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 390 毫秒
1.
The fluorescence polarization technique with 1,6-diphenyl 1,3,5-hexatriene as a probe was used to determine the lipid microviscosity, η, of isolated plasma membranes of mouse thymus-derived ascitic leukemia (GRSL) cells and of extracellular membraneous vesicles exfoliated from these cells and occurring in the ascites fluid. For comparison, η was also determined in isolated plasma cell supernatants.For isolated plasma membranes of thymocytes and GRSL cells η values at 25° C amounted to 4.67 and 3.28 P, respectively, which were higher than the microviscosities of the corresponding intact cells, 3.24 and 1.73 P, respectively.Microviscosities inextracellular membranes of thymocytes and GRSL cells were 5.96 and 5.83 P, respectively. The fluidity difference between these membranes and plasma membranes was most pronounced for the leukemic cells and was thereby correlated with a large difference in cholesterol/phospholipid molar ratio (1.19 for extracellular membranes and 0.37 for plasma membranes). It is proposed that extracellular membraneous vesicles are shed from the surface of GRSL cells similar to the budding process of viruses, that is by selection of the most rigid parts of the host cell membrane.Liposomes of total lipid extracts of plasma membranes and extracellular membranes of both cell types exhibited about the same microviscosity as the corresponding intact membranes, indicating virtually no contribution of (glyco)-protein to the lipid fluidity as measured by the fluorescence polarization technique. For both cell types η (25° C) values of liposomes consisting of membrane phospholipids varied between 1.5 and 1.9 P, much lower than the values for total lipids, indicating a significant rigidizing effect of cholesterol in each type of membrane.  相似文献   

2.
Nanosecond decays of the fluorescence anisotropy, r, were studied for the emission of 1,6-diphenyl-l,3,5-hexatriene (DPH) embedded in a series of mixed multilamellar liposomes containing egg yolk phosphatidylcholine, phosphatidylethanolamine and cholesterol in varying molar ratios, as well as in membranes of intact cells and in virus envelopes.The relative contributions of the fast and the infinitely slow decaying component to the steady-state value, r, of the fluorescence anisotropy were very similar for artifical and biological membranes.Angles, θ, of the cone, by which the motion of the fluorescent molecule is limited, were calculated from the intensity of the infinitely slow decaying anisotropy component and compared with steady-state fluorescence anisotropies and with ‘microviscosities’, 〈η〉. An increase in 〈η〉 from 1.5 to 5.2 P in our systems was accompanied by a decrease in θ from 49° to 30° while the decrease in the mean motional relaxation times, φf, of the label molecule was not more than 1 ns and due mainly to changes in the potential, by which the diffusion of DPH in the membrane is restricted. From these observations we conclude that differences in the steady-state fluorescence anisotropy and in ‘microviscosities’ of cholesterol-containing membranes (r > 0.15) represent changes in the degree of static orientational constraint rather than changes in diffusion rates of the label.  相似文献   

3.
A wide range of concentrated random coil polysaccharide solutions have been assessed for textural attributes by a trained sensory panel. The only textural terms invoked to describe these model systems were ‘thickness’ and ‘stickiness’, which were shown to be highly correlated, and essentially identical numerically, using a ratio scaling technique. Viscosity (η) measurements over a wide range of shear rates (γ) for all these samples gave flow curves (log η versus log γ) of the same form. Differences in flow behaviour between samples could then be characterised completely by two parameters, the maximum viscosity at low shear rates (η0), and the shear rate (γ?0·1) at which η = solη010. A simple linear relationship was demonstrated between these two parameters and perceived thickness (T) or stickiness (S), irrespective of polysaccharide type. For Newtonian liquids, log T (or log S) varied linearly with log η. Hence the effective ‘in-mouth’ thickness of random coil polysaccharide solutions, in normal viscosity units, may be predicted directly from η0 and γ?0·1 by the simple relationship: log ηN = 1·13 log η0 + 0·45 logγ?0·1 ? 1·72 where ηN is the viscosity of a Newtonian solution which would be perceived as identical in thickness (and stickiness) to the polysaccharide solution.  相似文献   

4.
5.
Mouse L cell interferon (IFN αβ) inhibited the differentiation of mouse 3T3-Li fibroblasts into adipocytes. IFN αβ also inhibited hexose monophosphate shunt (HMP) activity in these cells. HMP activity is required for the reducing power necessary for the conversion of 3T3-Li fibroblasts to adipocytes. Both IFN blockage of differentiation and HMP activity were reversed by the reducing agent 2-mercaptoethanol, probably through interaction with membrane receptors and not through direct inactivation of IFN. Several non-antiviral effects of IFN αβ on cellular function, including differentiation and immunoregulation, may be mediated at the biochemical level through blockage of HMP activity.  相似文献   

6.
A connection between the processes of cell death and differentiation is suggested by observations which show that chemical inducers of differentiation are cytotoxic to CCRF-CEM human leukaemic lymphoblasts, cells which have properties typical of immature lymphoid cells. Sodium n-butyrate, salts of other short-chain fatty acids, 5-azacytidine, hypoxanthine, L-ethionine and dimethyl sulphoxide were all cytotoxic to these cells at concentrations similar to those reported to produce reversible growth inhibition in more mature lymphocytes or growth inhibition and differentiation in other cell types. Only actively cycling cells were susceptible to killing by n-butyrate. Inhibitory effects of these compounds on DNA methylation are postulated to be responsible for their cytotoxic actions.  相似文献   

7.
The human embryonal carcinoma cell lines NT2D1 and NT2B9, clonally derived from Tera-2, differentiate extensively in vitro when exposed to retinoic acid. This differentiation is marked by the appearance of several morphologically distinct cell types and by changes in cell surface phenotype, particularly by the disappearance of stage-specific embryonic antigen-3 (SSEA-3), which is characteristically expressed by human EC cells. Among the differentiated cells are neurons, which form clusters interconnected by extended networks of axon bundles, and which express tetanus toxin receptors and neurofilament proteins. These observations constitute the first instance of extensive somatic differentiation of a clonal human EC cell line in vitro.  相似文献   

8.
The genesis and transmission of action potentials in epidermal cells of the newt (Cynops pyrrhogaster) embryo were investigated with special reference to cellular differentiation during development. Typical action potentials can be recorded from any of the epidermal cells at Stage 31. These potentials consist of a fast spike (18 msec) followed by a slow component (164 msec). The potential is graded with current intensity, and only the slow component initiates action potentials in adjacent cells and induces a transmission to other cells. The fast spike was found in all epidermal cells throughout the embryonic stages examined (Stages 26–47). The slow potential, however, appears at Stage 28, persists until Stage 3637 just before hatching and then disappears at Stage 3842. Electrical recordings from traumatic embryos (embryos without neural crest cells) or from cultured epidermal cell masses isolated from the pregastrula or the ventral region of the neurula, were compared with the intact embryo. No differences were observed in either the form of the action potential or its transmission. Thus these action potentials appear to be derived from epidermal cells, and are not of nervous origin. Evidence suggests that the transient establishment of excitable membranes in epidermal cells during differentiation is closely related to neural cell differentiation.  相似文献   

9.
HL-60, a human promyelocytic leukemia cell line, is induced to differentiate by retinoic acid to mature granulocytes. We have now found that after the addition of 1 μM retinoic acid to HL-60 cultures an increase in NAD+-glycohydrolase (NADase) activity is detected by 6 hr and after a 33-fold increase in activity reaches a plateau by 24 hr. Cycloheximide inhibits completely the retinoic acid-induced increase in NADase activity indicating that enzyme induction requires protein synthesis de, novo. An increase of NADase activity was found not only in HL-60 cells but also in two human monoblast cell lines (U-937 and THP-1) and fresh cells in primary culture from two patients with acute promyelocytic leukemia. An increase in synthesis de, novo of NADase does not appear to be obligatory for differentiation of HL-60 because there was no increase of NADase activity in HL-60 cells induced to differentiate with either dimethylsulfoxide, hypoxanthine, butyrate, or 1, 25-dihydroxycholecalciferol and there were marked increases in NADase activity at concentrations of retinoic acid having little or no effect on differentiation.  相似文献   

10.
Based on the knowledge that dimethyl sulfoxide (DMSO) can induce differentiation and function of several types of immature neoplastic cell lines, we examined the effects of DMSO on the function of normal pituitary cells with high activity and found that it stimulated the synthesis and release of growth hormone and prolactin invitro in the anterior pituitary from lactating mouse.  相似文献   

11.
The migration of intestinal epithelial cells from the crypts to the tips of villi is associated with progressive cell differentiation. The changes in Na+-pump levels during migration have been measured in epithelial cells isolated from rabbit small intestine. A significant proportion of ouabain-sensitive (Na++K+)-ATPase in the cell homogenates was latent but could be unmasked by detergent treatment. Highest detergent activation was observed in villus cells. The distribution of pumping sites was also assessed by measuring ouabain binding to intact cells. The kinetics of specific binding was consistent with the interaction of the cardiac glycoside with a single population of binding sites with an apparent Kd of around 10?7 M. Both enzyme assay and ouabain-binding measurements suggest that a 2–3-fold increase in the number of Na+-pumping sites accompanies cell differentiation in rabbit jejunal epithelium. This increase in pumping capacity might be an adaptation of the cells to their absorptive function.  相似文献   

12.
Inorganic selenium compounds are shown to be inducers of hemoglobin synthesis in malignant murine erythroleukemia (MEL) cells. SeO2 can induce hemoglobin synthesis at 120 the concentration of butyric acid and 15000 the concentration of dimethylsulfoxide (DMSO), two potent inducers of erythroid differentiation in MEL cells. SeO2 and H2SeO3 showed an equivalent capacity to stimulate hemoglobin synthesis in three different MEL cell lines. The incorporation of 3H-glycine into hemoglobin was demonstrated in lysates of SeO2-induced MEL cells.  相似文献   

13.
The accumulation of α- and β-globin mRNA sequences in murine erythroleukemia cells (MELC) treated with various inducers has been studied using specific α- and β-globin complementary DNAs (cDNAs). In cells cultured with dimethylsulfoxide (Me2SO), hexamethylene bisacetamide (HMBA) or butyric acid, accumulation of α-globin mRNA is detectable after 16, 12 and 8 hr of culture, respectively. An increase in β-globin mRNA sequences is not detected until 20–24 hr after culture. In cells exposed to hemin, both α- and β-globin mRNAs are detectable by 6 hr of culture, and a constant ratio of αβ-mRNA is maintained during induction. In maximally induced cells, the αβ-globin mRNA ratios are approximately 1 in cells induced by Me2SO and HMBA, and 0.66 and 0.3–0.50 in cells induced by butyric acid and hemin, respectively. Thus different inducers of erythroid differentiation in MELC lead to different times of onset of the expression of α- and β-like genes. In addition, the relative accumulation of α- and β-globin mRNAs in induced cells differs with various types of inducers.  相似文献   

14.
The molecular weight distribution of sinistrin (Inutest ®, Laevosan Ges., Linz, Austria), determined by analytical gel-permeation chromatography, using narrow fractions (MwMn< 1.07) obtained by preparative gel-permeation chromatography, covered the range 800–16,000 with Mn  2,500 and Mw  3,500. From viscosity measurements on dilute, aqueous solutions, the relation [η]  0.28 X M0.3 was obtained, indicating a branched molecular structure; the largest molecules can be described by a sphere with r  23 Å. Comparison of the content of glucose and reducing sugars in the fractions with the molecular weight determined by vapour-pressure osmometry indicated that a glucose end-group is present in the majority of the molecules. The percentage of glucose end-groups is higher in the fractions of lower molecular weight. From this finding, speculations on the biosynthesis of sinistrin are made. The specific optical rotation of sinistrin fractions decreases linearly with 1/Mn.  相似文献   

15.
Induction of erythroid differentiation in ouabain-resistant murine erythroleukemia cells by ouabain is reported. Ouabain induction results in the appearance of hemoglobin-containing cells 12–24 hr earlier than induction of the same clone by dimethyl sulfoxide. The levels of globin mRNA after ouabain induction are similar in amount to the globin mRNA levels observed after induction by dimethyl sulfoxide. The concentration of ouabain required to induce hemoglobin synthesis depends upon the K+ ion levels in the culture medium. Lowering the extracellular K+ ion concentration 2–4 fold reduced by 10–40 fold the ouabain concentration necessary for the induction of hemoglobin synthesis. In low K+ medium (1.8 mM), ouabain is an effective inducer of hemoglobin synthesis at a concentration of 0.02 mM. This K+ effect is specific for ouabain induction, since induction by other inducers, such as dimethyl sulfoxide and dimethyl acetamide, does not exhibit this marked sensitivity to the levels of K+ ions in the culture medium. These results suggest that the binding of ouabain to the plasma membrane enzyme, NaK ATPase, is required for the induction of erythroid differentiation by ouabain. A small but significant proportion of wild-type, ouabain-sensitive cells also can be induced by ouabain, below ouabain concentrations that are toxic to these cells. The observation that the binding of ouabain to the NaK ATPase induces hemoglobin synthesis suggests that changes in the intracellular concentration of K+ ions may be involved in the control of erythroid differentiation in Friend erythroleukemic cells.  相似文献   

16.
Dose-response relations for pokeweed mitogen (PWM)-induced B- and T-cell proliferation and differentiation of human peripheral blood B lymphocytes were derived. For each tested concentration of PWM used in stimulating mononuclear cells, proliferation, assayed by cell population size and distribution of cells with respect to cell cycle phases; and differentiation, assayed by incidence of cytoplasmic immunoglobulin, were determined as a function of time following PWM stimulation. Balanced T- and non-T-cell proliferation occurred without necessarily being associated with B-cell differentiation. Differentiation, in contrast, was not observed without proliferation. The onset of balanced T- and non-T-cell proliferation preceded the differentiation of B lymphocytes into plasmacytoid cells bearing detectable cytoplasmic immunoglobulin. The dose-response relations for PWM-induced proliferation and differentiation were dissimilar. Optimum proliferation occurred at a PWM concentration 1100 of that required to induce differentiation. The results indicate that while B- and T-cell proliferation may be necessary for B-cell differentiation, it is not sufficient. Proliferation can be uncoupled from differentiation. The dissimilarity of the dose-response relations for the two responses makes it improbable that PWM triggers a unique cellular process seminal to proliferation coupled inevitably to subsequent differentiation.  相似文献   

17.
We have investigated the interaction of three lectins, differing in their sugar specificities, with the surface of the three differentiation stages of Trypanosoma cruzi. The Scatchard constants for each lectin and parasite stage imply that differentiation of T. cruzi is accompanied by changes in the cell surface saccharides. Trypomastigotes obtained from two different sources do not differ appreciably as to the number and affinity of binding sites for the three lectins employed, suggesting a similar cell-surface saccharide composition. These conclusions are reinforced by sodium dodecyl sulfate-polyacrylamide gel electrophoretic analysis of the 131I-labeled surface glycoproteins, following isolation by affinity chromatography. The surface membrane of trypomastigotes, the infective stage to T. cruzi for mammalian cells, possesses a specific glycoprotein of apparent Mr 85 000 (Tc-85) which is absent from the other two stages and can be isolated by affinity chromatography on wheat germ agglutinin-Sepharose columns. This glycoprotein also binds to concanavalin A, but not to Lens culinaris lectin. The binding of Tc-85 to wheat germ agglutinin is unnafected by treatment of either the isolated glycoprotein or intact living trypomastigotes with neuraminidase. Since N-acetyl-d-glucosamine inhibits internalization of trypomastigotes by cultured mammalian cells, it is suggested that Tc-85 might be involved in adhesion and/or interiorization of T. cruzi into mammalian cells, possibly via recognition of an ubiquitous host-cell surface N-acetyl-d-glucosamine-specific receptor activity.  相似文献   

18.
The differentiation of body-wall muscle cells was studied in the nematode Caenorhabditis elegans. Specific antibodies to myosin and paramyosin, major protein constituents of differentiated muscle, react with mesodermal cells in wild-type embryos towards the end of the first half of embryogenesis. Immunoreactive cells (2–16) first appear in embryos with 400–450 of the 550 cells present at hatching. Such embryos have developed at 25.5°C for 3–412 hr beyond the two-cell stage. As development proceeds, a maximum of 81 immunoreactive cells forms four columns running anterior-posterior. Each column is composed of two lines of tightly opposed round cells, which then elongate into spindle-shaped cells. Mutant embryos in which cleavage arrests prematurely also generate cells that produce myosin and paramyosin. The initiation of muscle differentiation appears to be independent of the number of cell or nuclear divisions within a lineage or of the proliferation of other cells. These results suggest that the biosynthesis of muscle-specific proteins by nematode embryonic muscle cells is regulated by mechanisms intrinsic to these cells.  相似文献   

19.
Ultrastructural evidence indicates that Xenopus retinal ganglion cell axons differentiate early, between stages 28 and 32. Light microscope studies indicated the presence of argryophilic material in the ventral retina and optic stalk of early embryos. Ultrastructural analysis of this region confirmed the presence of axons in the stalk and interstices of ventral retinal cells. Axons containing aligned microtubules and neurofilaments and elongated mitochondria with a paucity of other cell inclusions are found with increasing frequency in the ventral retina from stages 28 through 3334. Central and dorsal regions of the retinas examined show little or no evidence of axons. A discrete, small bundle of axons is found in the optic stalk of stage 28 embryos and by stage 3031 the number of axons in bundles has increased, suggesting early fasciculation. Between stages 28 and 3334 (± 12 hr) extracellular space surrounding early axons diminishes and processes from neuroretinal cells in contact with axons surround developing axon bundles. The evidence presented suggests that axon initiation occurs in stages much earlier than previously reported. Other investigators have failed to detect ganglion cell differentiation prior to stage 32 possibly because they examined regions of the retina with few axons. Thus, experiments which rotate the retina in the orbit may have to be reevaluated since regenerating axons may use previously established pathways to organize and “home in” on tectal target cells.  相似文献   

20.
In order to assess changes in β-adrenergic receptors that occur with differentiation of myogenic cells, a method was developed to determine the number of β receptor sites in intact cells in situ. Conditions of the binding assay were established to maximize specific binding of [125I]iodohydroxybenzylpindolol to cells and to insure that neither trypsin nor EDTA used for passaging purposes interfered with the assessment of binding sites. With differentiation, a 3-fold increase in β-adrenergic receptors in intact cells was observed. Characteristics of the β receptors when compared to cells prior to differentiation were otherwise unchanged with similar association and dissociation kinetics and binding affinity. These results confirm previous observations in broken cell preparations, provide a reasonable estimate for the number of β receptors in these cells, both before and after differentiation, and provide support for the possibility that the β-adrenergic catecholamines are important in the differentiation of myogenic 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