond interactions of active center domain, was less similar to other TGases, except for red sea bream TGase. These findings suggests that the overall structure of fish TGase resembles tissue-type TGases, but has some unique structure.  相似文献   

14.
The (+)- and (?)-enantiomers of 3-isopropyl 5-(4-methylphenethyl) 1,4-dihydro-2,6-dimethyl-4-(2-pyridyl)-3,5-pyridinedicarboxylate were synthesized using an efficient highly enantioselective (ee ≥ 96%) variant of the Hantzsch dihydropyridine synthesis. The key step in this procedure involved the asymmetric Michael addition of a metalated chiral aminocrotonate, derived from D -valine or L -valine, respectively, to the Knoevenagel acceptor (Z)-2-isopropoxycarbonyl-1-(2-pyridyl)-but-1-en-3-one. Both enantiomers exhibited a dual cardioselective partial calcium channel agonist (positive inotropic)/smooth muscle selective calcium channel antagonist effect. The relative in vitro smooth muscle calcium channel antagonist activities of the (?):(+) enantiomers was 26:1. In contrast, the (+)-enantiomer exhibited a greater in vitro positive inotropic effect on guinea pig left atrium where the contractile force was maximally increased by 14.8% at a concentration of 1.63 × 10?8 M. © 1994 Wiley-Liss, Inc.  相似文献   

15.
Breeding location choice provides a mechanism by which individuals can directly influence their reproductive success. Location choice should therefore reflect individual condition, habitat features, and the intensity of competition; with these factors then influencing reproductive success. To test whether such patterns were detectable in the wild, we tagged 705 sockeye salmon (Oncorhynchus nerka) in a natural population, and monitored them from when they started breeding until they died. We evaluated the role of individual condition (size, secondary sexual traits, energy stores) in the acquisition of breeding locations that differed in the intensity of competition (female density, sex ratio) and habitat features (water depth, water velocity). We then evaluated the influence of breeding location on reproductive life span and energy stores. At a coarse level (20‐m stream sections), females consistently settled in certain locations, and these locations sustained high densities and held larger females. At a fine scale (0.5‐m breeding sites), (1) larger fish occupied deeper water (males, r2=0.072; females, r2=0.199), (2) higher levels of competition reduced reproductive life span for males (r2=0.139) but not females, and (3) fish with shorter reproductive life spans died with more energy remaining in their muscle tissue (males, r2=0.414; females, r2=0.440). These patterns were nested within a tendency for late breeding fish to have shorter reproductive life spans. Energy stores and secondary sexual traits did not influence breeding location choice, and larger fish did not acquire locations of higher intrinsic quality (i.e., those sections settled first and sustaining higher competition). Our study provides evidence that some aspects of individual condition influence breeding location choice, which then influences components of reproductive success.  相似文献   

16.
Cooling increases the twitch force of frog skeletal muscle (Rana temporaria; Rana pipiens), but decreases the twitch force of tropical toad muscle (Leptodactylus insularis). Action potentials and intramembranous charge movement in frog and toad fibers were slowed identically by cooling. Cooling increased the integral of twitch Ca2+ detected by aequorin in frog fibers (1.4-fold), while also decreasing the peak and slowing the rate of decay. Conversely, cooling decreased the integral (0.6-fold) and the peak of twitch Ca2+ in toad fibers, without affecting the rate of decay. The difference in entire Ca2+ transients may account for cold-induced twitch potentiation in frogs and twitch paralysis in toads. In sustained contractions of toad fibers, cooling markedly decreased maximum force caused by: (i) tetanic stimulation, (ii) two-microelectrode voltage clamp steps, (iii) high [K+], or (iv) caffeine. Maximum force in sustained contractions was decreased moderately by cooling frog fibers. Rapid rewarming and simultaneous removal of high [K+] or caffeine during a sustained contraction, caused toad muscle force to rise towards the value corresponding to the warm temperature. This did not occur after removing high [K+] or caffeine from toad fibers kept in the cold. Transmission electron micrographs showed no relevant structural differences. Parvalbumins are thought to promote relaxation of frog muscle in the cold. The unique parvalbumin isoforms in toad muscle apparently lack this property. Accepted: 27 August 1998  相似文献   

17.
Migratory animals, such as Pacific salmon, can significantly shape communities in recipient habitats both by altering the flux of resources, and changing community composition and subsequent trophic interactions. Here we mainly used paleoecological records from natural sockeye salmon nursery lakes to quantify the response of plankton communities to the influx of salmon‐derived nutrients and consumers (juvenile salmon). Our long‐term data show that increases in the density of spawning salmon often elevated influx of nutrients, and, in turn, zooplankton production over the past few centuries. In contrast, significant correlations were not detected in two lakes with extremely low or high average spawner densities (i.e. 1.5 and 34.7 × 103 spawners km?2 year?1 respectively). With increasing spawner densities across lakes, analysis of the size structure of subfossils in sediments revealed a strong decrease in body size of a main juvenile salmon prey item (Eubosmina longispina; r2= 0.36, p < 0.001, n = 67), consistent with an overriding effect of predation in lakes with high salmon densities. These long‐term data not only highlight the key role of salmon‐derived nutrients in stimulating plankton communities, but also suggest that the relative effect of nutrient and consumer subsidies varies along gradients of lake production, despite a single ultimate causal mechanism (migrating fish).  相似文献   

18.
The eradication of the haematophagous isopod ectoparasite Gnathia maxillaris was not achieved after several years of physical elimination of larvae by filtration and chemical treatment with trichlorfon in a captive marine fish population in exhibition aquaria. In this study, different in vitro laboratory assays were performed to find effective alternative anti‐parasitic treatments to trichlorfon (lufenuron, emamectin benzoate, cypermethrin and abamectin). The lethal concentration 50 (LC50) values at 96 hr were calculated for each compound using both larval stages (zuphea and praniza); morphological deformities generated throughout the life cycle, reduction in egg laying and survival of adult forms were also determined. Abamectin, cypermethrin and emamectin benzoate proved effective at limiting further Gnathia maxillaris development, but their efficacy at eradicating an infestation in large tanks was not supported by laboratory assays. However, the use of lufenuron proved to be a good substitute for trichlorfon when treating this type of infestation in large‐volume tanks (LC50 55.8 mg m?3 for praniza). Routine lufenuron treatment once a month via an oral dose at 10 mg kg?1 body weight eradicated G. maxillaris from the exhibition aquaria.  相似文献   

19.
The previous identification of 2,5-dimethyl-3-(3-methylbutyl)pyrazine as the mandibular alarm pheromone of the little fire ant Wasmannia auropunctata (Roger), has been found to be incorrect. Gas chromatography–mass spectrometry (GC/MS) of ant extracts suggested the correct structure to be the regioisomer 2,5-dimethyl-3-(2-methylbutyl)pyrazine, which was confirmed by comparison with the synthetic pyrazine. GC/MS analysis also revealed the presence of an additional disubstituted alkylpyrazine which was identified as 3-methyl-2-(2-methylbutyl)pyrazine. Headspace sampling of confined ants with SPME and Porapak Q followed by GC/MS analysis showed 2,5-dimethyl-3-(2-methylbutyl)pyrazine as the major volatile released by W. auropunctata workers while 3-methyl-2-(2-methylbutyl)pyrazine was only detected in trace amounts. In laboratory bioassays, W. auropunctata workers were attracted and arrested by both pyrazines, although the results were not always consistent. Synthetic pyrazines generally attracted as many W. auropunctata workers as were attracted to a single crushed ant. However, higher numbers of W. auropunctata were arrested by crushed ant treatments than by synthetic pyrazines in all bioassays but one.  相似文献   

20.
Objective: Mitochondrial activity is altered in skeletal muscle of obese, insulin‐resistant or type 2 diabetic patients. We hypothesized that this situation was associated with profound adaptations in resting muscle energetics. For that purpose, we used in vivo 31P‐nuclear magnetic resonance (31P‐NMR) in male sedentary Wistar rats fed with obesogenic diets known to induce alterations in muscle mitochondrial activity. Methods and Procedures: Two experimental diets (high sucrose and high fat) were provided for 6 weeks at two levels of energy (standard, N and high, H) and compared to control diet. The rates of the adenosine triphosphate (ATP) exchange between phosphocreatine (PCr) and γ ‐ATP (ka) and β ‐adenosine diphosphate ( β ‐ADP) to β ‐ATP (kb) were evaluated using 31P‐NMR in resting gastrocnemius muscle. Muscle contents in phosphorylated compounds as well as creatine, were assessed using 31P‐NMR and biochemical assays, respectively. Results: ATP content increased by 6.7–8.5% in standard‐energy high‐sucrose (NSU), high‐energy high‐fat (HF) and high‐energy high‐sucrose (HSU) groups compared to control (P < 0.05), whereas PCr content decreased by 4.2–6.4% (P < 0.01). Consequently, PCr to ATP ratio decreased in NSU, HF, and HSU groups, compared to control (P < 0.01). Furthermore in high‐energy groups (HF and HSU) compared to control, creatine contents were decreased by 14–19% (P < 0.001), whereas ka and kb fluxes were increased by 89–133% (P < 0.001) and 243–277% (P < 0.01), respectively. Discussion: Our in vivo data showed adaptations of resting skeletal muscle energetics in response to high‐energy diets. Increased activity of enzymes catalyzing ATP production may reflect a compensatory mechanism to face impaired mitochondrial ATP synthesis in order to preserve intracellular energy homeostasis.  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