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.
To examine the relationship between mitochondrial NADH (NADH(m)) and cardiac work output, NADH(m) and the amplitude and frequency of the contractile response of electrically paced rat heart cells were measured at 25 degrees C. With 5.4 mM glucose plus 2 mM beta-hydroxybutyrate, NADH(m) was reversibly decreased by 23%, and the amplitude of contraction was reversibly decreased by 27% during 4-Hz pacing. With glucose plus 2 mM pyruvate or with 10 mM 2-deoxy-D-glucose, NADH(m) was maintained during rapid pacing, and the contractile amplitude remained high. Phosphocreatine levels decreased with 2-deoxy-D-glucose administration but not with rapid pacing. Respiration increased to meet the increased ATP demand at 30 degrees C. The data suggest that 1) when NADH(m) is decreased during rapid pacing with defined substrates, the amplitude of contraction is decreased; 2) the amplitude of contraction during electrical pacing does not change with rate of pacing when both the ATP and NADH(m) levels are continuously replenished; and 3) the replenishment of NADH(m) during pacing with physiological substrates may be rate-limited by substrate supply to mitochondrial dehydrogenases. During activation of mitochondrial dehydrogenases, or a significant increase in free ADP induced by 2-deoxy-D-glucose, this rate limitation is bypassed or overcome.  相似文献   

2.
It was shown that the temperature sensitivity of shortening velocity of skeletal muscles is higher at temperatures below physiological (10-25 degrees C) than at temperatures closer to physiological (25-35 degrees C) and is higher in slow than fast muscles. However, because intact muscles invariably express several myosin isoforms, they are not the ideal model to compare the temperature sensitivity of slow and fast myosin isoforms. Moreover, temperature sensitivity of intact muscles and single muscle fibers cannot be unequivocally attributed to a modulation of myosin function itself, as in such specimen myosin works in the structure of the sarcomere together with other myofibrillar proteins. We have used an in vitro motility assay approach in which the impact of temperature on velocity can be studied at a molecular level, as in such assays acto-myosin interaction occurs in the absence of sarcomere structure and of the other myofibrillar proteins. Moreover, the temperature modulation of velocity could be studied in pure myosin isoforms (rat type 1, 2A, and 2B and rabbit type 1 and 2X) that could be extracted from single fibers and in a wide range of temperatures (10-35 degrees C) because isolated myosin is stable up to physiological temperature. The data show that, at the molecular level, the temperature sensitivity is higher at lower (10-25 degrees C) than at higher (25-35 degrees C) temperatures, consistent with experiments on isolated muscles. However, slow myosin isoforms did not show a higher temperature sensitivity than fast isoforms, contrary to what was observed in intact slow and fast muscles.  相似文献   

3.
The main goal of this study is to investigate the role of mitochondrial [Ca(2+)], [Ca(2+)](m), in the possible up-regulation of the NADH production rate during increased workload. Such up-regulation is necessary to support increased flux through the electron transport chain and increased ATP synthesis rates. Intact cardiac trabeculae were loaded with Rhod-2(AM), and [Ca(2+)](m) and mitochondrial [NADH] ([NADH](m)) were simultaneously measured during increased pacing frequency. It was found that 53% of Rhod-2 was localized in mitochondria. Increased pacing frequency caused a fast, followed by a slow rise of the Rhod-2 signal, which could be attributed to an abrupt increase in resting cytosolic [Ca(2+)], and a more gradual rise of [Ca(2+)](m), respectively. When the pacing frequency was increased from 0.25 to 2 Hz, the slow Rhod-2 component and the NADH signal increased by 18 and 11%, respectively. Based on a new calibration method, the 18% increase of the Rhod-2 signal was calculated to correspond to a 43% increase of [Ca(2+)](m). There was also a close temporal relationship between the rise (time constant approximately 25 s) and fall (time constant approximately 65 s) of [Ca(2+)](m) and [NADH](m) when the pacing frequency was increased and decreased, respectively, suggesting that increased workload and [Ca(2+)](c) cause increased [Ca(2+)](m) and consequently up-regulation of the NADH production rate.  相似文献   

4.
Our understanding of plant responses to drought has improved over the decades. However, the importance of the rate of drought imposition on the response is still poorly understood. To test the importance of the rate at which drought stress develops, whole-plant photosynthesis (P(net) ), respiration (R(dark) ), daily carbon gain (DCG), daily evapotranspiration (DET) and water use efficiency (WUE) of vinca (Catharanthus roseus), subjected to different drought imposition rates, were investigated. We controlled the rate at which the substrate dried out with an automated irrigation system that allowed pot weight to decrease gradually throughout the drying period. Fast, intermediate and slow drying treatments reached their final pot weight [500 g, substrate water content (θ) ≈ 0.10 m3 m(-3) ] after 3.1, 6.6 and 10 days, respectively. Although all drying treatments decreased P(net) and R(dark) , slow drying reduced P(net) and R(dark) less than fast drying. At a θ < 0.10 m3 m(-3) , DCG and DET in the slow drying treatment were reduced by ≈50%, whereas DCG and DET in the fast drying treatment were reduced by 85 and 70% at a θ of 0.16 m(3) m(-3) . Plants exposed to slow drought imposition maintained a high WUE, even at θ < 0.10 m3 m(-3) . Overall, physiological responses to low θ were less severe in plants subjected to slow drying as compared with fast drying, even though the final θ was lower for plants exposed to slow drying. This suggests that the rate at which drought stress develops has important implications for the level of acclimation that occurs.  相似文献   

5.
Takahashi K  Klinman JP 《Biochemistry》2006,45(14):4683-4694
The expression of a copper amine oxidase (CAO) from Hansenula polymorpha in Saccharomyces cerevisiae under differing culture conditions leads to the incorporation of varied levels of CAO-bound zinc. The presence of substantial amount of zinc results in two distinctive enzyme species, designated as the fast and slow enzymes. Both forms are rapidly reduced by substrate methylamine with a rate constant of 199 s(-1) but behave remarkably differently in their oxidation rates; the fast enzyme is oxidized by dioxygen at a rate of 22.1 s(-1), whereas the slow enzyme reacts at a rate of 1.8 x 10(-4) s(-1). The apparent kcat of the enzyme preparation is linearly proportional to the fraction of the fast enzyme, with an extrapolated value of 6.17 s(-1) when the enzyme is 100% in its "fast" form. A comparison of rate constants for cofactor reduction and reoxidation steps, measured in stopped flow experiments, to the extrapolated kcat implicates additional steps in the steady state reaction. Measurement of the proportion of oxidized (ETPQ(ox)) and reduced cofactor (ETPQ(red)) under steady state conditions indicates approximately 50% of each cofactor form at 0.8 or 2 mM methylamine. Kinetic isotope effect measurements using deuterated amine substrate lead to the following steady state values: (D)(k(red)) = 8.5 (0.5), (D)(kcat) = 1.7 (0.1), and (D)(kcat/K(m)) = 4.3 (0.2). The collective data allow the calculation of partially rate-determining constants during the reductive half-reaction (ca. 200 s(-1) for binding of substrate to ETPQ(ox) and 27.9 s(-1) for release of aldehyde product or a protein isomerization from ETPQ(red)); an additional step with a rate constant of 13.2 s(-1) is assigned to the oxidative half-reaction, most likely for the release of product hydrogen peroxide. These results, together with the sole detection of oxidized and reduced cofactor during rapid scanning stopped flow experiments, indicate that four steps contribute to kcat, with the first electron transfer from cofactor to O2 contributing ca. 29%. An investigation of the relationship between the copper content and the extent of the fast enzyme shows that only the copper-containing homodimer is capable of rapid reoxidation and that zinc-copper heterodimers are incapable of rapid turnover at either subunit. This implies communication between the metal sites of the two subunits per dimer that impacts O2 binding and/or electron transfer from reduced cofactor to bound O2.  相似文献   

6.
We have examined the rate of dissociation of internalized 125I-asialo-orosomucoid-receptor complexes in freshly isolated rat hepatocytes. Cell suspensions were washed with ethylene glycol bis (beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid at 0 degrees C to remove surface-bound ligand and then assessed for the retention of radioactive glycoprotein in the presence of digitonin, which permeabilized the cells and released the internal soluble contents. In cells which initially contained only surface-bound ligand, about 50% of the internalized ligand dissociated from receptor very rapidly (t1/2 less than or equal to 2.5 min, k greater than or equal to 0.28 min-1), at 37 degrees C, whereas the other 50% dissociated more slowly with apparent first order kinetics (t1/2 = 50 min, k = 0.014 min-1). This equal distribution of internalized ligand into two compartments, from which dissociation occurred with very different kinetics, did not depend on the extent of surface receptor occupancy and also occurred under non-steady state conditions of continuous exposure to ligand. Ligand entering both the rapid and slow dissociation compartments was eventually degraded with apparent first order kinetics (k = 0.0047 min-1), suggesting that the intracellular routing of ligand to lysosomes after dissociation from either compartment was via the same pathway. The fast and slow dissociation of receptor-ligand complexes were also distinguished by different temperature sensitivities; the slow dissociation process ceased below 18 degrees C, whereas the fast dissociation process still proceeded. The equal partition of internalized complexes into the two kinetic compartments did not change as a function of temperature but did change as cells continued to endocytose asialo-orosomucoid at 37 degrees C. As the internal receptor pool approached a steady state level of occupancy, there was an increase in the average time for receptor recycling and an increase in the fraction of incoming receptor-ligand complexes which dissociated rapidly (approximately 75%). In addition, under steady state conditions, the rate of the slow dissociation process increased (k = 0.026 min-1, t1/2 = 27 min).  相似文献   

7.
Twenty young men trained the right knee extensors and flexors on an isokinetic dynamometer three times weekly over a 10-week period. During each session, 10 men in the slow training group completed three sets of 8 maximal contractions at a rate of 1.05 rad s-1, whereas the other 10, the fast group, completed three sets of 20 contractions at a rate of 4.19 rad s-1. Subjects wer pre- and post-tested for peak torque and power on an isokinetic dynamometer at 1.05, 3.14, and 4.19 rad s-1. Proportions of muscle fibre-types and fibre cross-sectional areas were determined from biopsy specimens taken before and after training from the right vastus lateralis. When testing was conducted at 1.05 rad s-1, the slow group improved (P less than 0.05) peak torque by 24.5 N m (8.5%), but no change was noted for the fast group. Power increased (P less than 0.05) by 32.7 W (13.6%) in the slow group and 5.5 W (2.5%) in the fast. At 3.14 rad s-1, both groups increased (P less than 0.05) peak torque and power. At 4.19 rad s-1, the fast group increased (P less than 0.05) peak torque by 30.0 N m (19.7%), whereas no training effect was observed in the slow group. There was no significant change in power in either group at 4.19 rad s-1. No significant changes were observed over the 10-week training period in percentages of type I, IIa and IIb fibres, but both groups showed significant increases (P less than 0.05) in type I and IIa fibre areas.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The Huxley-Simmons phase 2 controls the kinetics of the first stages of tension recovery after a step-change in fiber length and is considered intimately associated with tension generation. It had been shown that phase 2 is comprised of two distinct unrelated phases. This is confirmed here by showing that the properties of phase 2(fast) are independent of fiber type, whereas those of phase 2(slow) are fiber type dependent. Phase 2(fast) has a rate of 1000-2000 s(-1) and is temperature insensitive (Q(10) approximately 1.16) in fast, medium, and slow speed fibers. Regardless of fiber type and temperature, the amplitude of phase 2(fast) is half (approximately 0.46) that of phase 1 (fiber instantaneous stiffness). Consequently, fiber compliance (cross-bridge and thick/thin filament) appears to be the common source of both phase 1 elasticity and phase 2(fast) viscoelasticity. In fast fibers, stiffness increases in direct proportion to tension from an extrapolated positive origin at zero tension. The simplest explanation is that tension generation can be approximated by two-state transition from attached preforce generating (moderate stiffness) to attached force generating (high stiffness) states. Phase 2(slow) is quite different, progressively slowing in concert with fiber type. An interesting interpretation of the amplitude and rate data is that reverse coupling of phase 2(slow) back to P(i) release and ATP hydrolysis appears absent in fast fibers, detectable in medium speed fibers, and marked in slow fibers contracting isometrically. Contracting slow and heart muscles stretched under load could employ this enhanced reversibility of the cross-bridge cycle as a mechanism to conserve energy.  相似文献   

9.
10.
The estimated lower limb length (0.761–0.793 m) of the partial skeleton of Australopithecus afarensis from Woranso‐Mille (KSD‐VP‐1/1) is outside the previously known range for Australopithecus and within the range of modern humans. The lower limb length of KSD‐VP‐1/1 is particularly intriguing when juxtaposed against the lower limb length estimate of the other partial skeleton of A. afarensis, AL 288‐1 (0.525 m). A sample of 36 children (age, >7 years, trochanteric height = 0.56–0.765 m) and 16 adults (trochanteric height = 0.77–1.00 m) walked at their self‐selected slow, preferred, and fast walking velocities, while their oxygen consumption was monitored. Lower limb length and velocity were correlated with slow (P < 0.001, r2 = 0.44), preferred (P < 0.001, r2 = 0.55), and fast (P < 0.001, r2 = 0.69) walking velocity. The relationship between optimal velocity and lower limb length was also determined and lower limb length explained 47% of the variability in optimal velocity. The velocity profile for KSD‐VP‐1/1 (slow = 0.73–0.75 m/s, preferred = 1.08–1.11 m/s, and fast = 1.48–1.54 m/s) is 36–44% higher than that of AL 288‐1 (slow = 0.53 m/s, preferred = 0.78 m/s, and fast = 1.07 m/s). The optimal velocity for AL 288‐1 is 1.04 m/s, whereas that for KSD‐VP‐1/1 is 1.29–1.33 m/s. This degree of lower limb length dimorphism suggests that members of a group would have had to compromise their preferences to walk together or to split into subgroups to walk at their optimal velocity. Am J Phys Anthropol, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

11.
Babesia rodhaini parasites in murine blood containing 1.5 m DMSO were frozen at two rates, as judged by the duration of the “freezing plateau”, then cooled to ?196 °C and rewarmed at two rates to detect interactions between the duration of the plateau and rates of subsequent cooling and rewarming. Infectivity tests showed that fast and slow freezing (plateau times of about 1 sec and 30 sec, respectively) had similar effects on parasite survival when cooling was at 130 °C/min and warming was at 800 °C/min. However, when either the cooling rate was increased to 3500 °C/min or the warming rate was decreased to 2.3 °C/min, fast freezing decreased parasite survival more than did slow freezing. It is suggested that fast freezing accentuated the damaging effects of fast cooling and slow warming by increasing intracellular ice formation.  相似文献   

12.
Hypertrophy and failure (H/F) in humans and large mammals are characterized by a change from a positive developed force-frequency relationship (+FFR) in normal myocardium to a flattened or negative developed force-frequency relationship (-FFR) in disease. Altered Ca(2+) homeostasis underlies this process, but the role of intracellular Na(+) concentration ([Na(+)](i)) in H/F and frequency-dependent contractility reserve is unclear. We hypothesized that altered [Na(+)](i) is central to the -FFR response in H/F feline myocytes. Aortic constriction caused left ventricular hypertrophy (LVH). We found that as pacing rate was increased, contraction magnitude was maintained in isolated control myocytes (CM) but decreased in LVH myocytes (LVH-M). Quiescent LVH-M had higher [Na(+)](i) than CM (LVH-M 13.3 +/- 0.3 vs. CM 8.9 +/- 0.2 mmol/l; P < 0.001) with 0.5-Hz pacing (LVH-M 14.9 +/- 0.5 vs. CM 10.8 +/- 0.4 mmol/l; P < 0.001) but were not different at 2.5 Hz (17.0 +/- 0.7 vs. control 16.0 +/- 0.7 mmol/l; not significant). [Na(+)](i) was altered by patch pipette dialysis to define the effect of [Na(+)](i) on contraction magnitude and action potential (AP) wave shape at slow and fast pacing rates. Using AP clamp, we showed that LVH-M require increased [Na(+)](i) and long diastolic intervals to maintain normal shortening. Finally, we determined the voltage dependence of contraction for Ca(2+) current (I(Ca))-triggered and Na(+)/Ca(2+) exchanger-mediated contractions and showed that there is a greater [Na(+)](i) dependence of contractility in LVH-M. These data show that increased [Na(+)](i) is essential for maintaining contractility at slow heart rates but contributes to small contractions at fast rates unless rate-dependent AP shortening is prevented, suggesting that altered [Na(+)](i) regulation is a critical contributor to abnormal contractility in disease.  相似文献   

13.
This study examined multifiber baroreceptor nerve activity (BNA) as a function of carotid sinus wall distension in 19 rabbits. Analysis estimated mechanical or viscoelastic properties of the sinus wall and their influence on BNA. In six sinuses, properties were altered by treatment with the enzyme protease to remove the endothelium and with nifedipine to passively relax smooth muscle. Properties were estimated from dynamic and steady state wall response to a 45 mm Hg step increase and decrease in intrasinus pressure (ISP) of 20 min. Control wall response had fast and slow (creep) portions with a viscosity increase from 1,370 N(s)/m to 17,864 N(s)/m during step-up in ISP. Wall elasticity averaged 77 N/m; which estimated the relationship of force and change in steady state response. Control BNA response also had fast and slow (resetting) portions. A BNA and wall response relationship (BNA/m) was defined as transduction-gain (T-G) with proportional and dynamic components. In the subgroup, wall creep and baroreceptor resetting were abolished by protease treatment, suggesting an endothelial mediator which influenced sinus smooth muscle. Histology data indicated enzyme damage was limited to tunica intima tissues, and nifedipine did not block Ca2+ channels on neural structures. By comparison of responses before and after treatments the proportional component of T-G was equated to an elastic influence (1/E), with E = 7.5 x 10(-6) m/BNA, while the dynamic component was equated to a viscous influence (1/V), with V = 1.53 x 10(-4) m(s)/BNA. A simple but fundamental relationship for baroreceptor-tissue linkages was estimated by BNA/m = 1/(Vs + E), a first-order transfer function.  相似文献   

14.
Dissection of a model for neuronal parabolic bursting   总被引:9,自引:0,他引:9  
We have obtained new insight into the mechanisms for bursting in a class of theoretical models. We study Plant's model [24] for Aplysia R-15 to illustrate our view of these so-called parabolic bursters, which are characterized by low spike frequency at the beginning and end of a burst. By identifying and analyzing the fast and slow processes we show how they interact mutually to generate spike activity and the slow wave which underlies the burst pattern. Our treatment is essentially the first step of a singular perturbation approach presented from a geometrical viewpoint and carried out numerically with AUTO [12]. We determine the solution sets (steady state and oscillatory) of the fast subsystem with the slow variables treated as parameters. These solutions form the slow manifold over which the slow dynamics then define a burst trajectory. During the silent phase of a burst, the solution trajectory lies approximately on the steady state branch of the slow manifold and during the active phase of spiking, the trajectory sweeps through the oscillation branch. The parabolic nature of bursting arises from the (degenerate) homoclinic transition between the oscillatory branch and the steady state branch. We show that, for some parameter values, the trajectory remains strictly on the steady state branch (to produce a resting steady state or a pure slow wave without spike activity) or strictly in the oscillatory branch (continuous spike activity without silent phases). Plant's model has two slow variables: a calcium conductance and the intracellular free calcium concentration, which activates a potassium conductance. We also show how bursting arises from an alternative mechanism in which calcium inactivates the calcium current and the potassium conductance is insensitive to calcium. These and other biophysical interpretations are discussed.  相似文献   

15.
ATP-dependent Ca2+ uptake by subfractions of skeletal muscle sarcoplasmic reticulum (SR) was studied with the Ca2+ indicator dye, antipyrylazo III. Ca2+ uptake by heavy SR showed two phases, a slow uptake phase and a fast uptake phase. By contrast, Ca2+ uptake by light SR exhibited a monophasic time course. In both fractions a steady state of Ca2+ uptake was observed when the concentration of free Ca2+ outside the vesicles was reduced to less than 0.1 microM. In the steady state, the addition of 5 microM Ca2+ to the external medium triggered rapid Ca2+ release from heavy SR but not from light SR, indicating that the heavy fraction contains a Ca2+-induced Ca2+ release channel. During Ca2+ uptake, heavy SR showed a constant Ca2+-dependent ATPase activity (1 mumol/mg protein X min) which was about 150 times higher than the rate of Ca2+ uptake in the slow uptake phase. Ruthenium red, an inhibitor of Ca2+-induced Ca2+ release, enhanced the rate of Ca2+ uptake during the slow phase without affecting Ca2+-dependent ATPase activity. Adenine nucleotides, activators of Ca2+ release, reduced the Ca2+ uptake rate. These results suggest that the rate of Ca2+ accumulation by heavy SR is not proportional to ATPase activity during the slow uptake phase due to the activation of the channel for Ca2+-induced Ca2+ release. In addition, they suggest that the release channel is inactivated during the fast Ca2+ uptake phase.  相似文献   

16.
Skeletal muscles are a mosaic of slow and fast twitch myofibers. During embryogenesis, patterns of fiber type composition are initiated that change postnatally to meet physiological demand. To examine the role of the protein phosphatase calcineurin in the initiation and maintenance of muscle fiber types, we used a "Flox-ON" approach to obtain muscle-specific overexpression of the modulatory calcineurin-interacting protein 1 (MCIP1/DSCR1), an inhibitor of calcineurin. Myo-Cre transgenic mice with early skeletal muscle-specific expression of Cre recombinase were used to activate the Flox-MCIP1 transgene. Contractile components unique to type 1 slow fibers were absent from skeletal muscle of adult Myo-Cre/Flox-MCIP1 mice, whereas oxidative capacity, myoglobin content, and mitochondrial abundance were unaltered. The soleus muscles of Myo-Cre/Flox-MCIP1 mice fatigued more rapidly than the wild type as a consequence of the replacement of the slow myosin heavy chain MyHC-1 with a fast isoform, MyHC-2A. MyHC-1 expression in Myo-Cre/Flox-MCIP1 embryos and early neonates was normal. These results demonstrate that developmental patterning of slow fibers is independent of calcineurin, while the maintenance of the slow-fiber phenotype in the adult requires calcineurin activity.  相似文献   

17.
Z Qu  D Chung 《PloS one》2012,7(8):e43587
In normal cardiac myocytes, the action potential duration (APD) is several hundred milliseconds. However, experimental studies showed that under certain conditions, APD could be excessively long (or ultralong), up to several seconds. Unlike the normal APD, the ultralong APD increases sensitively with pacing cycle length even when the pacing rate is very slow, exhibiting a sensitive slow rate-dependence. In addition, these long action potentials may or may not exhibit early afterdepolarizations (EADs). Although these phenomena are well known, the underlying mechanisms and ionic determinants remain incompletely understood. In this study, computer simulations were performed with a simplified action potential model. Modifications to the L-type calcium current (I(Ca,L)) kinetics and the activation time constant of the delayed rectifier K current were used to investigate their effects on APD. We show that: 1) the ultralong APD and its sensitive slow rate-dependence are determined by the steady-state window and pedestal I(Ca,L) currents and the activation speed and the recovery of the delayed rectifier K current; 2) whether an ultralong action potential exhibits EADs or not depends on the kinetics of I(Ca,L); 3) increasing inward currents elevates the plateau voltage, which in general prolongs APD, however, this can also shorten APD when the APD is already ultralong under certain conditions; and 4) APD alternans occurs at slow pacing rates due to the sensitive slow rate-dependence and the ionic determinants are different from the ones causing APD alternans at fast heart rates.  相似文献   

18.
A comparison of the transient kinetics of cardiac ventricular normal and hyperthyroid modified myosin subfragment-1 reveals substantial similarities between the two proteins. The nucleotide-binding kinetics are nonexponential for both proteins, but the large tryptophan fluorescence changes, 34% for ATP binding and 12% for ADP binding which are comparable to those of rabbit skeletal myosin subfragment-1, permit the kinetic data to be resolved into a sum of two exponentials. Both the fast and slow forms of the proteins reach limiting rate constants at high nucleotide concentration. The fast forms of normal and thyrotoxic cardiac subfragment-1 are kinetically identical for nucleotide binding at 20 degrees C and pH 7 and the slow forms differ by less than a factor of 2. The kinetic data for ADP release and the single turnover of ATP could neither be fit by a single exponential nor resolved into two components, which indicates a difference in the rate constants by a factor of 2 or less. The largest difference found was in the steady state turnover of ATP for which thyrotoxic subfragment-1 had a 2.5 times faster turnover as compared to normal subfragment-1. The fractions of fast and slow forms of the two proteins are dependent on the nucleotide concentration and the fractions as well as the rate constants are a function of the protein concentration. This is consistent with the kinetic heterogeneity of cardiac myosin subfragment-1 resulting from aggregation. The differences in the rate constant for the steady state turnover of ATP and in aggregation properties between normal and hyperthyroid cardiac subfragment-1 are consistent with the induction of a myosin isozyme by thyroxine treatment. Moreover, the increase in the steady state turnover of ATP is consistent with the increase in contractility of the muscle in the hyperthyroid state.  相似文献   

19.
Thirty healthy Sohagi ram lambs with an average age of 188.6±7.3 days were used to study the effect of pre-pubertal growth rate on some physiological parameters and sexual behavioral patterns at puberty. Ram lambs were divided into three groups (10 animals per each group) according to the previous growth rate until 6 months of age. Groups were marked as fast, medium and slow growing. Animal groups were housed in closed barns with access to an open area. Results showed that age and weight of ram lambs at puberty were significantly affected (P<0.05) by the pre-pubertal growth rate. Ram lambs in the fast growing group were reached to onset puberty firstly at 272.6 days with body weight (BW) 37.1 kg on average then ram lambs in medium group (284.8 days with BW 32.7 kg), while ram lambs in slow growing group were the last (314.1 days with BW 32.5 kg). Blood‎ testosterone‎ concentration at puberty was not significantly different among growing groups (1.494± 0.03 ng/ml on average, ranged from 1.287 to 1.902 ng/ml). Testes measurements from 6 months of age until puberty show that ram lambs in fast growing group had the highest values of testes length, circumference and volume followed by those in medium and slow growing group. Sexual behavioral observation showed that flehmen and mounting behavior were significantly higher for ram lambs in fast growing group (5.63 and 6.75 number/12h) than slow growing group (4.25 and 5.38 number/12h) while in medium growing group were intermediate (4.88 and 5.88 number/12h). From these findings, could be concluded that age, weight and sexual behavioral patterns of Sohagi ram lambs at puberty were affected by pre-pubertal growth rate, and the breeders should strive to achieve good growth rates for their lambs before puberty which led to improving reproductive performance.  相似文献   

20.
Depolarization induced charge movement associated currents, analogous to gating currents in channels, were recently demonstrated in G-protein coupled receptors (GPCRs), and were found to affect the receptor''s Agonist binding Affinity, hence denoted AA-currents. Here we study, employing a combined theoretical-experimental approach, the properties of the AA-currents using the m2-muscarinic receptor (m2R) as a case study. We found that the AA-currents are characterized by a “bump”, a distinct rise followed by a slow decline, which appears both in the On and the Off responses. The cumulative features implied a directional behavior of the AA-currents. This forced us to abandon the classical chemical reaction type of models and develop instead a model that includes anisotropic processes, thus producing directionality. This model fitted well the experimental data. Our main findings are that the AA-currents include two components. One is extremely fast, , at all voltages. The other is slow, at all voltages. Surprisingly, the slow component includes a process which strongly depends on voltage and can be as fast as at . The reason that it does not affect the overall time constant of the slow component is that it carries very little charge. The two fast processes are suitable candidates to link between charge movement and agonist binding affinity under physiological conditions.  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