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 0 毫秒
1.
Two-dimensional fluid phase connectivity is examined in mixed bilayers of dimyristoyl phosphatidylcholine and dipalmitoyl phosphatidylcholine as a function of composition and temperature at constant pressure using fluorescence recovery after photobleaching (FRAP). These isomorphous phospholipid mixtures exhibit nearly ideal mixing behavior. Dilauroyl phosphatidylethanolamine covalently linked through its amino function to NBD is used as the fluorescent probe in this study. These studies show the line of connectivity to be coincident with the line connecting the midpoints of all tie lines in the two-phase region of the phase diagram.  相似文献   

2.
The distribution of ganglioside in supported lipid bilayers has been studied by atomic force microscopy. Hybrid dipalmitoylphosphatidylcholine (DPPC)/dipalmitoylphosphatidylethanolamine (DPPE) and (2:1 DPPC/cholesterol)/DPPE bilayers were prepared using the Langmuir Blodgett technique. Egg PC and DPPC bilayers were prepared by vesicle fusion. Addition of ganglioside GM1 to each of the lipid bilayers resulted in the formation of heterogeneous surfaces that had numerous small raised domains (30--200 nm in diameter). Incubation of these bilayers with cholera toxin B subunit resulted in the detection of small protein aggregates, indicating specific binding of the protein to the GM1-rich microdomains. Similar results were obtained for DPPC, DPPC/cholesterol, and egg PC, demonstrating that the overall bilayer morphology was not dependent on the method of bilayer preparation or the fluidity of the lipid mixture. However, bilayers produced by vesicle fusion provided evidence for asymmetrically distributed GM1 domains that probably reflect the presence of ganglioside in both inner and outer monolayers of the initial vesicle. The results are discussed in relation to recent inconsistencies in the estimation of sizes of lipid rafts in model and natural membranes. It is hypothesized that small ganglioside-rich microdomains may exist within larger ordered domains in both natural and model membranes.  相似文献   

3.
In recent work [Vaz, W.L.C., Melo, E.C.C., & Thompson, T.E. (1989) Biophys. J. 56, 869-876] we have shown that translational diffusion studies using fluorescence recovery after photobleaching (FRAP) provide information concerning domain structures and fluid-phase connectivity in lipid bilayers in which solid and fluid phases coexist. In the present paper, translational diffusion of the fluid-phase-soluble, solid-phase-insoluble fluorescent lipid derivative N-(7-nitrobenzoxa-2,3-diazol-4-yl) dilauroyl-phosphatidylethanolamine and the fluid-phase connectivity are examined in lipid bilayers prepared from binary mixtures of 1-docosanoyl-2-dodecanoylphosphatidylcholine (C22:0C12:0PC) and 1,2-diheptadecanoylphosphatidylcholine (di-C17:0PC) by using FRAP. The phosphatidylcholine mixture used provides a eutectic system with a eutectic point at a composition of about 0.4 mole fraction of di-C17:0PC and a temperature of about 37 degrees C [Sisk, R.B., Wang, Z.Q., Lin, H.N., & Huang, C.H. (1990) Biophys. J. 58, 777-783]. Two regions in temperature and composition, respectively below and above 0.4 mole fraction of di-C17:0PC, where fluid and solid phases coexist in the same lipid bilayer, are available for examination of fluid-phase connectivity. In mixtures containing less than 0.4 mole fraction of di-C17:0PC the fluid phase coexists with a mixed interdigitated Lc gel phase composed mostly of C22:0C12:0PC, whereas in mixtures containing greater than 0.4 mole fraction of di-C17:0PC the fluid phase coexists with a P beta' gel phase mostly composed of di-C17:0PC. When the solid phase is a P beta' gel phase, the temperature of fluid-phase connectivity for the mixtures lies close to the fluidus, which means that a small (approximately 20%) mass fraction of solid phase can divide the large bulk of the bilayer that is fluid into nonconnected domains.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The average sizes of fluid and gel domains in the two-component, two-phase system formed from mixtures of dimyristoyl phosphatidylcholine and distearoyl phosphatidylcholine were determined from an analysis of the electron spin resonance spectral lineshapes of a dimyristoyl phosphatidylcholine-nitroxide spin label as a function of spin label concentration. The ratio, R, of the intensities measured at two magnetic field strengths was found to be diagnostic of a statistical distribution of spin labels in disconnected domains. R is defined as V'/2Vpp, where Vpp is the maximum intensity and V' is the intensity at a position in the wings of a first derivative electron spin resonance line that is a constant multiple of the peak-to-peak linewidth. The intensity ratio for Gaussian or Voigt lineshapes is less than or equal to the value for a Lorentzian lineshape. The intensity ratio was found to be greater than the value for a Lorentzian line when spectra from disconnected domains containing a statistical distribution of spin labels undergoing spin-spin interactions were summed. The intensity ratio, R, calculated by spectral simulations as a function of the average number of labels per domain, N, was found to increase to a maximum with increasing N and then to decrease. The dependence on spin label concentration of the experimentally measured intensity ratios paralleled this predicted behavior. A method is presented to calculate the average number of lipids per fluid or gel domain based on a knowledge of R, and of the distribution of the spin label between the fluid and gel phases determined from the phase diagram. The results demonstrate that the number of lipids per domain increases linearly from a fixed number of nucleation sites, as the fraction of the phase that is disconnected increases. At any given mole fraction of the particular phase, the gel domains are bigger than the fluid domains because they have a lower nucleation density. The results also suggest that the disconnected domains are, in most cases, nonrandomly distributed in the plane of the bilayer.  相似文献   

5.
W C Wimley  T E Thompson 《Biochemistry》1990,29(5):1296-1303
The rate and extent of spontaneous exchange of dimyristoylphosphatidylcholine (DMPC) from large unilamellar vesicles (LUV) composed of either DMPC or mixtures of DMPC/distearoylphosphatidylcholine (DSPC) have been examined under equilibrium conditions. The phase state of the vesicles ranged from all-liquid-crystalline through mixed gel/liquid-crystalline to all-gel. The exchange rate of DMPC between liquid-crystalline DMPC LUV, measured between 25 and 55 degrees C, was found to have an Arrhenius activation energy of 24.9 +/- 1.4 kcal/mol. This activation energy and the exchange rates are very similar to those obtained for the exchange of DMPC between DMPC small unilamellar vesicles (SUV). The extent of exchange of DMPC in LUV was found to be approximately 90%. This is in direct contrast to the situation in DMPC SUV where only the lipid in the outer monolayer is available for exchange. Thus, transbilayer movement (flip-flop) is substantially faster in liquid-crystalline DMPC LUV than in SUV. Desorption from gel-phase LUV has a much lower rate than gel-phase SUV with an activation energy of 31.7 +/- 3.7 kcal/mol compared to 11.5 +/- 2 kcal/mol reported for SUV. A defect-mediated exchange in gel-phase SUV, which is not the major pathway for exchange in LUV, is proposed on the basis of the thermodynamic parameters of the activation process. Surprisingly, the rates of DMPC exchange between DMPC/DSPC two-component LUV, measured over a wide range of compositions and temperatures, were found to exhibit very little dependence on the composition or phase configuration of the vesicles.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
J T Mason 《Biochemistry》1988,27(12):4421-4429
The mixing behavior of symmetric chain length and mixed chain length phosphatidylcholines in two-component multilamellar bilayers has been investigated by high-sensitivity differential scanning calorimetry. Phase diagrams have been constructed for two-component bilayers composed of C(18)C(18)PC and either C(18)C(16)PC, C(18)C(14)PC, C(18)C(12)PC, or C(18)C(10)PC. It is found that C(18)C(18)PC-C(18)C(16)PC and C(18)C(18)PC-C(18)C(14)PC mixed bilayers exhibit complete miscibility of the components in both the gel and liquid-crystalline phases. Whereas this mixing is observed to be nearly ideal for the C(18)C(18)PC-C(18)C(16)PC binary system, the intermixing of the lipids is highly nonideal in the gel phase of the C(18)C(18)PC-C(18)C(14)PC binary mixture. The C(18)C(18)PC-C(18)C(12)PC and C(18)C(18)PC-C(18)C(10)PC mixed bilayers are characterized by partial immiscibility of the phosphatidylcholine components in the bilayer gel phase. Over a large compositional range, these bilayers appear to consist of phase-separated regions of interdigitated and noninterdigitated gel phases. In addition, the C(18)C(18)PC-C(18)C(10)PC two-component bilayer displays a limited region of liquid-liquid immiscibility in the liquid-crystalline bilayer phase. The phase separation of the mixed chain length phosphatidylcholines revealed in these mixed bilayers may represent a three-dimensional phase separation of the lipid components where the phosphatidylcholines are both laterally separated within the plane of the bilayer and conformationally coupled across the bilayer. Such phase-separated domains could have profound effects on membrane structure and function if they were to occur in biological membranes.  相似文献   

7.
P F Almeida  W L Vaz  T E Thompson 《Biochemistry》1992,31(31):7198-7210
Fluorescence recovery after photobleaching (FRAP) has recently been used to examine the percolation properties of coexisting phases in two-component, two-phase phosphatidylcholine bilayers [Vaz, W. L. C., Melo, E. C. C., & Thompson, T. E. (1989) Biophys. J. 56, 869-876]. We now report the use of FRAP to study two additional problems in similar systems. The first is the effect of solid-phase obstacles on the lateral diffusion in the fluid phase. The second is the question of whether or not, in a single bilayer, solid-phase domains in one monolayer are exactly superimposed on solid domains in the apposing monolayer. To address the first problem, the lateral diffusion of N-(7-nitrobenzoxa-2,3-diazol-4-yl)-1-palmitoyl-2-oleoylphosp hatidylethanolamine (NBD-POPE), a probe soluble only in the fluid phase when solid and fluid phases coexist, has been studied in the mixture N-lignoceroyldihydrogalactosylceramide (LigGalCer)/dipalmitoylphosphatidylcholine (DPPC). Percolation of the fluid phase occurs at a high mass fraction of solid phase. This indicates that the solid domains have a centrosymmetric shape, a characteristic which makes this a good experimental system to test theoretical simulations of diffusion in an archipelago. It is shown that agreement between theory and experiment is poor, a result that had already been observed when the obstacles were integral membrane proteins. We develop an effective-medium model for diffusion in two-phase systems which explains both our results and those obtained with integral proteins. The distinctive feature of the model is the consideration of an annular region around the obstacles where the lipids are more ordered than in the bulk fluid phase. The diffusion coefficient is then calculated by extending the free area model to two-phase systems, taking these annuli into account. The second question, the organization of the solid-phase domains across the lipid bilayer, is examined in the systems LigGalCer/DPPC and dimyristoylphosphatidylcholine (DMPC)/distearoylphosphatidylcholine (DSPC) by comparing the diffusion of a fluid-phase-soluble, gel-phase-insoluble lipid derivative which spans the two monolayers of a bilayer (NBD-membrane-spanning-phosphatidylethanolamine, NBD-msPE) with that of a probe which is restricted to a single monolayer. In LigGalCer/DPPC, 20:80, the distribution of solid domains in one of the monolayers is independent of the distribution in the apposing monolayer. In contrast, in DMPC/DSPC, 50:50, the solid domains in one monolayer are exactly superimposed upon the solid domains existing in the apposing monolayer.  相似文献   

8.
9.
It has been repeatedly observed that lipid bilayers in the gel phase are solubilized by lower concentrations of Triton X-100, at least within certain temperature ranges, or other nonionic detergents than bilayers in the fluid phase. In a previous study, we showed that detergent partition coefficients into the lipid bilayer were the same for the gel and the fluid phases. In this contribution, turbidity, calorimetry, and 31P-NMR concur in showing that bilayers in the gel state (at least down to 13-20°C below the gel-fluid transition temperature) become saturated with detergent at lower detergent concentrations than those in the fluid state, irrespective of temperature. The different saturation may explain the observed differences in solubilization.  相似文献   

10.
Tilghman RW  Hoover RL 《FEBS letters》2002,518(1-3):83-87
The homeobox gene Cdx1 is a regulator of intestinal epithelial cell proliferation and differentiation. Using a transfection approach, we showed here that the oncogenic activation of the beta-catenin pathway stimulates the endogenous expression of the Cdx1 mRNA as well as the activity of the Cdx1 promoter in cancer cells of the human colon. Reciprocally, the paralogue homeobox gene Cdx2 exerts an inhibitory effect on the basal and on the beta-catenin-stimulated activity of the Cdx1 promoter. The inhibitory effect of CDX2 requires the intact homeodomain. It is not dependent on canonical CDX binding sites in the Cdx1 promoter nor on the cis-elements specifically targeted by the beta-catenin/Tcf complex. We conclude that the oncogenically activated beta-catenin and CDX2 have opposite and independent effects on the Cdx1 homeobox gene.  相似文献   

11.
The special physical state of the sphingolipid-enriched membranes with characteristic lipid composition, presently one of the most controversial foci in cell biology, provides the essential environment for the proteins inside to be involved in the related physiological processes. The role of gangliosides, an important component of the membranes, deserves attention. The present investigation using several biophysical techniques indicates that ganglioside GM(1) induces the phase separation in the sphingomyelin membrane with 5 mol% cholesterol and regulates the membrane structure. The results of differential scanning calorimetry show that a higher T(m), GM(1)-rich phase emerges behind the lower T(m), sphingomyelin-rich phase with the incorporation of GM(1) into the sphingomyelin/cholesterol bilayers; and the GM(1)-rich phase dominates the membrane when the proportion of GM(1) reaches about 20 mol%. Fluorescence quenching further shows that the separation of the two domains is independent of temperature, occurring both in the gel phase and in the liquid phase. Laser Raman spectroscopy and fluorescence polarization suggest that the order of hydrocarbon chains increases and membrane fluidity decreases with increase in GM(1) content. Use of the fluorescence probe merocyanine-540 and electron microscopy reveals that the insertion of GM(1) leads to an increase in the spatial density of the lipid headgroups and a decrease in the curvature of the sphingomyelin/cholesterol bilayers. In sums, both the hydrophilic sugar heads and the hydrophobic hydrocarbon chains of GM(1) contribute to the regulation of membrane architecture. We suggest that the convex curvature of ganglioside-enriched membrane could be involved in forming and maintaining the characteristic flask-shaped invagination of caveolae.  相似文献   

12.
Following injury of their left sciatic nerves by means of a standardized procedure, male rats received intravenous injections of a tritiated ganglioside. GM1, on different days during the process of regeneration. The rats were killed at two different times after the injection and the concentrations of the total radioactivity, nonvolatile radioactivity, and labelled GM1 were estimated in six segments of the crushed and intact sciatic nerves. The segments of the damaged nerves showed higher concentrations of radioactivity and a higher content of GM1 than the corresponding segments of the contralateral nerves. Within the immediate area of the lesion the highest levels were found on the 3rd and 6th days after the injury; the segments distal from the lesion showed the highest levels of activity on days 9 and 12. The nerve segments proximal to the site of the injury showed a low rate of radioactivity incorporation. The higher concentrations of [3H]GM1 in damaged nerves as well as the rate of incorporation as a function of time indicate that exogenous gangliosides may be involved in the processes of regeneration and have a bearing on the latter.  相似文献   

13.
GM1 ganglioside was found to increase the survival of PC12 cells exposed to H2O2, its action was blocked by Trk tyrosine kinase inhibitor K-252a. Thus, the inhibition of H2O2 cytotoxic action by GM1 constituted 52.8 ± 4.3%, but in the presence of 1.0 μM K-252a it was only 11.7 ± 10.8%, i.e. the effect of GM1 became insignificant. Exposure to GM1 markedly reduced the increased accumulation of reactive oxygen species (ROS) and diminished the inactivation of Na+,K+-ATPase induced in PC12 cells by H2O2, but in the presence of K-252a GM1 did not change these metabolic parameters. The inhibitors of extracellular signal-regulated protein kinase, phosphatidyl inositol 3-kinase and protein kinase C decreased the effects of GM1. A combination of these protein kinase inhibitors reduced inhibition of H2O2 cytotoxic action by GM1 to the larger extent than each of the inhibitors and practically abolished the ability of GM1 to decrease H2O2-induced ROS accumulation. The protective and antioxidative effects of GM1 in PC12 cells exposed to H2O2 appear to be mediated by activation of Trk receptor tyrosine kinase and the protein kinases downstream from this enzyme.  相似文献   

14.
We examined the action of porcine pancreatic and bee-venom phospholipase A2 towards bilayers of phosphatidylcholine as a function of several physical characteristics of the lipid-water interface. 1. Unsonicated liposomes of dimyristoyl phosphatidylcholine are degraded by both phospholipases in the temperature region of the phase transition only (cf. Op den Kamp et al. (1974) Biochim. Biophys. Acta 345, 253--256 and Op den Kamp et al. (1975) Biochim. Biophys. Acta 406, 169--177). With sonicates the temperature range in which hydrolysis occurs is much wider. This discrepancy between liposomes and sonicates cannot be ascribed entirely to differences in available substrate surface. 2. Below the phase-transition temperature the phospholipases degrade dimyristoyl phosphatidylcholine single-bilayer vesicles with a strongly curved surface much more effectively than larger single-bilayer vesicles with a relatively low degree of curvature. 3. Vesicles composed of egg phosphatidylcholine can be degraded by pancreatic phospholipase A2 at 37 degrees C, provided that the substrate bilayer is strongly curved. The bee-venom enzyme shows a similar, but less pronounced, preference for small substrate vesicles. 4. In a limited temperature region just above the transition temperature of the substrate the action of both phospholipases initially proceeds with a gradually increasing velocity. This stimulation is presumably due to an increase of the transition temperature, effectuated by the products of the phospholipase action. 5. Structural defects in the substrate bilayer, introduced by sonication below the phase-transition temperature (cf. Lawaczeck et al. (1976) Biochim. Biophys. Acta 443, 313--330) facilitate the action of both phospholipases. The results lead to the general conclusion that structural irregularities in the packing of the substrate molecules facilitate the action of phospholipases A2 on phosphatidylcholine bilayers. Within the phase transition and with bilayers containing structural defects these irregularities represent boundaries between separate lipid domains. The stimulatory effect of strong bilayer curvature can be ascribed to an overall perturbation of the lipid packing as well as to a change in the phase-transition temperature.  相似文献   

15.
Pulsed field gradient (pfg)-NMR measurements of the lipid lateral diffusion coefficients in several macroscopically aligned bilayer systems were summarized from previous and new studies. The aim was to carry out a comparison of the translational dynamics for bilayers with various mixtures of l,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC), l,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and chicken egg yolk sphingomyelin (eSM), with or without cholesterol. New useful information was obtained on the dynamics in these lipid bilayers that has not been previously appreciated. Thus, we were able to propose that the driving force behind the phase separation into l(d)and l(o)phases evolves from the increasing difficulty to incorpotate DOPC into a highly ordered phase. Our results suggest that DOPC has a preference to be located in a disordered phase, while DPPC and eSM prefer the ordered phase. Quite unexpectedly, CHOL seems to partition into both phases to roughly the same extent, indicating that CHOL has no particular preference for any of the l(d)or l(o) phases, and there are no specific interactions between CHOL and saturated lipids.  相似文献   

16.
We previously reported that 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) forms an interdigitated gel phase in the presence of 1-palmitoyl-sn-glycero-3-phosphocholine (16:0LPC) at concentrations below 30 mol%. In the present investigation, fluorescent probe 1,6-diphenyl-1,3,5-hexatriene (DPH), X-ray diffraction, and differential scanning calorimetry (DSC) were used to investigate the effect of cholesterol on the phase behavior of 16:0LPC/DPPC binary mixtures. At 25 degrees C, 30 mol% 16:0LPC significantly decreases the DPH fluorescence intensity during the transition of DPPC from the L(beta') phase to the L(betaI) phase. However, the addition of cholesterol to 16:0LPC/DPPC mixtures results in a substantial increase in fluorescence intensity. The changes in DPH fluorescence intensity reflect the probe's redistribution from an orientation parallel to the acyl chain to the center of the bilayer, suggesting a bilayer structure transition from interdigitation to noninterdigitation. The normal repeat period of small angle X-ray diffraction patterns can be restored and a reflection appears at 0.42 nm with a broad shoulder around 0.41 nm in wide angle X-ray diffraction patterns when 10 mol% cholesterol is incorporated into 30 mol% 16:0LPC/DPPC vesicles, indicating that the mixtures are in the gel phase (L(beta')). Moreover, DSC results demonstrate that 10 mol% cholesterol is sufficient to significantly decrease the main enthalpy, cooperativity and lipid chain melting of 30 mol% 16:0LPC/DPPC binary mixtures, which are L(betaI), indicating that the transition of the interdigitated phase is more sensitive to cholesterol than that of the noninterdigitated phase. Our data imply that the interdigitated gel phase induced by 16:0LPC is prevented in the presence of 10 mol% cholesterol, but unlike ethanol, an increasing concentration of 16:0LPC is not able to restore the interdigitation structure of the lipid mixtures.  相似文献   

17.
The effects of bovine alpha-lactalbumin on the thermotropic properties of dimyristoylphosphatidylcholine liposomes are studied by Raman spectroscopy, fluorescence polarization and differential scanning calorimetry. The Raman spectrum reveals the drastic effects of the protein on the phospholipid structure. The transition temperature shifts downwards and the inter- and intrachain order in the lipid matrix progressively diminish with increasing protein concentration. Up to a lipid to protein molar ratio R = 25, the bilayer structure however is maintained. From fluorescence polarization data we conclude that the protein restricts the mobility of the DPH probe. In view of the Raman results, the lower probe mobility obviously cannot be associated with a more rigid lipid matrix. Nevertheless the transition temperatures of the alpha-lactalbumin-phospholipid complex increases. DSC measurements give no decisive way out for this discrepancy. These results confirm that different types of lipid order are involved in lipid-protein interactions. Compared to the free protein, the alpha-helicity of the protein has increased in the complex.  相似文献   

18.
Ubiquinones (n = 1,2,3,4,7,9,10) and ubiquinols (n = 1,2,3,4,10) were incorporated into ordinary (protonated) or perdeuterated dimyristoyl phosphatidylcholine vesicles and were found to have significant local molecular motion. The motion of the quinone ring, as judged from the linewidth of the OCH3 proton resonances, decreased in longer-chain ubiquinones. Minimum values for the transverse mobility (flip-flop rates) of ubiquinones-1,2,3,4,10, measured with the aid of lanthanide shift reagents, suggest that they are all able to function in a protonmotive ‘Q cycle’ during electron transport. As the length of the side chain increases beyond 1 isoprenoid unit, the quinone/quinol ring tends to be deeper in the outer monolayer of small sonicated vesicles and in both monolayers of larger freeze-thaw vesicles, but little or no change in depth is observed in the inner monolayer of small vesicles. The ubiquinol rings are closer to the membrane surface than are the ubiquinone rings. For side chain n = 9 or 10, a second resonance from the OCH3 protons of ubiquinones and ubiquinols in vesicles appears in the 1H-NMR spectrum. This is due to the presence of two types of vesicles with different ubiquinone/phospholipid ratios.  相似文献   

19.
We have investigated the question of the presence of a cap structure located at the top of the F(1) alpha(3)beta(3) hexamer of the yeast mitochondrial F(1)F(0)-ATP synthase complex. Specifically, we sought to determine whether the putative cap has a rigid structure and occludes the central shaft space formed by the alpha(3)beta(3) hexamer or alternatively whether the cap is more flexible permitting access to the central shaft space under certain conditions. Thus, we sought to establish whether subunit gamma, an essential component of the F(1) central stalk housed within the central shaft space and whose N and C termini would both lie beneath a putative cap, could be fused at its C terminus to green fluorescent protein (GFP) without loss of enzyme function. The GFP moiety serves to report on the integrity and location of fusion proteins containing different length polypeptide linkers between GFP and subunit gamma, as well as being a potential occluding structure in itself. Functional incorporation of subunit gamma-GFP fusions into ATP synthase of yeast cells lacking native subunit gamma was demonstrated by the ability of intact complexes to hydrolyze ATP and retain sensitivity to oligomycin. Our conclusion is that the putative cap structure cannot be an inflexible structure, but must be of a more flexible nature consistent with the accommodation of subunit gamma-GFP fusions within functional ATP synthase complexes.  相似文献   

20.
It is known that gangliosides, being ubiquitous membrane components, play important roles in cell-cell recognition, differentiation and transmembrane signalling. GM3, GM1 and GD1a were detected in the rat oviduct as major gangliosides by thin-layer chromatography (TLC) analysis. The total amounts of gangliosides from the oviducts at various times after hormone injection were not much changed. In order to identify their distribution and possible changes during ovulation, frozen sections of the rat oviducts were stained with specific monoclonal antibodies (MAbs) against the ganglio-series gangliosides. GM3 and GM1 were expressed in a different manner, but GD1a and other gangliosides were not immunohistochemically detected. In the ampullar region, GM3 was expressed in all the stroma and epithelial cells, but not GM1. GM1 was also not observed in epithelial cells. Staining by anti-GM1 monoclonal antibodies revealed long and minute thread-like structures in some of the stroma cells, whereas anti-GM3 monoclonal antibodies stained the entire cytoplasm, but not the nucleus, of all the stroma and epithelial cells. Other ganglio-series gangliosides, including GD1a, were not detected to some extent in the ampullar region by immunohistochemistry. Thus, these data suggest that GM3 and GM1 are oviduct-specific gangliosides.  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