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.
The purpose of this study was to examine the effects of moment of antagonistic muscle on the resultant joint moment during isokinetic eccentric and concentric efforts of the knee extensors. Ten males performed maximum eccentric and concentric knee extension and flexion efforts on a Biodex dynamometer at 0.52 rad · s−1 (30° · s−1). Electromyographic (EMG) activity of vastus medialis and biceps femoris (hamstrings) was also recorded. The antagonistic moment of the hamstrings was determined by recording the integrated EMG (iEMG)/moment relationship at different levels of muscle effort. The iEMG/moment curves were fitted using second-degree polynomials. The polynomials were then used to predict the antagonistic moment exerted by the hamstrings from the antagonist iEMG. The antagonistic moment had a maximum of 42.92 Nm and 28.97 Nm under concentric and eccentric conditions respectively; paired t-tests indicated that this was a significant difference (P < 0.05). These results indicate that the resultant joint moment of knee extensors is the result of both agonist and antagonist muscle activation. The greater antagonist muscle activity under concentric activation conditions may be partly responsible for the lower resultant joint concentric moment of knee extensors compared with the corresponding eccentric activation. The antagonist moment significantly affects comparisons between the isokinetic moments and agonist EMG and in vitro force measurements under different testing (muscle action and angular velocity) conditions. Accepted: 25 February 1997  相似文献   

2.
The purpose of this study was to compare isokinetic peak torque and the patterns of responses for electromyographic (EMG) and mechanomyographic (MMG), amplitude (AMP) and mean power frequency (MPF) for bilateral (BL) versus unilateral (UL), maximal, isokinetic leg extensions. Eleven recreationally trained women (Mean ± SD: age 22.9 ± 0.9 yrs; body mass 60.5 ± 10.1 kg; height 167.2 ± 6.4 cm) performed 50 maximal, BL and UL isokinetic leg extensions at 60° s−1 on separate days. Electromyographic and MMG signals from the vastus lateralis of the nondominant leg were recorded. Five separate 2 (Condition [BL and UL]) × 10 (Repetitions [5–50]) repeated measures ANOVAs were performed to examine normalized EMG AMP, EMG MPF, MMG AMP, MMG MPF, and isokinetic torque. The results indicated no significant interactions or main effects for EMG AMP and MMG AMP. There were significant interactions for normalized isokinetic peak torque (p < 0.001, η2p = 0.493) and MMG MPF (p = 0.003, η2p = 0.234). For EMG MPF, there was no significant interaction, but significant main effects for Condition (p = 0.003, η2p = 0.607) and Repetitions (p < 0.001, η2p = 0.805). The current findings demonstrated greater performance fatigability for UL than BL leg extensions. Both modalities exhibited similar patterns of neuromuscular responses that were consistent with the Muscular Wisdom hypothesis.  相似文献   

3.
Protein tyrosine phosphatase (PTP) targeted, peptide based chemical probes are valuable tools for studying this important family of enzymes, despite the inherent difficulty of developing peptides targeted towards an individual PTP. Here, we have taken a rational approach to designing a SHP-2 targeted, fluorogenic peptide substrate based on information about the potential biological substrates of SHP-2. The fluorogenic, phosphotyrosine mimetic phosphocoumaryl aminopropionic acid (pCAP) provides a facile readout for monitoring PTP activity. By optimizing the amino acids surrounding the pCAP residue, we obtained a substrate with the sequence Ac-DDPI-pCAP-DVLD-NH2 and optimized kinetic parameters (kcat = 0.059 ± 0.008 s−1, Km = 220 ± 50 µM, kcat/Km of 270 M−1s−1). In comparison, the phosphorylated coumarin moiety alone is an exceedingly poor substrate for SHP-2, with a kcat value of 0.0038 ± 0.0003 s−1, a Km value of 1100 ± 100 µM and a kcat/Km of 3 M−1s−1. Furthermore, this optimized peptide has selectivity for SHP-2 over HePTP, MEG1 and PTPµ. The data presented here demonstrate that PTP-targeted peptide substrates can be obtained by optimizing the sequence of a pCAP containing peptide.  相似文献   

4.
Aquatic exercises are widely implemented into rehabilitation programs. However, both evaluating their mechanical demands on the musculoskeletal system and designing protocols to provide progressive loading are difficult tasks. This study reports for the first time shoulder joint kinetics and dynamics during underwater forward arm elevation performed at speeds ranging from 22.5 to 90°/s. Net joint moments projected onto anatomical axes of rotation, joint power, and joint work were calculated in 18 participants through a novel approach coupling numerical fluid flow simulations and inverse dynamics. Joint dynamics was revealed from the 3D angle between the joint moment and angular velocity vectors, identifying three main functions—propulsion, stabilization, and resistance. Speeds <30°/s necessitated little to no power at all, whereas peaks about 0.20 W⋅kg−1 were seen at 90°/s. As speed increased, peak moments were up to 61 × higher at 90 than at 22.5°/s, (1.82 ± 0.12%BW⋅AL vs 0.03 ± 0.01%BW⋅AL, P < 0.038). This was done at the expense of a substantial decrease in the joint moment contribution to joint stability though, which goes against the intuition that greater stabilization is required to protect the shoulder from increasing loads. Slow arm elevations (<30°/s) are advantageous for joint mobility gain at low mechanical solicitation, whereas the intensity at 90°/s is high enough to stimulate muscular endurance improvements. Simple predictive equations of shoulder mechanical loading are provided. They allow for easy design of progressive protocols, either for the postoperative shoulder or the conditioning of athlete targeting very specific intensity regions.  相似文献   

5.
This study tested the hypothesis that the effect of hip joint angle on concentric knee extension torque depends on knee joint angle during a single knee extension task. Twelve men performed concentric knee extensions in fully extended and 80° flexed hip positions with maximal effort. The angular velocities were set at 30° s−1 and 180° s−1. The peak torque and torques attained at 30°, 50°, 70° and 90° (anatomical position = 0°) of the knee joint were compared between the two hip positions. Muscle activations of the vastus lateralis, medialis, rectus femoris and biceps femoris were determined using surface electromyography. The peak torque was significantly greater in the flexed than in the extended hip position irrespective of angular velocity. The torques at 70° and 90° of the knee joint at both angular velocities and at 50° at 180° s−1 were significantly greater in the flexed than in the extended hip position, whereas corresponding differences were not found at 30° (at either angular velocity) and 50° (at 30° s−1) of the knee joint. No effect of hip position on muscle activation was observed in any muscle. These results supported our hypothesis and may be related to the force–length and force–velocity characteristics of the rectus femoris.  相似文献   

6.
A technique based on homogenisation of rapidly frozen tissue was used to investigate the regulation of intracellular pH (pHi) in freshwater and marine fish from diverse environmental temperatures. The following species were held at ambient temperatures of ca. 1°C (Notothenia coriiceps; Antarctica), 5°C (Pleuronectes platessa, Myoxocephalus scorpius; North Sea), and 26°C (Oreochromis niloticus; African lakes). The effects of seasonal acclimatisation to 4, 11 and 18°C were also examined in rainbow trout in the winter, autumn and summer, respectively. Extracellular (whole blood) pH (pHe) did not follow the constant relative alkalinity relationship, where pH+=pOH for any particular temperature, over a range of 1–26°C (overall δpHeT=0.009±0.002 U °C−1; P<0.001), apparently being regulated by ionic fluxes and ventilation. Intracellular pH (pHi) was also regulated independently of pN(=0.5 pK water) in all species of fish examined. The inverse relationship between pHi and environmental temperature gave an overall δpHiT of −0.010±0.001 U °C−1 (for both white and red muscle) and −0.004±0.003 U °C−1 (cardiac muscle). However, between 1 and 11°C δpHiT was much higher (P<0.001), −0.022±0.003 U °C−1 (white muscle) and −0.022±0.004 U °C−1 (red muscle). The possible adaptive roles for these different acid–base responses to environmental temperature variation among tissues and species, and the potential difficulties of estimating pHi, are discussed.  相似文献   

7.
Muscle force production and power output in active males, regardless of the site of measurement (hand, leg, or back), are higher in the evening than the morning. This diurnal variation is attributed to motivational, peripheral, and central factors and higher core and, possibly, muscle temperatures in the evening. This study investigated whether decreasing evening resting rectal temperatures to morning values, by immersion in a water tank, leads to muscle force production and power output becoming equal to morning values in motivated subjects. Ten healthy active males (mean?±?SD: age, 22.5?±?1.3 yrs; body mass, 80.1?±?7.8?kg; height, 1.72?±?0.05?m) completed the study, which was approved by the local ethics committee of the university. The subjects were familiarized with the techniques and protocol and then completed three sessions (separated by at least 48?h): control morning (07:30?h) and evening (17:30?h) sessions (with an active 5-min warm-up on a cycle ergometer at 150?W) and then a further session at 17:30?h but preceded by an immersion in cold water (~16.5?°C) to lower rectal temperature (Trec) to morning values. During each trial, three measures of grip strength, isokinetic leg strength measurements (of knee flexion and extension at 1.05 and 4.19?rad?s?1 through a 90° range of motion), and three measures of maximal voluntary contraction (MVC) on an isometric dynamometer (utilizing the twitch-interpolation technique) were performed. Trec, rating of perceived exertion (RPE), and thermal comfort (TC) were also measured after the subjects had reclined for 30?min at the start of the protocol and prior to the measures for grip, isokinetic, and isometric dynamometry. Muscle temperature was taken after the warm-up or water immersion and immediately before the isokinetic and MVC measurements. Data were analyzed using general linear models with repeated measures. Trec values were higher at rest in the evening (by 0.37?°C; p?<?0.05) than the morning, but values were no different from morning values immediately after the passive pre-cooling. However, Trec progressively decreased throughout the experiments, this being reflected in the subjects’ ratings of thermal comfort. Muscle temperatures also displayed significant diurnal variation, with higher values in the evening (by 0.39?°C; p?<?0.05). Right grip strength, isometric peak power, isokinetic knee flexion and extension for peak torque and peak power at 1.05?rad?s?1, and knee extension for peak torque at 4.19?rad?s?1 all showed higher values in the evening (a range of 3–14%), and all other measures of strength or power showed a statistical trend to be higher in the evening (0.10?>?p?>?0.05). Pre-cooling in the evening significantly reduced force or power variables towards morning values. In summary, effects of time of day were seen in some measures of muscle performance, in agreement with past research. However, in this population of motivated subjects, there was evidence that decreasing evening Trec to morning values by coldwater immersion decreased muscle strength to values similar to those found in the morning. It is concluded that diurnal changes in muscle performance are linked to diurnal changes in Trec. (Author correspondence: B.J.Edwards@ljmu.ac.uk)  相似文献   

8.
Zhang F W  Liu A H  Li Y N  Zhao L  Wang Q X  Du M Y 《农业工程》2008,28(2):453-462
Using the CO2 flux data measured by the eddy covariance method in the northeast of Qinghai-Tibetan Plateau in 2005, we analyzed the carbon flux dynamics in relation to meteorological and biotic factors. The results showed that the alpine wetland ecosystem was the carbon source, and it emitted 316.02 gCO2 · m−2 to atmosphere in 2005 with 230.16 gCO2 · m−2 absorbed in the growing season from May to September and 546.18 gCO2 · m−2 released in the non-growing season from January to April and from October to December. The maximum of the averaged daily CO2 uptake rates and release rates was (0.45 ± 0.0012) mgCO2 · m−2 · s−1 (Mean ± SE) in July and (0.22 ± 0.0090) mgCO2 · m−2 · s−1 in August, respectively. The averaged diurnal variation showed a single-peaked pattern in the growing season, but exhibited very small fluctuation in the non-growing season. Net ecosystem exchange (NEE) and gross primary production (GPP) were all correlated with some meteorological factors, and they showed a negatively linear correlation with aboveground biomass, while a positive correlation existed between the ecosystem respiration (Res) and those factors.  相似文献   

9.
10.
In humans, an inhibitory via Ia afferent pathway from the medial gastrocnemius (MG) to the soleus (SOL) motoneuron pool has been suggested. Herein, we examined the relation between MG fascicle length changes and the SOL H-reflex modulation during passive knee movement. Twelve subjects performed static and passive (5° s?1) knee movement tasks with the ankle immobilized using an isokinetic dynamometer in sitting posture. The maximal H- and M-waves were measured at four target angles (20°, 40°, 60°, and 80° flexion from full knee extension). The MG fascicles length and velocity were measured using a B-mode ultrasonic apparatus. Results demonstrated that the SOL Hmax/Mmax; i.e., ratio of the maximal H- to M-waves, was attenuated with increasing MG fascicle length in static tasks. The SOL Hmax/Mmax at 20° was significantly attenuated compared with 60° and 80° with increasing MG fascicle length and lengthening velocity in passive knee extension. However, no significant differences in the SOL Hmax/Mmax were found across the target angles in the passive knee flexion task. In conclusion, as muscle spindles increase their discharge with lengthening fascicle velocity, but keep silent when fascicles shorten, our data suggest that lengthening the MG facilitates an inhibitory Ia pathway from MG to SOL, and modulates SOL motoneuron activity during movements.  相似文献   

11.
β-Mannanase is the key enzyme in the hydrolysis of mannan which has been widely applied in diverse industrial fields such as biobleaching pulps, food and feed industry, bioethanol and pharmaceutical applications. In this study, a novel GH5 family β-mannanase gene (LrMan5B) with 381 amino acid residues was identified from Lichtheimia ramosa, and highly expressed in Pichia pastoris X33. The amino acid sequence shares the highest identity (64%) with the β-mannanase from Rhizomucor miehei. Purified recombinant LrMan5B showed the optimal activity at pH 5.0 and 65 °C. It had broad-range pH stability (retaining >65% activity after incubation at pH 3.0–8.0 at 37 °C for 24 h) and was highly thermostable (retaining >80% activity after incubation at 60 °C for 30 min). LrMan5B displayed the highest catalytic efficiency for locust bean gum and the kcat/Km value was 1357.47 mL·mg−1·s−1, followed by guar gum (512.82 mL·mg−1·s−1), konjac glucomannan (454.21 mL·mg−1·s−1), and palm kernel meal (137.00 mL·mg−1·s−1). In order to evaluate the synergistic effect of LrMan5B and α-galactosidase LrAgal36A from L. ramosa, LrAgal36A was supplemented to hydrolyze palm kernel meal with LrMan5B together, showing that the reducing sugar release significantly increased by 21% (compared with the sum of that by hydrolysis of single Lrman5B or LrAgal36A). Due to its favorable enzymatic properties, LrMan5B might own potential applications in the area of food and feed processing.  相似文献   

12.
Until now the reproducibility of the short latency stretch reflex of the internal rotator muscles of the glenohumeral joint has not been identified. Twenty-three healthy male participants performed three sets of external shoulder rotation stretches with various pre-activation levels on two different dates of measurement to assess test–retest reliability. All stretches were applied with a dynamometer acceleration of 104°/s2 and a velocity of 150°/s. Electromyographical response was measured via surface EMG. Reflex latencies showed a pre-activation effect (ƞ2 = 0,355). ICC ranged from 0,735 to 0,909 indicating an overall “good” relative reliability. SRD 95% lay between ±7,0 to ±12,3 ms. The reflex gain showed overall poor test–retest reproducibility. The chosen methodological approach presented a suitable test protocol for shoulder muscles stretch reflex latency evaluation. A proof-of-concept study to validate the presented methodical approach in shoulder involvement including subjects with clinically relevant conditions is recommended.  相似文献   

13.
We determined the maximum sustained swimming speed (Ucrit), and resting and maximum ventilation rates of the Antarctic fish Pagothenia borchgrevinki at five temperatures between −1°C and 8°C. We also determined resting metabolic rate (VO2) at −1°C, 2°C, and 4°C. Ucrit of P. borchgrevinki was highest at −1°C (2.7±0.1 BL s−1) and rapidly decreased with temperature, representing a thermal performance breadth of only 5°C. This narrow thermal performance supports our prediction that specialisation to the subzero Antarctic marine environment is associated with a physiological trade-off in performance at high temperatures. Resting oxygen consumption and ventilation rate increased by more than 200% across the temperature range, which most likely contribute to the decrease in aerobic swimming capabilities at higher temperatures.  相似文献   

14.
《BBA》2023,1864(3):148973
The mechanisms underlying cigarette smoke-induced mitochondrial dysfunction in skeletal muscle are still poorly understood. Accordingly, this study aimed to examine the effects of cigarette smoke on mitochondrial energy transfer in permeabilized muscle fibers from skeletal muscles with differing metabolic characteristics. The electron transport chain (ETC) capacity, ADP transport, and respiratory control by ADP were assessed in fast- and slow-twitch muscle fibers from C57BL/6 mice (n = 11) acutely exposed to cigarette smoke concentrate (CSC) using high-resolution respirometry. CSC decreased complex I-driven respiration in the white gastrocnemius (CONTROL:45.4 ± 11.2 pmolO2.s−1.mg−1 and CSC:27.5 ± 12.0 pmolO2.s−1.mg−1; p = 0.01) and soleus (CONTROL:63.0 ± 23.8 pmolO2.s−1.mg−1 and CSC:44.6 ± 11.1 pmolO2.s−1.mg−1; p = 0.04). In contrast, the effect of CSC on Complex II-linked respiration increased its relative contribution to muscle respiratory capacity in the white gastrocnemius muscle. The maximal respiratory activity of the ETC was significantly inhibited by CSC in both muscles. Furthermore, the respiration rate dependent on the ADP/ATP transport across the mitochondrial membrane was significantly impaired by CSC in the white gastrocnemius (CONTROL:-70 ± 18 %; CSC:-28 ± 10 %; p < 0.001), but not the soleus (CONTROL:47 ± 16 %; CSC:31 ± 7 %; p = 0.08). CSC also significantly impaired mitochondrial thermodynamic coupling in both muscles. Our findings underscore that acute CSC exposure directly inhibits oxidative phosphorylation in permeabilized muscle fibers. This effect was mediated by significant perturbations of the electron transfer in the respiratory complexes, especially at complex I, in both fast and slow twitch muscles. In contrast, CSC-induced inhibition of the exchange of ADP/ATP across the mitochondrial membrane was fiber-type specific, with a large effect on fast-twitch muscles.  相似文献   

15.
This study examines the hypothesis that Mystromys albicaudatus, a cricetine relic in southern Africa, has thermal characteristics typical of a rodent adapted to a cold temperature regime. Metabolic rate (oxygen consumption) of M. albicaudatus was measured using open-flow respirometry at ambient temperatures ranging from 5°C to 35°C. Lowest specific oxygen consumption was 1.352 ± 0.089 ml O2 g−1hr−1 (n = 8; body mass = 93.78 ± 6.27 g) at 25°C, equivalent to 121.8% of the predicted value of Kleiber (1975), 128.8% of the value predicted for eutherians and 113.7% of the value predicted for cricetidae (Hayssen and Lacy, 1985).  相似文献   

16.
We have used inelastic laser light scattering to study the kinetics of the spontaneous assembly of heads and tails of bacteriophage T4D to form noninfectious tail fiberless particles. For interpretation of the kinetics, it was first necessary to determine the physical properties of the strongly scattering phage parts. For heads, these are D20,w = 3.60 × 10−8cm2/s, 820,w = 1025 S, M = 1.76 × 108. For tail fiberless particles, D20,w = 3.14 × 10−8cm2/s, 820,w = 968 S, and M = 1.95 × 108. The kinetics of the head-tail joining process was followed by measuring the time variation of the homodyne scattering autocorrelation function. This was interpreted as a sum of exponentials whose decay constants were known from the scattering angle and the diffusion coefficients, and whose amplitudes were related to the concentrations of reactants and products. Scattering experiments at 22 °C gave a bimolecular rate constant of 1.02 × 107m−1 s−1, while infectivity assays at 30 °C gave a rate constant of 1.28 × 107. Adjustment of both rate constants to 20 °C, assuming diffusion controlled reaction, gave 0.97 × 107 and 0.98 × 107m−1 s−1, respectively. This rate is about 1500 that predicted by Smoluchowski theory for a diffusion controlled reaction between two spherical particles; the discrepancy is largely explicable from orientational factors.  相似文献   

17.
The reactions of NO2 with both oxidized and reduced cytochrome c at pH 7.2 and 7.4, respectively, and with N-acetyltyrosine amide and N-acetyltryptophan amide at pH 7.3 were studied by pulse radiolysis at 23 °C. NO2 oxidizes N-acetyltyrosine amide and N-acetyltryptophan amide with rate constants of (3.1±0.3)×105 and (1.1±0.1)×106 M−1 s−1, respectively. With iron(III)cytochrome c, the reaction involves only its amino acids, because no changes in the visible spectrum of cytochrome c are observed. The second-order rate constant is (5.8±0.7)×106 M−1 s−1 at pH 7.2. NO2 oxidizes iron(II)cytochrome c with a second-order rate constant of (6.6±0.5)×107 M−1 s−1 at pH 7.4; formation of iron(III)cytochrome c is quantitative. Based on these rate constants, we propose that the reaction with iron(II)cytochrome c proceeds via a mechanism in which 90% of NO2 oxidizes the iron center directly—most probably via reaction at the solvent-accessible heme edge—whereas 10% oxidizes the amino acid residues to the corresponding radicals, which, in turn, oxidize iron(II). Iron(II)cytochrome c is also oxidized by peroxynitrite in the presence of CO2 to iron(III)cytochrome c, with a yield of ~60% relative to peroxynitrite. Our results indicate that, in vivo, NO2 will attack preferentially the reduced form of cytochrome c; protein damage is expected to be marginal, the consequence of formation of amino acid radicals on iron(III)cytochrome c.  相似文献   

18.
The enthalpies of the hexokinase-catalyzed phosphorylation or glucose, mannose, and fructose by ATP to the respective hexose 6-phosphates have been measured calorimetrically in TRIS/TRIS HCl buffer at 25.0, 28.5, and 32.0°C. The effects on the measured enthalpy of the glucose/hexokinase reaction due to variation of pH (over the range 6.7 to 9.0) and ionic strength (over the range 0.02 to 0.25) have been examined. Correction for enthalpy of buffer protonation leads to δHo and δCpo values for the processes: eq-D-hexose + ATP4− = eq-D-hexose 6-phosphate2− + ADP3−+ H+. Results are δHo = −23.8 ± 0.7 kJ · mol−1 and δCpo = −156 ± 280 J·mol−1·K−1 for glucose. δHo = −21.9 ± 0.7 kJ·mol−1 and δCpo = 10 ± 140 J·mol−1·K−1 for mannose, and δHo = −15.0 ± 0.9 kJ·mol−1 and δCpo = −41 ± 160 J·mol−1·K−1 for fructose. Combination of these measured enthalpies with Gibbs energy data for hydrolysis of ATP4− and that for the hexose 6-phosphates lead to δSo values for the above hexokinase-catalyzed reactions.  相似文献   

19.
Analkali tolerant α-l-rhamnosidase has been purified to homogeneity from the culture filtrate of a new fungal strain, Fusarium moniliforme MTCC-2088, using concentration by ultrafiltration and cation exchange chromatography on CM cellulose column. The molecular mass of the purified enzyme has been found to be 36.0 kDa using SDS-PAGE analysis. The Km value using p-nitrophenyl-α-l-rhamnopyranoside as the variable substrate in 0.2 M sodium phosphate buffer pH10.5 at50 °C was 0.50 mM. The catalytic rate constant was15.6 s−1giving the values of kcat/Km is 3.12 × 104M−1 s−1. The pH and temperature optima of the enzyme were 10.5 and 50 °C, respectively. The purified enzyme had better stability at 10 °C in basic pH medium. The enzyme derhamnosylated natural glycosides like naringin to prunin, rutin to isoquercitrin and hesperidin to hesperetin glucoside. The purified α-l-rhamnosidase has potential for enhancement of wine aroma.  相似文献   

20.
Organisms employ a wide array of physiological and behavioral responses in an effort to endure stressful environmental conditions. For many marine invertebrates, physiological and/or behavioral performance is dependent on physical conditions in the fluid environment. Although factors such as water temperature and velocity can elicit changes in respiration and feeding, the manner in which these processes integrate to shape growth remains unclear. In a growth experiment, juvenile barnacles (Balanus glandula) were raised in dockside, once-through flow chambers at water velocities of 2 versus 19 cm s−1 and temperatures of 11.5 versus 14 °C. Over 37 days, growth rates (i.e., shell basal area) increased with faster water velocities and higher temperatures. Barnacles at high flows had shorter feeding appendages (i.e., cirri), suggesting that growth patterns are unlikely related to plastic responses in cirral length. A separate experiment in the field confirmed patterns of temperature- and flow-dependent growth over 41 days. Outplanted juvenile barnacles exposed to the faster water velocities (32±1 and 34±1 cm s−1; mean±SE) and warm temperatures (16.81±0.05 °C) experienced higher growth compared to individuals at low velocities (1±1 cm s−1) and temperatures (13.67±0.02 °C). Growth data were consistent with estimates from a simple energy budget model based on previously measured feeding and respiration response curves that predicted peak growth at moderate temperatures (15 °C) and velocities (20–30 cm s−1). Low growth is expected at both low and high velocities due to lower encounter rates with suspended food particles and lower capture efficiencies respectively. At high temperatures, growth is likely limited by high metabolic costs, whereas slow growth at low temperatures may be a consequence of low oxygen availability and/or slow cirral beating and low feeding rates. Moreover, these results advocate for approaches that consider the combined effects of multiple stressors and suggest that both increases and decreases in temperature or flow impact barnacle growth, but through different physiological and behavioral mechanisms.  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