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 31 毫秒
1.
The semi-empirical MSINDO method has been used to investigate the mode of adsorption of benzoic acid on the nano anatase TiO(2) (100) surface. The (100) surface is modeled with a Ti(36)O(90)H(36) cluster. Molecular dynamics simulations for the adsorption behavior of benzoic acid indicate it is linked to the TiO(2) surface through interactions from the oxygen atoms of the carboxylic acid moiety with surface titanium atoms. The benzoic acid may be positioned with its aromatic ring either parallel or perpendicular relative to the surface, however the perpendicular adsorption mode is more stable. The calculated substrate-surface interaction energy is influenced by the number of linkages between the substrate and the surface as well as the degree of hydrogen bonding between the acid hydrogen and lattice oxygen atom. The greater stability of the perpendicular adsorption orientation is ascribed to the higher number of linkages between the substrate and the surface. It is concluded that the simplified model is sufficiently detailed to elucidate surface interactions.  相似文献   

2.
Surfactant-like membranes containing the 21-residue peptide KLLLLKLLLLKLLLLKLLLLK (KL4), have been clinically tested as a therapeutic agent for respiratory distress syndrome in premature infants. The aims of this study were to investigate the interactions between the KL4 peptide and lipid bilayers, and the role of both the lipid composition and KL4 structure on the surface adsorption activity of KL4-containing membranes. We used bilayers of three-component systems [1,2-dipalmitoyl-phosphatidylcholine/1-palmitoyl-2-oleoyl-phosphatidylglycerol/palmitic acid (DPPC/POPG/PA) and DPPC/1-palmitoyl-2-oleoyl-phosphatidylcholine (POPC)/PA] and binary lipid mixtures of DPPC/POPG and DPPC/PA to examine the specific interaction of KL4 with POPG and PA. We found that, at low peptide concentrations, KL4 adopted a predominantly alpha-helical secondary structure in POPG- or POPC-containing membranes, and a beta-sheet structure in DPPC/PA vesicles. As the concentration of the peptide increased, KL4 interconverted to a beta-sheet structure in DPPC/POPG/PA or DPPC/POPC/PA vesicles. Ca2+ favored alpha<-->beta interconversion. This conformational flexibility of KL4 did not influence the surface adsorption activity of KL4-containing vesicles. KL4 showed a concentration-dependent ordering effect on POPG- and POPC-containing membranes, which could be linked to its surface activity. In addition, we found that the physical state of the membrane had a critical role in the surface adsorption process. Our results indicate that the most rapid surface adsorption takes place with vesicles showing well-defined solid/fluid phase co-existence at temperatures below their gel to fluid phase transition temperature, such as those of DPPC/POPG/PA and DPPC/POPC/PA. In contrast, more fluid (DPPC/POPG) or excessively rigid (DPPC/PA) KL4-containing membranes fail in their ability to adsorb rapidly onto and spread at the air-water interface.  相似文献   

3.
The adsorption of nitrogen-containing compounds (NCCs) including 2,4,6-trinitrotoluene (TNT), 2,4-dinitrotoluene (DNT), 2,4-dinitroanisole (DNAN), and 3-nitro-1,2,4-triazol-5-one (NTO) on kaolinite surfaces was investigated. The M06-2X and M06-2X-D3 density functionals were applied with the cluster approximation. Several different positions of NCCs relative to the adsorption sites of kaolinite were examined, including NCCs in perpendicular and parallel orientation toward both surface models of kaolinite. The binding between the target molecules and kaolinite surfaces was analyzed and bond energies were calculated applying the atoms in molecules (AIM) method. All NCCs were found to prefer a parallel orientation toward both kaolinite surfaces, and were bound more strongly to the octahedral than to the tetrahedral site. TNT exhibited the strongest interaction with the octahedral surface and DNAN with the tetrahedral surface of kaolinite. Hydrogen bonding was shown to be the dominant non-covalent interaction for NCCs interacting with the octahedral surface of kaolinite with a small stabilizing effect of dispersion interactions. In the case of adsorption on the tetrahedral surface, kaolonite–NCC binding was shown to be governed by the balance between hydrogen bonds and dispersion forces. The presence of water as a solvent leads to a significant decrease in the adsorption strength for all studied NCCs interacting with both kaolinite surfaces.  相似文献   

4.
Wymore T  Wong TC 《Biophysical journal》1999,76(3):1199-1212
Two neuropeptides, substance P (SP) and SP-tyrosine-8 (SP-Y8), have been studied by molecular dynamics (MD) simulation in a TIP3P water/CCl4 biphasic solvent system as a mimic for the water-membrane system. Initially, distance restraints derived from NMR nuclear Overhauser enhancements (NOE) were incorporated in the restrained MD (RMD) in the equilibration stage of the simulation. The starting orientation/position of the peptides for the MD simulation was either parallel to the water/CCl4 interface or in a perpendicular/insertion mode. In both cases the peptides equilibrated and adopted a near-parallel orientation within approximately 250 ps. After equilibration, the conformation and orientation of the peptides, the solvation of both the backbone and the side chain of the residues, hydrogen bonding, and the dynamics of the peptides were analyzed from trajectories obtained in the RMD or the subsequent free MD (where the NOE restraints were removed). These analyses showed that the peptide backbone of nearly all residues are either solvated by water or are hydrogen-bonded. This is seen to be an important factor against the insertion mode of interaction. Most of the interactions with the hydrophobic phase come from the hydrophobic interactions of the side chains of Pro-4, Phe-7, Phe-8, Leu-10, and Met-11 for SP, and Phe-7, Leu-10, Met-11 and, to a lesser extent, Tyr-8 in SP-Y8. Concerted conformational transitions took place in the time frame of hundreds of picoseconds. The concertedness of the transition was due to the tendency of the peptide to maintain the necessary secondary structure to position the peptide properly with respect to the water/CCl4 interface.  相似文献   

5.
ACTH (1-10), an adrenocorticotropin hormone fragment, was studied by molecular dynamics (MD) simulation in the NPT ensemble in an explicit sodium dodecylsulfate (SDS) micelle. Initially, distance restraints derived from NMR nuclear Overhauser enhancements were incorporated during the equilibration stage of the simulation. The analyses of the trajectories from the subsequent unrestrained MD showed that ACTH (1-10) does not conform to a helical structure at the micelle-water interface; however, the structure is amphipathic. The loss of the helical structure is due to decreased intramolecular hydrogen bonding accompanied by an increase of hydrogen bonding between the amide hydrogens of the peptide and the micelle head-groups. ACTH (1-10) was found to lie on the surface of the SDS micelle. Most of the hydrophobic interactions came from the side-chains of Met-4, Phe-7 and Trp-9. The peptide bonds were either hydrated or involved in intramolecular hydrogen bonding. Decreased hydration for the backbone of His-6 and Phe-7 was due to intermolecular hydrogen bonding with the SDS head-groups. The time correlation functions of the N-H bonds of the peptide in water and in the micelle showed that the motions of the peptide, except for the N- and C-termini, are significantly reduced when partitioned in the micelle.  相似文献   

6.
We have investigated the folding of polyalanine by combining discontinuous molecular dynamics simulation with our newly developed off-lattice intermediate-resolution protein model. The thermodynamics of a system containing a single Ac-KA(14)K-NH(2) molecule has been explored by using the replica exchange simulation method to map out the conformational transitions as a function of temperature. We have also explored the influence of solvent type on the folding process by varying the relative strength of the side-chain's hydrophobic interactions and backbone hydrogen bonding interactions. The peptide in our simulations tends to mimic real polyalanine in that it can exist in three distinct structural states: alpha-helix, beta-structures (including beta-hairpin and beta-sheet-like structures), and random coil, depending upon the solvent conditions. At low values of the hydrophobic interaction strength between nonpolar side-chains, the polyalanine peptide undergoes a relatively sharp transition between an alpha-helical conformation at low temperatures and a random-coil conformation at high temperatures. As the hydrophobic interaction strength increases, this transition shifts to higher temperatures. Increasing the hydrophobic interaction strength even further induces a second transition to a beta-hairpin, resulting in an alpha-helical conformation at low temperatures, a beta-hairpin at intermediate temperatures, and a random coil at high temperatures. At very high values of the hydrophobic interaction strength, polyalanines become beta-hairpins and beta-sheet-like structures at low temperatures and random coils at high temperatures. This study of the folding of a single polyalanine-based peptide sets the stage for a study of polyalanine aggregation in a forthcoming paper.  相似文献   

7.
The amyloid beta (1-40) peptide (A beta) is the main component of amyloid deposits found in the brain of patients afflicted with Alzheimer's disease. After treatment with hexafluoroisopropanol, commercial A beta is readily soluble in water and buffers at pH 7.4 and has an irregular secondary structure. The adsorption of A beta to the water-air interface and to the surface of the dipalmitoylphosphatidylethanolamine monolayer at a surface pressure pi close to zero leads to an increase in pressure up to 17 mN/m. When being adsorbed, the molecules of the peptide occupy a part of the monolayer surface, which leads to the compression of lipid molecules forming the monolayer. Further compression of the monolayer composed of the molecules of the lipid and peptide leads to the extrusion of the peptide from the monolayer. If the lipid monolayer is preliminarily (prior to the addition of the peptide to the liquid phase) compressed to pi = 30 mN/m, no adsorption of the peptide to the monolayer occurs. No changes in the structure of the dipalmitoylphosphatidylethanolamine monolayer were detected by the sliding X-ray diffraction method, indicating the absence of specific interactions. The method of reflection and absorption infrared spectroscopy makes it possible to determine the conformation of the adsorbed peptide and its orientation in the lipid monolayer. It was found that A beta has the conformation of a beta-fold oriented parallel to the interface, as it is the case with the adsorption of peptide molecules to the lipid monolayer at pi < 30 mN/m and upon adsorption to the interface that is not occupied by the lipid.  相似文献   

8.
We have performed molecular dynamics simulations of the interactions of the peptide SP-B(1-25), which is a truncated version of the full pulmonary surfactant protein SP-B, with dipalmitoylphosphatidylcholine monolayers, which are the major lipid components of lung surfactant. Simulations of durations of 10-20 ns show that persistent hydrogen bonds form between the donor atoms of the protein and the acceptors of the lipid headgroup and that these bonds determine the position, orientation, and secondary structure of the peptide in the membrane environment. From an ensemble of initial conditions, the most probable equilibrium orientation of the alpha-helix of the peptide is predicted to be parallel to the interface, matching recent experimental results on model lipid mixtures. Simulations of a few mutated analogs of SP-B(1-25) also suggest that the charged amino acids are important in determining the position of the peptide in the interface. The first eight amino acids of the peptide, also known as the insertion sequence, are found to be essential in reducing the fluctuations and anchoring the peptide in the lipid/water interface.  相似文献   

9.
Gao X  Wong TC 《Biopolymers》2001,58(7):643-659
Adrenocorticotropin (ACTH) (1-10), an adrenocorticotropin hormone fragment, has been studied by molecular dynamics (MD) simulation in an NPT ensemble in an explicit dodecylphosphocholine (DPC) micelle. Two starting configurations of the peptide/micelle system, corresponding to the insertion and surface-binding modes, were used. A common equilibrated configuration, in which the peptide lies parallel to the micellar surface, was reached from both simulations. In the initial part of the simulations, distance restraints derived from NMR nuclear Overhauser enhancements were incorporated before the peptide reached an equilibrium configuration with respect to the micelle. Analyses of the trajectories from the subsequent free (unrestrained) MD simulation showed that ACTH (1-10) does not conform strictly to a helical structure. The loss of the helical structure is due to decreased intramolecular hydrogen bonding accompanied by an increase of hydrogen bonding between the amide protons of the peptide and the micellar head groups. However, the extent of the latter interaction is less pronounced than in the negatively charged SDS micelle. The final structure enhances the amphipathic nature of the peptide, facilitating better interactions at the water-hydrophobic interface. The primary hydrophobic interactions with the micelle came from the side chains of Met4, Phe7, and Trp9. All peptide bonds were either hydrated or were involved in intramolecular hydrogen bonding. The interactions with the DPC micelle, the conformation of the bound peptide, and the dynamics of the peptide, as revealed by the time correlation functions of the N-H bonds, were compared with those of the ACTH (1-10)/SDS system studied previously by MD simulations.  相似文献   

10.
Cell membranes provide an environment for several types of molecular processes and we are attempting to mimic the cell membranes' environment on a chromatography solid support. Chromatography solid supports utilizing lecithin as the bonded phase were synthesized and the HPLC behavior of hydrophilic peptides evaluated. A diC14 lecithin containing a terminal carboxy group on the C2 fatty acid chain was amidated with the surface amines of Nucleosil-300 (7NH2) silica particles. Based on elemental analysis, lecithin was coupled to Nucleosil-300 (7NH2) at a surface density near that of lecithin found in biological membranes and this novel chromatographic support material is denoted as Nucleosil-lecithin, the prototype immobilized artificial membrane. Infrared difference spectra of Nucleosil-lecithin minus Nucleosil-300 (7NH2) clearly showed amide I (1653.1 cm-1) and amide II (1550.9 cm-1) bands, giving direct spectroscopic evidence for the amide linkage. Spectral deconvolution resolved two peaks for the amide I band, and three peaks for the amide II band. This demonstrates lecithin interchain amide hydrogen bonding and/or hydrogen bonds between the lecithin amide link and unreacted silica surface amines. Nucleosil-lecithin as a solid phase mimics membranes and can be used to study the interactions of biomolecules with membranes. Our primary objective is to develop HPLC methods for studying the interaction between cell membranes and peptide sequences found near the interfaces of cell membranes. A frequency distribution of amino acids bracketing approximately 400 transmembrane peptide sequences showed Cys to be the least frequently occurring amino acid at this putative interfacial membrane region. Hydrophilic peptide analogs bearing Cys were used as model compounds to test Nucleosil-lecithin solid supports. Small peptides, six to eight amino acids in length, containing Cys bind approximately 2X tighter to Nucleosil-lecithin compared to identical peptides without the Cys residue. Thus, Cys at the interface of cells may stabilize protein-lipid interactions.  相似文献   

11.
The penetratin peptide, a 16 amino acid sequence extracted from Antennapedia homeodomain, is able to translocate across a neural cell membrane through an unknown mechanism, most likely a non-specific interaction with membrane lipids. Beyond its potential application as vector targeting small hydrophilic molecules and enabling them to reach a cell nucleus, this observation raises intriguing questions concerning the physico-chemistry of peptide-lipid interactions. Here we present a study of the role of lipid surface pressure and head charge on the mechanism of interaction. This was performed using optical techniques: surface infrared spectroscopy and ellipsometry, applied to a monolayer of phospholipids deposited at the air-water interface. Determination of the structure and orientation of peptides and lipids (separately or together) evidenced that electrostatic rather than amphiphilic interactions determine the peptide adsorption and its action on lipids.  相似文献   

12.
This review is focused on peptide molecules which exhibit a limited solubility in the aqueous phase and bind to the lipid membrane from the aqueous medium. Surface adsorption, membrane insertion, and specific binding are usually accompanied by changes in the heat content of the system and can be measured conveniently with isothermal titration calorimetry, avoiding the necessity of peptide labeling. The driving forces for peptide adsorption and binding are hydrophobicity, electrostatics, and hydrogen bonding. An exclusively hydrophobic interaction is exemplified by the immunosuppressant drug cyclosporine A. Its insertion into the membrane can be described by a simple partition equilibrium X(b)=K(0)C(eq). If peptide and membrane are both charged, electrostatic interactions are dominant leading to nonlinear binding curves. The concentration of the peptide near the membrane interface can then be much larger than its bulk concentration. Electrostatic effects must be accounted for by means of the Gouy-Chapman theory before conventional binding models can be applied. A small number of peptides and proteins bind with very high affinity to a specific lipid species only. This is illustrated for the lantibiotic cinnamycin (Ro 09-0198) which forms a 1:1 complex with phosphatidyethanolamine with a binding constant of 10(8) M(-1). Membrane adsorption and insertion can be accompanied by conformational transitions facilitated, in part, by hydrogen bonding mechanisms. The two membrane-induced conformational changes to be discussed are the random coil-to-alpha-helix transition of amphipathic peptides and the random coil-to-beta-structure transition of Alzheimer peptides.  相似文献   

13.
Contrary to the widespread view that hydrogen bonding and its entropy effect play a dominant role in protein folding, folding into helical and hairpin-like structures is observed in molecular dynamics (MD) simulations without hydrogen bonding in the peptide-solvent system. In the widely used point charge model, hydrogen bonding is calculated as part of the interaction between atomic partial charges. It is removed from these simulations by setting atomic charges of the peptide and water to zero. Because of the structural difference between the peptide and water, van der Waals (VDW) interactions favor peptide intramolecular interactions and are a major contributing factor to the structural compactness. These compact structures are amino acid sequence dependent and closely resemble standard secondary structures, as a consequence of VDW interactions and covalent bonding constraints. Hydrogen bonding is a short range interaction and it locks the approximate structure into the specific secondary structure when it is included in the simulation. In contrast to standard molecular simulations where the total energy is dominated by charge-charge interactions, these simulation results will give us a new view of the folding mechanism.  相似文献   

14.
The acid-catalyzed hydrogen exchange rate constants kH, and the base-catalyzed rate constants kOH, have been determined (in the preceding paper) for the 25 most rapidly exchanging NH groups of bovine pancreatic trypsin inhibitor. Most of these NH groups are at the protein-solvent interface. The correlation of kH, but not kOH, with the static accessibility and hydrogen bonding of the peptide carbonyl O atom indicates that the mechanism of acid catalysis in proteins involves O-protonation. Agreement between the ionic strength dependence observed for kH and kOH and the ionic strength dependence calculated for an O-protonation mechanism supports this conclusion. N-protonation for acid catalysis, as well as N-deprotonation for base catalysis, have traditionally been assumed in the mechanism of the chemical step in peptide amide proton exchange. A preference for the alternative O-protonation mechanism has far-reaching implications in the interpretation of protein hydrogen exchange kinetics. With an O-protonation mechanism, acid-catalyzed rates of surface NH groups are primarily a function of the average solvent accessibility of the carbonyl O atoms in the dynamic solution structure, while base-catalyzed rates of surface NH groups measure solvent accessibility of the peptide N. The relative dynamic accessibilities of peptide O atoms, as measured by relative values of kH (corrected for electrostatic effects), correlate with O static accessibilities in the crystal structure. A lower correlation of static accessibility of N atoms with kOH is observed for surface NH groups in peptide groups in which the carbonyl O is not hydrogen bonded. For some surface NH groups, the observed pH of minimum rate, pHmin, deviates widely from the pHmin of model compounds. This is explained as the combined result of electrostatic effects and of the differences in accessibility of the carbonyl O and N atoms that result in a change in the relative values of kH and kOH as compared to those of model peptides. A mechanism whereby exchange of interior sites is catalyzed by interactions of catalysis ions with protein surface atoms via charge transfer is suggested.  相似文献   

15.
Tobi D  Elber R  Thirumalai D 《Biopolymers》2003,68(3):359-369
The conformational equilibrium of a blocked valine peptide in water and aqueous urea solution is studied using molecular dynamics simulations. Pair correlation functions indicate enhanced concentration of urea near the peptide. Stronger hydrogen bonding of urea-peptide compared to water-peptide is observed with preference for helical conformation. The potential of mean force, computed using umbrella sampling, shows only small differences between urea and water solvation that are difficult to quantify. The changes in solvent structure around the peptide are explained by favorable electrostatic interactions (hydrogen bonds) of urea with the peptide backbone. There is no evidence for significant changes in hydrophobic interactions in the two conformations of the peptide in urea solution. Our simulations suggest that urea denatures proteins by preferentially forming hydrogen bonds to the peptide backbone, reducing the barrier for exposing protein residues to the solvent, and reaching the unfolded state.  相似文献   

16.
The accurate design of new protein–protein interactions is a longstanding goal of computational protein design. However, most computationally designed interfaces fail to form experimentally. This investigation compares five previously described successful de novo interface designs with 158 failures. Both sets of proteins were designed with the molecular modeling program Rosetta. Designs were considered a success if a high‐resolution crystal structure of the complex closely matched the design model and the equilibrium dissociation constant for binding was less than 10 μM. The successes and failures represent a wide variety of interface types and design goals including heterodimers, homodimers, peptide‐protein interactions, one‐sided designs (i.e., where only one of the proteins was mutated) and two‐sided designs. The most striking feature of the successful designs is that they have fewer polar atoms at their interfaces than many of the failed designs. Designs that attempted to create extensive sets of interface‐spanning hydrogen bonds resulted in no detectable binding. In contrast, polar atoms make up more than 40% of the interface area of many natural dimers, and native interfaces often contain extensive hydrogen bonding networks. These results suggest that Rosetta may not be accurately balancing hydrogen bonding and electrostatic energies against desolvation penalties and that design processes may not include sufficient sampling to identify side chains in preordered conformations that can fully satisfy the hydrogen bonding potential of the interface.  相似文献   

17.
Monolayers of oleanolic acid (OLA) mixed with stearic acid (SA) were studied at the air-water interface. The surface pressure-area (pi-A) isotherms, measured over the whole composition range, and BAM observations were used to investigate the phase behaviour and self-organization of these components in a two-dimensional structure. Pure OLA forms a very compressible monolayer, and BAM observation revealed the coexistence of large and irregular solid domains of different thickness dispersed in a gas matrix, compatible with the two most probable orientations of the OLA molecule at the interface. Mixtures of OLA/SA form condensed monolayers from low surface pressures and the thermodynamic analysis indicates that OLA molecules, in the presence of the long-chain SA, orient with the major axis almost perpendicular to the interface. Langmuir-Blodgett (LB) monolayers of pure SA and mixtures were further characterized by atomic force microscopy (AFM) and Fourier transform infrared spectroscopy (FTIR). AFM images of LB mixed monolayers evidenced microphase separation, not observable by BAM. The SA rich domains are 4-6A thicker than those rich in OLA. The FTIR spectra of mixed LB films on CaF2 substrates showed that OLA does not perturb the all-trans conformation of the SA long alkyl chains, up to a mole fraction of 0.4. The carbonyl-stretching band of OLA suggests that the carboxylic groups of neighbour OLA molecules are involved in hydrogen bonds, forming dimers, as in pure solid phase OLA. These interactions seem to prevail over the OLA-water hydrogen bonds.  相似文献   

18.
HscA, a specialized bacterial Hsp70-class molecular chaperone, interacts with the iron-sulfur cluster assembly protein IscU by recognizing a conserved LPPVK sequence motif. We report the crystal structure of the substrate-binding domain of HscA (SBD, residues 389-616) from Escherichia coli bound to an IscU-derived peptide, ELPPVKIHC. The crystals belong to the space group I222 and contain a single molecule in the asymmetric unit. Molecular replacement with the E.coli DnaK(SBD) model was used for phasing, and the HscA(SBD)-peptide model was refined to Rfactor=17.4% (Rfree=21.0%) at 1.95 A resolution. The overall structure of HscA(SBD) is similar to that of DnaK(SBD), although the alpha-helical subdomain (residues 506-613) is shifted up to 10 A relative to the beta-sandwich subdomain (residues 389-498) when compared to DnaK(SBD). The ELPPVKIHC peptide is bound in an extended conformation in a hydrophobic cleft in the beta-subdomain, which appears to be solvent-accessible via a narrow passageway between the alpha and beta-subdomains. The bound peptide is positioned in the reverse orientation of that observed in the DnaK(SBD)-NRLLLTG peptide complex placing the N and C termini of the peptide on opposite sides of the HscA(SBD) relative to the DnaK(SBD) complex. Modeling of the peptide in the DnaK-like forward orientation suggests that differences in hydrogen bonding interactions in the binding cleft and electrostatic interactions involving surface residues near the cleft contribute to the observed directional preference.  相似文献   

19.
The orientation of lipid headgroups may serve as a powerful sensor of electrostatic interactions in membranes. As shown previously by 2H NMR measurements, the headgroup of phosphatidylcholine (PC) behaves like an electrometer and varies its orientation according to the membrane surface charge. Here, we explored the use of solid-state 14N NMR as a relatively simple and label-free method to study the orientation of the PC headgroup in model membrane systems of varying composition. We found that 14N NMR is sufficiently sensitive to detect small changes in headgroup orientation upon introduction of positively and negatively charged lipids and we developed an approach to directly convert the 14N quadrupolar splittings into an average orientation of the PC polar headgroup. Our results show that inclusion of cholesterol or mixing of lipids with different length acyl chains does not significantly affect the orientation of the PC headgroup. In contrast, measurements with cationic (KALP), neutral (Ac-KALP), and pH-sensitive (HALP) transmembrane peptides show very systematic changes in headgroup orientation, depending on the amount of charge in the peptide side chains and on their precise localization at the interface, as modulated by varying the extent of hydrophobic peptide/lipid mismatch. Finally, our measurements suggest an unexpectedly strong preferential enrichment of the anionic lipid phosphatidylglycerol around the cationic KALP peptide in ternary mixtures with PC. We believe that these results are important for understanding protein/lipid interactions and that they may help parametrization of membrane properties in computational studies.  相似文献   

20.
Prion diseases are neurodegenerative disorders characterized by the aggregation of an abnormal form of prion protein. The interaction of prion protein and cellular membrane is crucial to elucidate the occurrence and development of prion diseases. Its fragment, residues 106–126, has been proven to maintain the pathological properties of misfolded prion and was used as a model peptide. In this study, explicit solvent molecular dynamics (MD) simulations were carried out to investigate the adsorption, folding and aggregation of PrP106–126 with different sizes (2-peptides, 4-peptides and 6-peptides) on the surface of both pure neutral POPC (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine) and negatively charged POPC/POPG (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol) (3:1) lipids. MD simulation results show that PrP106–126 display strong affinity with POPC/POPG but does not interact with pure POPC. The positively charged and polar residues participating hydrogen bonding with membrane promote the adsorption of PrP106–126. The presence of POPC and POPC/POPG exert limited influence on the secondary structures of PrP106–126 and random coil structures are predominant in all simulation systems. Upon the adsorption on the POPC/POPG surface, the aggregation states of PrP106–126 have been changed and more small oligomers were observed. This work provides insights into the interactions of PrP106–126 and membranes with different compositions in atomic level, which expand our understanding the role membrane plays in the development of prion diseases. This article is part of a Special Issue entitled: Protein Aggregation and Misfolding at the Cell Membrane Interface edited by Ayyalusamy Ramamoorthy.  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