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 43 毫秒
1.
This work advances bottom‐up design of bioinspired materials built from peptide‐amphiphiles, which are a class of bioconjugates in which a biofunctional peptide is covalently attached to a hydrophobic moiety that drives self‐assembly in aqueous solution. Specifically, this work highlights the importance of peptide contour length in determining the equilibrium secondary structure of the peptide as well as the self‐assembled (i.e., micelle) geometry. Peptides used here repeat a seven‐amino acid sequence between one and four times to vary peptide contour length while maintaining similar peptide‐peptide interactions. Without a hydrophobic tail, these peptides all exhibit a combination of random coil and α‐helical structure. Upon self‐assembly in the crowded environment of a micellar corona, however, short peptides are prone to β‐sheet structure and cylindrical micelle geometry while longer peptides remain helical in spheroidal micelles. The transition to β‐sheets in short peptides is rapid, whereby amphiphiles first self‐assemble with α‐helical peptide structure, then transition to their equilibrium β‐sheet structure at a rate that depends on both temperature and ionic strength. These results identify peptide contour length as an important control over equilibrium peptide secondary structure and micelle geometry. Furthermore, the time‐dependent nature of the helix‐to‐sheet transition opens the door for shape‐changing bioinspired materials with tunable conversion rates. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 573–581, 2013.  相似文献   

2.
The crystal structure of a tripeptide Boc‐Leu‐Val‐Ac12c‐OMe ( 1 ) is determined, which incorporates a bulky 1‐aminocyclododecane‐1‐carboxylic acid (Ac12c) side chain. The peptide adopts a semi‐extended backbone conformation for Leu and Val residues, while the backbone torsion angles of the Cα,α‐dialkylated residue Ac12c are in the helical region of the Ramachandran map. The molecular packing of 1 revealed a unique supramolecular twisted parallel β‐sheet coiling into a helical architecture in crystals, with the bulky hydrophobic Ac12c side chains projecting outward the helical column. This arrangement resembles the packing of peptide helices in crystal structures. Although short oligopeptides often assemble as parallel or anti‐parallel β‐sheet in crystals, twisted or helical β‐sheet formation has been observed in a few examples of dipeptide crystal structures. Peptide 1 presents the first example of a tripeptide showing twisted β‐sheet assembly in crystals. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
The capacity to form β‐sheet structure and to self‐organize into amyloid aggregates is a property shared by many proteins. Severe neurodegenerative pathologies such as Alzheimer's disease are thought to involve the interaction of amyloidogenic protein oligomers with neuronal membranes. To understand the experimentally observed catalysis of amyloid formation by lipid membranes and other water‐hydrophobic interfaces, we examine the physico‐chemical basis of peptide adsorption and aggregation in a model membrane using atomistic molecular simulations. Blocked octapeptides with simple, repetitive sequences, (Gly‐Ala)4, and (Gly‐Val)4, are used as models of β‐sheet‐forming polypeptide chains found in the core of amyloid fibrils. In the presence of an n‐octane phase mimicking the core of lipid membranes, the peptides spontaneously partition at the octane‐water interface. The adsorption of nonpolar sidechains displaces the peptides' conformational equilibrium from a heterogeneous ensemble characterized by a high degree of structural disorder toward a more ordered ensemble favoring β‐hairpins and elongated β‐strands. At the interface, peptides spontaneously aggregate and rapidly evolve β‐sheet structure on a 10 to 100 ns time scale, while aqueous aggregates remain amorphous. Catalysis of β‐sheet formation results from the combination of the hydrophobic effect and of reduced conformational entropy of the polypeptide chain. While the former drives interfacial partition and displaces the conformational equilibrium of monomeric peptides, the planar interface further facilitates β‐sheet organization by increasing peptide concentration and reducing the dimensionality of self‐assembly from three to two. These findings suggest a general mechanism for the formation of β‐sheets on the surface of globular proteins and for amyloid self‐organization at hydrophobic interfaces. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
The aggregation behavior of peptides Ac‐VQIVYK‐amide (AcPHF6) and Ac‐QIVYK‐amide (AcPHF5) from the amyloidogenic protein tau was examined by atomic force microscopy (AFM) and fluorescence microscopy. Although AcPHF5 did not show enhancement of thioflavin T (ThT) fluorescence in aqueous buffer, distinct aggregates were discernible when peptide was dissolved in organic solvents such as methanol (MeOH), trifluoroethanol (TFE), and hexafluoroisopropanol (HFIP) dried on mica and examined by AFM. Self‐association was evident even though the peptide did not have the propensity to form secondary structures in the organic solvents. In dried films, the peptide adopts predominantly β‐conformation which results in the formation of distinct aggregates. ThT fluorescence spectra and fluorescence images indicate the formation of fibrils when AcPHF6 solutions in organic solvents were diluted into buffer. AcPHF6 had the ability to organize into fibrillar structures when AFM samples were prepared from peptide dissolved in MeOH, TFE, HFIP, and also when diluted into buffer. AcPHF6 showed propensity for β‐structure in aqueous buffer. In MeOH and TFE, AcPHF6 showed helical and β‐structure. Morphology of the fibrils was dependent on peptide conformation in the organic solvents. The structures observed for AcPHF6 are formed rapidly and long incubation periods in the solvents are not necessary. The structures with varying morphologies observed for AcPHF5 and AcPHF6 appear to be mediated by surfaces such as mica and the organic solvents used for dissolution of the peptides. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
The right‐handed α‐helix is the dominant helical fold of α‐peptides, whereas the left‐handed 314‐helix is the dominant helical fold of β‐peptides. Using molecular dynamics simulations, the properties of α‐helical α‐peptides and 314‐helical β‐peptides with different C‐terminal protonation states and in the solvents water and methanol are compared. The observed energetic and entropic differences can be traced to differences in the polarity of the solvent‐accessible surface area and, in particular, the solute dipole moments, suggesting different reasons for their stability.  相似文献   

6.
Lin Shen  Hua Wan 《Biopolymers》2015,103(10):597-608
Botulinum neurotoxins (BoNTs) are known as the most toxic natural substances. Synaptic vesicle protein 2 (SV2) has been proposed to be a protein receptor for BoNT/A. Recently, two short peptides (BoNT/A‐A2 and SV2C‐A3) were designed to inhibit complex formation between the BoNT/A receptor‐binding domain (BoNT/A‐RBD) and the synaptic vesicle protein 2C luminal domain (SV2C‐LD). In this article, the two peptide complex systems are studied by molecular dynamics (MD) simulations. The structural stability analysis indicates that BoNT/A‐A2 system is more stable than SV2C‐A3 system. The conformational analysis implies that the β‐sheet in BoNT/A‐A2 system maintains its secondary structure but the two β‐strands in SV2C‐A3 system have remarkable conformational changes. Based on the calculation of hydrogen bonds, hydrophobic interactions and cation‐π interactions, it is found that the internal hydrogen bonds play crucial roles in the structural stability of the peptides. Because of the stable secondary structure, the β‐sheet in BoNT/A‐A2 system establishes effective interactions at the interface and inhibits BoNT/A‐RBD binding to SV2C‐LD. In contrast, without other β‐strands forming internal hydrogen bonds, the two isolated β‐strands in SV2C‐A3 system become the random coil. This conformational change breaks important hydrogen bonds and weakens cation‐π interaction in the interface, so the complex formation is only partially inhibited by the two β‐strands. These results are consistent with experimental studies and may be helpful in understanding the inhibition mechanisms of peptide inhibitors. © 2015 Wiley Periodicals, Inc. Biopolymers 103: 597–608, 2015.  相似文献   

7.
Chiral self‐assembly of peptides is of fundamental interest in the field of biology and material science. Protamine, an alkaline biomacromolecule which is ubiquitous in fish and mammalian, plays crucial roles in directing the helical twisting of DNA. Inspired by this, we reported a bioinspired pathway to direct the hierarchical chiral self‐assembly of a short synthetic dipeptide. The peptide could self‐assemble into negatively charged chiral micelles in water that spontaneously formed a nematic liquid crystalline phase. By incorporation with protamine, the micelles condensed with the protamine into large helical bundles with precisely controlled diameter. Furthermore, to simulate the intracellular environments, we investigated macromolecular crowding on the coassembly of peptide and protamine, which leads to the formation of much thinner helical structures. The results highlight the roles of highly charged biomacromolecules and macromolecular crowding on peptide self‐assembly, which are beneficial for the practical applications of self‐assembling peptides in biomedicine and sensing.  相似文献   

8.
Protein spin labeling to yield the nitroxide‐based R1 side chain is a powerful method to measure protein dynamics and structure by electron spin resonance. However, R1 measurements are complicated by the flexibility of the side chain. While analysis approaches for solvent‐exposed α‐helical environment have been developed to partially account for flexibility, similar work in β‐sheets is lacking. The goal of this study is to provide the first essential steps for understanding the conformational preferences of R1 within edge β‐strands using X‐ray crystallography and double electron electron resonance (DEER) distance measurements. Crystal structures yielded seven rotamers for a non‐hydrogen‐bonded site and three rotamers for a hydrogen‐bonded site. The observed rotamers indicate contextual differences in R1 conformational preferences compared to other solvent‐exposed environments. For the DEER measurements, each strand site was paired with the same α‐helical site elsewhere on the protein. The most probable distance observed by DEER is rationalized based on the rotamers observed in the crystal structure. Additionally, the appropriateness of common molecular modeling methods that account for R1 conformational preferences are assessed for the β‐sheet environment. These results show that interpretation of R1 behavior in β‐sheets is difficult and indicate further development is needed for these computational methods to correctly relate DEER distances to protein structure at edge β‐strand sites.  相似文献   

9.
Secondary structural transitions from α‐helix to β‐sheet conformations are observed in several misfolding diseases including Alzheimer's and Parkinson's. Determining factors contributing favorably to the formation of each of these secondary structures is therefore essential to better understand these disease states. β‐hairpin peptides form basic components of anti‐parallel β‐sheets and are suitable model systems for characterizing the fundamental forces stabilizing β‐sheets in fibrillar structures. In this study, we explore the free energy landscape of the model β‐hairpin peptide GB1 and its E2 isoform that preferentially adopts α‐helical conformations at ambient conditions. Umbrella sampling simulations using all‐atom models and explicit solvent are performed over a large range of end‐to‐end distances. Our results show the strong preference of GB1 and the E2 isoform for β‐hairpin and α‐helical conformations, respectively, consistent with previous studies. We show that the unfolded states of GB1 are largely populated by misfolded β‐hairpin structures which differ from each other in the position of the β‐turn. We discuss the energetic factors contributing favorably to the formation of α‐helix and β‐hairpin conformations in these peptides and highlight the energetic role of hydrogen bonds and non‐bonded interactions. Proteins 2014; 82:2394–2402. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
Acylation of proteins is known to mediate membrane attachment and to influence subcellular sorting. Here, we report that acylation can stabilize secondary structure. Circular dichroism spectroscopy showed that N‐terminal attachment of acyl chains decreases the ability of an intrinsically flexible hydrophobic model peptide to refold from an α‐helical state to β‐sheet in response to changing solvent conditions. Acylation also stabilized the membrane‐embedded α‐helix. This increase of global helix stability did not result from decreased local conformational dynamics of the helix backbone as assessed by deuterium/hydrogen‐exchange experiments. We concluded that acylation can stabilize the structure of intrinsically dynamic helices and may thus prevent misfolding.  相似文献   

11.
Elastin, a core protein of the elastic fibers, exhibits the coacervation (temperature‐dependent reversible association/dissociation) under physiological conditions. Because of this characteristic, elastin and elastin‐derived peptides have been considered to be useful as base materials for developing various biomedical products, skin substitutes, synthetic vascular grafts, and drug delivery systems. Although elastin‐derived polypeptide (Val‐Pro‐Gly‐Val‐Gly)n also has been known to demonstrate coacervation property, a sufficiently high (VPGVG)n repetition number (n > 40) is required for coacervation. In the present study, a series of elastin‐derived peptide (Phe‐Pro‐Gly‐Val‐Gly)5 dimers possessing high coacervation potential were newly developed. These novel dimeric peptides exhibited coacervation at significantly lower concentrations and temperatures than the commonly used elastin‐derived peptide analogs; this result suggests that the coacervation ability of the peptides is enhanced by dimerization. Circular dichroism (CD) measurements indicate that the dimers undergo similar temperature‐dependent and reversible conformational changes when coacervation occurs. The molecular dynamics calculation results reveal that the sheet‐turn‐sheet motif involving a type II β‐turn‐like structure commonly observed among the dimers and caused formation of globular conformation of them. These synthesized peptide dimers may be useful not only as model peptides for structural analysis of elastin and elastin‐derived peptides, but also as base materials for developing various temperature‐sensitive biomedical and industrial products. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Ups1 forms a complex with Mdm35 and is critical for the transport of phosphatidic acid (PA) from the mitochondrial outer membrane to the inner membrane. We report the crystal structure of the Ups1‐Mdm35‐PA complex and the functional characterization of Ups1‐Mdm35 in PA binding and transfer. Ups1 features a barrel‐like structure consisting of an antiparallel β‐sheet and three α‐helices. Mdm35 adopts a three‐helical clamp‐like structure to wrap around Ups1 to form a stable complex. The β‐sheet and α‐helices of Ups1 form a long tunnel‐like pocket to accommodate the substrate PA, and a short helix α2 acts as a lid to cover the pocket. The hydrophobic residues lining the pocket and helix α2 are critical for PA binding and transfer. In addition, a hydrophilic patch on the surface of Ups1 near the PA phosphate‐binding site also plays an important role in the function of Ups1‐Mdm35. Our study reveals the molecular basis of the function of Ups1‐Mdm35 and sheds new light on the mechanism of intramitochondrial phospholipid transport by the MSF1/PRELI family proteins.  相似文献   

13.
《Chirality》2017,29(1):5-9
Amyloid beta‐protein 42 plays an important role in the onset and progression of Alzheimer's disease. Familial mutations have identified the glutamate residue 22 as a hotspot with regard to peptide neurotoxicity. We introduce an approach to study the influence of systematic sidechain modification at this residue, employing chirality as a structural probe. Circular dichroism experiments reveal that charge‐preserving alterations of the amino acid sidechain attenuate the characteristic random coil to β‐sheet transition associated with the wildtype peptide. Removal of the negative charge from residue 22, a trait observed with all known familial mutations at this residue, gives rise to a peptide with limited random coil propensity and high β‐sheet characteristics. Our approach can be extended to other residues of Aβ, as well as further amyloidogenic peptides.  相似文献   

14.
Aggregation of β‐amyloid peptides into senile plaques has been identified as one of the hallmarks of Alzheimer's disease. An attractive therapeutic strategy for Alzheimer's disease is the inhibition of the soluble β‐amyloid aggregation using synthetic β‐sheet breaker peptides that are capable of binding Aβ but are unable to become part of a β‐sheet structure. As the early stages of the Aβ aggregation process are supposed to occur close to the neuronal membrane, it is strategic to define the β‐sheet breaker peptide positioning with respect to lipid bilayers. In this work, we have focused on the interaction between the β‐sheet breaker peptide acetyl‐LPFFD‐amide, iAβ5p, and lipid membranes, studied by ESR spectroscopy, using either peptides alternatively labeled at the C‐ and at the N‐terminus or phospholipids spin‐labeled in different positions of the acyl chain. Our results show that iAβ5p interacts directly with membranes formed by the zwitterionic phospholipid dioleoyl phosphatidylcholine and this interaction is modulated by inclusion of cholesterol in the lipid bilayer formulation, in terms of both peptide partition coefficient and the solubilization site. In particular, cholesterol decreases the peptide partition coefficient between the membrane and the aqueous medium. Moreover, in the absence of cholesterol, iAβ5p is located between the outer part of the hydrophobic core and the external hydrophilic layer of the membrane, while in the presence of cholesterol it penetrates more deeply into the lipid bilayer. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
Ligand binding of neutral progesterone, basic propranolol, and acidic warfarin to human α1‐acid glycoprotein (AGP) was investigated by Raman spectroscopy. The binding itself is characterized by a uniform conformational shift in which a tryptophan residue is involved. Slight differences corresponding to different contacts of the individual ligands inside the β‐barrel are described. Results are compared with in silico ligand docking into the available crystal structure of deglycosylated AGP using quantum/molecular mechanics. Calculated binding energies are ?18.2, ?14.5, and ?11.5 kcal/mol for warfarin, propranolol, and progesterone, respectively. These calculations are consistent with Raman difference spectroscopy; nevertheless, minor discrepancies in the precise positions of the ligands point to structural differences between deglycosylated and native AGP. Thermal dynamics of AGP with/without bounded warfarin was followed by Raman spectroscopy in a temperature range of 10–95 °C and analyzed by principal component analysis. With increasing temperature, a slight decrease of α‐helical content is observed that coincides with an increase in β‐sheet content. Above 45 °C, also β‐strands tend to unfold, and the observed decrease in β‐sheet coincides with an increase of β‐turns accompanied by a conformational shift of the nearby disulfide bridge from high‐energy trans‐gauche‐trans to more relaxed gauche‐gauche‐trans. This major rearrangement in the vicinity of the bridge is not only characterized by unfolding of the β‐sheet but also by subsequent ligand release. Hereby, ligand binding alters the protein dynamics, and the more rigid protein–ligand complex shows an improved thermal stability, a finding that contributes to the reported chaperone‐like function of AGP. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

16.
Peptide P11‐4 (QQRFEWEFEQQ) was designed to self‐assemble to form β‐sheets and nematic gels in the pH range 5–7 at concentrations ≥12.6 mM in water. This self‐assembly is reversibly controlled by adjusting the pH of the solvent. It can also self‐assemble into gels in biological media. This together with its biocompatibility and biodegradability make P11‐4 an attractive building block for the fabrication of nanoscale materials with uses in, for example, tissue engineering. A limitation to large‐scale production of such peptides is the high cost of solid phase chemical synthesis. We describe expression of peptide P11‐4 in the bacterium Escherichia coli from constructs carrying tandem repeats of the peptide coding sequence. The vector pET31b+ was used to express P11‐4 repeats fused to the ketosteroid isomerase protein which accumulates in easily recoverable inclusion bodies. Importantly, the use of auto‐induction growth medium to enhance cell density and protein expression levels resulted in recovery of 2.5 g fusion protein/L culture in both shake flask and batch fermentation. Whole cell detergent lysis allowed recovery of inclusion bodies largely composed of the fusion protein. Cyanogen bromide cleavage followed by reverse phase HPLC allowed purification of the recombinant peptide with a C‐terminal homoserine lactone (rP11‐4(hsl)). This recombinant peptide formed pH dependent hydrogels, displayed β‐structure measured by circular dichroism and fibril formation observed by transmission electron microscopy. Biotechnol. Bioeng. 2009;103: 241–251. © 2009 Wiley Periodicals, Inc.  相似文献   

17.
β‐Cardiotoxin is a novel member of the snake venom three‐finger toxin (3FTX) family. This is the first exogenous protein to antagonize β‐adrenergic receptors and thereby causing reduction in heart rates (bradycardia) when administered into animals, unlike the conventional cardiotoxins as reported earlier. 3FTXs are stable all β‐sheet peptides with 60–80 amino acid residues. Here, we describe the three‐dimensional crystal structure of β‐cardiotoxin together with the identification of a molten globule intermediate in the unfolding pathway of this protein. In spite of the overall structural similarity of this protein with conventional cardiotoxins, there are notable differences observed at the loop region and in the charge distribution on the surface, which are known to be critical for cytolytic activity of cardiotoxins. The molten globule intermediate state present in the thermal unfolding pathway of β‐cardiotoxin was however not observed during the chemical denaturation of the protein. Interestingly, circular dichroism (CD) and NMR studies revealed the presence of α‐helical secondary structure in the molten globule intermediate. These results point to substantial conformational plasticity of β‐cardiotoxin, which might aid the protein in responding to the sometimes conflicting demands of structure, stability, and function during its biological lifetime.  相似文献   

18.
Fibril formation is the hallmark of pathogenesis in Alzheimer's disease and other amyloid disorders caused by conformational alterations leading to the aggregation of soluble monomers. Aβ40 self‐associates to form amyloid fibrils. Its central seven‐residue segment KLVFFAE (Aβ16–22), which is thought to be crucial for fibril formation of the full‐length peptide, forms fibrils even in isolation. Context‐dependent induction of amyloid formation by such sequences in peptides, which otherwise do not have that propensity, is of considerable interest. We have examined the effect of introducing the Aβ16–22 sequence at the N‐terminus of two amphipathic helical 18‐residue peptides Ac‐WYSEMKRNVQRLERAIEE‐am and Ac‐KQLIRFLKRLDRNLWGLA‐am, which have high average hydrophobic moment <μH> values but have net charges of 0 and +4, respectively, at neutral pH. Upon incubation in aqueous buffer, fibril‐like aggregates were discernible by transmission electron microscopy for the peptide with only 0 net charge, which also displayed ThT binding and β‐structure. Although both the sequences have been derived from amphipathic helical segments in globular proteins and possess high average hydrophobic moments, the +4 charge peptide lacks the ability to form fibrils, while the peptide with 0 charge has the tendency to form fibrillar structures. Variation in the net charge and the presence of several glutamic acids in the sequence of the peptide with net charge 0 appear to favor the formation of fibrils when the Aβ16–22 sequence is attached at the N‐terminus. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Examples of homomeric β‐helices and β‐barrels have recently emerged. Here we generalize the theory for the shear number in β‐barrels to encompass β‐helices and homomeric structures. We introduce the concept of the “β‐strip,” the set of parallel or antiparallel neighboring strands, from which the whole helix can be generated giving it n‐fold rotational symmetry. In this context, the shear number is interpreted as the sum around the helix of the fixed register shift between neighboring identical β‐strips. Using this approach, we have derived relationships between helical width, pitch, angle between strand direction and helical axis, mass per length, register shift, and number of strands. The validity and unifying power of the method is demonstrated with known structures including α‐hemolysin, T4 phage spike, cylindrin, and the HET‐s(218‐289) prion. From reported dimensions measured by X‐ray fiber diffraction on amyloid fibrils, the relationships can be used to predict the register shift and the number of strands within amyloid protofilaments. This was used to construct models of transthyretin and Alzheimer β(40) amyloid protofilaments that comprise a single strip of in‐register β‐strands folded into a “β‐strip helix.” Results suggest both stabilization of an individual β‐strip helix and growth by addition of further β‐strip helices can involve the same pair of sequence segments associating with β‐sheet hydrogen bonding at the same register shift. This process would be aided by a repeat sequence. Hence, understanding how the register shift (as the distance between repeat sequences) relates to helical dimensions will be useful for nanotube design.  相似文献   

20.
We describe a method that can thoroughly sample a protein conformational space given the protein primary sequence of amino acids and secondary structure predictions. Specifically, we target proteins with β‐sheets because they are particularly challenging for ab initio protein structure prediction because of the complexity of sampling long‐range strand pairings. Using some basic packing principles, inverse kinematics (IK), and β‐pairing scores, this method creates all possible β‐sheet arrangements including those that have the correct packing of β‐strands. It uses the IK algorithms of ProteinShop to move α‐helices and β‐strands as rigid bodies by rotating the dihedral angles in the coil regions. Our results show that our approach produces structures that are within 4–6 Å RMSD of the native one regardless of the protein size and β‐sheet topology although this number may increase if the protein has long loops or complex α‐helical regions. Proteins 2010. © Published 2009 Wiley‐Liss, Inc.  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