首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The influence of the physical state of the membrane on the swimming behaviour of Tetrahymena pyriformis was studied in cells with lipid-modified membranes. When the growth temperature of Tetrahymena cells was increased from 15 degrees C to 34 degrees C or decreased from 39 degrees C to 15 degrees C, their swimming velocity changed gradually in a similar to the adaptive change in membrane lipid composition. Therefore, such adaptive changes in swimming velocity were not observed during short exposures to a different environment. Tetrahymena cells adapted to 34 degrees C swam at 570 microns/s. On incubation at 15 degrees C these cells swam at 100 microns/s. When the temperature was increased to 34 degrees C after a 90-min incubation at 15 degrees C, the initial velocity was immediately recovered. On replacement of tetrahymanol with ergosterol, the swimming velocity of 34 degrees C-grown cells decreased to 210 microns/s, and the cells ceased to move when the temperature was decreased to 15 degrees C. To investigate the influence of the physical state of the membrane on the swimming velocity, total phospholipids were prepared from Tetrahymena cells grown under these different conditions. The fluidities of liposomes of these phospholipid were measured using stearate spin probe. The membrane fluidity of the cells cooled to 15 degrees C increased gradually during incubation at 15 degrees C. On the other hand, the fluidity of the heated cell decreased during incubation at 34 degrees C. Replacement of tetrahymanol with ergosterol decreased the membrane fluidity markedly. Consequently, a good correlation was observed between swimming velocity and membrane fluidity; as the membrane fluidity increased, the swimming velocity increased linearly up to 600 microns/s. These results provide evidence for the regulation of the swimming behaviour by physical properties of the membrane.  相似文献   

2.
The fatty acid distribution pattern of lipids extracted from different subcellular components of Tetrahymena pyriformis was found to be significantly different from one type of membrane to another.The growth-temperature shift caused alterations in fatty acid composition. The ratio of palmitoleic to palmitic acid, especially, showed a sharp linear decline with increase of temperature in all of the membrane fractions.The spin labels were rapidly incorporated into Tetrahymena membranes. The order parameter of 5-nitroxide stearate spin label incorporated into various membrane fractions was found to be different for the different membrane fractions, suggesting the following order of the fluidity; microsomes > pellicles > cilia.The fluidity of the surface membranes, cilia and pellicles isolated from Tetrahymena cells grown at 15°C was noticeably higher than that of the membranes from cells grown at 34°C but was not so different with microsomal fractions.The motion of the spin label in the pellicular membrane was more restricted than in its extracted lipids, thus indicating the assumption that in Tetrahymena membranes the proteins influence the fluidity.It was also suggested that a sterol-like triterpenoid compound, tetrahymanol, which is principally localized in the surface membranes, would be involved in the membrane fluidity.  相似文献   

3.
The effect of temperature on native microsomal membrane vesicles isolated from Tetrahymena is investigated by wide angle X-ray diffraction. A 4.2 Å reflection, typical for lipids in the crystalline state, can be recorded in the temperature range between 0°C and 35°C. Quantitative evaluation of this reflection reveals a broad thermotropic ‘two-stage’ liquid crystallinecrystalline lipid phase separation with a ‘breakpoint’ at approx. 18°C. This ‘breakpoint’ coincides with the emergence of lipid-protein segregations in endomembranes of intact Tetrahymena cells as previously visualized by freeze-etch electron microscopy.  相似文献   

4.
A specific effect of cardiolipin on fluidity of mitochondrial membranes was demonstrated in Tetrahymena cells acclimated to a lower temperature in the previous report (Yamauchi, T., Ohki, K., Maruyama, H. and Nozawa, Y. (1981) Biochim. Biophys. Acta 649, 385–392). This study was further confirmed by the experiment using fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene (DPH). Anisotropy of DPH for microsomal and pellicular total lipids from Tetrahymena cells showed that membrane fluidity of these lipids increased gradually as the cells were incubated at 15°C after the shift down of growth temperature from 39°C. However, membrane fluidity of mitochondrial total lipids was kept constant up to 10 h. This finding is compatible with the result obtained using spin probe in the previous report. Additionally, the break-point temperature of DPH anisotropy was not changed in mitochondrial lipids whereas those temperatures in pellicular and microsomal lipids lowered during the incubation at 15°C. Interaction between cardiolipins and various phospholipids, which were isolated from Tetrahymena cells grown at 39°C or 15°C and synthesized chemically, was investigated extensively using a spin labeling technique. The addition of cardiolipins from Tetrahymena cells grown at either 39°C or 15°C did not change the membrane fluidity (measured at 15°C) of phosphatidylcholine from whole cells grown at 39°C. On the other hand, both cardiolipins of 39°C-grown and 15°C-grown cells decreased the membrane fluidity of phosphatidylcholine from Tetrahymena cells grown at 15°C. The same results were obtained for phosphatidylcholines of mitochondria and microsomes. Membrane fluidity of phosphatidylethanolamine, isolated from cells grown at 15°C, was reduced to a small extent by Tetrahymena cardiolipin whereas that of 39°C-grown cells was not changed. Representative molecular species of phosphatidylcholines of cells grown at 39°C and 15°C were synthesized chemically; 1-palmitoyl-2-oleoylphosphatidylcholine for 39°C-grown cells and dipalmitoleoylphosphatidylcholine for 15°C-grown ones. By the addition of Tetrahymena cardiolipin, the membrane fluidity of 1-palmitoyl-2-oleoylphosphatidylcholine was not changed but that of dipalmitoleoylphosphatidylcholine was decreased markedly. These phenomena were caused by Tetrahymena cardiolipin. However, bovine heart cardiolipin, which has a different composition of fatty acyl chains from the Tetrahymena one, exerted only a small effect.  相似文献   

5.
The growth rate of Tetrahymena setosa cells is stimulated significantly by as little as 0.1 μg and optimally by about 1 μg of ergosterol per ml of medium. Cell yields in the stationary phase are, however, not perceptibly affected by increasing sterol concentrations. Ergosterol, in concentrations that stimulate growth optimally, does not cause a reduction of tetrahymanol synthesis. The latter process is impaired only at much higher ergosterol concentrations. Epicholesterol and coprostanol inhibit ergosterol-stimulated growth competitively. It is concluded that the trace amounts of sterol needed by T. setosa do not serve to replace tetrahymanol but function in some other manner, probably unrelated to the control of membrane fluidity. This conclusion supports the views advanced earlier by Holz, Erwin, Wagner & Rosenbaum.  相似文献   

6.
The ciliated protozoan Tetrahymena pyriformis has been used to study the biochemistry of cellular injury induced by rapid cooling (cold shock). Cellular viability was found to depend on the time and temperature of cold exposure, and the rate of cooling. During cooling to −7.5 °C, in the absence of ice, an optimal rate of cooling of 2.5 °C min−1 was observed; at both faster and slower cooling the recovery decreased. Following acclimation at a reduced temprature (10 °C) the viability following rapid cooling was significantly different from that of cultures maintained at 20 °C. Analysis of the phospholipid fatty acids from cells grown at 10 °C demonstrated that, at the reduced temperature, there was an increase in the average degree of fatty acyl unsaturation. Cold-shock injury in Tetrahymena is associated with membrane thermotropic events which are determined by temperature per se, whereas viability is a function of the rate of cooling. A hypothesis of injury is presented in which the presence of gel-phase lipid within the membrane is not the critical event, but it is the pattern of nucleation within the membrane which ultimately determines the extent of cellular injury.  相似文献   

7.
In the plasma membrane of various eucaryotic cell types, in particular blood platelets and erythrocytes, it is known that phospholipids are asymmetrically distributed between the two leaflets of the lipid bilayer and that this transverse asymmetry is controlled by an aminophospholipid translocase activity. In this respect, it was of interest to check whether there are differential transbilayer movements between amino- and neutral phospholipids in the apical plasma membrane of vascular endothelial cells which form the inner nonthrombogenic lining of the large blood vessel. In the first step we compared the transbilayer localization and also the rate of lateral motion of two fluorescent analogs of phosphatidylcholine and phosphatidylethanolamine, namely C6-NBD-PC and C6-NBD-PE, inserted into the apical plasma membrane of bovine aortic endothelial cells, in vitro. By the use of back-exchange experiments we have found that C6-NBD-PC could be removed from the cell membrane toward the culture medium regardless of the incubation conditions used, i.e., just after cell labeling at 0°C or even after further cell incubation for 1 h at 0 or 20°C. In contrast, C6-NBD-PE could be removed only when the cells were maintained at 0°C. After incubation for 1 h at 20°C, 85% of the probe molecules remained nonexchangeable, indicating probe translocation from the outer to the inner leaflet of the lipid bilayer. This "flip" process, which occurred at 20°C, was abolished when the endothelial cells were preincubated with N-ethylmaleimide, diamide, vanadate (VO3-4) and vanadyl (VO2+) ions, a set of substances which inhibit aminophospholipid translocase activity in various systems, and with a combination of sodium azide and 2-deoxyglucose which led to nearly complete ATP depletion in the cells. Fluorescence recovery after photobleaching experiments were also carried out to specify more precisely the localization and dynamics of the probes in the two leaflets of the plasma membrane lipid bilayer. They produced lateral diffusion coefficients D of 1.2 ± 0.05 × 10-9 cm2/s for C6-NBD-PC and 2.8 ± 0.3 × 10-9 cm2/s for C6-NBD-PE, when the two probes were located in the outer leaflet of the plasma membrane, just after cell labeling at 0°C. After cell incubation for one hour at 20°C, i.e., when C6-NBD-PC was still in the outer leaflet whereas C6-NBD-PE was translocated in the inner leaflet, D was observed to slightly increase for C6-NBD-PC (D = 1.9 ± 0.06 × 10-9 cm2/s) and to greatly increase by at least a factor of 3 for C6-NBD-PE (D = 9.1 ± 0.9 × 10-9 cm2/s). These results show that the plasma membrane of bovine aortic endothelial cells is equipped with a protein-dependent and energy-mediated phosphatidylethanolamine translocase activity and that the lateral diffusion rate of this phospholipid is much faster in the inner than in the outer leaflet of the lipid bilayer, thus indicating large differences in the fluidity of the two halves of this membrane.  相似文献   

8.
Upon stimulation with either concanavalin A or the tuberculin antigen, purified protein derivative, human peripheral blood lymphocytes, purified on Ficoll-Hypaque, did not exhibit a concomitant lipid fluidity alteration as measured by fluorescence polarization (P) of the lipid probe, 1,6-diphenyl-1,3,5-hexatriene (DPH). This result was independent of the incubation period, ranging from 10 min to 72 h. However, a general reduction in polarization value, from P = 0.287 (maintained for up to 2 h of incubation) to P = 0.225 after 20 h was observed for both experimental and control samples. Moreover, fluorescence polarization studies of the nonpenetrating modified DPH cationic lipid probe, 1-[4′-trimethylaminophenyl]-6-phenyl-1,3,5-hexatriene (TMA-DPH), also failed to show any change in lipid fluidity subsequent to a 1–3 h incubation of lymphocytes with concanavalin A. Cell electrophoretic mobility, however, was altered (mean cell mobility increased by 10–15%) in a fast response to stimulation and was observed within several hours of in vitro application of concanavalin A and purified protein derivative. This initial response disappeared with further incubation at 37°C (>3 h) and was followed by a decline of cellular mobility of the concanavalin A-exposed cells after 48 and 72 h of incubation. The unstimulated control cells did not change in mobility as a function of incubation time. The slow decline in mean cell mobility of the experimental cells is believed to be associated with blastogenesis. It is concluded that neither blastogenic transformation nor short term membrane alterations associated with human lymphocyte activation lead to lipid fluidity changes as measured in steady state by the fluorescence polarization of both DPH and TMA-DPH.  相似文献   

9.
The content of adenosine 3′,5′-monophosphate in human mononuclear leukocytes was enhanced 3–5-times by venoms obtained from African toad (Bufo africanus), American toad (Bufo americanus), Colorado river toad (Bufo arenarum) and Marine toad (Bufo marinus) at 25 μg/ml for 5 min of incubation at 37°C. The maximum stimulation was observed after 1–5 min of incubation. The half-maximal stimulation was observed at 0.1 μg/ml venom obtained from Colorado river toad (Bufo arenarum). The increased content of adenosine 3′,5′-monophosphate in the mononuclear leukocytes persisted without significant change for at least 30 min of incubation at 37°C.  相似文献   

10.
The distribution and mobility of cell surface anions was investigated on low passage cultures of secondary BALB/c embryonic fibroblasts and their SV40-transformed counterparts (VLM cells) using polycationized ferritin (PCF) as a label. While the absolute number of anions/nm2 of membrane was equivalent on the two cell types, the topographical distribution and mobility of these anions was strikingly different. After pulse-labeling with PCF at 37 °C, anions on BALB/c fibroblasts occurred in large piled-up clusters separated by extensive areas of membrane ( = 0.47 μm) free of negative charges. Labeling at 4 °C reduced the degree of “piling up” within the clusters, but the intercluster spacing was maintained indicating that the anions have short-range mobility in the membrane and can be cross-linked into a tight lattice at 37 °C. These anions do not, however, demonstrate any long-range mobility during a 20 min post-label incubation in PBS. In contrast, anions on VLM cells are inherently present in a random configuration of microclusters and single anions with relatively small ( = 0.09 μm) intervening areas of low charge density. Short-range mobility of surface anions is not displayed, presumably as a result of the inter-site distance, but long-range mobility is indicated by the formation of large site clusters following a 20 min incubation in PBS after pulse-labeling. Very mild proteolysis of BALB/c fibroblasts induces a change in the topography of surface anions toward the random configuration typical of VLM cells. These data are discussed in relation to altered social interactions between tumor cells which may be influenced by cell-cell adhesion characteristics.  相似文献   

11.
A large number of viral materials are associated with the surface of cells after cell fusion with HVJ at 37 °C for 30 min. This is due to fusion of viral envelopes with the cell membrane. Studies were made on the process from viral adsorption to cell-cell, or cell-viral envelope fusion. On incubation at low temperatures, such as 0–15 °C, no envelope fusion or cell fusion was observed, although there was some interaction between the virus and cells. This interaction resulted in loss of hemadsorption (HA) activity of the cells and partial damage of the ion barrier of the cell membrane. The viral particles seem to come close to the lipid layer of the cell membrane at the low temperatures and to distort the non-flexible membrane structure. On incubation of the cell-virus complex at 37 °C, the cells rapidly became HA-positive and the HA activity was maximal within 5 min. At this stage there was much leakage of ions through the cell membrane. On further incubation the damage to the ion barrier of the cell membrane was repaired completely with completion of cell fusion. This process may be correlated with fusion of viral envelopes with cell membranes and restoration of the cell membrane fused with them.  相似文献   

12.
This paper offers the suggestion that heat shock inhibition of tubulin synthesis accounts for the molecular mechanism by which periodic heat shocks induce cell synchrony in Tetrahymena. Each heat shock (34 °C) represses tubulin synthesis and blocks the division cycle at the point when the oral structure, rich in microtubules, would normally begin to assemble. Recovery (at 28 °C) from each heat shock is characterized by parallel derepression of tubulin synthesis and of oral development. Changes in protein synthesis patterns are complex when the temperature is shifted up and down between 28 and 34 °C and further experimental support is required in support of the hypothesis here forwarded.  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.
Methylhippuric acid isomers (MHAs), urinary metabolites of xylenes, were determined, after clean-up by C18-SPE and esterification with hexafluoroisopropanol and diisopropylcarbodiimide, by GC with ECD detection, on an SPB-35 capillary column (30 m, 0.32 mm I.D., 0.25 μm film thickness, β=320). S-benzyl-mercapturic acid was used for internal standardization. Chromatographic conditions were: oven temperature 162°C, for 14.2 min; ramp by 30°C/min to 190°C, for 3.5 min; ramp by 30°C/min to 250°C, for 4 min; helium flow rate: 1.7 ml/min; detector and injector temperature: 300°C. The sample (1 μl) was injected with a split injection technique (split ratio 5:1). MHA recovery was >95% in the 0.5–20 μmol/l range; the limit of detection was <0.25 μmol/l; day-to-day precision, at 2 μmol/l, was Cv<10%. Urinary MHAs were determined in subjects exposed to different low-level sources of xylenes: (a) tobacco smoking habit and (b) BTX urban air pollution (airborne xylene ranging from 0.1 to 3.7 μmol/m3). Study (a) showed a significant difference between urinary MHA median excretion values of nonsmokers and smokers (4.6 μmol/l vs. 8.1 μmol/l, p<0.001). Study (b) revealed a significant difference between indoor workers and outdoor workers (4.3 μmol/l vs. 6.9 μmol/l, p<0.001), and evidenced a relationship between MHAs (y, μmol/mmol creatinine) and airborne xylene (x, μmol/m3) (y=0.085+0.34x; r=0.82, p<0.001, n=56). Proposed biomarkers could represent reliable tools to study very low-level exposure to aromatic hydrocarbons such as those observed in the urban pollution due to vehicular traffic or in indoor air quality evaluation.  相似文献   

15.
The ConA-mediated interaction of yeast cells with macrophages was brought about in two steps. The first step involved the interaction of either macrophages or yeast cells with ConA, MConA or YConA in brief, respectively. The second step consisted of interacting the ConA-coated cells with their non-coated counterpart, yielding MConA-Y or M-YConA. The extent of yeast cell attachment to macrophages depends on the degree of saturation of ConA binding sites on the cell coated with ConA in the first step and on the temperature at which the two cell types interact. The temperature dependence in the range of 10–25 °C implies that cell-cell attachment is sensitive to the physical state of membrane lipids as reflected in increased lateral mobility of ConA receptors in the membrane plane. The extent of ConA-mediated cell association is not influenced significantly by colchicine, cytochalasin B (CB) or hydrocortisone. A mild treatment of macrophages with glutaraldehyde reduces, however, the association of yeast cells, further indicating a need for lateral mobility of ConA receptors. ConA-mediated yeast cell attachment could be totally reversed by α-methyl mannoside in the case of MConA-Y and only partially in the case of M-YConA. Yeast cell ingestion is highly temperature-dependent; in MConA-Y a 50% interiorization of the associated yeast cells is reached at 32 °C and detectable interiorization starts only above 19 °C, while in M-YConA a 50% value of interiorization is reached at 18 °C and about 15% of yeast cells are interiorized already at 5 °C. Interiorization of attached yeast cells is not affected by colchicine. Cytochalasin B (CB) (10 μg/ml) inhibits 82% of yeast interiorization in MConA-Y and only 12% in M-YConA. Hydrocortisone has a similar differential effect of inhibition of ingestion; at 25 °C inhibition in MConA-Y amounts to 78% and in M-YConA to 22%. Sodium azide inhibits 90% of interiorization of yeast cells in both MConA-Y and M-YConA. The following working hypothesis was proposed to explain both the characteristics of attachment and the remarkable difference in ingestion pattern of yeast cells in MConA-Y and M-YConA. ConA-mediated yeast cell attachment to macrophages involves multipoint interaction between the two cells achieved by a certain clustering of ConA receptors in the membrane plane. To achieve interiorization a higher extent of bridge formation between the cells is required, and a higher number of ConA-membrane receptors have to be recruited to the area of apposition of the two membranes. This requires lateral mobility of either ConA receptors conjugates (in the case of MConA) or of mobile non-crosslinked ConA receptors in macrophages interacting with YConA). Mobility of ConA receptor conjugates is more sensitive to membrane fluidity than that of non-crosslinked receptors and hence the differential temperature-dependence of ingestion. The effect of CB suggests an involvement of the cytoskeleton in the reorganization of ConA receptors at the membrane level.  相似文献   

16.
The present study was conducted to assess the capacitation status of fresh and frozen-thawed buffalo spermatozoa and its relationship with sperm cholesterol level, membrane fluidity and intracellular calcium. Semen from seven buffalo bulls (eight ejaculates each) was divided into two parts. Part I was used as fresh semen and part II was extended in Tris–egg yolk extender, equilibrated (4 °C for 4 h) and frozen at −196 °C in LN2. The fresh and frozen-thawed spermatozoa were assessed for capacitation status using chlortetracycline (CTC) fluorescent assay, membrane fluidity using merocyanine 540/Yo-Pro-1 assay and intracellular calcium using Fluo-3 AM with flowcytometry. Results revealed a significant (P < 0.01) increase in capacitated sperm population in frozen-thawed semen compared to fresh semen (42.21% vs 14.32%). Similarly, a significantly (P < 0.01) higher proportion of frozen-thawed live spermatozoa showed high membrane fluidity (53.62% vs 25.67%) and high intracellular calcium (43.68% vs 11.72%) compared to fresh semen. The sperm cholesterol was significantly (P < 0.01) reduced after freezing–thawing as compared to fresh semen. The proportion of capacitated spermatozoa (CTC pattern B) was positively correlated with the proportion of sperm with high intracellular calcium (r = 0.81) and high membrane fluidity (r = 0.65), and negatively correlated with cholesterol level (r = −0.56) in frozen-thawed semen. The membrane fluidity was also strongly associated with the cholesterol level and intracellular calcium. The study concluded that changes in buffalo spermatozoa and established the relationship among capacitation status, sperm cholesterol level, membrane fluidity and intracellular calcium concentration in frozen-thawed spermatozoa.  相似文献   

17.
Classes and mechanisms of calcium waves   总被引:3,自引:0,他引:3  
The best known calcium waves move at about 5–30 μm/s (at 20°C) and will be called fast waves to distinguish them from slow (contractile) ones which move at 0.1-1 μm/s as well as electrically propagated, ultrafast ones. Fast waves move deep within cells and seem to underlie most calcium signals. Their velocity and hence mechanism has been remarkably conserved among all or almost all eukaryotic cells. In fully active (but not overstimulated) cells of all sorts, their mean speeds lie between about 15–30 μm/s at 20°C. Their amplitudes usually lie between 3–30 μM and their frequencies from one per 10–300 s. They are propagated by a reaction diffusion mechanism governed by the Luther equation in which Ca2+ ions are the only diffusing propagators, and calcium induced calcium release, or CICR, the only reaction; although this reaction traverses various channels which are generally modulated by IP3 or cADPR. However, they may be generally initiated by a second, lumenal mode of CICR which occurs within the ER. Moreover, they are propagated between cells by a variety of mechanisms. Slow intracellular waves, on the other hand, may be mechanically propagated via stretch sensitive calcium channels.  相似文献   

18.
The present study reports on effects of different light:dark periods, light intensities, N:P ratios and temperature on the specific growth rate of flagellated cells of Phaeocystis pouchetii in culture. The specific growth rate was estimated by diel changes in cellular DNA content. The cellular DNA content and cell cycle of flagellated cells of P. pouchetii are shown, and the importance of light:dark period in cell division is demonstrated. Diel patterns of the cellular DNA content showed that cell division was confined to the dark period. The cells dealt with more than one division per day by rapid divisions shortly after each other.The specific growth rates (μDNA) based on the DNA cell cycle model were in close agreement with specific growth rates (μCell) determined from cell counts. The temperature affected the specific growth rates (multiple regression, p < 0.01) and were higher at 5 °C (μ ≤ 2.2 d−1) than at 10 °C (μ ≤1.6 d−1). Increasing the light:dark period from 12:12 h to 20:4 h affected the specific growth rate of P. pouchetii at the lower temperature tested (5 °C) (multiple regression, p < 0.01), resulting in higher specific growth rates than at 10 °C. At 10 °C, the effect of light:dark period was severely reduced. Neither light nor nutrients could compensate the reduction in specific growth rates caused by elevated temperature. The specific growth rates was not affected by the N:P ratios tested (multiple regression, p = 0.21). The experiments strongly suggest that the flagellated cells have a great growth potential and could play a dominating role in northern areas at increased day length.  相似文献   

19.
The freshwater crayfish, Orconectes virilis, shows good anoxia tolerance, enduring 20 h in N2-bubbled water at 15°C. Metabolic responses to anoxia by tolerant species often include reversible phosphorylation control over selected enzymes. To analyze the role of serine/threonine kinases and phosphatases in signal transduction during anoxia in O. virilis, changes in the activities of cAMP-dependent protein kinase (PKA) and protein phosphatases 1, 2A, and 2C were measured in tail muscle and hepatopancreas over a time course of exposure to N2-bubbled water. A strong increase in the percentage of PKA present as the free catalytic subunit (% PKAc) occurred between 1 and 2 h of anoxia exposure whereas phosphatase activities were strongly reduced. This suggests that PKA-mediated events are important in the initial response by tissues to declining oxygen availability. As oxygen deprivation became severe and prolonged (5–20 h) these changes reversed; the % PKAc fell to below control values and activities of phosphatases returned to or rose above control values. Subcellular fractionation also showed a decrease in PKA associated with the plasma membrane after 20 h anoxia whereas cytosolic PKA content increased. PKAc purified from tail muscle showed a molecular weight of 43.8±0.4 kDa, a pH optimum of 6.8, a high affinity for Mg ATP (Km=131.0±14.4 μM) and Kemptide (Km=31.6±5.2 μM). Crayfish PKAc was sensitive to temperature change; a break in the Arrhenius plot occurred at approximately 15°C with a 2.5-fold rise in activation energy at temperatures <15°C. These studies demonstrate a role for serine/threonine protein kinases and phosphatases in the metabolic adjustments to oxygen depletion by crayfish organs.  相似文献   

20.
A mammalian plasma membrane protein(s) which catalyzes ATP-dependent transbilayer movement (flip-flop) of phosphatidylserine (PS) has been suggested to be involved in the formation and maintenance of membrane lipid asymmetry. Flip-flop of PS in the cell surface of nucleated cells was first described by O. C. Martin and R. E. Pagano (1987,J. Biol. Chem.262, 5890–5898). It has been suggested that flip-flop is involved in the internalization of exogenous PS in cultured cells. In the present study we report that incubation with an excess amount of PS is cytotoxic to Chinese hamster ovary (CHO) cells, while the same amount of phosphatidylcholine gives no effect. This effect allowed us to obtain PS-resistant cells among mutagenized CHO cells. Endocytosis-independent internalization of exogenous fluorescent PS analog was defective in 40% of the PS-resistant mutants. One of the mutants, PSR (phosphatidylserine resistant) 406 was further characterized. Unlike wild-type CHO cells, this mutant did not transport fluorescent PS significantly at 15°C. Fluorescent PS was not metabolized at 15°C in either wild-type or mutant cells. These results suggest that transbilayer movement of cell surface PS is defective in PS-resistant cells.  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