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 46 毫秒
1.
Fluorescent fatty acid labels have been incorporated into the palmitoylation sites of rhodopsin and used to probe the membrane accessibility and location of these sites. The fluorescence properties of anthroyloxy and pyrenyl fatty acids bound to rhodopsin were investigated in a reconstituted vesicle system. Collisional quenching of fluorescence by stearic acid (DSA) labeled with doxyls in the 16, 12, and 5 positions was used to determine the membrane accessibility and disposition of the modifying fatty acids. To properly determine the membrane concentration of these quenchers, the dependence of the Stern-Volmer parameters on both quencher and vesicle concentration was determined. An analysis of these dependences provided a correction for partitioning of the quencher between the aqueous phase and the membrane. After this correction, the relative effectiveness of doxyl quenchers was 16-DSA > 12-DSA > 5-DSA. Parallel studies on free anthroyloxy and pyrenyl fatty acids incorporated into the reconstituted system showed the same dependence on quencher position. These results indicate that the labels at the palmitoylation sites of rhodopsin are situated in the membrane much as a free fatty acid. This anchoring of the palmitates in the membrane results in the formation of a fourth cytoplasmic loop.  相似文献   

2.
Amphotericin B (AmB) is the most widely used polyene antibiotic to treat systemic fungal infections which affect an increasing number of immunocompromised patients. It is generally thought that AmB forms pores within the fungi membranes by interacting with ergosterol, the main sterol of fungi. However, it also interacts with the cholesterol contained in mammalian cells, hence its toxicity. In order to have a better understanding of the interactions prevailing between AmB and sterols, differential scanning calorimetry was used to study various mixtures incorporating from 6.5 to 25 mol% of AmB in pure dipalmitoylphosphatidylcholine (DPPC) vesicles and in ergosterol- or cholesterol-containing DPPC vesicles. The sterol concentration was kept constant at 12.5 mol% with respect to the phospholipid. Our results show that three phases co-exist when AmB is dispersed in the pure phospholipid. One corresponds to the phospholipid phase alone. The two others are characterised by a broad transition at temperatures higher than the main transition temperature of the pure phospholipid, corresponding to the drug in interaction with the aliphatic chains of the lipid. The fact that the transition temperatures of these additional components are higher than that of the pure phospholipid suggests that AmB interacts strongly with the aliphatic chains of the lipid, consistent with the idea prevailing in the literature that AmB by itself may form pores in a lipid matrix. When AmB interacts with cholesterol-containing bilayers the thermograms also present three components. Upon increasing the concentration of AmB, though, an important broadening of these components is observed which is explained in terms of destabilisation of the organisation of the aliphatic chains. The situation is strikingly different if ergosterol is present in the lipid matrix. The thermograms remain unmodified as the concentration of AmB is increased and a broad transition, now involving only two components when the thermograms are decomposed, is observed. An analysis of the results shows that various interacting units, e.g. AmB+DPPC and (AmB+ergosterol)+DPPC, are present within the membrane. These units involve the phospholipid and hence contribute to its structurisation. The important differences between the thermograms obtained with the ergosterol- as compared to the cholesterol-containing bilayers, in spite of the structural similarity of these two sterols, provides strong evidence for the selectivity of interaction of AmB with ergosterol as compared to cholesterol. It is thus clear that the action of AmB on cholesterol- as compared to ergosterol-containing membranes results from different mechanisms. Finally, UV-visible spectra of AmB in pure as well as sterol-containing DPPC vesicles show the presence of absorption bands that give support to the interpretation derived from the calorimetric data.  相似文献   

3.
Pulmonary surfactant-associated protein B (SP-B) has been isolated from porcine lungs and reconstituted in bilayers of dipalmitoylphosphatidylcholine (DPPC) or egg yolk phosphatidylcholine (PC) to characterize the extent of insertion of the protein into phospholipid bilayers. The parameters for the interaction of SP-B with DPPC or PC using different reconstitution protocols have been estimated from the changes induced in the fluorescence emission spectrum of the single protein tryptophan. All the different reconstituted SP-B-phospholipid preparations studied had similar Kd values for the binding of the protein to the lipids, on the order of a few micromolar. The depth of penetration of SP-B into phospholipid bilayers has been estimated by the parallax method, which compares the relative efficiencies of quenching of the protein fluorescence by a shallow or a deeper spin-labeled phospholipid probe. SP-B tryptophan was found to be located 10-13 A from the center of bilayers, which is consistent with a superficial location of SP-B in phosphatidylcholine membranes. Parallax experiments, as well as resonance energy transfer from SP-B tryptophan to an acceptor probe located in the center of the bilayer, indicate that there are significant differences in the extent of insertion of the protein, depending on the method of reconstitution. SP-B reconstituted from lipid/protein mixtures in organic solvents is inserted more deeply in PC or DPPC bilayers than the protein reconstituted by addition to preformed phospholipid vesicles. These differences in the extent of insertion lead to qualitative and quantitative differences in the effect of the protein on the mobility of the phospholipid acyl chains, as studied by spin-label electron spin resonance (ESR) spectroscopy, and could represent different functional stages in the surfactant cycle.  相似文献   

4.
Complexes of alpha-lactalbumin (alpha-LA)1 with dimyristoylphosphatidylcholine (DMPC) or dipalmitoylphosphatidylcholine (DPPC) liposomes at pH 8 and at pH 2 have been obtained by means of gel filtration. Thermal denaturation of alpha-LA complexes of DMPC or DPPC at pH 8 was found to depend on the saturation of protein by metal cations. The intrinsic fluorescence of DMPC-alpha-LA and DPPC-alpha-LA was sensitive to two thermal transitions. The first transition corresponded to the Tc of the lipid vesicles, while the second transition arose from the denaturation of the protein. Fluorescence spectrum position suggested that at low temperature tryptophan accessibility increases upon protein-DMPC or protein-DPPC association. At temperatures above the protein transition (70 degrees C) tryptophan appears to interact significantly with the apolar phase of DMPC and DPPC, evidenced by spectral blue shifts. Whereas the free protein at pH 2 adopts the molten globule (MG) state and is characterized by the absence of a thermal transition, the rapidly-isolated DMPC-alpha-LA complex was characterized by the appearance of a distinct fluorescence thermal transition between 50 and 60 degrees C. This result is consistent with a model of a partially-inserted form of alpha-LA which may possess some degree of tertiary structure and therefore unfolds cooperatively.  相似文献   

5.
Fluorescence quenching of perylene by Co2+ and Ni2+ ions has been investigated both below and above the phase transition temperature in small unilamellar DPPC vesicles. Classic F?rster type energy transfer was observed for perylene quenched by Co2+ ions below the phase transition when the effects of donor-donor energy transfer are taken into account. In the liquid crystalline phase a simple diffusion theory incorporating energy transfer was found to model the system well. For quenching by Ni2+ ions both above and below the phase transition temperature in lipid bilayers and in glycerol the data did not follow classic F?rster type energy transfer but indicated that an additional quenching mechanism was present. A mechanism other than F?rster behaviour was also observed for the quenching by Cr3+ ions in glycerol.  相似文献   

6.
A lectin-induced orientation change of a helical glycopeptide in lipid bilayer membranes was studied. Glycopeptides composed of hydrophobic nona-(G8) and pentapeptide (G4) with a fluorescent probe at the N-terminal and a lactose unit at the C-terminal were synthesized. The glycopeptides were incorporated into lipid bilayer membranes with the lactose unit exposed to the aqueous phase and the peptide chain buried in the membrane. G8 takes a partially helical structure in the membrane, while G4 an irregular structure. Upon binding of lectin to G8 held in the membrane of DPPC liposome, enhancement of fluorescence intensity of the N-terminal anthryl group, reduction of fluorescence quenching of the anthryl group with acrylamide, and increase of CF-leakage from the DPPC liposome were observed. G8', which lacks the O-anthryrlmethylserine residue from G8, formed a voltage-dependent ion channel in BLM experiments. The frequency of single current fluctuations induced by G8' incorporation increased with addition of lectin. These results indicate that the peptide segment of G8 prefers taking a more perpendicular orientation to the membrane upon association with lectin.  相似文献   

7.
The measurements of diphenylhexatriene (DPH) and trimethylammonium diphenylhexatriene (TMA-DPH) fluorescence anisotropy in egg yolk lecithin (EYL) and of DPH anisotropy in dipalmitoylphosphatidylcholine (DPPC) liposomes containing different concentrations of oxidized and reduced ubiquinone (UQ) and plastoquinone (PQ) homologues have been performed. All the oxidized UQ homologues strongly induced ordering of EYL membrane structure, whereas in DPPC liposomes, above the phase transition temperature, the most pronounced effect showed UQ-4. PQ-2 and PQ-9 were less effective than the corresponding ubiquinones in this respect. The reduced forms of UQ and PQ homologues increased the order of membrane lipids to a smaller extent than the corresponding quinones both in the interior of the membrane and closer to its surface. Nevertheless, the investigated prenylquinols showed stronger increase in the membrane order than alpha-tocopherol or alpha-tocopherol acetate, which could be connected with binding of prenylquinol head groups to phospholipid molecules by hydrogen bonds. The strong ordering influence of ubiquinones on the membrane structure was attributed to methoxyl groups of the UQ quinone rings.  相似文献   

8.
Clostridium perfringens perfringolysin O (PFO or theta-toxin) is a cytolytic toxin that binds to cholesterol-containing membranes and then self-associates to spontaneously form aqueous pores of varying size in the bilayer. In this study, a membrane-spanning domain has been identified in PFO by a combination of fluorescence spectroscopic methods using the fluorescent dye N, N'-dimethyl-N-(iodoacetyl)-N'-(7-nitrobenz-2-oxa-1, 3-diazolyl)ethylenediamine (NBD) whose emission properties are sensitive to water. PFO was substituted with a single cysteine at most of the residues between amino acids K189 and N218, and then each cysteine was modified with NBD. Each purified NBD-labeled PFO was then bound to membranes, and the probe's environment was ascertained by measuring its fluorescence lifetime, emission intensity, and collisional quenching with either aqueous (iodide ions) or nonaqueous (nitroxide-labeled phospholipids) quenchers. Lifetime and intensity measurements revealed that the amino acid side chains in this region of the membrane-bound PFO polypeptide alternated between being in an aqueous or a nonaqueous environment. This pattern indicates that this portion of the membrane-bound PFO spans the membrane in an antiparallel beta-sheet conformation. The alternating exposure of these residues to the hydrophobic interior of the bilayer was demonstrated by their susceptibility to quenching by nitroxide moieties attached to phospholipid acyl chains. Residues K189-N218 therefore form a two-stranded, amphipathic beta-sheet in the membrane-bound PFO that creates a stable interface between the pore and the membrane. This same region packs as three short alpha-helices in the soluble, monomeric form of PFO, and therefore, the cholesterol-dependent conversion of PFO to a membrane-bound oligomer involves a major structural transition in which three alpha-helices unfold to form a membrane-spanning amphipathic beta-sheet.  相似文献   

9.
Spin-labeled stearic acid derivatives (N-DS) can be used to determine the rate at which lipid-derived drugs can cross a phospholipid bilayer (flip-flop). The flip-flop rate of N-DS (where N=5, 6, 7, 9, 10, 12, 16), was measured using vectorial photoreduction of nitroxides to their corresponding hydroxylamine by FMN, a charged, membrane-impermeable flavin, by hydrogen atom transfer from EDTA. From the time difference in the photoreduction rates of N-DS located in the outer and inner half of the bilayer, the flip-flop rate of N-DS across the bilayer can be determined. The results show that at pH 8.0 or lower, the photoreduction of 5-DS on one side of the membrane by FMN is slower than the flip-flop rate of 5-DS across phospholipid bilayers. For 5-DS at pH 7.0, this rate is at least 33.8+/-4.24 s or faster. Stearic acids with the spin label at different positions along the acyl chain (N=5, 6, 7, 9, 10, 12) have similar flip-flop rates in the liposomes at pH 7.0 although 16-DS is slower, probably due to the inaccessibility of the nitroxide moiety to FMN. It is most likely that the fast distribution of 5-DS in cells is due to the fast movement of acidic form, but not the salt form, of 5-DS across membrane bilayers. The oxazolidine (nitroxide moiety) does not seem to affect the pKa ( approximately 8.3) of stearic acid at air-water interface. Thus, N-DS are good probes for studying the distribution kinetics of stearic acid derivatives in biological systems.  相似文献   

10.
Steady-state and time-resolved fluorescence data are reported for the local anesthetic dibucaine in the absence and presence of phospholipid vesicles. These vesicles were comprised of dimyristylphosphatidyl choline and approximately 10% dimyristylphosphatidyl glycerol. Solute quenching studies show the bound drug to be protected from collision with iodide ion. The fluorescence lifetime of dibucaine is not significantly changed upon binding to vesicles. The fluorescence anisotropy of dibucaine increases upon association with the vesicles. Anisotropy decay measurements show that the rotational correlation time, phi, of bound dibucaine is increased about one hundred fold over that for free dibucaine. This indicates that the rotational motion of bound dibucaine is slowed by its interaction with the phospholipids. However, we find no evidence that the rotational motion of bound dibucaine is anisotropic.  相似文献   

11.
We report on the surface behavior of a hydrophobic, cationic peptide, [lysine-(leucine)4]4-lysine (KL4), spread at the air/water interface at 25 degrees C and pH 7.2, and its effect at very low molar ratios on the surface properties of the zwitterionic phospholipid 1,2-dipalmitoylphosphatidylcholine (DPPC), and the anionic forms of 1-palmitoyl-2-oleoylphosphatidylglycerol (POPG) and palmitic acid (PA), in various combinations. Surface properties were evaluated by measuring equilibrium spreading pressures (pi(e)) and surface pressure-area isotherms (pi-A) with the Wilhelmy plate technique. Surface phase separation was observed with fluorescence microscopy. KL4 itself forms a single-phase monolayer, stable up to a surface pressure pi of 30 mN/m, and forms an immiscible monolayer mixture with DPPC. No strong interaction was detected between POPG and KL4 in the low pi region, whereas a stable monolayer of the PA/KL4 binary mixture forms, which is attributed to ionic interactions between oppositely charged PA and KL4. KL4 has significant effects on the DPPC/POPG mixture, in that it promotes surface phase separation while also increasing pi(e) and pi(max), and these effects are greatly enhanced in the presence of PA. In the model we have proposed, KL4 facilitates the separation of DPPC-rich and POPG/PA-rich phases to achieve surface refinement. It is these two phases that can fulfill the important lung surfactant functions of high surface pressure stability and efficient spreading.  相似文献   

12.
The spontaneous transfer between membranes of GM3, a ganglioside present in a vesicular form of aggregation instead of micellar form like the majority of gangliosides in aqueous medium, has been studied. Upon incubation of GM3 in the presence of dipalmitoylphosphatidylcholine (DPPC) large unilamellar vesicles at 50 degrees C, mixed GM3/DPPC vesicles are formed. The maximum amount of GM3 that can be inserted into vesicles is about 8%. The temperature dependence of the kinetics has been followed by the excimer formation technique, using the fluorescent analogue pyrenyldodecanoyl-GM3. The transfer of ganglioside from its vesicles to DPPC vesicles depends on the physicochemical characteristics of both the donor and of the acceptor vesicles and increases with the temperature (k = 0.006 0.012, 0.037 at 30, 41 and 50 degrees C, respectively), with a major break point at 41 degrees C and a minor one at 35 degrees C. These temperatures correspond to the gel- to liquid-crystalline transition of DPPC (Tm = 41.3 degrees C), and to a temperature transition displayed by GM3 ganglioside. Similar experiments performed with erythrocyte ghosts yielded a rate constant of 0.04 at 37 degrees C. For the transfer of ganglioside from DPPC (donor) to DMPC (acceptor) the rate constants were 0 at 15 degrees C (both phospholipids in the gel phase), 0.005 at 37 degrees C (donor in the gel phase, acceptor in the fluid phase) and 0.04 at 50 degrees C (both phospholipids in the fluid phase).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Nuclear magnetic relaxation rates for water protons in aqueous palmitoyloleoylphosphatidylcholine vesicle suspensions containing different nitroxide free radical spin labels are reported as a function of magnetic field strength corresponding to proton Larmor frequencies from 10 kHz to 30 MHz. Under these conditions the water proton relaxation rate is determined by the magnetic coupling between the water protons and the paramagnetic nitroxide fixed on the phospholipid. This coupling is made time-dependent by the relative translational motion of the water proton spins past the nitroxide radical. Using theories developed by Freed and others, we interpret the NMR relaxation data in terms of localized water translational motion and find that the translational diffusion constant for water within approximately 10 A of the phospholipid surface is 6 x 10(-10) m2 s(-1) at 298 K. Similar results are obtained for three different nitroxide labels positioned at different points on the lipid. The diffusion is a thermally activated process with an activation energy only slightly higher than that for bulk water.  相似文献   

14.
In the mixture of lipids and proteins which comprise pulmonary surfactant, the dominant protein by mass is surfactant protein A (SP-A), a hydrophilic glycoprotein. SP-A forms octadecamers that interact with phospholipid bilayer surfaces in the presence of calcium. Deuterium NMR was used to characterize the perturbation by SP-A, in the presence of 5 mM Ca(2+), of dipalmitoyl phosphatidylcholine (DPPC) properties in DPPC/egg-PG (7:3) bilayers. Effects of SP-A were uniformly distributed over the observed DPPC population. SP-A reduced DPPC chain orientational order significantly in the gel phase but only slightly in the liquid-crystalline phase. Quadrupole echo decay times for DPPC chain deuterons were sensitive to SP-A in the liquid-crystalline mixture but not in the gel phase. SP-A reduced quadrupole splittings of DPPC choline beta-deuterons but had little effect on choline alpha-deuteron splittings. The observed effects of SP-A on DPPC/egg-PG bilayer properties differ from those of the hydrophobic surfactant proteins SP-B and SP-C. This is consistent with the expectation that SP-A interacts primarily at bilayer surfaces.  相似文献   

15.
2-Methoxy-5-(2',3',4'-trimethoxy)-2,4,6-cycloheptatrien-1-one (MTC) is a colchicine analogue that lacks the B ring. 2-Methoxy-5-(2',4'-dimethoxyphenyl)-2,4,6-cycloheptatrien-1-one (MD) is an A-ring analogue of MTC, in which one methoxy group is replaced by a hydrogen atom. This paper describes the kinetic features of MDC binding to tubulin, and compares its behaviour with MTC to analyse the effect of the A-ring modification on the recognition process by tubulin. Binding is accompanied by a strong enhancement of MDC fluorescence and quenching of protein fluorescence. The kinetic and thermodynamic parameters were obtained from fluorescence stopped-flow measurements. The kinetics are described by a single exponential, indicating that this drug does not discriminate between the different tubulin isotypes. The observed pseudo-first-order rate constant of the fluorescence increase upon binding increases in a non-linear way, indicating that this ligand binds with a similar overall mechanism as colchicine and MTC, consisting of a fast initial binding of low affinity followed by a slower isomerisation step leading to full affinity. The K1 and k2 values for MDC at 25 degrees C were 540 +/- 65 M(-1) and 70 +/- 6 s(-1) respectively. From the temperature dependence, a reaction enthalpy change (deltaH(o)1) of the initial binding of 49 +/- 11 kJ/mol(-1) and an activation energy for the second step of 28 +/- 9 kJ/mol(-1) were calculated. Displacement experiments of bound MDC by MTC allowed the determination of a rate constant of reverse isomerisation of 0.60 +/- 0.07 s(-1) at 25 degrees C and the activation energy of 81 +/- 6 kJ/mol(-1). The overall binding constant was (6.3 +/- 0.2) x 10(4) M(-1) at 25 degrees C. Combination of these results with the kinetic parameters for association gives a full characterisation of the enthalpy pathway for the binding of MDC. The pathway of MDC is shown to differ considerably from that of MTC binding. Since its structural difference is located in ring A, this result indicates the use of ring A in the first step. The kinetics of the binding of MDC in the presence of some A-ring colchicine analogues (podophyllotoxin, 3',4',5'-trimethoxyacetophenone and N-acetylmescaline) and a C-ring analogue (tropolone methyl ether) suggest that the A and C rings are involved in the binding of MDC.  相似文献   

16.
The unfolding kinetics of horse cytochrome c in the oxidized state has been studied at 10, 22, and 34 degreesC as a function of guanidine hydrochloride (GdnHCl) concentration. Rapid (millisecond) measurements of far-UV circular dichroism (CD) as well as fluorescence quenching due to tryptophan to heme excitation energy transfer have been used to monitor the unfolding process. At 10 degreesC, the decrease in far-UV CD signal that accompanies unfolding occurs in two phases. The unobservable burst phase is complete within 4 ms, while the slower phase occurs over tens to hundreds of milliseconds. The burst phase unfolding amplitude increases cooperatively with an increase in GdnHCl concentration, exhibiting a transition midpoint of 3.2 M at 10 degreesC. In contrast, no burst phase change in fluorescence occurs during unfolding at 10 degreesC. At 22 and 34 degreesC, both the fluorescence-monitored unfolding kinetics and the far-UV CD-monitored unfolding kinetics are biphasic. At both temperatures, the two probes yield burst phase unfolding transitions that are noncoincident with respect to the transition midpoints as well as the dependency of the burst phase amplitudes on GdnHCl concentration. The results suggest that at least two kinetic unfolding intermediates accumulate during unfolding. One burst phase intermediate, IU1, has lost virtually all the native-state secondary structure, while the other burst phase intermediate, IU2, has lost both secondary structure and native-like compactness. The presence of kinetic unfolding intermediates is also indicated by the nonlinear dependence of the logarithm of the apparent unfolding rate constant on GdnHCl concentration, which is particularly pronounced at 10 and 22 degreesC. Analysis of the burst phase unfolding transitions obtained using the two probes shows that the stabilities of IU1 and IU2 decrease steadily with an increase in temperature from 10 to 34 degreesC, suggesting that the structures present in them are stabilized principally by hydrogen bonding interactions.  相似文献   

17.
A first order calcium-dependent transition can be monitored by a decrease in the intrinsic fluorescence of the isolated "pro" (Fragment1) region of prothrombin. The maximum fluorescence change is -40% for Fragment 1, and only about -6% for whole prothrombin. The most remarkable features of this transition are its rate and activation energy. The half-life for the transition at 0 degrees is about 100 min, and the temperature dependence shows an activation energy of 21 kcal/mol. The rate constant for the forward reaction is zero order in calcium and is not affected by the presence of phospholipid membranes. The equilibrium for the transition, however, is affected by phospholipid. At 30 degrees, [Ca]eq (the calcium concentration where half of the protein has undergone the transition) is 0.4 mM and the Hill coefficient is 2.6. Under the same conditions but in the presence of phospholipid [Ca]eq is 0.24 mM and the Hill coefficient is about 4.5. The transition is triggered by binding 3 or 4 calcium ions. The rate of Fragment 1 binding to phospholipid vesicles was tested using gel filtration techniques at 0 degrees. The rate constants, activation energy, and [Ca]eq values for this process were shown to correspond to the properties of the fluorescence change. The rate constants, activation energy, and Hill coefficients for binding of whole prothrombin to phospholipid correspond to the same parameters for Fragment 1 but the [Ca]eq values are lower. At 0 degrees, the [Ca]eq is 0.19 mM for the prothrombin transition and 0.1 mM for the transition in the presence of phospholipid. These results demonstrate that Fragment 1 and prothrombin undergo a transition when exposed to calcium ions which necessarily precedes protein-phospholipid interactions. In addition to its role in determining the correct protein structure, calcium plays a second role in prothrombin-phos-pholipid interaction which is in the actual formation of the protein-phospholipid bond. The [Ca]eq for binding protein (after its transition) to phospholipid is about 0.06 mM.  相似文献   

18.
Hemoglobin quenching of the fluorescence intensity of 12-(9-anthroyl)stearic acid (AS) embedded in the red blood cell membrane occurs through an energy transfer mechanism and can be used to measure the binding of hemoglobin to the membrane. The binding of hemoglobin to red cell membranes was found to be reversible and electrostatic in nature. Using a theory of energy transfer based on F?rster formulation, the quantitative data for the binding were derived. The number of binding sites was found to be 1.4 +/- 0.2 X 10(6) molecules per cell and the binding constant was 0.85 X 10(8) M-1.  相似文献   

19.
Divalent cations are required for two roles in prothrombin-phospholipid interaction. The first role, catalysis of a prothrombin protein transition has a reaction half-life of 100 min at 0 degrees and is a prerequisite to phospholipid binding. The binding sites required for the transition have a very low cation specificity. All di- and trivalent cations tested were effective in this role with the exception of beryllium. Barium catalyzed the transition but only at high concentrations (6.6 mM was required for half-reaction). Blood-clotting Factor X, another gamma-carboxyglutamic acid-containing protein, also undergoes a cation-catalyzed protein transition which is a prerequisite to Factor X-phospholipid binding. In both proteins, the transition can be monitored by a decrease in the protein's intrinsic fluorescence. Compared to prothrombin, the Factor X transition occurs much more rapidly, has a somewhat greater specificity for cations, and requires higher concentrations of cations. This indicates that the cation binding sites provided by gamma-carboxyglutamic acid are not completely uniform in all proteins. The second role of divalent cations in prothrombin-phospholipid interaction is in the actual protein-phospholipid binding. This interaction was studied by protein fluorescence quenching resulting from excitation energy transfer to a chromophore attached to the phospholipid membrane. Only strontium and barium satisfactorily replaced calcium in this role. A number of other cations form protein-phospholipid complexes but of the wrong structure. These cations inhibit the prothrombinase complex (Factor Xa, calcium, phospholipid, Factor V). The cation specificity for Factor X-phospholipid binding is the same as for prothrombin except that higher concentrations of cations are required. Factor Xa (generated by action of Russell's viper venom on Factor X) displayed the same calcium requirements for the protein transition and phospholipid interaction as Factor X. The cation requirements of the prothrombinase complex correlate with the cation requirements of prothrombin and Factor X-phospholipid binding. Strontium is the only cation that will singly replace calcium. Barium is ineffective alone because the concentrations required to catalyze the protein transitions cause precipitation of the phospholipid. Combination of certain other cations with barium will, however, substitute for calcium. The other cations (specifically magnesium or manganous ion) catalyze the protein transitions and barium forms the correct protein-phospholipid complexes.  相似文献   

20.
The thermotropic properties of triolein-rich, low-cholesterol dipalmitoyl phosphatidylcholine (DPPC) emulsion particles with well-defined chemical compositions (approximately 88% triolein, 1% cholesterol, 11% diacyl phosphatidylcholine) and particle size distributions (mean diameter, approximately 1000-1100 A) were studied in the absence and presence of apolipoprotein-A1 by a combination of differential scanning and titration calorimetry. The results are compared to egg yolk PC emulsions of similar composition and size. Isothermal titration calorimetry at 30 degrees C was used to saturate the emulsion surface with apo-A1 and rapidly quantitate the binding constants (affinity Ka = 11.1 +/- 3.5 x 10(6) M-1 and capacity N = 1.0 +/- 0.09 apo-A1 per 1000 DPPC) and heats of binding (enthalpy H = -940 +/- 35 kcal mol-1 apo-A1 or -0.92 +/- 0.12 kcal mol-1 DPPC). The entropy of association is -3070 cal deg-1 mol-1 protein or -3 cal deg-1 mol-1 DPPC. Without protein on the surface, the differential scanning calorimetry heating curve of the emulsion showed three endothermic transitions at 24.3 degrees C, 33.0 degrees C, and 40.0 degrees C with a combined enthalpy of 1.53 +/- 0.2 kcal mol-1 DPPC. With apo-A1 on the surface, the heating curve showed the three transitions more clearly, in particular, the second transition became more prominent by significant increases in both the calorimetric and Van't Hoff enthalpies. The combined enthalpy was 2.70 +/- 0.12 kcal mol-1 DPPC and remained constant upon repeated heating and cooling. Indicating that the newly formed DPPC emulsion-Apo-A1 complex is thermally reversible during calorimetry. Thus there is an increase in delta H of 1.17 kcal mol-1 DPPC after apo-A1 is bound, which is roughly balanced by the heat released during binding (-0.92 kcal) of apo-A1. The melting entropy increase, +3.8 cal deg-1 mol-1 DPPC of the three transitions after apo-A1 binds, also roughly balances the entropy (-3 cal deg-1 mol-1 DPPC) of association of apo-A1. These changes indicate that apo-A1 increases the amount of ordered gel-like phase on the surface of DPPC emulsions when added at 30 degrees C. From the stoichiometry of the emulsions we calculate that the mean area of DPPC at the triolein/DPPC interface is 54.5 A2 at 41 degrees C and 54.2 A2 at 30 degrees C. The binding of apo-A1 at 30 degrees C to the emulsion reduces the surface area per DPPC molecule from 54.2 A2 to 50.8 A2. At 30 degrees apo-A1 binds with high affinity and low capacity to the surface of DPPC emulsions and increases the packing density of the lipid domain to which it binds. Apo-A1 was also titrated onto DPPC emulsions at 45 degrees C. This temperature is above the gel liquid crystal transition. No heat was released or adsorbed. Furthermore, egg yolk phosphatidylcholine emulsions of nearly identical composition were also titrated at 30 degrees C with apo-A1 and were euthermic. Association constants were previously measured using a classical centrifugation assay and were used to calculate the entropy of apo-A1 binding (+28 cal deg-1 mol-1 apo-A1). This value indicates that apo-A1 binding to a fluid surface like egg yolk phosphatidylcholine or probably DPPC at 45 degrees C is hydrophobic and is consistent with hydrocarbon lipid or protein moities coming together and excluding water. Thus the binding of apo-A1 to partly crystalline surfaces is entropically negative and increases the order of the already partly ordered phases, whereas binding to liquid surfaces is mainly an entropically driven hydrophobic process.  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