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.
Pressure vs. distance relationships have been obtained for phosphatidylglycerol bilayers, in both charged and uncharged states. Water was removed from the lipid multilayers by the application of osmotic pressures in the range of 0-2.7 x 10(9) dyn/cm2, and the distance between adjacent bilayers was obtained from Fourier analysis of lamellar x-ray diffraction data. For phosphatidylglycerol bilayers made electrically neutral either by lowering the pH or by adding equimolar concentrations of the positively charged lipid stearylamine, the pressure-distance data could be fit with a single exponential. The measured decay lengths were 1.1 A at low pH and 1.5 A with stearylamine, which are similar to decay lengths of the hydration pressure found for gel phases of other neutral bilayers. In addition, the magnitude of this repulsive pressure was proportional to the square of the Volta potential (equivalent to the dipole potential for electrically neutral bilayers) measured in monolayers in equilibrium with bilayers, in agreement with results previously found for the hydration pressure between phosphatidylcholine bilayers. For charged phosphatidylglycerol bilayers, the pressure-distance relation had two distinct regions. For bilayer separations greater than 10 A, the pressure-distance data had an exponential decay length (11 A) and a magnitude consistent with that expected for electrostatic repulsion from double-layer theory. For bilayer separations of 2-10 A, the pressure decayed much more rapidly with increasing bilayer separation (decay length less than 1 A). We interpret these data at low bilayer separations in terms of a combination of hydration repulsion and steric hindrance between the lipid head groups and the sodium ions trapped between apposing bilayers.  相似文献   

2.
Hydration force and bilayer deformation: a reevaluation   总被引:20,自引:0,他引:20  
T J McIntosh  S A Simon 《Biochemistry》1986,25(14):4058-4066
The hydration repulsive force between lipid bilayers and the deformability of both gel and liquid-crystalline bilayers have been quantitated by an X-ray diffraction analysis of osmotically stressed liposomes. Both sampling theorem reconstructions and electron density distributions were calculated from diffraction data obtained from multilayers with applied osmotic pressures of 0-50 atm. The bilayer thickness and area per lipid molecule remain nearly constant (to within about 4%) in this pressure range, as adjacent bilayers move from their equilibrium separation in excess water to within 2-4 A of each other. This analysis indicates that the bilayers are relatively incompressible. This results differs from previously published X-ray diffraction studies of bilayer compressibility but agrees with direct mechanical measurements of the bilayer compressibility modulus. It is also found that the hydration repulsive force decays exponentially with separation between bilayers with a decay constant of 1.4 A for gel-state dipalmitoylphosphatidylcholine and 1.7 A for liquid-crystalline egg phosphatidylcholine bilayers. This implies that the exponential decay constant is not necessarily equal to the diameter of a water molecule, as has been previously suggested on experimental and theoretical grounds.  相似文献   

3.
We have measured the microscopic isothermal compressibility of dioleoyl- and dimyristyl-phosphatidylcholine multilayers and bilayers as a function of membrane depth by the pressure dependence of the polarization of a series of anthroyloxy fatty acids. In both systems, within experimental error, the compressibility did not change with membrane depth. The magnitudes of the compressibilities matched those of organic solids and those reported for dipalmitoylphosphatidylcholine multilayers from neutron diffraction measurements (Braganza, L. F., and D. L. Worcester. 1986. Biochemistry, 25:7484-7488). The bilayer compressibility decreased with temperature and this decrease was similar with membrane depth consistent with the isotropic thermal expansion of membranes previously observed (Scarlata, S. 1989. Biophys. J. 55:1215-1223). The vertical compressibility in the z direction is much lower than the horizontal (xy planes) for probes that lie parallel to the hydrocarbon chains which is consistent with an increase in bilayer thickness. The compressibility for probes that lie perpendicular to the hydrocarbon chains is more isotropic due to their limited spatial access to the z plane.  相似文献   

4.
Partial compressibilities of globular proteins in water are reviewed. Contribution of hydrational and of intrinsic compressibilities to experimental partial quantity have been evaluated from ultrasonic data using two independent methods: (a) additive calculation of the hydrational contributions of the surface atomic groups and (b) an analysis of correlation between partial compressibility and molecular surface area. The value (14 ± 3) × 10?6 bar ?1 for the isothermal compressibility coefficient of the protein interior at 25°C was obtained as an average value for variety of globular proteins. This value is similar to that of solid organic polymers. Possible relaxation contribution to partial compressibility is roughly estimated from comparison of thermodynamic with x-ray data on protein compressibility. The average compressibility of water in the hydration shell of proteins was found to be 35 × 10?6 bar ?1, which is 20% less than that of pure water. © 1993 John Wiley & Sons, Inc.  相似文献   

5.
The swelling behavior of gelatin gels containing proteoglycans (sulphated proteoglycans from bovine intervertebral discs and a hyaluronate proteoglycan from bovine synovial fluid) when immersed in osmotically active solutions of dextran have been measured. The presence of the proteoglycans markedly affects the internal osmotic contribution to the swelling pressure of the gel. These internal osmotic pressures are considerably in excess of the sum of the osmotic activities of the individual components. This behavior is understood in terms of an entropic interaction between the gelatin and the proteoglycan molecules. By use of the “dilute solution” treatment of Flory, the osmotic pressure excesses are related to the volumes and hence dimensions of the interact acting species. A comparison of these values with those calculated by other means shows good agreement. The osmotic behavior of the complex gels can be understood on a mechanistic basis, if we regard the gelatin and sulphated proteoglycans as spheres and the hyaluronate proteoglycan as a rod.  相似文献   

6.
This study focuses on dioleoylphosphatidylcholine (DOPC) bilayers near full hydration. Volumetric data and high-resolution synchrotron x-ray data are used in a method that compares DOPC with well determined gel phase dipalmitoylphosphatidylcholine (DPPC). The key structural quantity obtained is fully hydrated area/lipid A0 = 72.2 +/- 1.1 A2 at 30 degrees C, from which other quantities such as thickness of the bilayer are obtained. Data for samples over osmotic pressures from 0 to 56 atmospheres give an estimate for the area compressibility of KA = 188 dyn/cm. Obtaining the continuous scattering transform and electron density profiles requires correction for liquid crystal fluctuations. Quantitation of these fluctuations opens an experimental window on the fluctuation pressure, the primary repulsive interaction near full hydration. The fluctuation pressure decays exponentially with water spacing, in agreement with analytical results for soft confinement. However, the ratio of decay length lambda(fl) = 5.8 A to hydration pressure decay length lambda = 2.2 A is significantly larger than the value of 2 predicted by analytical theory and close to the ratio obtained in recent simulations. We also obtain the traditional osmotic pressure versus water spacing data. Our analysis of these data shows that estimates of the Hamaker parameter H and the bending modulus Kc are strongly coupled.  相似文献   

7.
Sodium hyaluronate samples of various molecular weights were prepared and characterized by size exclusion chromatography and dilution viscometry. Densitometry was used for determination of the densities of sodium hyaluronate solutions and the results expressed as partial specific volumes in the respective solvents. Solution adiabatic compressibilities were measured. Hydration parameters of sodium hyaluronate were consequently determined and the values obtained discussed in relation to existing hydration models for sodium hyaluronate in water.  相似文献   

8.
The apparent specific volumes and isentropic compressibilities of hen egg white lysozyme were measured in aqueous guanidinium chloride solutions at 25 degrees C by means of a vibrational densimeter and a sing-around ultrasonic velocimeter. Little transition attributable to a protein unfolding was detected in the partial specific volume, while the partial specific isentropic compressibility decreased slightly around the transition region. The pressure-assisted unfolding was also investigated in aqueous guanidinium chloride solutions by means of ultraviolet spectroscopy. Assuming a two-state transition model, it was found that the free energy change of unfolding depends almost linearly on pressure and the unfolding reaction is accompanied by a small decrease in volume. The compressibility behavior is in conflict with the notion that a protein structure is almost completely unfolded by guanidinium chloride and most of the amino acid residues in the protein interior are exposed to solvent. These results support the current view that globular proteins have some residual structures even in the unfolded state induced by a strong denaturant.  相似文献   

9.
The ultrasonic absorption spectra of proteins in solution generally show relaxational behaviour. There will be a corresponding dispersion of sound velocities accompanying each relaxation. The compressibility of the protein as measured by sound velocity techniques will therefore include a relaxational contribution. We have evaluated this contribution for a number of proteins and found that in some cases the relaxational contribution is a significant fraction of the total compressibility. The relaxational contribution will be large only if the molecule has a large number of degrees of freedom with low force constants. However, such motions are likely to be those involved in the biological functioning of the molecule. Care is needed in interpreting the relaxation spectrum since proton transfer processes give large apparent relaxational compressibilities.  相似文献   

10.
Abstract. Regulation of the concentration of osmotic solutes was studied in Chlorella emersonii grown at external osmotic pressures (II) ranging between 0.08 and 1.64MPa. NaCl was used as osmoticum. The total solute content of the cells was manipulated by applying 2 mol m−3 3- O -methylglucose (MG), which was not metabolized, and accumulated at concentrations ranging between 60 and 230 mol m−3 within 4 h after its addition to the medium. Methylglucose uptake resulted in decreases in concentrations of proline and sucrose, the two solutes mainly responsible for osmotic adaptation of C. emersonii to high external II. The responses were consistent with the hypothesis that proline and sucrose concentrations are controlled by a system of osmotic regulation, with turgor and/or volume as a primary signal. Short-term experiments showed that even very small increases in turgor and/or volume, due to accumulation of methylglucose, resulted in large decreases in proline and sucrose. Over the first 30-60 min the total solute concentration in the cells increased by at most 15 osmol m−3 which would represent an increase in turgor pressure of at most 0.04 M Pa. Yet, the decreases in proline and sucrose were as fast as those in cells exposed to a sudden decrease of 0.25 MPa in external II, when the turgor pressure would have increased by at least 0.15 MPa. High concentrations of methylglucose in cells grown at high II did not affect the rapid synthesis of proline and sucrose which started when the cells were transferred to yet higher II. Thus, methylglucose had no direct effects on proline and sucrose metabolism, and it has been assumed that it acted solely as an inert osmotic solute within the cell.  相似文献   

11.
Binding of hyaluronate to the surface of cultured cells   总被引:7,自引:0,他引:7       下载免费PDF全文
The binding of hyaluronate to SV-3T3 cells was measured by incubating a suspension of cells (released from the substratum with EDTA) with 3H-labeled hyaluronate and then applying the suspension to glass fiber filters which retained the cells and the bound hyaluronate. The extent of binding was a function of both the concentration of labeled hyaluronate and the cell number. Most of the binding took place within the first 2 min of the incubation and was not influenced by the presence or absence of divalent cations. The binding of labeled hyaluronate to SV-3T3 cells could be prevented by the addition of an excess of unlabeled hyaluronate. High molecular weight preparations of hyaluronate were more effective in preventing binding than low molecular weight preparations. The binding of [3H]hyaluronate was inhibited by high concentrations of oligosaccharide fragments of hyaluronate consisting of six sugars or more, and by chondroitin. The sulfated glycosaminoglycans (chondroitin-4-sulfate, chondroitin-6-sulfate, dermatan sulfate, heparin, and heparan sulfate) had little or no effect on the binding. The labeled hyaluronate bound to the cells could be totally removed by incubating the cells with testicular hyaluronidase, streptomyces hyaluronidase, or trypsin, indicating that the hyaluronate-binding sites are located on the cell surface.  相似文献   

12.
13.
The mechanical response of a TIM-barrel protein to an applied pressure has been studied. We generated structures under an applied pressure by assuming the volume change to be a linear function of normal mode variables. By Delaunay tessellation, the space occupied by protein atoms is divided uniquely into tetrahedra, whose four vertices correspond to atomic positions. Based on the atoms that define them, the resulting Delaunay tetrahedra are classified as belonging to various secondary structures in the protein. The compressibility of various regions identified with respect to secondary structural elements in this protein is obtained from volume changes of respective regions in two structures with and without an applied pressure. We found that the β barrel region located at the core of the protein is quite soft. The interior of the β barrel, occupied by side chains of β strands, is the softest. The helix, strand, and loop segments themselves are extremely rigid, while the regions existing between these secondary structural elements are soft. These results suggest that the regions between secondary structural elements play an important role in protein dynamics. Another aspect of tetrahedra, referred to as bond distance, is introduced to account for rigidities of the tetrahedra. Bond distance is a measure of separation of the atoms of a tetrahedron in terms of number of bonds along the polypeptide chain or side chains. Tetrahedra with longer bond distances are found to be softer on average. From this behavior, we derive a simple empirical equation, which well describes the compressibilities of various regions. © 1997 Wiley-Liss Inc.  相似文献   

14.
Hyaluronate in cultured skin fibroblasts derived from patients with Werner's syndrome, who excrete large amounts of urinary hyaluronate, was investigated. The amount of hyaluronate secreted into the medium by Werner's fibroblasts was 2-3-times that of normal fibroblasts, whereas no difference in enzyme activities related to the degradation of hyaluronate was found. Werner's fibroblasts were then cultured in the presence of [3H]glucosamine, and the amount of [3H]hyaluronate and its chain lengths in the medium and matrix (trypsinate) fractions were compared with those of normal cells. No significant difference in the chain length of hyaluronate was observed between normal and Werner's fibroblasts. On the other hand, a significant increase of hyaluronate was found in the matrix fraction of Werner's fibroblasts when the cells reached confluency. In addition, a hyaluronate of small chain length was found in the matrix fraction of Werner's fibroblasts, although this was absent from that of normal cells. It was concluded that the constituents of the extracellular matrix of Werner's fibroblasts differed from those of normal cells, characterized by the presence of a large amount of hyaluronate and a relatively small hyaluronate chain.  相似文献   

15.
Effect of Osmotic Stress on Turgor Pressure in Mung Bean Root Cells   总被引:3,自引:0,他引:3  
Turgor pressure in cells of the elongating region of intactmung bean roots was directly measured by using the pressure-probetechnique. After the external osmotic pressure had been increasedfrom 0 MPa to 0.5 MPa, turgor pressure rapidly decreased byabout 0.5 MPa from 0.65 MPa to 0.14 MPa and root elongationstopped. Subsequent turgor regulation was clearly confirmed,which followed the osmotic adjustment to maintain a constantdifference in the osmotic pressure between root-cell sap andthe external medium ( II). It took at least 6 h for turgor pressureto recover to an adjusted constant level of about 0.5 MPa dueto turgor regulation, but rootelongation resumed within onlyan hour after the osmotic treatment. Therefore, the resumptionof root elongation under osmotic stress could not have beendirectly connected with turgor regulation. Furthermore, sincethe amounts of decrease in turgor pressure just after applicationsof various degrees of osmotic stress could be interpreted inrelation to those in II, hydraulic conductivity between theinside and the outside of root cells must be large enough toattain water potential equilibrium rapidly in response to osmoticstress. We conclude that turgor pressure in the cells of theelongating region of mung bean roots is determined mainly by II because of water potential equilibrium. (Received January 27, 1987; Accepted May 21, 1987)  相似文献   

16.
T J McIntosh  A D Magid  S A Simon 《Biochemistry》1989,28(19):7904-7912
Well-ordered multilamellar arrays of liquid-crystalline phosphatidylcholine and equimolar phosphatidylcholine-cholesterol bilayers have been formed in the nonaqueous solvents formamide and 1,3-propanediol. The organization of these bilayers and the interactions between apposing bilayer surfaces have been investigated by X-ray diffraction analysis of liposomes compressed by applied osmotic pressures up to 6 X 10(7) dyn/cm2 (60 atm). The structure of egg phosphatidylcholine (EPC) bilayers in these solvents is quite different than in water, with the bilayer thickness being largest in water, 3 A narrower in formamide, and 6 A narrower in 1,3-propanediol. The incorporation of equimolar cholesterol increases the thickness of EPC bilayers immersed in each solvent, by over 10 A in the case of 1,3-propanediol. The osmotic pressures of various concentrations of the neutral polymer poly(vinylpyrrolidone) dissolved in formamide or 1,3-propanediol have been measured with a custom-built membrane osmometer. These measurements are used to obtain the distance dependence of the repulsive solvation pressure between apposing bilayer surfaces. For each solvent, the solvation pressure decreases exponentially with distance between bilayer surfaces. However, for both EPC and EPC-cholesterol bilayers, the decay length and magnitude of this repulsive pressure strongly depend on the solvent. The decay length for EPC bilayers in water, formamide, and 1,3-propanediol is found to be 1.7, 2.4, and 2.6 A, respectively, whereas the decay length for equimolar EPC-cholesterol bilayers in water, formamide, and 1,3-propanediol is found to be 2.1, 2.9, and 3.1 A, respectively. These data indicate that the decay length is inversely proportional to the cube root of the number of solvent molecules per unit volume.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Hydraulic conductivity of chondroitin sulfate proteoglycan solutions   总被引:1,自引:0,他引:1  
The hydraulic conductivity of solutions of Swarm rat chondrosarcoma proteoglycan subunit and of chondroitin 4- and 6-sulfate up to concentrations of 80 mg ml-1 have been measured under physiological conditions using sedimentation velocity and membrane ultrafiltration techniques. This study establishes the very high flow resistance of the proteoglycan and that this resistance is due to its constituent chondroitin sulfate chains. We have also demonstrated little difference in the hydraulic conductivity of chondroitin 4-sulfate as compared to chondroitin 6-sulfate. Studies of hydraulic conductivity of chondroitin sulfate and proteoglycan subunit over a range of salt concentrations demonstrate that the chondroitin sulfates exhibit only a small degree of electrolyte dissipation indicating that their constituent charge groups do not significantly contribute to flow resistance at high mechanical pressures. It appears that the shape and conformation of the polysaccharide backbone and its glycosidic linkages are the factors that primarily govern flow resistance. This is also consistent with the fact that hydraulic conductivity of the proteoglycans and chondroitin sulfates is considerably lower than that of its more charged counterpart heparin but has similar values to hyaluronate. Qualitative agreement between sedimentation analysis and ultrafiltration measurements is also established although the latter technique suffers from not knowing over what distance, adjacent to the membrane, ultrafiltration takes place. It is predicted that the proteoglycans will significantly contribute to flow resistance of cartilagenous tissues which confirms the Maroudas correlation that high proteoglycan concentration in cartilage yields high flow resistance. Further, we establish through a comparison of hydraulic conductivity measurements on hyaluronate, desulfated chondroitin sulfate, chondroitin sulfate, and proteoglycan subunit and osmotic pressure measurements of hyaluronate and proteoglycan that the sulfate groups of the chondroitin sulfate chain play only a small role in the net movement of water relative to the proteoglycan.  相似文献   

18.
Repulsive pressure in the A-band filament lattice of relaxed frog skeletal muscle has been measured as a function of interfilament spacing using an osmotic shrinking technique. Much improved chemical skinning was obtained when the muscles were equilibrated in the presence of EGTA before skinning. The lattice shrank with increasing external osmotic pressure. At any specific pressure, the lattice spacing in relaxed muscle was smaller than that of muscle in rigor, except at low pressures where the reverse was found. The lattice spacing was the same in the two states at a spacing close to that found in vivo. The data were consistent with an electrostatic repulsion over most of the pressure range. For relaxed muscle, the data lay close to electrostatic pressure curves for a thick filament charge diameter of approximately 26 nm, suggesting that charges stabilizing the lattice are situated about midway along the thick filament projections (HMM-S1). At low pressures, observed spacings were larger than calculated, consistent with the idea that thick filament projections move away from the filament backbone. Under all conditions studied, relaxed and rigor, at short and very long sarcomere lengths, the filament lattice could be modeled by assuming a repulsive electrostatic pressure, a weak attractive pressure, and a radial stiffness of the thick filaments (projections) that differed between relaxed and rigor conditions. Each thick filament projection could be compressed by approximately 5 or 2.6 nm requiring a force of 1.3 or 80 pN for relaxed and rigor conditions respectively.  相似文献   

19.
SWELLING OF FISH MITOCHONDRIA   总被引:2,自引:0,他引:2       下载免费PDF全文
The physical properties of fish liver and rat liver mitochondria were compared as a function of temperature and osmotic pressure. The data indicate that fish mitochondria are more flexible and swell at a more rapid rate over a 0 to 30°C temperature range, whereas the rates of swelling at 30 to 40°C are comparable. The swelling rates of both fish and rat mitochondria vary with temperature and approximate the Arrhenius relationship. Apparent energies of activation for swelling averaged 26.5 kcal and 12.9 kcal for rat and fish, respectively. Fish mitochondria were less stable than rat mitochondria to osmotic variation, and the disparity in initial swelling rates became increasingly greater with lower osmotic pressure. The hypotonic swelling of both fish and rat mitochondria was readily reversed osmotically; however, there was a very rapid decay of reversal in fish mitochondria and only a very slow decay in the case of rat. All the data indicate that under comparable conditions the fish mitochondrial membranes are more flexible and presumably more permeable and labile than rat mitochondrial membranes. The findings are discussed in relation to the general metabolic implications and the possible contributions of the membrane constituents to membrane behavior.  相似文献   

20.
The sound velocity and density of suspensions of large unilamellar liposomes from dimyristoylphosphatidylcholine with admixed cholesterol have been measured as a function of temperature around the chain melting temperature of the phospholipid. The cholesterol-to-phospholipid molar ratio xc has been varied over a wide range (0 </= xc </= 0.5). The temperature dependence of the sound velocity number, of the apparent specific partial volume of the phospholipid, and of the apparent specific adiabatic compressibility have been derived from the measured data. These data are particularly discussed with respect to the volume fluctuations within the samples. A theoretical relation between the compressibility and the excess heat capacity of the bilayer system has been derived. Comparison of the compressibilities (and sound velocity numbers) with heat capacity traces display the close correlation between these quantities for bilayer systems. This correlation appears to be very useful as it allows some of the mechanical properties of membrane systems to be calculated from the specific heat capacity data and vice versa.  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