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.
Water transport across plant cell membranes is difficult to measure. We present here a model assay, based on chlorophyll (Chl) a fluorometry, with which net water transport across the cell membrane of freshwater cyanobacterium Synechococcus sp. PCC7942 (S7942) can be followed kinetically with millisecond-time resolution. In cyanobacteria, the phycobilisome (PBS)-sensitized Chl a fluorescence increases when cells expand (e.g., in hypo-osmotic suspension) and decreases when cells contract (e.g., in hyper-osmotic suspension). The osmotically-induced Chl a fluorescence changes are proportional to the reciprocal of the suspension osmolality (ΔF ∝ Osm−1; Papageorgiou GC and Alygizaki-Zorba A (1997) Biochim Biophys Acta 1335: 1–4). In our model assay, S7942 cells were loaded with NaCl (passively penetrating solute) and shrunk in hyper-osmotic glycine betaine (nonpenetrating solute). Upon injecting these cells into hypo-osmotic medium, the PBS-sensitized Chl a fluorescence rose to a maximum due to the osmotically-driven water uptake. The rise of Chl a fluorescence (water uptake) was partially inhibited by HgCl2, at micromolar concentrations. Arrhenius plots of the water uptake rates gave activation energies of EA=4.9 kcal mol−1, in the absence of HgCl2, and EA=11.9 kcal mol−1 in its presence. These results satisfy the usual criteria for facilitated water transport through protein water pores of plasma membranes (aquaporins), namely sensitivity to Hg2+ ions and low activation energy.  相似文献   

2.
1. The fat mouse Steatomys pratensis natalensis (mean body mass 37.4±0.43 (se)) has a low euthermic body temperature Tb=30.1–33.8 °C and a low basal metabolic rate (BMR)=0.50 ml O2 g−1 h−1.
2. Below an ambient temperature (Ta)=15 °C, the mice were hypothermic.
3. The lowest survivable Ta=10 °C.
4. Torpor is efficient in conserving energy between Ta=15–30 °C, below Ta=15 °C, the mice arouse.
5. Euthermic and torpid mice were hyperthermic at Ta=35 °C.
6. Thermal conductance was 0.159 ml O2 g−1 h−1 °C−1, 98.8% of the expected value.
7. Non-shivering thermogenesis (NST) was 2.196 ml O2 g−1 h−1 (3.69×BMR).
8. Maximal oxygen consumption, however, was 3.83 ml O2 g−1 h−1 (6.44×BMR), indicating that other methods of heat production are additive.
9. Because fat mice conserve energy by torpor only between Ta=15–30 °C, we suggest that torpor may be a more important mechanism for surviving food shortages than for surviving cold weather.
Keywords: Steatomys pratensis natalensis; Metabolism; Torpor; Fat mouse  相似文献   

3.
In this study, the hydraulic conductivity (Lp), Me2SO permeability ( Me2SO), and the reflection coefficients (ς) and their activation energies were determined for Metaphase II (MII) mouse oocytes by exposing them to 1.5 M Me2SO at temperatures of 30, 20, 10, 3, 0, and −3°C. These data were then used to calculate the intracellular concentration of Me2SO at given temperatures. Individual oocytes were immobilized using a holding pipette in 5 μl of an isosmotic PBS solution and perfused with precooled or prewarmed 1.5 M Me2SO solutions. Oocyte images were video recorded. The cell volume changes were calculated from the measurement of the diameter of the oocytes, assuming a spherical shape. The initial volume of the oocytes in the isoosmotic solution was considered 100%, and relative changes in the volume of the oocytes after exposure to the Me2SO were plotted against time. Mean (means ± SEM) Lpvalues in the presence of Me2SO ( Me2SOp) at 30, 20, 10, 3, 0, and −3°C were determined to be 1.07 ± 0.03, 0.40 ± 0.02, 0.18 ± 0.01, 7.60 × 10−2± 0.60 × 10−2, 5.29 × 10−2± 0.40 × 10−2, and 3.69 × 10−2± 0.30 × 10−2μm/min/atm, respectively. The Me2SOvalues were 3.69 × 10−3± 0.3 × 10−3, 1.07 × 10−3± 0.1 × 10−3, 2.75 × 10−4± 0.15 × 10−4, 7.83 × 10−5± 0.50 × 10−5, 5.24 × 10−5± 0.50 × 10−5, and 3.69 × 10−5± 0.40 × 10−5cm/min, respectively. The ς values were 0.70 ± 0.03, 0.77 ± 0.04, 0.81 ± 0.06, 0.91 ± 0.05, 0.97 ± 0.03, and 1 ± 0.04, respectively. The estimated activation energies (Ea) for Me2SOp, Me2SO, and ς were 16.39, 23.24, and −1.75 Kcal/mol, respectively. These data may provide the fundamental basis for the development of more optimal cryopreservation protocols for MII mouse oocytes.  相似文献   

4.
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   

5.
A comparison of the thermoregulation of water foraging wasps (Vespula vulgaris, Polistes dominulus) under special consideration of ambient temperature and solar radiation was conducted. The body surface temperature of living and dead wasps was measured by infrared thermography under natural conditions in their environment without disturbing the insects’ behaviour. The body temperature of both of them was positively correlated with Ta and solar radiation. At moderate Ta (22–28 °C) the regression lines revealed mean thorax temperatures (Tth) of 35.5–37.5 °C in Vespula, and of 28.6–33.7 °C in Polistes. At high Ta (30–39 °C) Tth was 37.2–40.6 °C in Vespula and 37.0–40.8 °C in Polistes. The thorax temperature excess (TthTa) increased at moderate Ta by 1.9 °C (Vespula) and 4.4 °C (Polistes) per kW−1 m−2. At high Ta it increased by 4.0 °C per kW−1 m−2 in both wasps. A comparison of the living water foraging Vespula and Polistes with dead wasps revealed a great difference in their thermoregulatory behaviour. At moderate Ta (22–28 °C) Vespula exhibited distinct endothermy in contrast to Polistes, which showed only a weak endothermic activity. At high Ta (30–39 °C) Vespula reduced their active heat production, and Polistes were always ectothermic. Both species exhibited an increasing cooling effort with increasing insolation and ambient temperature.  相似文献   

6.
The Cl conductance in isolated skin of frogs (Rana catesbeiana) acclimated to 30 mM solutions of NaCl, Na2SO4, MgCl2 and distilled water (DW) was studied. Transepithelial potential difference (PDtrans), short-circuit current (ISC) and total conductance (Gt) were measured under conditions such that there was Cl flux in the presence and absence of Na+ transport. The Cl content of the mucosal solution was acutely replaced with SO42− or gluconate to evaluate the effect of removal of Cl conductance on electrophysiological parameters. Mitochondria-rich cell density (DMRC) was also measured. Skins from frogs acclimated to NaCl and Na2SO4 showed the lowest and the highest DMRC, respectively, but no difference could be found between the skins from frogs acclimated to DW and MgCl2 indicating that DMRC is not unconditionally dependent on environmental Cl in this species. Frogs acclimated to NaCl showed marked differences when compared to the other groups: the highest Gt, probably represented by a higher paracellular conductance; the lowest transepithelial electrical potential difference which remained invariant after replacement of mucosal Cl with SO42− or replacement of mucosal Cl with gluconate and an inwardly oriented positive current in the absence of bilateral Na+.  相似文献   

7.
Summary Changes in extracellular pH (pHo) in red cell suspensions were monitored in a stopped-flow rapid reaction apparatus under conditions wheredpHo/dt was determined by the rate of HCO 3 /Cl exchange across the membrane. Experiments were performed at 5°C<T<40°C using either untreated cells or cells exposed to 0.11mm SITS (4-acetamido-4-isothiocyanostilbene-2,2-disulfonic acid). Although SITS exposure reduced the rate of exchange by 90%, both untreated and SITS-treated cells are similarly affected by changes in pH0 and temperature. The rate of HCO 3 /Cl exchange exhibits a minimum at about pHo 5 and a maximum at about pH0 7.4 at all temperatures. A transition temperature of 17°C was observed in the Arrhenius relationship for all pH0. The activation energies (E a) in kcal/mol are 19.6 below and 11.7 above 17°C for 50<8. These findings, similar to those reported for Cl self-exchange, suggest that: (i) a change in the rate-limiting step for HCO 3 /Cl exchange occurs at 17°C, possibly due to an altered interaction between the transport pathway and membrane lipids; (ii) the carrier system can be titrated by either H+ or SITS from the outside of the membrane, but the untitrated sites continue to transport normally; (iii) the pH0 dependence of the rate of exchange is consistent with the titratable carrier having its most alkaline pK in the range expected for amino groups; and (iv) below pH0 5, the nature of the exchange is markedly altered.  相似文献   

8.
Glycerol diffusional permeabilities through the cytoplasmic cell membrane of Dunaliella salina, the cell envelope of pig erythrocyte and egg phosphattidylcholine vesicles were measured by NMR spectroscopy employing the spin-echo method and nuclear T1 relaxation. The following permeability coefficients (P) and corresponding enthalpies of activation (ΔH) were determined for glycerol at 25°C: for phosphatidylcholine vesicles 5·10−6 cm/s and 11±2 kcal/mol; for pig erythrocytes 7·10−8 cm/s and 18±3 kcal/mol, respectively; for the cytoplasmic membrane of D. salina the permeability at 17°C was found to be exceptionally low and only a lower limit (P<5·10−11cm/s) could be calculated. At temperatures above 50°C a change in membrane permeability occurred leading to rapid leakage of glycerol accompanied by cell death. The data reinforce the notion that the cytoplasmic membrane of Dunaliella represents a genuine anomaly in its exceptional low permeability to glycerol.  相似文献   

9.
The “two-step” low-temperature microscopy (equilibrium and dynamic) freezing methods and a differential scanning calorimetry (DSC) technique were used to assess the equilibrium and dynamic cell volumes in Rana sylvatica liver tissue during freezing, in Part I of this study. In this study, the experimentally determined dynamic water transport data are curve fit to a model of water transport using a standard Krogh cylinder geometry (Model 1) to predict the biophysical parameters of water transport: Lpg = 1.76 μm/min-atm and ELp = 75.5 kcal/mol for control liver cells and Lpg[cpa] = 1.18 μm/min-atm and ELp[cpa] = 69.0 kcal/mol for liver cells equilibrated with 0.4 M glucose. The DSC technique confirmed that R. sylvatica cells in control liver tissue do not dehydrate completely when cooled at 5°C/min but do so when cooled at 2°C/min. Cells also retained twice as much intracellular fluid in the presence of 0.4 M glucose than in control tissue when cooled at 5°C/min. The ability of R. sylvatica liver cells to retain water during fast cooling (≥5°C/min) appears to be primarily due to its liver tissue architecture and not to a dramatically lower permeability to water, in comparison to mammalian (rat) liver cells which do dehydrate completely when cooled at 5°C/min. A modified Krogh model (Model 2) was constructed to account for the cell–cell contact in frog liver architecture. Using the same biophysical permeability parameters obtained with Model 1, the modified Krogh model (Model 2) is used in this study to qualitatively explain the experimentally measured water retention in some cells during freezing on the basis of different volumetric responses by cells directly adjacent to vascular space versus cells at least one cell removed from the vascular space. However, at much slower cooling rates (1–2°C/h) experienced by the frog in nature, the deciding factor in water retention is the presence of glucose and the maintenance of a sufficiently high subzero temperature (≥−8°C).  相似文献   

10.
Active transport of Cl accounts for 90% of the short-circuit current (s.c.c.) in the isolated frog cornea. 1·10−5 M furosemide produced a 50% reversible inhibition of this s.c.c. 1·10−4 M ethacrynic acid reduced the corneal s.c.c. to 32% of the control. In the isolated frog skin epithelium furosemide had no effect on the s.c. at a concentration of 1·10−4 M and a small stimulation at a concentration of 1·10−3 M. The furosemide inhibitory effects seems to be specific for Cl, as it also inhibits Cl transport in the ascending limb of the loop of Henle (Burg, M.B. (1972) Proc. 5th Int. Congr. Nephrol., p. 50, Abstr.).  相似文献   

11.
For both MoO42− and WO42− the maximum rate of uptake by the small intestine of the rat (studied in vitro using the everted sac technique) occurs in the lower ileum. Kinetic constants, derived by a least squares procedure, are compared with those previously obtained for SO42− transport. For both V and Ka, SO42− > MoO42− > WO42−, with only small differences between sacs IV and V. Mutual inhibition of MoO42− and WO42− transport and inhibition of both by SO42− are competitive processes. This is shown by the generally good agreement between Ka values and derived Ki values and by V values in the presence and absence of the inhibiting species. The three ions SO42−, MoO42− and WO42− are probably transferred across the intestine by a common carrier system. Implications for the sulphate-molybdenum interaction in molybdosis are discussed.  相似文献   

12.
Acid hydrolysis of sugarcane bagasse for lactic acid production   总被引:3,自引:0,他引:3  
In order to use sugarcane bagasse as a substrate for lactic acid production, optimum conditions for acid hydrolysis of the bagasse were investigated. After lignin extraction, the conditions were varied in terms of hydrochloric (HCl) or sulfuric (H2SO4) concentration (0.5–5%, v/v), reaction time (1–5 h) and incubation temperature (90–120 °C). The maximum catalytic efficiency (E) was 10.85 under the conditions of 0.5% of HCl at 100 °C for 5 h, which the main components (in g l−1) in the hydrolysate were glucose, 1.50; xylose, 22.59; arabinose, 1.29; acetic acid, 0.15 and furfural, 1.19. To increase yield of lactic acid production from the hydrolysate by Lactococcus lactis IO-1, the hydrolysate was detoxified through amberlite and supplemented with 7 g l−1 of xylose and 7 g l−1 of yeast extract. The main products (in g l−1) of the fermentation were lactic acid, 10.85; acetic acid, 7.87; formic acid, 6.04 and ethanol, 5.24.  相似文献   

13.
We measured Na+/K+ ATPase activity in homogenates of gill tissue prepared from field caught, winter and summer acclimatized yellow perch, Perca flavescens. Water temperatures were 2–4°C in winter and 19–22°C in summer. Na+/K+ ATPase activity was measured at 8, 17, 25, and 37°C. Vmax values for winter fish increased from 0.48±0.07 μmol P mg−1 protein h−1 at 8°C to 7.21±0.79 μmol P mg−1 protein h−1 at 37°C. In summer fish it ranged from 0.46±0.08 (8°C) to 3.86±0.50 (37°C) μmol P mg−1 protein h−1. The Km for ATP and for Na+ at 8°C was ≈1.6 and 10 mM, respectively and did not vary significantly with assay temperature in homogenates from summer fish. The activation energy for Na+/K+ ATPase from summer fish was 10 309 (μmol P mg−1 h−1) K−1. In winter fish, the Km for ATP and Na+ increased from 0.59±0.08 mM and 9.56±1.18 mM at 8°C to 1.49±0.11 and 17.88±2.64 mM at 17°C. The Km values for ATP and Na did not vary from 17 to 37°C. A single activation energy could not be calculated for Na/K ATPase from winter fish. The observed differences in enzyme activities and affinities could be due to seasonal changes in membrane lipids, differences in the amount of enzyme, or changes in isozyme expression.  相似文献   

14.
The kinetics of formation of the complex ion, μ-carbonato-di-μ-hydroxo-bis((1,5-diamino-3-aza-pentane) cobalt(III), from the tri-μ-hydroxo-bis((1,5-diamino-3-aza-pentane(III)cobalt(III)) ion in aqueous buffered carbonate solution have been studied spectrophotometrically at 295 nm over the ranges 20.0θ°C34.8, 8.03pH9.44, 5 mM [CO32−35 mM and at an ionic strength of 0.1 M (LiClO4). On the basis of the kinetic results a mechanism, involving rapid cleavage of an hydroxo bridge followed by carbon dioxide uptake with subsequent bridge formation, has been proposed. At 25 °C, the rate of the carbon dioxide uptake is 0.58 M−1 s−1 with ΔH≠ = (13.2±0.7) kcal mol−1 and ΔS≠ = (−15.1 ± 0.7) cal deg−1 mol−1. The results are composed with those obtained for several mononuclear cobalt(III) and one dinuclear cobalt(III) complexes.  相似文献   

15.
In vitro studies were undertaken to determine the effect of pH, temperature, water availability and carbon dioxide (CO2) concentration on germination and growth of Colletotrichum musae, the causal pathogen of anthracnose of bananas. The optimum pH for germination and growth varied between 4·0 and 5·0 depending on temperature. At low pH (< 3·0) and 15°C, both germination and growth were significantly reduced, with a marked increase in the lag time, in days, prior to growth. C. musae germinated and grew over a wide range of water activities (aw; 0·995−0·94 and 0·995−0·92, respectively) at 20, 25 and 30°C. In all cases where germination occurred appresoria were subsequently produced. Optimum growth occurred at 30°C and 0·995 aw, although this changed to 0·98 aw at 35°C. Increasing CO2 concentration to 15% or reducing oxygen concentration to 1% resulted in a significant (P < 0·05) reduction in growth, but did not inhibit growth completely.  相似文献   

16.
1. 1. The Michaelis-Menten parameters of labelled d-glucose exit from human erythrocytes at 2°C into external solution containing 50 mM d-galactose were obtained. The Km is 3.4 ± 0.4 mM, V 17.3 ± 1.4 mmol · 1−1 cell water · min−1 for this infinite-trans exit procedure.
2. 2. The kinetic parameters of equilibrium exchange of d-glucose at 2°C are Km = 25 ± 3.4 mM, V 30 ± 4.1 mmol · 1−1 cell water · min−1.
3. 3. The Km for net exit of d-glucose into solutions containing zero sugar is 15.8 ± 1.7 mM, V 9.3 ± 3.3 mol 9.3 ± 3.3 mol · 1−1 cell water · min−1.
4. 4. This experimental evidence corroborates the previous finding of Hankin, B.L., Lieb, W.R. and Stein, W.D. [(1972) Biochim. Biophys. Acta 255, 126–132] that there are sites with both high and low operational affinities for d-glucose at the inner surface of the human erythrocyte membrane. This result is inconsistent with current asymmetric carrier models of sugar transport.
Keywords: d-Glucose transport; Asymmetric carrier; Pore kinetics; (Erythrocyte)  相似文献   

17.
The thermal coefficient of expansion of egg lecithin bilayer thickness, αd1, was measured as a function of its cholesterol content up to mole ratio lecithin/cholesterol of 1:1, and over the temperature range 0–40 °C. At all cholesterol contents αd1 changes abruptly at approximately 12 °C indicating a structural transition at this temperature. Above 12 °C, αd1 decreases monotonically from −2·10−3 for pure egg lecithin to −1·10–3 at mole ratio 1:1. Below 12 °C αd1 is walways higher than above 12 °C and shows a sharp, anomalously high value of −6·10−3 at the mole ratio 2:1. The results have been interpreted as the movement of cholesterol into the bilayer or the formation of lecithin-cholesterol “complexes” at temperatures below 12 °C. Similar studies with phosphatidylinositol containing cholesterol showed no structural transition and lysolecithin containing cholesterol behaved differently giving two lamellar phases in equilibrium.  相似文献   

18.
The rate constants k12n for isomerization of the E1H isomer (pKa 8 in H2O) of ribonuclease-A to the E2H isomer (pKa = 6.1 in H2O), determined from proton-uptake measurements by the temperature-jump technique, in mixtures of protium and deuterium oxides (atom fraction of deuterium n), are described by the equation k12n = (733 ± 16)(1 − n + [0.46 ± 0.04]n)(1 − n + 0.69n)2sec−1 at 25°C. On the basis of the absolute magnitude of the rate constant, the magnitude of the solvent isotope effect and the proton inventory, it appears that the rate-determining step is proton transfer to a water molecule from the imidazolium form of a histidine residue, with a product-like activated complex resembling a hydronium ion. The subsequent motion of the protein structure to generate the new isomer (conformation change) must then occur in a time approaching a vibrational period. Alternative but less likely mechanisms include rate-limiting protein reorganization concerted with proton transfer to water, rate-limiting diffusion of hydronium ion away from the enzyme, or “solvation catalysis” of protein reorganization.  相似文献   

19.
Malate enzyme (l-malate : NADP+ oxidoreductase (oxalacetate-decarboxylating, EC 1.1.1.40)) has been purified from Pseudomonas putida to 99 per cent homogeneity by heat, ammonium suphate fractionation, gel filtration and anion exchange chromatography. Sodium dodecylsulphate-(SDS)-polyacrylamide disc gel electrophoresis analysis showed an approximate tetrameric subunit with a molecular weight of 52,000. The purified enzyme showed a pH optimum between 8.0 and 8.5 (for Tris-HCl buffer) and required bivalent cations for catalysis ; monovalent ions like K+ and NH4+ acted as very effective activators. The temperature-activity relationship for the malate enzyme from 35–80 °C showed broken Arrhenius plots with an inflexion at 65 °C. The enzyme halflife was 30s at 85 °C.The enzyme showed hyperbolic kinetics for both substrates with apparent Km values of 4.0 × 10−4 M and 2.3 × 10−5 M for l-malate and NADP+ respectively. From the study of the effects of some compounds on the enzyme, the physiological significance of those produced by fumarate, succinate and oxalacetate can be emphasized.  相似文献   

20.
T1 nuclear relaxation measurements of 1H and 17O of water have been applied to study the kinetics of the diffusional transport of water across the cytoplasmic cell membrane of Dunaliella salina and Dunaliella bardawil. The water permeability coefficients at 25°C were found to be 1.5·10−3 cm/s and 1.8·10−3 cm/s, respectively, with an activation energy of 3.7 kcal/mol. The results indicate that the cell membrane of Dunaliella exhibits high diffusional permeability to water, similar in magnitude to that found for other cells and model membranes, and a relatively low activation energy. This regularity is in contrast to the exceptionally low glycerol permeability of the membrane (Brown, F.F., Sussman, I., Avron, M. and Degani, H. (1982) Biochim. Biophys. Acta 690, 165–173).  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