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.
S Malinchik  S Xu    L C Yu 《Biophysical journal》1997,73(5):2304-2312
By using synchrotron radiation and an imaging plate for recording diffraction patterns, we have obtained high-resolution x-ray patterns from relaxed rabbit psoas muscle at temperatures ranging from 1 degree C to 30 degrees C. This allowed us to obtain intensity profiles of the first six myosin layer lines and apply a model-building approach for structural analysis. At temperatures 20 degrees C and higher, the layer lines are sharp with clearly defined maxima. Modeling based on the data obtained at 20 degrees C reveals that the average center of the cross-bridges is at 135 A from the center of the thick filament and both of the myosin heads appear to wrap around the backbone. At 10 degrees C and lower, the layer lines become very weak and diffuse scattering increases considerably. At 4 degrees C, the peak of the first layer line shifts toward the meridian from 0.0047 to 0.0038 A(-1) and decreases in intensity approximately by a factor of four compared to that at 20 degrees C, although the intensities of higher-order layer lines remain approximately 10-15% of the first layer line. Our modeling suggests that as the temperature is lowered from 20 degrees C to 4 degrees C the center of cross-bridges extends radially away from the center of the filament (135 A to 175 A). Furthermore, the fraction of helically ordered cross-bridges decreases at least by a factor of two, while the isotropic disorder (the temperature factor) remains approximately unchanged. Our results on the order/disordering effects of temperature are in general agreement with earlier results of Wray [Wray, J. 1987. Structure of relaxed myosin filaments in relation to nucleotide state in vertebrate skeletal muscle. J. Muscle Res. Cell Motil. 8:62a (Abstr.)] and Lowy et al. (Lowy, J., D. Popp, and A. A. Stewart. 1991. X-ray studies of order-disorder transitions in the myosin heads of skinned rabbit psoas muscles. Biophys. J. 60:812-824). and support Poulsen and Lowy's hypothesis of coexistence of ordered and disordered cross-bridge populations in muscle (Poulsen, F. R., and J. Lowy. 1983. Small angle scattering from myosin heads in relaxed and rigor frog skeletal muscle. Nature (Lond.). 303:146-152.). However, our results added new insights into the disordered population. Present modeling together with data analysis (Xu, S., S. Malinchik, Th. Kraft, B. Brenner, and L. C. Yu. 1997. X-ray diffraction studies of cross-bridges weakly bound to actin in relaxed skinned fibers of rabbit psoas muscle. Biophys. J. 73:000-000) indicate that in a relaxed muscle, cross-bridges are distributed in three populations: those that are ordered on the thick filament helix and those that are disordered; and within the disordered population, some cross-bridges are detached and some are weakly attached to actin. One critical conclusion of the present study is that the apparent order <--> disorder transition as a function of temperature is not due to an increase/decrease in thermal motion (temperature factor) for the entire population, but a redistribution of cross-bridges among the three populations. Changing the temperature leads to a change in the fraction of cross-bridges located on the helix, while changing the ionic strength at a given temperature affects the disordered population leading to a change in the relative fraction of cross-bridges detached from and weakly attached to actin. Since the redistribution is reversible, we suggest that there is an equilibrium among the three populations of cross-bridges.  相似文献   

2.
Previously it was shown that gramicidin can induce HII phase formation in diacylphosphatidylcholine model membranes only when the lipid acyl chain length exceeds 16 carbon atoms (Van Echteld, C.J.A., De Kruijff, B., Verkleij, A.J., Leunissen-Bijvelt, J. and De Gier, J. (1982) Biochim. Biophys. Acta 692, 126-138). Using 31P-NMR and small angle X-ray diffraction we now demonstrate that upon increasing the length of gramicidin, the peptide loses its ability to induce HII phase formation in di-C18:1c-PC but not in the longer chained di-C22:1c-PC. It is concluded that a mismatch in length between gramicidin and the lipid acyl chains, when the latter would provide excess bilayer thickness, is a prerequisite for HII phase formation in phosphatidylcholine model membranes.  相似文献   

3.
Induction of lateral phase separations in binary lipid mixtures by alcohol   总被引:5,自引:0,他引:5  
E S Rowe 《Biochemistry》1987,26(1):46-51
It has previously been shown that alcohol has different effects on the gel to liquid-crystal phase transition of phosphatidylcholines (PC's) and phosphatidylethanolamines (PE's) [Rowe, E. S. (1985) Biochim. Biophys. Acta 813, 321-330]. In this investigation, the thermotropic properties of binary PE-PC mixtures were studied in the presence of ethanol in order to determine whether the differential interactions of alcohol with PC and PE would lead to lateral phase separations. Phase diagrams of the dilaurylphosphatidylethanolamine-dipalmitoylphosphatidylcholine [PE(12:0)-PC(16:0)] system were constructed in the presence and absence of ethanol. It was shown that lateral phase separations occur in the gel phase over a certain composition range in the presence of 100 mg/mL ethanol. In the absence of alcohol these two lipids are miscible in both the gel and liquid-crystal states. The data suggest that in the presence of ethanol these lateral phase separations involve the coexistence of regular bilayer gel and the fully interdigitated gel phase, which has previously been shown to occur in pure PC(16:0) under these conditions [Simon, S. A., & McIntosh, T. J. (1984) Biochim. Biophys. Acta 773, 169-172]. The biological implications of these findings are discussed.  相似文献   

4.
S Xu  J Gu  T Rhodes  B Belknap  G Rosenbaum  G Offer  H White    LC Yu 《Biophysical journal》1999,77(5):2665-2676
The thick filaments of mammalian and avian skeletal muscle fibers are disordered at low temperature, but become increasingly ordered into an helical structure as the temperature is raised. Wray and colleagues (Schlichting, I., and J. Wray. 1986. J. Muscle Res. Cell Motil. 7:79; Wray, J., R. S. Goody, and K. Holmes. 1986. Adv. Exp. Med. Biol. 226:49-59) interpreted the transition as reflecting a coupling between nucleotide state and global conformation with M.ATP (disordered) being favored at 0 degrees C and M.ADP.P(i) (ordered) at 20 degrees C. However, hitherto this has been limited to a qualitative correlation and the biochemical state of the myosin heads required to obtain the helical array has not been unequivocally identified. In the present study we have critically tested whether the helical arrangement of the myosin heads requires the M.ADP.P(i) state. X-ray diffraction patterns were recorded from skinned rabbit psoas muscle fiber bundles stretched to non-overlap to avoid complications due to interaction with actin. The effect of temperature on the intensities of the myosin-based layer lines and on the phosphate burst of myosin hydrolyzing ATP in solution were examined under closely matched conditions. The results showed that the fraction of myosin mass in the helix closely followed that of the fraction of myosin in the M.ADP.P(i) state. Similar results were found by using a series of nucleoside triphosphates, including CTP and GTP. In addition, fibers treated by N-phenylmaleimide (Barnett, V. A., A. Ehrlich, and M. Schoenberg. 1992. Biophys. J. 61:358-367) so that the myosin was exclusively in the M.ATP state revealed no helical order. Diffraction patterns from muscle fibers in nucleotide-free and in ADP-containing solutions did not show helical structure. All these confirmed that in the presence of nucleotides, the M.NDP.P(i) state is required for helical order. We also found that the spacing of the third meridional reflection of the thick filament is linked to the helical order. The spacing in the ordered M.NDP.P(i) state is 143.4 A, but in the disordered state, it is 144. 2 A. This may be explained by the different interference functions for the myosin heads and the thick filament backbone.  相似文献   

5.
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   

6.
The nonexponential closed-time distributions observed for ionic channels have been explained recently by quasi-one-dimensional models of structural diffusion (Millhauser, G. L., E. E. Salpeter, and R. E. Oswald. 1988. Proc. Natl. Acad. Sci. USA. 85: 1503-1507; Condat, C. A., and J. J?ckle. 1989. Biophys. J. 55: 915-925; Levitt, D. G. 1989. Biophys. J. 55: 489-498). We generalize this treatment by allowing for more complex trajectories using percolation theory. We assume that the gating transition depends on marginally connected conformational states leading to the observed spread in time scales.  相似文献   

7.
Direct measurements of phosphorylation of the Ca2+ ATPase of the sarcoplasmic reticulum (SR) have shown that the lifetime of the first phosphorylated intermediate in the Ca2+ transport cycle, E1 approximately P, increases with decreasing [Mg2+] (Dupont, Y. 1980. Eur. J. Biochem. 109:231-238). Previous x-ray diffraction work (Pascolini, D., and J.K. Blasie. 1988. Biophys. J. 54:669-678) under high [Mg2+] conditions (25 mM) indicated that changes in the profile structure of the SR membrane could be responsible for the low-temperature transient trapping of E1 approximately P that occurs at temperatures below 2-3 degrees C, the upper characteristic temperature th for lipid lateral phase separation in the membrane. We now present results of our study of the Ca2+ uptake kinetics and of the structure of the SR membrane at low [Mg2+] (less than or equal to 100 microM). Our results show a slowing in the kinetics of both phases of the Ca2+ uptake process and an increase in the duration of the plateau of the fast phase before the onset of the slow phase, indicating an increase in the lifetime (transient trapping) of E1 approximately P. Calcium uptake kinetics at low [Mg2+] and moderately low temperature (approximately 0 degree C) are similar to those observed at much lower temperatures (approximately -10 degrees C) at high [Mg2+]. The temperature-induced structural changes that we observed at low [Mg2+] are much more pronounced than those found to occur at higher [Mg2+]. Also, at the lower [Mg2+] the upper characteristic temperature th for lipid lateral phase separation was found to be higher, at approximately 8-10 degrees C. Our studies indicate that both temperature and [Mg2+] affect the structure and the functionality (as measured by changes in the kinetics of Ca2+ uptake) of the SR membrane. Membrane lipid phase behavior and changes in the Ca2+ ATPase profile structure seem to be related, and we have found that structural changes are responsible for the slowing of the kinetics of the fast phase of Ca2+ uptake, and could also mediate the effect that [Mg2+] has on E1 approximately P lifetime.  相似文献   

8.
M Tarek  K Tu  M L Klein    D J Tobias 《Biophysical journal》1999,77(2):964-972
Molecular dynamics simulations have been used to investigate the structure of hybrid bilayers (HB) formed by dipalmitoylphosphatidylcholine (DPPC) lipid monolayers adsorbed on a hydrophobic alkanethiol self-assembled monolayer (SAM). The HB system was studied at 20 degrees C and 60 degrees C, and the results were compared with recent neutron reflectivity measurements (Meuse, C. W., S. Krueger, C. F. Majkrzak, J. A. Dura, J. Fu, J. T. Connor, and A. L. Plant. 1998. Biophys. J. 74:1388) and previous simulations of hydrated multilamellar bilayers (MLB) of DPPC (Tu, K., D. J. Tobias, and M. L. Klein. 1995. Biophys. J. 69:2558; and 1996. 70:595). The overall structures of the HBs are in very good agreement with experiment. The structure of the SAM monolayer is hardly perturbed by the presence of the DPPC overlayer. The DPPC layer presents characteristics very similar to the MLB gel phase at low temperature and to the liquid crystal phase at high temperature. Subtle changes have been found for the lipid/water interface of the HBs compared to the MLBs. The average phosphatidylcholine headgroup orientation is less disordered, and this produces changes in the electric properties of the HB lipid/water interface. These changes are attributed to the fact that the aqueous environment of the lipids in these unilamellar films is different from that of MLB stacks. Finally, examination of the intramolecular and whole-molecule dynamics of the DPPC molecules in the fluid phase HB and MLB membranes revealed that the reorientations of the upper part of the acyl chains (near the acyl ester linkage) are slower, the single molecule protrusions are slightly damped, and the lateral rattling motions are significantly reduced in the HB compared with the MLB.  相似文献   

9.
The general unireactant modifier mechanism in the absence of product can be described by the following linked reactions: E + S k1 in equilibrium k-1 ES k3----E + P; E + I k5 in equilibrium k-5 EI; EI + S k2 in equilibrium k-2 ESI k4----EI + P; and ES + I k6 in equilibrium k-6 ESI where S is a substrate and I is an effector. A full steady state treatment yields a velocity equation that is second degree in both [S] and [I]. Two different conditions (or assumptions) permit reduction of the velocity equation to one that is first degree in [S] and [I]. These are (a) that k-2k3 = k-1k4 (Frieden, C., J. Biol. Chem. 239, pp. 3522-3531, (1964)) and (b) that the I-binding reactions are at equilibrium (Reinhart, G. D., Arch. Biochem. Biophys. 224, pp. 389-401 (1983)). It is shown that each condition gives rise to the other (i.e., if the I-binding reactions are at equilibrium, then k-2k3 must equal k-1k4 and vice-versa). If one assumes equilibrium for the I-binding steps, the velocity equation derived by the method of Cha (J. Biol. Chem. 243, pp. 820-825 (1968)) is apparently second degree in [I] (Segel, I. H., Enzyme Kinetics, p. 838, Wiley-Interscience (1975)), but reduces to a first degree equation when the relationship derived by Frieden is inserted. If one starts by assuming a single equilibrium condition for I binding, e.g., k-5[EI] = k5[E][I] or k-6[ESI] = k6[ES][I], then a traditional algebraic manipulation of the remaining steady state equations provides first degree expressions for the concentrations of all enzyme species and also discloses the Frieden relationship.  相似文献   

10.
The depolarization signal of the diffraction patterns from muscle fibers includes information that differs from that of transmission birefringence experiments. Although both the birefringence studies and the phase shift studies of Yeh et al. (Yeh, Y, and G. Pinsky, 1983, Biophys. J., 42:83-90; Yeh, Y., M. E. Corcoran, R. J. Baskin, and R. L. Lieber, 1983, Biophys. J., 44:343-351) include inseparable intrinsic and form contributions, the present analysis shows that the magnitude of the E-field components of diffracted light is affected only by the intrinsic contribution. We have analyzed the amplitude portion of the data of which the phase shift portion had previously been reported (Yeh, Y., M. E. Corcoran, R. J. Baskin, and R. L. Lieber, 1983, Biophys. J., 44:343-351). For the relaxed-to-rigor transition, these field amplitudes also exhibit changes when ATP concentration is decreased. The observed decrease in optical depolarization upon rigor is consistent with the idea that optically anisotropic elements move away from the myosin thick filament under such conditions.  相似文献   

11.
It was recently shown that a 25-residue peptide, Dk-(61-85), derived from the alpha 1 domain of a murine major histocompatibility class I molecule (H-2Dk), affects insulin receptor functions (Hansen, T., Stagsted, J., Pedersen, L., Roth, R. A., Goldstein, A., and Olsson, L. (1989) Proc. Natl. Acad. Sci. U. S. A. 86, 3123-3126; Stagsted, J., Reaven, G. M., Hansen, T., Goldstein, A., and Olsson, L. (1990) Cell 62, 297-307). We now report that this peptide can reversibly assume a biologically active or inactive state as measured in the rat adipocyte glucose uptake assay, implying that the peptide has at least two interconvertible conformations. The peptide has an ordered conformation in 0.1 M HCl or 0.1 M NaCl stock solution as shown by circular dichroism, but has a disordered molecular structure and is inactive when dissolved in H2O. The biologically active peptide forms liquid crystals at the stock solution concentration (1 mM), so the CD spectra do not provide information on the secondary structure. Under all conditions tested, biological activity (measured after transfer to assay buffer) is associated with an ordered conformation in stock solution. Biological activity and an ordered conformation of the peptide in H2O stock solution can be induced by increasing ionic strength (greater than 100 mM NaCl for maximal effect) or increasing pH (greater than 5 for maximal effect). The induction rate of the ordered conformation is slow with a half-maximal value obtained after approximately 20 min. Both biological activity and the ordered structure are lost upon heating of stock solution to 90 degrees C or upon transfer to assay buffer. A similar correlation of ordered structure with biological activity was observed with two truncated peptides derived from Dk-(61-85). It is inferred from these results that the Dk-(61-85) peptide and related peptides only affect insulin-stimulated glucose uptake in rat adipocytes if they have assumed an ordered conformation in stock solution prior to transfer to assay buffer and exposure to cells.  相似文献   

12.
Xu S  Offer G  Gu J  White HD  Yu LC 《Biochemistry》2003,42(2):390-401
Mammalian myosin filaments are helically ordered only at higher temperatures (>20 degrees C) and become progressively more disordered as the temperature is decreased. It had previously been suggested that this was a consequence of the dependence of the hydrolytic step of myosin ATPase on temperature and the requirement that hydrolysis products (e.g., ADP.P(i)) be bound at the active site. An alternative hypothesis is that temperature directly affects the conformation of the myosin heads and that they need to be in a particular conformation for helical order in the filament. To discriminate between these two hypotheses, we have studied the effect of temperature on the helical order of myosin heads in rabbit psoas muscle in the presence of nonhydrolyzable ligands. The muscle fibers were overstretched to nonoverlap such that myosin affinity for nucleotides was not influenced by the interaction of myosin with the thin filament. We show that with bound ADP.vanadate, which mimics the transition state between ATP and hydrolysis products, or with the ATP analogues AMP-PNP or ADP.BeF(x)() the myosin filaments are substantially ordered at higher temperatures but are reversibly disordered by cooling. These results reinforce recent studies in solution showing that temperature as well as ligand influence the equilibrium between multiple myosin conformations [Málnási-Csizmadia, A., Pearson, D. S., Kovács, M., Woolley, R. J., Geeves, M. A., and Bagshaw, C. R. (2001) Biochemistry 40, 12727-12737; Málnási-Csizmadia, A., Woolley, R. J., and Bagshaw, C. R. (2000) Biochemistry 39, 16135-16146; Urbanke, C., and Wray, J. (2001) Biochem. J. 358, 165-173] and indicate that helical order requires the myosin heads to be in the closed conformation. Our results suggest that most of the heads in the closed conformation are ordered, and that order is not produced in a separate step. Hence, helical order can be used as a signature of the closed conformation in relaxed muscle. Analysis of the dependence on temperature of helical order and myosin conformation shows that in the presence of these analogues one ordered (closed) conformation and two disordered conformations with distinct thermodynamic properties coexist. Low temperatures favor one disordered conformation, while high temperatures favor the ordered (closed) conformation together with a second disordered conformation.  相似文献   

13.
A simple model of a double-headed crossbridge is introduced to explain the retardation of force decay after an imposed stretch in skeletal muscle fibers under equilibrium conditions. The critical assumption in the model is that once one of the heads of a crossbridge is attached to one of the available actin sites, the attachment of the second head will be restricted to a level of strain determined by the attachment of the first head. The crossbridge structure, namely the connection of both heads of a crossbridge to the same tail region, is assumed to impose this constraint on the spatial configurations of crossbridge heads. The unique feature of the model is the prediction that, in the presence of a ligand (PPi, ADP, AMP-PNP) and absence of Ca2+, the halftime of force decay is many times larger than the inverse rate of detachment of a crossbridge head measured in solution. This prediction is in agreement with measured values of half-times of force decay in fibers under similar conditions (Schoenberg, M., and E. Eisenberg. 1985. Biophys. J. 48:863-871f). It is predicted that a crossbridge head is more likely to re-attach to its previously strained position than remain unattached while the other head is attached, leading to the slow decay of force. Our computations also show that the apparent cooperativity in crossbridge binding observed in experiments (Brenner, B., L. C. Yu, L. E. Greene, E. Eisenberg, and M. Schoenberg. 1986. Biophys. J. 50:1101-1108) can be partially accounted by the double-headed crossbridge attachment.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Motional changes of the dipalmitoyl lecithin molecule associated with the pretransition in multibilayers are investigated by proton-enhanced 13C-NMR and proton spin-locking experiments. The nitrogen-bound methyl groups of the polar head exhibit faster motion and more disorder in the intermediate phase compared with the gel phase. Although little or no change occurs in the hydrocarbon chain order at the pretransition, the corresponding motional correlation time changes by one order of magnitude. This is consistent with a model involving rotational motion of the hydrocarbon chains about their long axes: in the gel phase the motion is such that neighboring chains make an oscillating disrotatory motion, while in contrast, in the intermediate phase a quasi-free chain rotation takes place. Earlier contradicting results of Davies, J., 1979, Biophys. J., 27:339-358, and ourselves, Trahms, L., and E. Boroske, 1979, Biochim, Biophys. Acta. 552:189-193, are explained by this model.  相似文献   

15.
Phase diagrams of 3-component lipid bilayer mixtures containing cholesterol reveal major differences among the different types of lipids. Here we report that mixtures of cholesterol together with POPC and a high-melting temperature PC or sphingomyelin show different phase behavior from similar mixtures that contain DOPC or di-phytanoyl-PC instead of POPC. In particular, only one region of macroscopic phase coexistence occurs with POPC, a region of coexisting liquid disordered and solid phases, {Lalpha+Lbeta}. Fluorescence microscopy imaging is useful for these studies, but is subject to artifactual light-induced domain formation, as reported by Ayuyan and Cohen [A.G. Ayuyan, F.S. Cohen, Lipid peroxides promote large rafts: Effects of excitation of probes in fluorescence microscopy and electrochemical reactions during vesicle formation, Biophys. J. 91 (2006) 2172-2183.]. This artifact can be attenuated by decreased illumination and low dye concentration. The use of the free radical scavenger n-propyl gallate can reduce the artifact, but this molecule enters the bilayer and itself perturbs the phase behavior. We suggest that the light-induced domain separation artifact might actually arise from pre-existing lipid clusters that are induced to coalesce, and therefore indicates highly nonrandom mixing of the lipid components.  相似文献   

16.
The intriguing decompositions of nitro-containing explosives have been attracting interest. While theoretical investigations have long been concentrated mainly on unimolecular decompositions, bimolecular reactions have received little theoretical attention. In this paper, we investigate theoretically the bimolecular reactions between nitromethane (CH3NO2)—the simplest nitro-containing explosive—and its decomposition products, such as NO2, NO and CO, that are abundant during the decomposition process of CH3NO2. The structures and potential energy surface (PES) were explored at B3LYP/6-31G(d), B3P86/6-31G(d) and MP2/6-311?+?G(d,p) levels, and energies were refined using CCSD(T)/cc-pVTZ methods. Quantum chemistry calculations revealed that the title reactions possess small barriers that can be comparable to, or smaller than, that of the initial decomposition reactions of CH3NO2. Considering that their reactants are abundant in the decomposition process of CH3NO2, we consider bimolecular reactions also to be of great importance, and worthy of further investigation. Moreover, our calculations show that NO2 can be oxidized by CH3NO2 to NO3 radical, which confirms the conclusion reached formerly by Irikura and Johnson [(2006) J Phys Chem A 110:13974–13978] that NO3 radical can be formed during the decomposition of nitramine explosives.  相似文献   

17.
Pulsed field gradient (pfg)-NMR spectroscopy was utilized to determine lipid lateral diffusion coefficients in oriented bilayers composed of 25 mol % sterol and equimolar amounts of dioleoylphosphatidylcholine and sphingomyelin. The occurrence of two lipid diffusion coefficients in a bilayer was used as evidence of lateral phase separation into liquid ordered and liquid disordered domains. It was found that cholesterol, ergosterol, sitosterol, and lathosterol induced domains, whereas lanosterol, stigmasterol, and stigmastanol resided in homogeneous membranes in the temperature interval of 24-70 degrees C. Among the domain-forming sterols, differences in the upper miscibility temperature indicated that the stability of the liquid ordered phase could be modified by small changes in the sterol structure. The domain-forming capacity for the different sterols is discussed in terms of the ordering effect of the sterols on the lipids, and it is proposed that the driving force for the lateral phase separation is the reduced solubility of the unsaturated lipid in the highly ordered phase.  相似文献   

18.
J A Killian  B de Kruijff 《Biochemistry》1985,24(27):7881-7890
The effect of gramicidin incorporation on the thermodynamic properties of phosphatidylcholine (PC) and phosphatidylethanolamine (PE) dispersions was investigated by differential scanning calorimetry. The results show that incorporation of gramicidin in PC systems results in a decrease of the energy content of the gel to liquid-crystalline phase transition. When incorporated in PE systems, however, the peptide does not affect the properties of the gel to liquid-crystalline phase transition with the exception that at high gramicidin concentrations the onset of the melting process is shifted to a slightly lower temperature. We therefore assume that in the lamellar gel state of PE aggregation of the peptide occurs. To get more insight into the nature of the gramicidin-PE interaction, we studied the motional and structural details of HII phase formation in gramicidin/PE systems with the use of 31P and 13C nuclear magnetic resonance (NMR) and small-angle X-ray diffraction. In agreement with earlier results [Van Echteld, C. J. A., Van Stigt, R., de Kruijff, B., Leunissen-Bijvelt, J., Verkleij, A. J., & De Gier, J. (1981) Biochim. Biophys. Acta 648, 287-291] it was shown that gramicidin incorporation lowers and broadens the bilayer to hexagonal HII phase transition in PE systems. 31P NMR chemical shift anisotropy (CSA) measurements indicated that a phase separation occurs between a gramicidin-poor lamellar phase and a gramicidin-rich HII phase. From combined CSA and spin-lattice relaxation time (T1) measurements it was suggested that in the HII phase gramicidin decreases the molecular order and increases the rate of motion of the phosphate moiety of PE. In addition, 13C NMR line width measurements indicated that the acyl chains are more disordered in the HII phase than in the lamellar phase and that a similar disorder occurs in the HII phase of the pure PE as in the gramicidin-rich HII phase. This interpretation was supported by the X-ray diffraction data, which show similar first-order repeat distances in both types of HII phase. From saturation-transfer NMR experiments in PE and gramicidin-PE mixtures it was shown that no exchange occurs between the lamellar and the HII phases in the time scale of 1-2 s, suggesting a macroscopic phase separation. Finally, we discussed the gramicidin-lipid interaction and in particular the HII phase formation by gramicidin in PE and in PC systems. It is proposed that aggregation of the peptide plays a crucial role in HII phase formation.  相似文献   

19.
P Nambi  E S Rowe  T J McIntosh 《Biochemistry》1988,27(26):9175-9182
It is now well established that a number of amphiphilic molecules such as ethanol can induce the formation of the fully interdigitated gel phase in phosphatidylcholines. We have shown earlier that alcohols such as ethanol induce biphasic melting behavior in phosphatidylcholines [Rowe, E. S. (1983) Biochemistry 22, 3299-3305] but not in phosphatidylethanolamines [Rowe, E. S. (1985) Biochim. Biophys. Acta 813, 321-330]. Simon and McIntosh [(1984) Biochim. Biophys. Acta 773, 169-172] showed that the alcohol-induced biphasic melting behavior in phosphatidylcholines is a consequence of acyl chain interdigitation. In the present study we demonstrate the detection of the transition of DPPC and DSPC to the interdigitated phase in the presence of ethanol using the fluorescence properties of the commonly used fluorophore 1,6-diphenyl-1,3,5-hexatriene (DPH). By correlating fluorescence and X-ray diffraction results, we have demonstrated the use of fluorescence to study the phase transition from the noninterdigitated to the interdigitated phase. Using this method, we have investigated the temperature and ethanol concentration dependence of the induction of the interdigitated phase in DSPC and DPPC and shown that the induction of interdigitation by ethanol is temperature dependent, with higher temperature favoring interdigitation. The temperature-ethanol phase diagrams have been determined for DPPC and DSPC.  相似文献   

20.
The kinetic mechanisms of the NAD- and NADP-linked reactions catalyzed by glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides were examined using product inhibition, dead-end inhibition and alternate substrate experiments. The results are consistent with a steady-state random mechanism for the NAD-linked and an ordered, sequential mechanism with NADP+ binding first for the NADP-linked reaction. Thus, the enzyme can bind NADP+, NAD+, and glucose 6-phosphate, but the enzyme-glucose 6-phosphate complex can react only with NAD+, not with NADP+. This affects the rate equation for the NADP-linked reaction by introducing a term for a dead-end enzyme-glucose 6-phosphate complex. The kinetic mechanisms represent revisions of those proposed previously (C. Olive, M.E. Geroch, and H.R. Levy, 1971, J. Biol. Chem. 246, 2047-2057) and provide a kinetic basis for the regulation of coenzyme utilization of the enzyme by glucose 6-phosphate concentration (H.R. Levy, and G.H. Daouk, 1979, J. Biol. Chem. 254, 4843-4847) and NADPH/NADP+ concentration ratios (H.R. Levy, G.H. Daouk, and M.A. Katopes, 1979, Arch, Biochem. Biophys. 198, 406-413). The kinetic mechanisms were found to be the same at pH 6.2 and pH 7.8. The kinetics of ATP inhibition of the NAD- and NADP-linked reactions were examined at pH 6.2 and pH 7.8. The results are interpreted in terms of ATP addition to binary enzyme-coenzyme and enzyme-glucose 6-phosphate complexes.  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