首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
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.
α‐Helical coiled coil structures, which are noncovalently associated heptad repeat peptide sequences, are ubiquitous in nature. Similar amphipathic repeat sequences have also been found in helix‐containing proteins and have played a central role in de novo design of proteins. In addition, they are promising tools for the construction of nanomaterials. Small‐angle X‐ray scattering (SAXS) has emerged as a new biophysical technique for elucidation of protein topology. Here, we describe a systematic study of the self‐assembly of a small ensemble of coiled coil sequences using SAXS and analytical ultracentrifugation (AUC), which was correlated with molecular dynamics simulations. Our results show that even minor sequence changes have an effect on the folding topology and the self‐assembly and that these differences can be observed by a combination of AUC, SAXS, and circular dichroism spectroscopy. A small difference in these methods was observed, as SAXS for one peptide and revealed the presence of a population of longer aggregates, which was not observed by AUC. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
4.
Tropoelastin is the primary component of elastin, which forms the elastic fibers that make up connective tissues. The hydrophobic domains of tropoelastin are thought to mediate the self‐assembly of elastin into fibers, and the temperature‐mediated self‐assembly (coacervation) of one such repetitive peptide sequence (VPGVG) has been utilized in various bio‐applications. To elucidate a mechanism for coacervation activity enhancement and to develop more potent coacervatable elastin‐derived peptides, we synthesized two series of peptide analogs containing an aromatic amino acid, Trp or Tyr, in addition to Phe‐containing analogs and tested their functional characteristics. Thus, position 1 of the hydrophobic pentapeptide repeat of elastin (X1P2G3V4G5) was substituted by Trp or Tyr. Eventually, we acquired a novel, short Trp‐containing elastin‐derived peptide analog (WPGVG)3 with potent coacervation ability. From the results obtained during this process, we determined the importance of aromaticity and hydrophobicity for the coacervation potency of elastin‐derived peptide analogs. Generally, however, the production of long‐chain synthetic polypeptides in quantities sufficient for commercial use remain cost‐prohibitive. Therefore, the identification of (WPGVG)3, which is a 15‐mer short peptide consisting simply of five natural amino acids and shows temperature‐dependent self‐assembly activity, might serve as a foundation for the development of various kinds of biomaterials. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
Peptide‐based hydrogels are of interest for their potential use in regenerative medicine. Combining these hydrogels with materials that may enhance their physical and biological properties, such as glycosaminoglycans, has the potential to extend their range of biomedical applications, for example in the repair of early cartilage degeneration. The aim of this study was to combine three self‐assembling peptides (P11‐4, P11‐8, and P11‐12) with chondroitin sulphate at two molar ratios of 1:16 and 1:64 in 130 and 230 mM Na+ salt concentrations. The study investigates the effects of mixing self‐assembling peptide and glycosaminoglycan on the physical and mechanical properties at 37°C. Peptide alone, chondroitin sulphate alone, and peptide in combination with chondroitin sulphate were analysed using Fourier transform infrared spectroscopy to determine the β‐sheet percentage, transmission electron microscopy to determine the fibril morphology, and rheology to determine the elastic and viscous modulus of the materials. All of the variables (peptide, salt concentration, and chondroitin sulphate molar ratio) had an effect on the mechanical properties, β‐sheet formation, and fibril morphology of the hydrogels. P11‐4 and P11‐8‐chondroitin sulphate mixtures, at both molar ratios, were shown to have a high β‐sheet percentage, dense entangled fibrillar networks, as well as high mechanical stiffness in both (130 and 230 mM) Na+ salt solutions when compared with the P11‐12/chondroitin sulphate mixtures. These peptide/chondroitin sulphate hydrogels show promise for biomedical applications in glycosaminoglycan depleted tissues.  相似文献   

6.
Oligopeptides are well‐known to self‐assemble into a wide array of nanostructures including β‐sheet‐rich fibers that when present above a critical concentration become entangled and form self‐supporting hydrogels. The length, quantity, and interactions between fibers influence the mechanical properties of the hydrogel formed and this is typically achieved by varying the peptide concentration, pH, ionic strength, or the addition of a second species or chemical cross‐linking agent. Here, we outline an alternative, facile route to control the mechanical properties of the self‐assembling octa‐peptide, FEFEFKFK (FEKII); simply doping with controlled quantities of its double length peptide, FEFEFKFK‐GG‐FKFKFEFE (FEKII18). The structure and properties of a series of samples were studied here (0–100 M% of FEKII18) using Fourier transform infrared, small angle X‐ray scattering, transmission electron microscopy, and oscillatory rheology. All samples were found to contain elongated, flexible fibers and all mixed samples contained Y‐shaped branch points and parallel fibers which is attributed to the longer peptide self‐assembling within two fibers, thus creating a cross‐link in the network structure. Such behavior was reflected in an increase in the elasticity of the mixed samples with increasing quantity of double peptide. Interestingly the elastic modulus increased up to 30 times the pure FEKII value simply by adding 28 M% of FEKII18. These observations provide an easy, off‐the‐shelf method for an end‐user to control the cross‐linked network structure of the peptide hydrogel, and consequently strength of the hydrogel simply by physically mixing pre‐determined quantities of two similar peptide molecules. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 669–680, 2014.  相似文献   

7.
A systematic study was performed of the fiber forming properties and polymerization characteristics of two peptide amphiphiles containing a diacetylene functionality in the alkyl tail comprising 23 and 25 C atoms, respectively. Both diyne containing peptide amphiphiles were able to form stable beta-sheet fibers of micrometers length in an aqueous solution. However, there was a large difference between the stability of the two amphiphiles. This was shown by a large difference in assembly and disassembly temperature and by different behavior during polymerization. Because the monomers were preorganized with a tight molecular packing, the polymerization could be carried out using wavelengths up to 532 nm. For both amphiphiles, the fiber structure did not change when the polymerization was carried out at an elevated temperature. The degree of polymerization, however, barely decreased for the longer amphiphile (2) but showed a gradual decline for the shorter one (1) when the temperature was raised from room temperature to the melting temperature of the fibers. Furthermore, the pH did not influence the fiber assembly for 2, but hampered it for 1 at alkaline pH. The fiber structure was, for both of the amphiphiles, not dependent on the pH. After polymerization, the molecular packing of the amphiphiles was only slightly influenced by an increase in temperature, as indicated by the small color change of polymerized fibers, which was also reversible. Additionally, pH had no influence on the assembly structure, as indicated by the color of the polymer which was the same at all pH values. Thus, both fibers increased in stability upon polymerization. The large difference in assembly and polymerization behavior of the two similar-looking amphiphiles 1 and 2, with a 23 or 25 carbon tail, is indicative of the subtlety of the assembly and disassembly processes in these fibrous architectures.  相似文献   

8.
Abstract We report the first demonstration of nonionic detergent micelle conjugation and phase separation using purpose‐synthesized, peptide amphiphiles, C10‐(Asp)5 and C10‐(Lys)5. Clustering is achieved in two different ways. Micelles containing the negatively charged peptide amphiphile C10‐(Asp)5 are conjugated (a) via a water‐soluble, penta‐Lys mediator or (b) to micelles containing the C10‐(Lys)5 peptide amphiphile. Both routes lead to phase separation in the form of oil‐rich globules visible in the light microscope. The hydrophobic nature of these regions leads to spontaneous partitioning of hydrophobic dyes into globules that were found to be stable for weeks to months. Extension of the conjugation mechanism to micelles containing a recently discovered, light‐driven proton pump King Sejong 1‐2 (KS1‐2) demonstrates that a membrane protein may be concentrated using peptide amphiphiles while preserving its native conformation as determined by characteristic UV absorption. The potential utility of these peptide amphiphiles for biophysical and biomedical applications is discussed.  相似文献   

9.
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   

10.
A novel amphiphilic cyclic peptide composed of two β‐glucosamino acids and one trans‐2‐aminocyclohexylcarboxylic acid was synthesized and investigated on assembly formation. The cyclic tri‐β‐peptide was self‐assembled into rodlike crystals or nanofibers depending on preparative conditions. The rodlike crystals showed a layer spacing of 4.8 Å along the long axis, and columnar spacings of 10.8 and 21.5 Å by electron diffraction analysis along the short axis. The former confirms the columnar structure upon molecular stacking, and the latter indicates triple bundle formation of the columnar assemblies. Fourier transform infrared (FT‐IR) measurement of the fibrous assembly showed formation of homogeneous hydrogen bonds among amide groups, also supporting the molecular stacking of cyclic β‐peptides. Straight nanofibers with uniform diameter were also uniquely obtained. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
It has proven challenging to obtain collagen‐mimetic fibrils by protein design. We recently reported the self‐assembly of a mini‐fibril showing a 35 nm, D‐period like, axially repeating structure using the designed triple helix Col108. Peptide Col108 was made by bacterial expression using a synthetic gene; its triple helix domain consists of three pseudo‐identical units of amino acid sequence arranged in tandem. It was postulated that the 35 nm d‐period of Col108 mini‐fibrils originates from the periodicity of the Col108 primary structure. A mutual staggering of one sequence unit of the associating Col108 triple helices can maximize the inter‐helical interactions and produce the observed 35 nm d‐period. Based on this unit‐staggered model, a triple helix consisting of only two sequence units is expected to have the potential to form the same d‐periodic mini‐fibrils. Indeed, when such a peptide, peptide 2U108, was made it was found to self‐assemble into mini‐fibrils having the same d‐period of 35 nm. In contrast, no d‐periodic mini‐fibrils were observed for peptide 1U108, which does not have long‐range repeating sequences in its primary structure. The findings of the periodic mini‐fibrils of Col108 and 2U108 suggest a way forward to create collagen‐mimetic fibrils for biomedical and industrial applications.  相似文献   

12.
Slc11a1 is an integral membrane protein with 12 putative transmembrane domains (TMDs) and functions as a pH‐coupled divalent metal cation transporter. The conservation of three negatively charged residues in the TMD3 of Slc11 protein family implies the important role of this domain in the function of the proteins. However, aggregation of the transmembrane peptide in micelles prevents structural study of the peptide in these membrane‐mimetic environments by NMR spectroscopy. Here, we characterized the structure, position, and assembly model of Slc11a1‐TMD3 (Lys128‐Ile151) in SDS micelles by the NMR study of its Leu‐substituted peptide. It was found that the two‐site substitutions of Ala for Leu residues at positions 136 and 140 of TMD3 disrupt the aggregation without altering the secondary structure of the peptide. The Leu‐substituted peptide folds as an α‐helix spanning from Leu133 to Gly144 and embedded in the micelles. A Leu zipper is suggested to account for the self‐assembly of the wild‐type peptide in SDS micelles. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
In this work we have probed the interactions of the amyloid Aβ(1–42) peptide with self‐assembled nanospheres. The nanospheres were formed by self‐assembly of a newly developed bolaamphiphile bis(N‐alpha‐amido‐methionine)‐1,8 octane dicarboxylate under aqueous conditions. It was found that the interactions of the Aβ(1–42) peptide with the nanospheres were concentration as well as pH dependent and the peptide largely adopts a random coil structure upon interacting with the nanospheres. Further, upon incorporation with the nanospheres, we observed a relative diminution in the aggregation of Aβ(1–42) at low concentrations of Aβ(1–42). The interactions between the nanospheres and the Aβ(1–42) peptide were investigated by atomic force microscopy, transmission electron microscopy, circular dichroism, FTIR and fluorescence spectroscopy, and the degree of fibrillation in the presence and absence of nanospheres was monitored by the Thioflavine T assay. We believe that the outcome from this work will help further elucidate the binding properties of Aβ peptide as well as designing nanostructures as templates for further investigating the nucleation and fibrillation process of Aβ‐like peptides. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
We utilised a simple bidirectional (N→C and C→N) solid phase synthesis strategy entailing conventional solid phase peptide synthesis and fragment condensation with a water‐soluble carbodiimide to synthesise a model anionic glycylglycine bolaamphiphile containing a suberic acid linker moiety, namely N,N′‐suberoyldiglycylglycine. The synthetic suberoyldiglycylglycine was purified using its inherent ability to rapidly self‐assemble in an aqueous acidic solution (0.1% trifluoroacetic acid). Monitoring of the rapid assembly process corroborated our visual observation and confirmed packing‐directed self‐assembly rather than non‐specific aggregation or precipitation. The progress of suberoyldiglycylglycine self‐assembly was observed to be via the formation of oligomers in the solution, which then self‐assembled to form layered β‐sheet type macrostructures. Within 24 h, nanotubes grew from these macrostructures and eventually combined to formed microtubes, which we isolated after 5–7 days. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
Self‐assembly of PAs composed of palmitic acid and several repeated heptad peptide sequences, C15H31CO‐(IEEYTKK)n‐NH2 (n = 1–4, represented by PA1–PA4), was investigated systematically. The secondary structures of the PAs were characterized by CD. PA3 and PA4 (n = 3 and 4, respectively) showed an α‐helical structure, whereas PA1 and PA2 (n = 1 and 2, respectively) did not display an α‐helical conformations under the tested conditions. The morphology of the self‐assembled peptides in aqueous medium was studied by transmission electron microscopy. As the number of heptad repeats in the PAs increased, the nanostructure of the self‐assembled peptides changed from nanofibers to nanovesicles. Changes of the secondary structures and the self‐assembly morphologies of PA3 and PA4 in aqueous medium with various cations were also studied. The critical micelle concentrations were determined using a pyrene fluorescence probe. In conclusion, this method may be used to design new peptide nanomaterials. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
It is well known that GnRH analogs can self‐assemble into amyloid fibrils and that the duration of action of GnRH analogs depends on the ability of the amyloid to slowly release active peptides. The aim of this study was to investigate the influence of the amino acid residues at position 7 of GnRH analogues on peptide self‐assembly. It was found that the dominant shape of the nanostructure can be changed when the structures of the residues at position 7 differ significantly from that of leucine in Degarelix. When the backbone length was extended (peptide 9), or the side chain of the residue at position 7 was replaced by an aromatic ring (peptide 6), or the rotation of the amide bond was restricted (peptide 8), the nanostructure changed from fibrils to vesicles. The results also indicate that the increasing hydrophilicity had little influence on the nanostructure morphology. In addition, a suitable release rate was found to play a more important role for the duration of the peptide action by maintaining the equilibrium between the drug concentration and the persistent release time, while the nanostructure shape was found to exert little influence on the duration of the peptide action. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
The amyloid β‐peptide with a sequence of 42 amino acids is the major constituent of extracellular amyloid deposits in Alzheimer's disease plaques. The control of the peptide self‐assembly is difficult to achieve because the process is fast and is affected by many variables. In this paper, we describe the effect of different charged and non‐charged surfactants on Aβ(1‐42) fibrillation to define common alternate aggregation pathways. The characterization of the peptide‐surfactant interactions by ultra‐structural analysis, thioflavin T assay and secondary structure analysis, suggested that charged surfactants interact with Aβ(1‐42) through electrostatic interactions. Charged micelles slow down the aggregation process and stabilize the peptide in the oligomeric state, whereas non‐charged surfactants promote the Aβ(1‐42) fibril formation. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
Many bioactive peptides are featured by their unique amino acid compositions such as argine/lysine‐rich peptides. However, histidine‐rich bioactive peptides are hardly found. In this study, histidine‐containing peptides were constructed by selectively replacing the corresponded lysine residues in a lytic peptide LL‐1 with histidines. Interestingly, all resulting peptides demonstrated pH‐dependent activities. The cell lysis activities of these peptides could be increased up to four times as the solution pHs dropped from pH = 7.4 to pH = 5.5. The pH sensitivity of a histidine‐containing peptide was determined by histidine substitution numbers. Peptide derivatives with more histidines were associated with increased pH sensitivity. Results showed that not the secondary structures but pH‐affected cell affinity changes were responsible for the pH‐dependent activities of histidine‐containing peptides. The histidine substitution approach demonstrated here may present a general strategy to construct bioactive peptides with desired pH sensitivity for various applications. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The morphology of structures formed by the self‐assembly of short N‐terminal t‐butyloxycarbonyl (Boc) and C‐terminal methyl ester (OMe) protected and Boc‐deprotected hydrophobic peptide esters was investigated. We have observed that Boc‐protected peptide esters composed of either only aliphatic hydrophobic amino acids or aliphatic hydrophobic amino acids in combination with aromatic amino acids, formed highly organized structures, when dried from methanol solutions. Transmission and scanning electron microscopic images of the peptides Boc‐Ile‐Ile‐OMe, Boc‐Phe‐Phe‐Phe‐Ile‐Ile‐OMe and Boc‐Trp‐Ile‐Ile‐OMe showed nanotubular structures. Removal of the Boc group resulted in disruption of the ability to form tubular structures though spherical aggregates were formed. Both Boc‐Leu‐Ile‐Ile‐OMe and H‐Leu‐Ile‐Ile‐OMe formed only spherical nanostructures. Dynamic light scattering studies showed that aggregates of varying dimensions were present in solution suggesting that self‐assembly into ordered structures is facilitated by aggregation in solution. Fourier transform infrared spectroscopy and circular dichroism spectroscopy data show that although all four of the protected peptides adopt well‐defined tertiary structures, upon removal of the Boc group, only H‐Phe‐Phe‐Phe‐Ile‐Ile‐OMe had the ability to adopt β‐structure. Our results indicate that hydrophobic interaction is a very important determinant for self‐assembly and presence of charged and aromatic amino acids in a peptide is not necessary for self‐assembly. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
The synthesis of difficult peptide sequences has been a challenge since the very beginning of SPPS. The self‐assembly of the growing peptide chains has been proposed as one of the causes of this synthetic problem. However, there is an increasing need to obtain peptides and proteins that are prone to aggregate. These peptides and proteins are generally associated with diseases known as amyloidoses. We present an efficient SPPS of two homologous peptide fragments of HuPrP (106–126) and MoPrP105–125 based on the use of the PEGA resin combined with proper coupling approaches. These peptide fragments were also studied by CD and TEM to determine their ability to aggregate. On the basis of these results, we support PEG‐based resins as an efficient synthetic tool to prepare peptide sequences prone to aggregate on‐resin. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  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