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 296 毫秒
1.
We applied precise densimetry and ultrasound velocimetry methods to study the interaction of a synthetic α-helical transmembrane peptide, acetyl-K2-L24-K2-amide (L24), with model bilayer lipid membranes. The large unilamellar vesicles (LUVs) utilized were composed of a homologous series of n-saturated diacylphosphatidylcholines (PCs). PCs whose hydrocarbon chains contained from 13 to 16 carbon atoms, thus producing phospholipid bilayers of different thicknesses and gel to liquid-crystalline phase transition temperatures. This allowed us to analyze how the difference between the hydrophobic length of the peptide and the hydrophobic thickness of the lipid bilayer influences the thermodynamical and mechanical properties of the membranes. We showed that the incorporation of L24 decreases the temperature and cooperativity of the main phase transition of all LUVs studied. The presence of L24 in the bilayer also caused an increase of the specific volume and of the volume compressibility in the gel state bilayers. In the liquid crystalline state, the peptide decreases the specific volume at relatively higher peptide concentration (mole ratio L24:PC = 1:50). The overall volume compressibility of the peptide-containing lipid bilayers in the liquid-crystalline state was in general higher in comparison with pure membranes. There was, however, a tendency for the volume compressibility of these lipid bilayers to decrease with higher peptide content in comparison with bilayers of lower peptide concentration. For one lipid composition, we also compared the thermodynamical and mechanical properties of LUVs and large multilamellar vesicles (MLVs) with and without L24. As expected, a higher cooperativity of the changes of the thermodynamical and mechanical parameters took place for MLVs in comparison with LUVs. These results are in agreement with previously reported DSC and 2H NMR spectroscopy study of the interaction of the L24 and structurally related peptides with phosphatidylcholine bilayers. An apparent discrepancy between 2H NMR spectroscopy and compressibility data in the liquid crystalline state may be connected with the complex and anisotropic nature of macroscopic mechanical properties of the membranes. The observed changes in membrane mechanical properties induced by the presence of L24 suggest that around each peptide a distorted region exists that involves at least 2 layers of lipid molecules.  相似文献   

2.
The binding of polymyxin-B to lipid bilayer vesicles of synthesis phosphatidic acid was studied using fluorescence, ESR spectroscopy and electron microscopy. 1,6-Diphenylhexatriene (which exhibits polarized fluorescence) and pyrene decanoic acid (which forms excimers) were used as fluorescene probes to study the lipid phase transition.The polymyxin binds strongly to negatively charged lipid layers. As a result of lipid/polymyxin chain-chain interactions, the transition temperature of the lipid. This can be explained in terms of a slight expansion of the crystalline lipid lattice (Lindeman's rule). Upon addition of polymyxin to phosphatidic acid vesicles two rather sharp phase transitions (with ΔT = 5°C) are observed. The upper transition (at Tu) is that of the pure lipid and the lower transition (at T1) concerns the lipids bound to the peptide. The sharpness of these transitions strongly indicates that the bilayer is characterized by a heterogeneous lateral distribution of free and bound lipid regions, one in the crystalline and the other in the fluid state. Such a domain structure was directly observed by electron microscopy (freeze etching technique). In (1:1) mixtures of dipalmitoyl phosphatidic acid and egg lecithin, polymyxin induces the formation of domains of charged lipid within the fluid regions of egg lecithin.With both fluorescence methods the fraction of lipid bound to polymxin-B as a function of the peptide concentration was determined. S-shaped binding curves were obtained. The same type of binding curve is obtained for the interaction action of Ca2+ with phosphatidic acid lamellae, while the binding of polylysine to such membranes is characterized by a linear or Langmuir type binding curve. The S-shaped binding curve can be explained in terms of a cooperative lipid-ligand (Ca2+, polymyxin) interaction.A model is proposed which explains the association of polymyxing within the membrane plane in terms of elastic forces caused by the elastic distortion of the (liquid crystalline) lipid layer by this highly asymmetric peptide.  相似文献   

3.
Glucagon can interact rapidly with multilamellar vesicles of dimyristoyl glycerophosphocholine over a narrow temperature range around or above the phase transition temperature of the pure phospholipid. The temperature dependence of the rates arises, in large part, from glucagon-induced alterations in the phase transition properties of the phospholipid. Similar effects are observed with dilaury glycerophosphocholine but the rate of reaction of glucagon with multilamellar dipalmitoyl glycerophosphocholine is too slow to measure.The rate of reaction of glucagon with equimolar mixtures of two phospholipid molecules has also been studied. Mixtures of dilauryl glycerophosphocholine and distearoyl glycerophosphocholine are known to exhibit lateral phase separation in the gel state. The presence of distearoyl glycerophosphocholine has no effect on the rate of reaction with glucagon, despite the increased number of phase boundaries present. In the case of mixtures of dilauryl glycerophosphocholine and dimyristoyl glycerophosphocholine, glucagon appears to induce some lateral phase separation. This is demonstrated by the ability of glucagon to react rapidly with this lipid mixture, even at temperatures well below the phase transition temperature of the mixture and by differential scanning calorimetry.The thermodynamics of the binding of glucagon to dimyristoyl glycerophosphocholine and dilauryl glycerophosphocholine were analyzed with Scatchard plots calculated from measurements of the fluorescence enhancement caused by lipids. Equilibrium binding constants of glucagon to dimyristoyl glycerophosphocholine and dilauryl glycerophosphocholine are 1·105 and 5·104 M?1, respectively. These values are relatively insensitive to temperature, indicating that the equilibrium being measured is between lipid-bound glucagon and free lipid which has had its phase transition properties altered. The number of moles of lipid bound per mole of glucagon decreases markedly above the phase transition temperature. In the water-soluble complex formed between glucagon and dimyristoyl glycerophosphocholine, the peptide binds directly to only 40% of the lipid molecules but, nevertheless, is able to modify the phase transition properties of all of the lipid in the particle.  相似文献   

4.
The interaction of zwitterionic lipid DMPC and DPPC with cyclic hexapeptide, cyclo (D-Ala-L-Pro-L-Ala)2 was studied using circular dichroism (CD) and differential scanning calorimetry (DSC). Preliminary membrane conductance results showed that the peptide has a tendency to form channels inside the lipid bilayer. CD studies indicated that as the lipid/peptide (L/P) ratio (DMPC/peptide) was increased, the magnitude of the negative CD band having a lambda(max) around 200 nm decreased. At a L/P ratio of 210:1, this band disappeared completely, indicating dramatic conformational changes in the peptide on interaction with the lipid bilayer. Reduction of the phase transition temperature and the maximum heat capacity of the lipid bilayer (DPPC) for gel-to-liquid crystalline phase transition indicates a strong interaction of the peptide with the lipid bilayer.  相似文献   

5.
X-ray small-angle diffraction, differential scanning calorimetry (DSC), and temperature scanning densitometry (TSD) were used to study the effect of -lysin on the phase transitions of lipid assemblies from 1,2-0-dixehadecyl-sn-glycero-3-phosphoholine (DHPC). The experiments were carried out in excess of water in a temperature range of 0–55 °C, and at low peptide concentrations between 10-4 and 10-2 moles peptide per mole phospholipid. The incorporation of -lysin into lipid assemblies alters the lipid structure without significant changes on the temperatures of phase transition from gel to liquid crystalline phase. The temperature of the main transition was nearly unaffected. A reduction in the transition volume of the lipids with increasing concentrations of -lysin was observed. The minor changes in these parameters were interpreted as long-range structural changes caused by the peptide incorporation. The results are discussed in terms of the concept of cooperative phase transition of entire clusters occurring within a membrane implying that relative stable domains of gel phase, and liquid crystalline phase co-exist.  相似文献   

6.
The influence of epidermal growth factor (EGF), 0.75 μg g?1; insulin, 1.5 μg g?1; glucagon, 1.25 ygg?1 and their combinations on the activities of hepatic pyruvate kinase (PK) and malic enzymes (ME) was monitored. Male CD2F1 mice were treated toward the end of the light or dark periods, 9 or 23 /tours after /ights on (9 or 23 HALO), and subgroups of six mice were killed at 4,8 or 12 hr post-treatment. PK and ME activities from control mice were well characterized by cosine curves. The PK activity was maximal when ME activity was minimal at the transition from light to dark (9 HALO plus 4 hr) and PK was at a minimum when ME was highest (23 HALO plus 4 hr). Both enzymes were influenced by at least one peptide hormone, and the effects were strongly circadian -stage dependent. The only effect attributed to EGF was an increase of PK activity (23%) 12 hr after injection at 23 HALO. PK activity was increased by insulin (23%) at 23 HALO (4 hr after injection), but not at 9 HALO, and decreased (17%) by glucagon 12 hr after injection at 9 HALO. Several reductions in PK activity in response to various combinations of peptides were observed, and appeared to be caused by glucagon but influenced by insulin. The activity of ME was decreased (33%) in response to insulin 4 hr after injection at 23 HALO but not at 9 HALO and increased (60-70%) by glucagon alone or in combinations with insulin or EGF, or both, at 4 hr after injection at 9 HALO but not at 23 HALO. In general, when ME activity was altered by either insulin or glucagon, PK activity was also altered in the opposite direction, and the effects of glucagon were opposed by insulin.  相似文献   

7.
A novel bilirubin oxidase (BOD), from the rice blast fungus Magnaporthe oryzae, has been identified and isolated. The 64-kDa protein containing four coppers was successfully overexpressed in Pichia pastoris and purified to homogeneity in one step. Protein yield is more than 100?mg for 2?L culture, twice that of Myrothecium verrucaria. The k cat/K m ratio for conjugated bilirubin (1,513?mM?1?s?1) is higher than that obtained for the BOD from M. verrucaria expressed in native fungus (980?mM?1?s?1), with the lowest K m measured for any BOD highly desirable for detection of bilirubin in medical samples. In addition, this protein exhibits a half-life for deactivation >300?min at 37?°C, high stability at pH?7, and high tolerance towards urea, making it an ideal candidate for the elaboration of biofuel cells, powering implantable medical devices. Finally, this new BOD is efficient in decolorizing textile dyes such as Remazol brilliant Blue R, making it useful for environmentally friendly industrial applications.  相似文献   

8.
The DNA helix–coil transition has been studied in the presence of high concentrations of manganese ions (about 10?3M), which corresponds to the conditions close to equal stability of the A+T and G+C pairs, at the ionic strengths of 10?1, 10?2, and 1.6 × 10?3M Na+. With the Mn2+ ion effect, the transition range is significantly reduced to not more than 0.2°C at 1.2 × 10?3M Mn2+ and 1.6 × 10?3M Na+. The melting curves display a sharp kink at the end of the helix–coil transition, which is interpreted as an indication of the second-order phase transition. It is shown that the melting curves obtained can be approximated by a simple analytical expression 1 – θ = exp[–a(tc - t)], where θ is the DNA helix fraction, tc is the phase transition temperature, and a is an empirical parameter characterizing the breadth of the melting range and responsible for the magnitude of a jump of the helicity derivative with respect to the temperature at the phase transition point.  相似文献   

9.
1. The properties of 3,5-di-tert-butyl-4-hydroxybenzylidenemalononitrile (SF 6847) were studied chemically and spectroscopically. Two molecular species of SF6847 were identified: the undissociated form (SFH; ?363, 10 mM?1) and the dissociated form (SF?; ?454, 35 mM?1). The pKa value of the molecule was determined to be 6.9.2. On the basis of these properties the interactions of SF6847 with liposomes and valinomycin · K+ were studied. The partition constants of SFH (Knp and SF? (K?p) to liposomes were determined separately; Knp was 56 mM?1 and was independent of the pH of the medium, whereas K?p dependend greatly on the pH, being 1.2 mM?1 at pH 7.0 and 2.9 mM?1 at pH 8.0. Using these values, the partition constant of total SF6847 (Kp) was calculated and found to be essentially the same as that calculated from the kinetics of proton uptake. It was concluded that the amount of SF? bound to liposomes is rate limiting for proton uptake.3. The effects of membrane potential on partition constants were studied. The K?p decreased greatly upon generation of a membrane potential negative inside the liposomes but increased upon generation of a membrane potential positive inside the liposomes.4. The interaction of SF6847 with valinomycin in aqueous solution and in liposomes was demonstrated only in the presence of potassium ion. Potassium ion could not be replaced by sodium ion. Evidence was obtained for the formation of the ternary complex valinomycin · K+ · SF? in liposomes and in hexane. It was concluded that SF? became more soluble in the liposomal membranes on formation of this ternary complex. All these results support our proposed mechanism for the proton uptake cycle (Yamaguchi, A. and Anraku, Y. (1978) Biochim. Biophys. Acta 501, 136–149).  相似文献   

10.
In this research, ecological risks for eight individual polycyclic aromatic hydrocarbons (PAHs) and ∑PAH8 in surface sediments from middle and lower reaches of Yellow River are evaluated using overlapping areas of probability density curves and margin of safety (MOS), based on the toxicity data and the exposure concentrations of PAHs in sediments collected from 23 sites. In the overlapping areas of probability density curves, the risk of Ant and Pyr are the highest, then the risk level is in the order of Flua > Nap > Phe > BaP > Flu > Ace. The values of MOS10 present that Pyr (4.62 × 10?4), Ant (5.60 × 10?3), and Flua (6.4 × 10?3) have a significantly high ecological risk level, while Nap and Phe have middle-level ecological risk. As for Ace, BaP, and Flu, they pose limited risk to the ecological system with MOS10 greater than 1.0. The ∑PAH8 (2.66 × 10?5) is a higher risk level than that of any individual PAHs, where the probabilities of ∑PAH8 in excess of the 10th percentile of the toxicity data were 86%.  相似文献   

11.
The hydrophobic lung surfactant SP-B is essential for respiration. SP-B promotes spreading and adsorption of surfactant at the alveolar air-water interface and may facilitate connections between the surface layer and underlying lamellar reservoirs of surfactant material. SP-B63–78 is a cationic and amphipathic helical peptide containing the C-terminal helix of SP-B. 2H NMR has been used to examine the effect of SP-B63–78 on the phase behavior and dynamics of bicellar lipid dispersions containing the longer chain phospholipids DMPC-d 54 and DMPG and the shorter chain lipid DHPC mixed with a 3∶1∶1 molar ratio. Below the gel-to-liquid crystal phase transition temperature of the longer chain components, bicellar mixtures form small, rapidly reorienting disk-like particles with shorter chain lipid components predominantly found around the highly curved particle edges. With increasing temperature, the particles coalesce into larger magnetically-oriented structures and then into more extended lamellar phases. The susceptibility of bicellar particles to coalescence and large scale reorganization makes them an interesting platform in which to study peptide-induced interactions between lipid assemblies. SP-B63–78 is found to lower the temperature at which the orientable phase transforms to the more extended lamellar phase. The peptide also changes the spectrum of motions contributing to quadrupole echo decay in the lamellar phase. The way in which the peptide alters interactions between bilayered micelle structures may provide some insight into some aspects of the role of full-length SP-B in maintaining a functional surfactant layer in lungs.  相似文献   

12.
The effect of phospholipid structure on the interaction between small peptides and phospholipid membranes has been studied by high-sensitivity differential scanning calorimetry. The peptides used, N-Boc-beta-Ala-Trp-Met-Arg-Phe-NH2 and N-Boc-beta-Ala-Trp-Met-Lys-Phe-NH2, are basic analogs of the hormone pentagastrin. These peptides split the gel-to-liquid crystalline phase transition of synthetic phosphatidylcholines into two components. For dimyristoyl (DMPC), dipalmitoyl (DPPC) and 1-stearoyl-2-oleoyl (SOPC) phosphatidylcholines, one component remains at the temperature corresponding to that of pure lipid and the other one is shifted towards higher temperatures. With increasing peptide concentration there is a gradual increase in the enthalpy of the high-temperature component at the expense of the low-temperature one, and there is also an increase in the total enthalpy of the transition. A mixture of the peptide with distearoylphosphatidylcholine (DSPC) behaves differently, with the transition occurring at a temperature below that of the pure lipid increasing with peptide concentration. The susceptibility of various phosphatidylcholines to perturbation by the peptides increases in the order DMPC greater than SOPC greater than DPPC greater than DSPC. The effect of these peptides on the phase transitions of acidic phosphatidylglycerols is generally greater than with the corresponding phosphatidylcholines, but the dependence on the length of lipid hydrocarbon chains is similar. Perturbation of the thermotropic phase transition is strongest for dimyristoylphosphatidylglycerol, followed by the dipalmitoyl and the distearoyl analogs. The effect of the peptides on the phase transition of dimyristoylphosphatidylserine is significantly smaller compared to that observed with dimyristoylphosphatidylglycerol and it is further reduced for dimyristoylphosphatidic acid. The phase transition of this latter lipid remains virtually unchanged, even in the presence of high concentrations of the peptide. Similar resistance to the perturbation of the phase transitions by the peptides is observed for synthetic phosphatidylethanolamine. The different susceptibility of various phospholipids to perturbation by the peptides is suggested to be related to different degrees of intermolecular interaction between phospholipid molecules, and particularly to different abilities of phospholipids to form intermolecular hydrogen bonding.  相似文献   

13.
Glucagon is found to interact with dimyristoyl glycerophosphocholine both above and below the phase transition temperature of the lipid. Above the phase transition temperature the interaction is manifested by an increase in the rate of vesicle aggregation and by an increased permeability of unilamellar vesicles to Eu3+ and to Fe(CN)63−. However, no stable lipoprotein complex can be detected by gel filtration. Below the phase transition glucagon can form stable complexes with dimyristoyl glycerophosphocholine vesicles but cannot rapidly rearrange these vesicles to disk-shaped particles until the phase transition temperature is approached. The energy of activation for the dissociation of glucagon from the disk-shaped lipoprotein particle is 29 kcal/mol at temperatures above 36°C but increases markedly at lower temperatures, as the region of the lipid phase transition is approached. This increase in energy of activation at lower temperatures is most probably due to the larger amount of energy required to rearrange gel-state lipid in the transition state and provides an explanation for the unusual kinetic stability of the glucagon-dimyristoyl glycerophosphocholine lipoprotein complex only at temperatures below the phase transition of the lipid.  相似文献   

14.
Ca-polygalacturonate is a demethoxylated component of pectins which are constitutive of plant root mucigel. In order to define the role of root mucigel in myrosinase immobilization and activity at root level, a myrosinase enzyme which had been isolated from Sinapis alba seeds was immobilized into Ca-polygalacturonate. The activity profile for the immobilized and free enzyme was evaluated using the pH-Stat method as a function of time, temperature, and pH. The Michaelis-Menten kinetic parameters change between the immobilized (V max ?=?127?±?13 U mg?1 protein; K M ?=?6.28?±?0.09?mM) and free (V max ?=?17?±?1 U mg?1 protein; K M ?=?0.96?±?0.01?mM) forms of myrosinase, probably due to conformational changes involving the active site as a consequence of enzyme immobilization. Immobilized enzyme activity evaluated as a function of different substrates gave the highest value with nasturtin, the glucosinolate that is typical of several brassicaceae plant roots containing the glucosinolate-myrosinase defensive system. No feedback regulation mechanism was found in the presence of an excess of enzymatic reaction products (i.e. allyl isothiocyanate or sulphate). The high enzyme immobilization yield into Ca-polygalacturonate and its activity preservation under different conditions suggest that the enzyme released by plants at root level could be entrapped in root mucigel in order to preserve its activity.  相似文献   

15.
Despite the presence of a similar number of glucagon and VIP receptors in liver membranes, VIP induces a negligeable stimulation of adenylate cyclase when compared with glucagon effect. In order to elucidate these discrepancies, the effects of guanine nucleotides on the VIP and glucagon-responsive adenylate cyclase of liver were compared using pure ATP as substrate. 10?8 M VIP accounted for a 1.5-fold increase of basal activity. In the presence of GTP or Gpp(NH)p (10?9 to 10?5 M), the level of cAMP production induced by VIP was no more than additive. In contrast, Gpp(NH)p potentiated the effect of glucagon on liver adenylate cyclase. These discrepancies are not explained by a difference in the peptide binding process. These data suggest that, in liver membranes, a GTP-binding protein N2 is associated with the glucagon-sensitive adenylate cyclase, but is not detected for VIP. It is suggested that N2 appears to be specific for the peptidic receptor.  相似文献   

16.
Serine hydroxymethyltransferase (SHMT) catalyzes the interconversion of serine and tetrahydrofolate (THF) to glycine and methylenetetrahydrofolate. cDNA encoding Bombyx mori SHMT (bmSHMT) was cloned and sequenced. The deduced amino acid sequence consisted of 465 amino acids and was found to share homology with other SHMTs. Recombinant bmSHMT was overexpressed in Escherichia coli and purified to homogeneity. The enzyme showed optimum activity at pH 3.0 and 30°C and was stable under acidic conditions. The Km and kcat/Km values for THF in the presence of Nicotinamide adenine dinucleotide phosphate (NADP+) were 0.055 mM and 0.081 mM?1 s?1, respectively, whereas those toward NADP+ were 0.16 mM and 0.018 mM?1 s?1 and toward l ‐serine were 1.8 mM and 0.0022 mM?1 s?1, respectively. Mutagenesis experiments revealed that His119, His132, and His135 are important for enzymatic activity. Our results provide insight into the roles and regulation mechanism of one‐carbon metabolism in the silkworm B. mori.  相似文献   

17.
A new approach to intentionally induce phase transition of Li‐excess layered cathode materials for high‐performance lithium ion batteries is reported. In high contrast to the limited layered‐to‐spinel phase transformation that occurred during in situ electrochemical cycles, a Li‐excess layered Li[Li0.2Mn0.54Ni0.13Co0.13]O2 is completely converted to a Li4Mn5O12‐type spinel product via ex situ ion‐exchanges and a post‐annealing process. Such a layered‐to‐spinel phase conversion is examined using in situ X‐ray diffraction and in situ high‐resolution transmission electron microscopy. It is found that generation of sufficient lithium ion vacancies within the Li‐excess layered oxide plays a critical role for realizing a complete phase transition. The newly formed spinel material exhibits initial discharge capacities of 313.6, 267.2, 204.0, and 126.3 mAh g?1 when cycled at 0.1, 0.5, 1, and 5 C (1 C = 250 mA g?1), respectively, and can retain a specific capacity of 197.5 mAh g?1 at 1 C after 100 electrochemical cycles, demonstrating remarkably improved rate capability and cycling stability in comparison with the original Li‐excess layered cathode materials. This work sheds light on fundamental understanding of phase transitions within Li‐excess layered oxides. It also provides a novel route for tailoring electrochemical performance of Li‐excess layered cathode materials for high‐capacity lithium ion batteries.  相似文献   

18.
The interaction between glucagon and dicaprylphosphatidylcholine (DCPC) was studied by fluorescence, circular dichroism and calorimetry, as well as by 1H- and 31P-nuclear magnetic resonance. The water-soluble lipid-protein complex was also characterized by gel filtration and ultracentrifugation. The complex appeared to be monodisperse by sedimentation equilibrium measurements, with a molecular weight of (4.55 ± 0.57)·104. This complex contained approximately 7 molecules of glucagon and 35 molecules of phospholipid. Proton-decoupled 31P-NMR spectra of the phospholipid in the lipid-protein complex display narrower resonances than those of sonicated vesicles of DCPC, and 1H-31P coupling could be detected in proton coupled spectra. These NMR results, together with gel-filtration results, suggest that glucagon ‘solubilizes’ phospholipid aggregates, forming a lipid-protein complex which is smaller than sonicated preparations of DCPC. 1H-NMR resonance of both the methionine methyl group (met-27) and the aromatic envelope of glucagon are broadened by the phospolipid, indicating that the C-terminal region and the aromatic residues are involved in the interaction with the phospholipid. Nuclear magnetic resonance titrations of the imidazole ring C(2) and C(4) protons of the histidine residue of glucagon show that DCPC lowers the pK of the imidazole. The alterations caused by the phospholipid in the far and near ultraviolet CD spectra of glucagon reflect, respectively, the increased helix content of the hormone and the fact that the aromatic residues are located in a more structured environment. The phospholipid also alters the fluorescence properties of glucagon, shifting the fluorescence emission maximum of the hormone to shorter wavelength, and enhancing its relative intensity. This suggests that the fluorophore is experiencing a more hydrophobic environment in the presence of the lipid. Binding of glucagon to the phospholipid was analysed by Scatchard plots of the enhancement of fluorescence caused by the phospholipid and showed that the equilibrium binding constants of glucagon to DCPC are (4.4 ± 0.5)·104M?1 and (7.5±0.5)·104M?1, at 15°C and 25°C, respectively. The average number of moles of phospholipid bound per mole of glucagon is 4.4±0.6. The isothermal enthalpy of reaction of glucagon with DCPC is ?20.5 kcal/mol of glucagon at 25°C and ?32.5 kcal/mol of glucagon at 15°C. The observed enthalpies can arise from glucagon-induced cyrstallization of the phospholipid, from the non-covalent interactions between the peptide and lipid as well as from the lipid-induced conformational change in the protein. These results demonstrate that, unlike the complexes formed between glucagon and phospholipids which form more stable bilayers, the complex formed between glucagon and DCPC is stable over a wide range of temperatures, including temperatures well above the phase transition.  相似文献   

19.
Apolipophorin III (apoLp-III) from Locusta migratoria was employed as a model apolipoprotein to gain insight into binding interactions with lipid vesicles. Differential scanning calorimetry (DSC) was used to measure the binding interaction of apoLp-III with liposomes composed of mixtures of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and sphingomyelin (SM). Association of apoLp-III with multilamellar liposomes occurred over a temperature range around the liquid crystalline phase transition (Lα). Qualitative and quantitative data were obtained from changes in the lipid phase transition upon addition of apoLp-III. Eleven ratios of DMPC and SM were tested from pure DMPC to pure SM. Broadness of the phase transition (T1/2), melting temperature of the phase transition (Tm) and enthalpy were used to determine the relative binding affinity to the liposomes. Multilamellar vesicles composed of 40% DMPC and 60% SM showed the greatest interaction with apoLp-III, indicated by large T1/2 values. Pure DMPC showed the weakest interaction and liposomes with lower percentage of DMPC retained domains of pure DMPC, even upon apoLp-III binding indicating demixing of liposome lipids. Addition of apoLp-III to rehydrated liposomes was compared to codissolved trials, in which lipids were rehydrated in the presence of protein, forcing the protein to interact with the lipid system. Similar trends between the codissolved and non-codissolved trials were observed, indicating a similar binding affinity except for pure DMPC. These results suggested that surface defects due to non-ideal packing that occur at the phase transition temperature of the lipid mixtures are responsible for apolipoprotein-lipid interaction in DMPC/SM liposomes.  相似文献   

20.
S. Krimm 《Biopolymers》1983,22(1):217-225
A vibrational force field for the polypeptide chain has been developed for normal-mode analysis of such molecules. It can reproduce observed frequencies of known structures to within about 5 cm?1. We review the application of this technique to conformational problems in peptides (β-turns and their model compounds), polypeptides [the αII-helix and crystalline poly(glycine II)], and proteins (bacteriorhodopsin and glucagon).  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