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.
Biodegradation of Methyl red by Galactomyces geotrichum MTCC 1360   总被引:1,自引:0,他引:1  
Galactomyces geotrichum MTCC 1360 can decolorize triphenylmethane, azo and reactive high exhaust textile dyes. At shaking condition this strain showed 100% decolorization of a toxic azo dye Methyl red (100 m gl−1) within 1 h in deionized water at 30 °C. The degradation of Methyl red was possible through a broad pH (3–12) and temperature (5–50 °C) range. Glucose and mycelium concentration had increased the decolorization rate, but the addition of 1 gl−1 molasses in deionized water made decolorization possible in only 10 min. Induction in the NADH–dichloro phenol indophenol (NADH–DCIP) reductase, Malachite green reductase, laccase and lignin peroxidase (Lip) activities were observed in the cells obtained after complete decolorization, showing that there is direct involvement in the degradation of Methyl red. The absence of N-N′-dimethyl-p-phenylenediamine (DMPD) in 5 °C, 2-aminobenzoic acid (ABA) in 50 °C and both the compounds in 30 °C sample have shown the differences in the metabolic fate of Methyl red at different temperatures. The untreated dye at 300 mg l−1 concentration showed 88% germination inhibition in Sorghum bicolor, whereas it was 72% in Triticum aestivum. There was no germination inhibition for both the plants by Methyl red metabolites at 300 mg l−1 concentration.

The scientific relevance of the paper

The azo dye Methyl red (100 mg l−1) was decolorized by G. geotrichum MTCC 1360 within 1 h at shaking condition in deionized water. This organism could decolorize Methyl red at wide pH and temperature ranges. Decolorization time was reduced to 10 min by the addition of molasses to deionized water. There was induction in laccase and Lip, NADH–DCIP reductase and Malachite green reductase activities. The metabolic fate of Methyl red changes with temperature which can be evidenced by the formation of 2-ABA at 5 °C, N-N′-DMPD at 50 °C and both the compounds were absent at 30 °C. Phytotoxicity showed that metabolites of dye had induced shoot and root length of both the tested plants.  相似文献   

2.
The reaction of the monofunctional [Pt(Gly-Gly-N,N′,O)I] complex, in which Gly-Gly is the dipeptide glycyl-glycine coordinated through two nitrogen and oxygen atoms, with the N-acetylated dipeptide l-methionyl-l-histidine (MeCOMet-His) studied by 1H NMR spectroscopy. All reactions were carried out in 50 mM phosphate buffer at pD 7.4 and at 25 °C. In the initial stage of the reaction, the platinum(II) complex forms the kinetically favored [Pt(Gly-Gly-N,N′,O)(MeCOMet-His-S)] complex, with unidentate coordination of the MeCOMet-His dipeptide through the sulfur atom of the methionine residue. In the second stage of the reaction, complete intramolecular migration of the [Pt(Gly-Gly-N,N′,O)] unit from the sulfur to the N3 nitrogen atom of imidazole was observed and a new platinum(II)-peptide complex, [Pt(Gly-Gly-N,N′,O)(MeCOMet-His-N3)] was formed. In comparison with previous results obtained for the reaction of [Pt(dien)Cl]+ with different methionine- and histidine-containing peptides, this migration reaction was sufficiently fast and strongly selective to the N3 atom of the imidazole ring of the histidine side chain. This study is an important step in the development of new platinum(II) complexes for selective covalent modification of peptides and proteins.  相似文献   

3.
Several synthetic adeonosine analogs: 8-fluoro-, 8-azido-, 8-iodo-, 8-methylthioadenosine; 8-bromo-2′-deoxyadenosine, 8-bromoxylofuranosyladenine, 5′-benzoly-8-bromoadenosine; 8,2′-S-, 8,2′-O-, 8,2′-NH-, 8,2′-N-CH3-, 8,3′,-S-, 8,3′-O-, 8,5′-S- and 8,5′O-cycloadenosine; 1-deaza- and 3-deazaadenosine, as well as tubercidine (7-deazaadenosine), were tested as substrates of calf intestine adenosine deaminase.It was found that the adenine base of adenosine should be in the range φrmCN = 0–120° (anti to syn-anti) and 8-fluoroadenosine was hydroylzed very slowly. The purine base should have N1, N3 or N7 atoms for the hydrolysis and only 1-deazaadenosine revealed an inhibitory effect toward the hydrolysis of adenosine.5′-OH group should be in the position of S-configuration and must not be substituted.  相似文献   

4.
Three phenolic glycosides 5-O-{[5′′-O-E-(4′′′-O-threo-guaiacylglycerol)-feruloyl]-β-apiofuranosyl-(1→2)-β-xylopyranosyl} gentisic acid, 5-O-[(5′′-O-vanilloyl)-β-apiofuranosyl-(1→2)-β-xylopyranosyl] gentisic acid and 1-O-[E-(4′′′-O-threo-guaiacylglycerol)-feruloyl]-3-O-β-galacturonopyranosyl glycerol were isolated and identified from the roots of Medicago truncatula together with four known 5-O-β-xylopyranosyl gentisic acid, vicenin-2, hovetrichoside C and pterosupin identified for the first time in this species. Structural elucidation was carried out on the basis of UV, mass, 1H and 13C NMR spectral data.  相似文献   

5.
Three series of new cannabinoids were prepared and their affinities for the CB1 and CB2 cannabinoid recptors were determined. These are the 1-methoxy-3-(1′,1′-dimethylalkyl)-, 1-deoxy-11-hydroxy-3-(1′,1′-dimethylalkyl)- and 11-hydroxy-1-methoxy-3-(1′,1′-dimethylalkyl)-Δ8-tetrahydrocannabinols, which contain alkyl chains from dimethylethyl to dimethylheptyl appended to C-3 of the cannabinoid. All of these compounds have greater affinity for the CB2 receptor than for the CB1 receptor, however only 1-methoxy-3-(1′,1′-dimethylhexyl)-Δ8-THC (JWH-229, 6e) has effectively no affinity for the CB1 receptor (Ki=3134±110 nM) and high affinity for CB2 (Ki=18±2 nM).  相似文献   

6.
Thyroid hormone (T3) has been demonstrated to inhibit the action of aldosterone on sodium transport in toad urinary bladder and rat kidney. We have exammined the effect of T3 on aldosterone action and specific nuclear binding in cultured epithelial cells derived from toad urinary bladder. In cell line TB6-C, addition of 5·10−8 M T3 to culture media for up to 3 days results in no change in short-circuit current or transepithelial resistance. This concentration of T3 completely inhibits the maximal increase in short-circuit current in response to 1·10−7 M aldosterone. The inhibition can be demonstrated with 18 h preincubation or with simultaneous addition of T3 and aldosterone. The half-maximal concentration for the inhibition of the aldosterone effect is approx. 5·10−9 M T3. T3 has no effect on cyclic AMP-stimulated short-circuit current in these cells. The effect of T3 on nuclear binding of [3H]aldosterone was examined using a filtration assay with data analysis by at least-squares curve-fitting program. Best fit was obtained with a model for two binding sites. The dissociation constants for the binding were Kd1 = (0.82 ± 0.36)·10−10 M and Kd2 = (3.2±0.60)·10−8 M.The half-maximal concentration for aldosterone-stimulated sodium transport in these cells is approx. 1·10−8 M. Analysis of nuclear aldosterone binding in cells preincubated for 18 h with 5·10−8 M T3 showed a Kd1 = (0.15 ± 0.10)·10−10 M and Kd2 = (3.5 ± 0.10)·10−8 M. We conclude that T3 i action of aldosterone on sodium transport at a site after receptor binding in the nucleus.  相似文献   

7.
A panel of six complementary monodeoxy and mono-O-methyl congeners of methyl β-d-mannopyranosyl-(1→2)-β-d-mannopyranoside (1) were synthesized by stereoselective glycosylation of monodeoxy and mono-O-methyl monosaccharide acceptors with a 2-O-acetyl-glucosyl trichloroacetimidate donor, followed by a two-step oxidation–reduction sequence at C-2′. The β-manno configurations of the final deprotected congeners 2–7 were confirmed by measurement of 1JC1,H1 heteronuclear and 3J1′,2′ homonuclear coupling constants. These disaccharide derivatives will be used to map the protective epitope recognized by a protective anti-Candida albicans monoclonal antibody C3.1 (IgG3) and to determine its key polar contacts with the binding site.  相似文献   

8.
An intracellular β-xylosidase from the thermophilic fungus Sporotricum thermophile strain ATCC 34628 was purified to homogeneity by Q-Sepharose and Mono-Q column chromatographies. The protein properties correspond to molecular mass and pI values of 45 kDa and 4.2, respectively. The enzyme is optimally active at pH 7.0 and 50 °C. The purified β-xylosidase is fully stable at pH 6.0–8.0 and temperatures up to 50 °C and retained over 58% of its activity after 1 h at 60 °C. The enzyme hydrolyzes β-1,4-linked xylo-oligosaccharides with chain lengths from 2 to 6, releasing xylose from the non-reducing end, but is inactive against xylan substrates. The apparent Km and Vmax values from p-nitrophenyl β-d-xylopyranoside are 1.1 mM and 114 μmol p-nitrophenol min−1 mg−1, respectively. Alcohols inactivate the enzyme, ethanol at 10% (v/v) yields a 30% decrease of its activity. The enzyme is irreversibly inhibited by 2,3-epoxypropyl β-d-xylobioside while alkyl epoxides derived from d-xylose were not inhibitors of the enzyme. The enzyme catalyses the condensation reaction using high donor concentration, up to 60% (w/v) xylose.  相似文献   

9.
Net ecosystem exchange of CO2 (NEE) was measured during 2005 using the eddy covariance (EC) technique over a reed (Phragmites australis (Cav.) Trin. ex Steud.) wetland in Northeast China (121°54′E, 41°08′N). Diurnal NEE patterns varied markedly among months. Outside the growing season, NEE lacked a diurnal pattern and it fluctuated above zero with an average value of 0.07 mg CO2 m−2 s−1 resulting from soil microbial activity. During the growing season, NEE showed a distinct V-like diel course, and the mean daily NEE was −7.48 ± 2.74 g CO2 m−2 day−1, ranging from −13.58 g CO2 m−2 day−1 (July) to −0.10 g CO2 m−2 day−1 (October). An annual cycle was also apparent, with CO2 uptake increasing rapidly in May, peaking in July, and decreasing from August. Monthly cumulative NEE ranged from −115 ± 24 g C m−2 month−1 (the reed wetland was a CO2 sink) in July to 75 ± 16 g C m−2 month−1 (CO2 source) in November. The annual CO2 balance suggests a net uptake of −65 ± 14 g C m−2 year−1, mainly due to the gains in June and July. Cumulative CO2 emission during the non-growing season was 327 g C m−2, much greater than the absolute value of the annual CO2 balance, which proves the importance of the wintertime CO2 efflux at the study site. The ratio of ecosystem respiration (Reco) to gross primary productivity (GPP) for this reed ecosystem was 0.95, indicating that 95% of plant assimilation was consumed by the reed plant or supported the activities of heterotrophs in the soil. Daytime NEE values during the growing season were closely related to photosynthetically active radiation (PAR) (r2 > 0.63, p < 0.01). Both maximum ecosystem photosynthesis rate (Amax) and apparent quantum yield (α) were season-dependent, and reached their peak values in July (1.28 ± 0.11 mg CO2 m−2 s−1, 0.098 ± 0.027 μmol CO2 μmol−1 photon, respectively), corresponding to the observed maximum NEE in July. Ecosystem respiration (Reco) relied on temperature and soil water content, and the mean value of Q10 was about 2.4 with monthly variation ranging from 1.8 to 4.1 during 2005. Annual methane emission from this reed ecosystem was estimated to be about 3 g C m−2 year−1, and about 5% of the net carbon fixed by the reed wetland was released to the atmosphere as CH4.  相似文献   

10.
We studied the seasonal variation on aerobic metabolism and the response of oxidative stress parameters in the digestive glands of the subpolar limpet Nacella (P.) magellanica. Sampling was carried out from July (winter) 2002 to July 2003 in Beagle Channel, Tierra del Fuego, Argentina. Whole animal respiration rates increased in early spring as the animals spawned and remained elevated throughout summer and fall (winter: 0.09 ± 0.02 μmol O2 h− 1 g− 1; summer: 0.31 ± 0.06 μmol O2 h− 1 g− 1). Oxidative stress was assessed at the hydrophilic level as the ascorbyl radical content / ascorbate content ratio (A / AH). The A / AH ratio showed minimum values in winter (3.7 ± 0.2 10− 5 AU) and increased in summer (18 ± 5 10− 5 AU). A similar pattern was observed for lipid radical content (122 ± 29 pmol mg− 1 fresh mass [FW] in winter and 314 ± 45 pmol mg− 1 FW in summer), iron content (0.99 ± 0.07 and 2.7 ± 0.6 nmol mg− 1 FW in winter and summer, respectively) and catalase activity (2.9 ± 0.2 and 7 ± 1 U mg− 1 FW in winter and summer, respectively). Since nitrogen derived radicals are thought to be critically involved in oxidative metabolism in cells, nitric oxide content was measured and a significant difference in the content of the Fe–MGD–NO adduct in digestive glands from winter and summer animals was observed. Together, the data indicate that both oxygen and nitrogen radical generation rates in N. (P.) magellanica are strongly dependent on season.  相似文献   

11.
Egg yolk phosphatidylcholine monolamellar liposomes (1000 Å in diameter) loaded with cytochrome c were placed into an external solution, in which superoxide radicals, O2, were generated by a xanthine-xanthine oxidase system. The penetration of the superoxide radicals across the liposomal membrane was detected by cytochrome c reduction in the inner liposome compartment. The effects of modifiers and temperature on this process were studied. The permeability of liposomal membrane for O2(PO2 = (7.6 ± 0.3) · 10-8 cm/s), or HO2 (PHO2 = 4.9 · 10-4 cm/s) were determined. The effect of the transmembrane electric potential (K+ concentration gradient, valinomycin) on the permeability of liposomal membranes for O2 were investigated. It was found that O2 can penetrate across liposomal membrane in an uncharged form. The feasibility of penetration of superoxide radicals through liposomal membrane, predominantly via anionic channels, was demonstrated by the use of an intramolecular cholesterol-amphotericin B complex.  相似文献   

12.
Ty1 is a retrotransposon of the yeast Saccharomyces cerevisiae whose transposition at new locations in the host genome is activated by stress conditions, such as exposure to UV light, X-rays, nitrogen starvation. In this communication, we supply evidence that cooling for 2 h at +4 °C followed by freezing for 1 h at −10 °C and 16 h at −20 °C also increased Ty1 transposition. The mobility of Ty1 was induced by cooling at slow rates (3 °C/min) and the accumulation of trehalose inside cells or the cooling at high rates (100 °C/min) inhibited significantly the induction of the transposition. The freeze-induced Ty1 transposition did not occur in mitochondrial mutants (rho) and in cells with disrupted SCO1 gene (Δsco1 cells) evidencing that the Ty1 transposition induced by cooling depends on the mitochondrial oxidative phosphorylation. We also found that the freeze induced Ty1 transposition is associated with increased synthesis and accumulation of superoxide anions (O2) into the cells. Accumulation of O2 and activation of Ty1 transposition were not observed after cooling of cells with compromised mitochondrial functions (rho, Δsco1), or in cells pretreated with O2 scavengers. It is concluded that (i) elevated levels of reactive oxygen species (ROS) have a key role in activation the transposition of Ty1 retrotransposon in yeast cells undergoing freezing and (ii) given the deleterious effect of increased ROS levels on cells, special precautions should be taken to avoid ROS production and accumulation during cryopreservation procedures.  相似文献   

13.
The compound [Cu2(bipy)2(OH)2](C4O4)·5.5H2O, where bipy and C4O42− correspond to 2,2′-bipyridyl and squarate (dianion of 3,4-dihydroxy-3-cyclo- butene-1,3-dione) respectively, has been synthesized. Its magnetic properties have been investigated in the 2–300 K temperature range. The ground state is a spin-triplet state, with a singlet-triplet separation of 145 cm−1. The EPR powder spectrum confirms the nature of the ground state.Well-formed single crystals of the tetrahydrate, [Cu2(bipy)2(OH)2](C4O4)·4H2O, were grown from aqueous solutions and characterized by X-ray diffraction. The system is triclinic, space group P , with a = 9.022(2), b = 9.040(2), c = 8.409(2) Å, α = 103.51(2), β = 103.42(3), γ = 103.37(2)°, V = 642.9(3) Å3, Z = 1, Dx = 1.699 g cm−3, μ(Mo Kα) = 17.208 cm−1, F(000) = 336 and T= 295 K. A total of 2251 data were collected over the range 1θ25°; of these, 1993 (independent and with I3σ(I)) were used in the structural analysis. The final R and Rw residuals were 0.034 and 0.038 respectively. The structure contains squarato-O1, O3-bridged bis(μ-hydroxo)bis[(2,2′-bipyridyl)copper(II)] units forming zigzag one-dimensional chains. Each copper atom is in a square-pyramidal environment with the two nitrogen atoms of 2,2′-bipyridyl and the two oxygen atoms of the hydroxo groups building the basal plane and another oxygen atom of the squarate lying in the apical position.The magnetic properties are discussed in the light of spectral and structural data and compared with the reported ones for other bis(μ-hydroxo)bis[(2,2′-bipyridyl)copper(II)] complexes.  相似文献   

14.
Changes in guanosine cyclic 3′,5′-monophosphate associated with adenosine cyclic 3′,5′-monophosphate and folic acid addition in the presence of ATP have been examined in Dictyostelium discoideum. Preincubation with 1 mM ATP had no effect on the basal cyclic GMP level but increased the cycli GMP accumulation in response to cylci AMP (5·10−8 M) or folic acid (5·10−6 M) 40–50%. ATP could not be replaced by ADP of 5′-adenylyliminodiphosphate. Because ATP has no effect on cyclic AMP receptor binding these results indicate that structural membrane alterations (e.g. membrane phosphorylation) may control the transduction of a chemotactic signal.  相似文献   

15.
Chitosan (Ch) was chemically modified with ethylenesulfide (Es) under solvent-free conditions to give (ChEs), displaying a high content of thiol groups due to opening of the three member cyclic reagent. Elemental analysis showed a decrease in nitrogen content. This result indicated the incorporation of two ethylenesulfide molecules for each unit of the polymeric structure of the precursor biopolymer. Infrared spectroscopy, thermogravimetry, and 13C NMR in the solid state demonstrated the effectiveness of the reaction, with signals at 30 ppm for ChEs due to the change in the methylene group environment. Divalent metal uptake by chemically modified biopolymer gave the order Cu > Ni > Co > Zn, reflecting the corresponding acidity of these cations in bonding to the sulfur and the basic nitrogen atoms available on the pendant chains. The equilibrium data were fitted to Freundlich, Temkin, and Langmuir models. The maximum monolayer adsorption capacity for the cations was found to be 1.54 ± 0.02, 1.25 ± 0.03, 1.13 ± 0.01, and 0.83 ± 0.03 mmol g−1, respectively. The Langmuir model best explained the cation–sulfur bond interactions at the solid–liquid interface. The thermodynamics for these interactions gave exothermic enthalpic values of −43.02 ± 0.03, −28.72 ± 0.02, −26.27 ± 0.04, and −17.32 ± 0.02 kJ mol−1, respectively. The spontaneity of the systems is given by negative Gibbs free energies of −31.2 ± 0.1, −32.7 ± 0.1, −31.7 ± 0.1, and −32.2 ± 0.1 kJ mol−1, respectively, in spite of the unfavorable negative entropic values of −39 ± 1, −13 ± 1, −18 ± 1, and −49 ± 1 J K−1 mol−1 due to solvent ordering in the course of complexation. This newly synthesized biopolymer is presented as a chemically useful material for cation removal from aqueous solution.  相似文献   

16.
Suspension cultures of Vitis vinifera were found to produce catechins and stilbenes. When cells were grown in a medium inducing polyphenol synthesis, (−)-epicatechin-3-O-gallate, dimeric procyanidin B-2 3′-O-gallate and two resveratrol diglucosides were isolated, together with a new natural compound that was identified as cis-resveratrol-3,4′-O-β-diglucoside by spectroscopical methods.  相似文献   

17.
The effect of several inhibitors of the enzyme cyclic 3′,5′-AMP phosphodiesterase as chemoattractants in Physarum polycephalum was examined. Of the compounds tested, 4-(3-butoxy-4-methoxybenzyl)-2-imidazolidinone (Roche 20-1724/001) and 1-ethyl-4-(isopropylidinehydrazino)-1H-pyrazolo-(3,4-b)-pyridine-5-carboxylic acid ethyl ester, hydrochloride (Squibb 20009) were the most potent attractants. 3-Isobutyl-1-methyl xanthine, theophylline, and morin (a flavanoid) were moderate attractants and sometimes gave negative chemotaxis at high concentrations. Cyclic 3′,5′-AMP was an effective, but not potent attractant. A repellent effect following the positive chemotactic action was sometimes observed with cyclic 3′,5′-AMP at concentrations as high as 1 · 10−2 M. Dibutyryl cyclic AMP appeared to be a somewhat more potent attractant than cyclic 3′,5′-AMP. The 8-thiomethyl and 8-bromoderivatives of cyclic AMP, which are poorly hydrolyzed by the phosphodiesterase, were not attractants in Physarum. Possible participation of cyclic 3′,5′-AMP in the directional movement in P. polycephalum is discussed.  相似文献   

18.
The Rhodococcus erythropolis strain (N′4) possesses the ability to convert 4-chloro-3-hydroxybutyronitrile into the corresponding acid. This conversion was determined to be performed by its nitrile hydratase and amidase. Ammonium sulfate fractionation, DEAE ion exchange chromatography, and phenyl chromatography were used to partially purify nitrile hydratase from cell-free extract. A SDS-PAGE showed that the partially purified enzyme had two subunits and gel filtration chromatography showed that it consisted of four subunits of α2β2. The purified enzyme had a high specific activity of 860 U mg−1 toward methacrylonitrile. The enzyme was found to have high activity at low temperature range, with a maximum activity occurring at 25 °C and be stable in the presence of organic acids at higher temperatures. The enzyme exhibited a preference for aliphatic saturated nitrile substrates over aliphatic unsaturated or aromatic ones. It was inhibited by sulfhydryl, oxidizing, and serine protease inhibitors, thus indicating that essential cysteine and serine residues can be found in the active site.The purified nitrile hydratase was able to convert 4-chloro-3-hydroxybutyronitrile into the corresponding amide at 15 °C. GC analysis showed that the initial conversion rate of the reaction was 215 mg substrate consumed min−1 mg−1. This demonstrated that this enzyme could be used in conjunction with a stereoselective amidase to synthesize ethyl (S)-4-chloro-3-hydroxybutyrate, an intermediate for a hypercholesterolemia drug, Atorvastatin.  相似文献   

19.
In this study we investigated the ability of Chara intermedia to acclimate to different irradiances (i.e. “low-light” (LL): 20–30 μmol photons m−2 s−1 and “high-light” (HL): 180–200 μmol photons m−2 s−1) and light qualities (white, yellow and green), using morphological, photosynthesis, chlorophyll fluorescence and pigment analysis.Relative growth rates increased with increasing irradiance from 0.016 ± 0.003 (LL) to 0.024 ± 0.005 (HL) g g−1 d−1 fresh weight and were independent of light quality. A growth-based branch orientation towards high-light functioning as a mechanism to protect the plant from excessive light was confirmed. It was shown that the receptor responsible for the morphological reaction is sensitive to blue-light.C. intermedia showed higher oxygen evolution (up to 10.5 (HL) vs. 4.5 (LL) nmol O2 mg Chl−1 s−1), photochemical and energy-dependent Chl fluorescence quenching and a lower Fv/Fm after acclimation to HL. With respect to qP, the acclimation of the photosynthetic apparatus depended on light quality and needed the blue part of the spectrum for full development. In addition, pigment composition was influenced by light and the Chl a/Car and Antheraxanthin (A) + Zeaxanthin (Z)/Violaxanthin (V) + A + Z (DES) ratios revealed the expected acclimation behaviour in favour of carotenoid protection under HL (i.e. decrease of Chl a/Car from 3.41 ± 0.48 to 2.30 ± 0.35 and increase of DES from 0.39 ± 0.05 to 0.87 ± 0.03), while the Chl a/Chl b ratios were not significantly affected. Furthermore it was shown that morphological light acclimation mechanisms influence the extent of the physiological modifications.  相似文献   

20.
This study compared the mass-specific routine metabolic rate (RMR) of similar sized mulloway (Argyrosomus japonicus), a sedentary species, and yellowtail kingfish (Seriola lalandi), a highly active species, acclimated at one of several temperatures ranging from 10–35 °C. Respirometry was carried out in an open-top static system and RMR corrected for seawater–atmosphere O2 exchange using mass-balance equations. For both species RMR increased linearly with increasing temperature (T). RMR for mulloway was 5.78T − 29.0 mg O2 kg− 0.8 h− 1 and for yellowtail kingfish was 12.11T − 39.40 mg O2 kg− 0.8 h− 1. The factorial difference in RMR between mulloway and yellowtail kingfish ranged from 2.8 to 2.2 depending on temperature. The energetic cost of routine activity can be described as a function of temperature for mulloway as 1.93T − 9.68 kJ kg− 0.8 day− 1 and for yellowtail kingfish as 4.04T − 13.14 kJ kg− 0.8 day− 1. Over the full range of temperatures tested Q10 values were approximately 2 for both species while Q10 responses at each temperature increment varied considerably with mulloway and yellowtail kingfish displaying thermosensitivities indicative of each species respective niche habitat. RMR for mulloway was least thermally dependent at 28.5 °C and for yellowtail kingfish at 22.8 °C. Activation energies (Ea) calculated from Arrhenius plots were not significantly different between mulloway (47.6 kJ mol− 1) and yellowtail kingfish (44.1 kJ mol− 1).  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