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.
To extend our studies on peptides and proteins with amphiphilic secondary structures, a series of peptides designed to form amphiphilic beta-strand structures was designed, synthesized, and characterized by circular dichroism and infrared spectroscopy. Amphiphilic beta-strand conformations may be likely to appear in a variety of surface-active proteins, including apolipoprotein B and fibronectin. In a beta-strand conformation, the synthetic peptides will possess a hydrophobic face composed of valine side chains and a hydrophilic face composed of alternating acidic (glutamic acid) and basic (ornithine or lysine) residues. The peptides studied had a variety of chain lengths (5, 9, and 13 residues), and had the amino groups either free or protected with the trifluoroacetyl group. While the peptides did not possess a high potential for beta-sheet formation based on the Chou Fasman parameters, they possessed significant beta-sheet content, with up to 90% beta-sheet calculated for the 13-residue protected peptide. The driving force for beta-sheet formation is the potential amphiphilicity of this conformation. The beta-strand conformation of the 13-residue deprotected peptide was stable in 50% trifluoroethanol, 6 M guanidine hydrochloride, and octanol. The peptides are strongly self-associating in water, which would reduce the unfavorable contacts of the hydrophobic residues with water. It is clear that small peptides can be designed to form stable beta-strand conformations.  相似文献   

2.
Synthetic peptides of different size, reproducing the proteolytic processing site of proocytocin, were studied by CD under several experimental conditions in order to ascertain the ability of different solvents to stabilize secondary structural motifs, such as α-helix tracts and β-turns. A combination of deconvolution methods and empirical calculations subtracting the contributions due to unordered structures from the spectra suggests that in solution (a) mainly two distinct families of ordered conformers containing structurally different β-turns are present, (b) the relative stability of the different conformers depends from the nature of the solvent, and (c) in the case of the larger peptides, a population containing an α-helical conformation is also present. From the biological point of view the presence of at least two families of ordered conformers could be in line with current theories assuming that the catalytic effect of the receptor microenvironment may be determinant in shifting the equilibrium toward the active conformation. © 1997 John Wiley & Sons, Inc. Biopoly 41 : 461–479, 1997  相似文献   

3.
The conformation of synthetic HA317-329-NH2representing the major B- and T-cell epitopic region of influenza virus hemagglutinin, its palmitoylated derivative (HA317-329-Thr(Pal)-NH2), and the intersubunit peptide (HA317-341-NH2) comprising also the fusion peptide, were studied in aqueous buffer and in the presence of neutral and negatively charged liposomes. The free peptide is unordered in aqueous solution, even in the presence of liposomes. However, grafting the palmitic acid or the fusion peptide onto the C-terminus of the peptide enables the hydrophilic HA317-329to adopt folded (turn) and β-strand structure on the surface of neutral and negatively charged liposomes, respectively. The results emphasize the importance of some kind of anchor for achieving a specific conformation of epitopic peptide HA317-329-NH2on the surface of liposomes.  相似文献   

4.
Circular dichroism spectra of proteins are sensitive to protein secondary structure. The CD spectra of alpha-rich proteins are similar to those of model alpha-helices, but beta-rich proteins exhibit CD spectra that are reminiscent of CD spectra of either model beta-sheets or unordered polypeptides. The existence of these two types of CD spectra for beta-rich proteins form the basis for their classification as betaI- and betaII-proteins. Although the conformation of beta-sheets is largely responsible for the CD spectra of betaI-proteins, the source of betaII-protein CD, which resembles that of unordered polypeptides, is not completely understood. The CD spectra of unordered polypeptides are similar to that of the poly(Pro)II helix, and the poly(Pro)II-type (P2) structure forms a significant fraction of the unordered conformation in globular proteins. We have compared the beta-sheet and P2 structure contents in beta-rich proteins to understand the origin of betaII-protein CD. We find that betaII-proteins have a ratio of P2 to beta-sheet content greater than 0.4, whereas for betaI-proteins this ratio is less than 0.4. The beta-sheet content in betaI-proteins is generally higher than that in betaII-proteins. The origin of two classes of CD spectra for beta-rich proteins appears to lie in their relative beta-sheet and P2 structure contents.  相似文献   

5.
We have recently designed a host-guest peptide system that allows us to quantitatively measure the energetics of interaction of viral fusion peptides with lipid bilayers. Here, we show that fusion peptides of influenza hemagglutinin reversibly associate with one another at membrane surfaces above critical surface concentrations, which range from one to five peptides per 1000 lipids in the systems that we investigated. It is further demonstrated by using circular dichroism and Fourier transform infrared spectroscopy that monomeric peptides insert into the bilayers in a predominantly alpha-helical conformation, whereas self-associated fusion peptides adopt predominantly antiparallel beta-sheet structures at the membrane surface. The two forms are readily interconvertible and the equilibrium between them is determined by the pH and ionic strength of the surrounding solution. Lowering the pH favors the monomeric alpha-helical conformation, whereas increasing the ionic strength shifts the equilibrium towards the membrane-associated beta-aggregates. The binding data are interpreted in terms of a cooperative binding model that yields free energies of insertion and free energies of self-association for each of the peptides studied at pH 7.4 and pH 5. At pH 5 and 35 mM ionic strength, the insertion energy of the 20 residue influenza hemagglutinin fusion peptide is -7.2 kcal/mol and the self-association energy is -1.9 kcal/mol. We propose that self-association of fusion peptides could be a major driving force for recruiting a small number of hemagglutinin trimers into a fusion site.  相似文献   

6.
Amyloid beta (Abeta) peptides are one of the classes of amphiphilic molecules that on dissolution in aqueous solvents undergo interesting conformational transitions. These conformational changes are known to be associated with their neuronal toxicity. The mechanism of structural transition involved in the monomeric Abeta to toxic assemblage is yet to be understood at the molecular level. Early results indicate that oriented molecular crowding has a profound effect on their assemblage formation. In this work, we have studied how different microenvironments affect the conformational transitions of one of the active amyloid beta-peptide fragments (Abeta(25-35)). Spectroscopic techniques such as CD and Fourier transform infrared spectroscopy were used. It was observed that a stored peptide concentrates on dissolution in methanol adopts a minor alpha-helical conformation along with unordered structures. On changing the methanol concentration in the solvated film form, the conformation switches to the antiparallel beta-sheet structure on the hydrophilic surface, whereas the peptide shows transition from a mixture of helix and unordered structure into predominantly a beta-sheet with minor contribution of helix structure on the hydrophobic surface. Our present investigations indicate that the conformations induced by the different surfaces dictate the gross conformational preference of the peptide concentrate.  相似文献   

7.
Pediocin AcH is a 44-residue antimicrobial peptide with bactericidal potency against Gram-positive bacteria such as Listeria. It belongs to a family of bacteriocins that, when membrane-associated, is predicted to contain beta-sheet and alpha-helical regions. All bacteriocins in this family have a conserved N-terminal disulfide bond. An additional C-terminal disulfide bond in pediocin AcH is thought to confer enhanced potency and broader specificity range against sensitive bacteria. The C-terminal disulfide bond may also affect the conformation of the C-terminus. The secondary structures of pediocin AcH in aqueous solution and vesicles from susceptible cells, as well as the ability of trifluoroethanol (TFE) and detergent systems to induce secondary structures like those induced in vesicles, were studied by circular dichroism (CD) spectroscopy. Like related peptides, pediocin AcH was highly unordered in aqueous solution, 56%. However, it also contained 20% beta-strand and 15% beta-turn structures. Upon complete binding to vesicles, 32% alpha-helical structure formed, the unordered structure decreased to 32%, and the beta-strand and beta-turn structures remained largely unchanged. Thus, a betaalpha domain structure formed in vesicles. The helical structure likely forces the C-terminal tail to loop back on the helix so that the C24-C44 disulfide bond can form. Detergent micelles were superior to TFE in their ability to induce secondary structural fractions in pediocin AcH comparable to those observed in vesicles. This demonstrates the importance of a hydrocarbon-water interface to pediocin AcH structure induction and suggests that it is preferable to use detergent micelles as solvents in NMR studies of pediocin AcH structure.  相似文献   

8.
Changes in the conformation of apoliprotein B-100 in the early stages of copper-mediated low density lipoprotein oxidation have been monitored by infrared spectroscopy. During the lag phase no variation in structure is observed, indicating that copper binding to the protein does not significantly affect its structure. In the propagation phase, while hydroperoxides are formed but the protein is not modified, no changes in secondary structure are observed, but the thermal profile of the band corresponding to alpha-helix is displaced in frequency, indicating changes in tertiary structure associated with this conformation but not with beta-sheet components. When aldehyde formation starts, a decrease of approximately 3% in the area of bands corresponding to alpha-helix and beta-sheet is produced, concomitantly with an increase in beta-turns and unordered structure. The two bands corresponding to beta-turns vary as well under these conditions, indicating changes in these structures. Also at this stage the thermal profile shows variations in frequency for the bands corresponding to both alpha-helix and beta-sheet.The results are consistent with the hypothesis that as soon as the polyunsaturated fatty acids from the particle core are modified, this change is reflected at the surface, in the alpha-helical components contacting the monolayer.  相似文献   

9.
A Fourier transform infrared spectrometer has been interfaced with a surface balance and a new external reflection infrared sampling accessory, which permits the acquisition of spectra from protein monolayers in situ at the air/water interface. The accessory, a sample shuttle that permits the collection of spectra in alternating fashion from sample and background troughs, reduces interference from water vapor rotation-vibration bands in the amide I and amide II regions of protein spectra (1520-1690 cm-1) by nearly an order of magnitude. Residual interference from water vapor absorbance ranges from 50 to 200 microabsorbance units. The performance of the device is demonstrated through spectra of synthetic peptides designed to adopt alpha-helical, antiparallel beta-sheet, mixed beta-sheet/beta-turn, and unordered conformations at the air/water interface. The extent of exchange on the surface can be monitored from the relative intensities of the amide II and amide I modes. Hydrogen-deuterium exchange may lower the amide I frequency by as much as 11-12 cm-1 for helical secondary structures. This shifts the vibrational mode into a region normally associated with unordered structures and leads to uncertainties in the application of algorithms commonly used for determination of secondary structure from amide I contours of proteins in D2O solution.  相似文献   

10.
Amelogenin, the major extracellular enamel matrix protein, plays critical roles in controlling enamel mineralization. This generally hydrophobic protein self-assembles to form nanosphere structures under certain solution conditions. To gain clearer insight into the mechanisms of amelogenin self-assembly, we first investigated the occurrences of secondary structures within its sequence. By applying isothermal titration calorimetry (ITC), we determined the thermodynamic parameters associated with protein-protein interactions and with conformational changes during self-assembly. The recombinant porcine full length (rP172) and a truncated amelogenin lacking the hydrophilic C-terminal (rP148) were used. Circular dichroism (CD) measurements performed at low concentrations (<5 microM) revealed the presence of the polyproline-type II (PPII) conformation in both amelogenins in addition to alpha-helix and unordered conformations. Structural transition from PPII/unordered to beta-sheet was observed for both proteins at higher concentrations (>62.5 microM) and upon self-assembly. ITC measurements indicated that the self-assembly of rP172 and rP148 is entropically driven (+DeltaS(A)) and energetically favorable (-DeltaG(A)). The magnitude of enthalpy (DeltaH(A)) and entropy changes of assembly (DeltaS(A)) were smaller for rP148 than rP172, whereas the Gibbs free energy change of assembly (DeltaG(A)) was not significantly different. It was found that rP172 had higher PPII content than rP148, and the monomer-multimer equilibrium for rP172 was observed in a narrower protein concentration range when compared to rP148. The large positive enthalpy and entropy changes in both cases are attributed to the release of ordered water molecules and the associated entropy gain (due to the hydrophobic effect). These findings suggest that PPII conformation plays an important role in amelogenin self-assembly and that rP172 assembly is more favorable than rP148. The data are direct evidence for the notion that hydrophobic interactions are the main driving force for amelogenin self-assembly.  相似文献   

11.
The A4 or beta-peptide (39 to 43 amino acid residues) is the principal proteinaceous component of amyloid deposits in Alzheimer's disease. Using circular dichroism (c.d.), we have studied the secondary structures and aggregational properties in solution of 4 synthetic amyloid beta-peptides: beta-(1-28), beta-(1-39), beta-(1-42) and beta-(29-42). The natural components of cerebrovascular deposits and extracellular amyloid plaques are beta-(1-39) and beta-(1-42), while beta-(1-28) and beta-(29-42) are unnatural fragments. The beta-(1-28), beta-(1-39) and beta-(1-42) peptides adopt mixtures of beta-sheet, alpha-helix and random coil structures, with the relative proportions of each secondary structure being strongly dependent upon the solution conditions. In aqueous solution, beta-sheet structure is favored for the beta-(1-39) and beta-(1-42) peptides, while in aqueous solution containing trifluoroethanol (TFE) or hexafluoroisopropanol (HFIP), alpha-helical structure is favored for all 3 peptides. The alpha-helical structure unfolds with increasing temperature and is favored at pH 1 to 4 and pH 7 to 10; the beta-sheet conformation is temperature insensitive and is favored at pH 4 to 7. Peptide concentration studies showed that the beta-sheet conformation is oligomeric (intermolecular), whereas the alpha-helical conformation is monomeric (intramolecular). The rate of aggregation to the oligomeric beta-sheet structure (alpha-helix----random coil----beta-sheet) is also dependent upon the solution conditions such as the pH and peptide concentration; maximum beta-sheet formation occurs at pH 5.4. These results suggest that beta-peptide is not an intrinsically insoluble peptide. Thus, solution abnormalities, together with localized high peptide concentrations, which may occur in Alzheimer's disease, may contribute to the formation of amyloid plaques. The hydrophobic beta-(29-42) peptide adopts exclusively an intermolecular beta-sheet conformation in aqueous solution despite changes in temperature or pH. Therefore, this segment may be the first region of the beta-peptide to aggregate and may direct the folding of the complete beta-peptide to produce the beta-pleated sheet structure found in amyloid deposits. Differences between the solution conformations of the beta-(1-39) and beta-(1-42) peptides suggests that the last 3 C-terminal amino acids are crucial to amyloid deposition.  相似文献   

12.
Peptides of 12, 16 and 24 amino acids length corresponding to the NH2 terminal sequence of SIV gp32 were synthesized. Fluorescence energy transfer studies have shown that those peptides can induce lipid mixing of SUV (Small Unilamellar Vesicles) of various compositions at pH 7.4 and 37 degrees C. LUV (Large Unilamellar Vesicles) were shown to undergo fusion, provided they contained PE in their lipid composition. This work is an attempt to determine how the fusogenic activity depends on the structure of the peptide inserted into a lipidic environment. The peptides secondary structure and orientation in the lipid bilayer were determined using Fourier Transform infrared spectroscopy (FTIR). They adopt mainly a beta-sheet conformation in the absence of lipids. After interaction with DOPC SUV, the beta-sheet is partly converted into alpha-helix oriented obliquely with respect to the membrane interface. We bring here evidence that this oblique orientation is a prerequisite to the fusion process.  相似文献   

13.
The study is focused on the secondary structure of the external N-terminal segment of human erythrocyte glycophorin A (NN) which was determined by applying methods of CHOU et FASMAN and LIM. This hydrophilic glycophorin segment is assumed to consist of 48.5% ordered (alpha-helix, beta-sheet, beta-turn) and 51.5% unordered sequences. From the secondary structure suggestions are made concerning (i) peptide interaction and (ii) binding to the lipid bilayer of the N-terminal segment.  相似文献   

14.
Ganesh S  Jayakumar R 《Biopolymers》2003,70(3):336-345
Self-assembly of two tripeptide derivatives containing three nonpolar isoleucine moieties and polar oxyethylene groups are studied in methanol. Peptide A [CH3(OCH2CH2)3OCH2CO(Ile)3OCH3] and peptide B [CH3(OCH2CH2)3OCH2CO(Ile)3NH (CH2CH2O)3CH3] take a mixture of unordered and helical conformation at low concentration (8.5 x 10(-4) M). However, at high concentration (2 x 10(-3) M), both the peptide showed significant increase in the helical conformation. An interesting conformational transition of peptides A and B at various methanol contents was observed in the solvated films of these compounds by spectroscopic methods like the far-uv circular dichroism and Fourier transform infrared (FT-IR) techniques. Peptide B, which contains more polar oxyethylene groups than A, showed a highly cooperative conformational transition when the methanol content was decreased. This transition was characterized by a large increase of beta-sheet, retaining a alpha-helical contribution. Peptide A showed a conformational transition resulting in a beta-sheet in the aggregated state. From the CD spectra, the ratio in the ellipticity indicates that peptide B forms twisted antiparallel beta-sheet conformation, whereas peptide A takes a parallel beta-sheet conformation. The results obtained in this work indicates the role of polar derivatization on the conformational preference of peptides having similar sequence.  相似文献   

15.
Liu D  Xu Y  Feng Y  Liu H  Shen X  Chen K  Ma J  Jiang H 《Biochemistry》2006,45(36):10963-10972
Abeta peptides cleaved from the amyloid precursor protein are the main components of senile plaques in Alzheimer's disease. Abeta peptides adopt a conformation mixture of random coil, beta-sheet, and alpha-helix in solution, which makes it difficult to design inhibitors based on the 3D structures of Abeta peptides. By targeting the C-terminal beta-sheet region of an Abeta intermediate structure extracted from molecular dynamics simulations of Abeta conformational transition, a new inhibitor that abolishes Abeta fibrillation was discovered using virtual screening in conjunction with thioflavin T fluorescence assay and atomic force microscopy determination. Circular dichroism spectroscopy demonstrated that the binding of the inhibitor increased the beta-sheet content of Abeta peptides either by stabilizing the C-terminal beta-sheet conformation or by inducing the intermolecular beta-sheet formation. It was proposed that the inhibitor prevented fibrillation by blocking interstrand hydrogen bond formation of the pleated beta-sheet structure commonly found in amyloid fibrils. The study not only provided a strategy for inhibitor design based on the flexible structures of amyloid peptides but also revealed some clues to understanding the molecular events involved in Abeta aggregation.  相似文献   

16.
Dennison SM  Greenfield N  Lenard J  Lentz BR 《Biochemistry》2002,41(50):14925-14934
Helical instability induced by gly residues in the transmembrane domain (TMD) of G protein, the fusion protein of vesicular stomatitis virus (VSV), was speculated to aid in the later steps of the fusion process, because G protein with ala's substituted for the two TMD gly's was inactive (Cleverley, D. Z., and Lenard, J. (1998) Proc. Natl. Acad. Sci. U. S. A. 95, 3425-30). Here we examine the conformations of synthetic peptides corresponding to fusion-active (GGpep) and inactive (AApep; G's replaced by A's) TMDs by CD spectroscopy, and then their effects on the kinetics of poly (ethyleneglycol) (PEG)-mediated fusion of small unilamellar vesicles. GGpep and AApep both assumed history-dependent, non-interconvertible ordered structures. Both peptides were largely helical under all conditions if derived from trifluoroethanol solutions, and aggregated in a beta-sheet form if derived from acetonitrile solutions. In solvent, detergents or lipid bilayers, GGpep showed a greater range of secondary structural features than did AApep. The two peptides had large but different effects on PEG-mediated fusion. Both enhanced the rate but not the extent of lipid mixing. AApep significantly inhibited the extent of fusion pore formation while GGpep had no effect. The initial rate of fusion was enhanced 6-fold by GGpep and less than 2-fold by AApep. Addition of 5 mol % hexadecane overrode all peptide-induced effects. We suggest that both GGpep and hexadecane promote pore formation by stabilizing the nonlamellar structures in fusion intermediates or initial small pores. AApep, which had fewer nonhelical features in its CD spectrum than GGpep, actually inhibited fusion pore formation.  相似文献   

17.
The amyloid beta-protein (Abeta) constitutes the major peptide component of the amyloid plaque deposits of Alzheimer's disease in humans. The Abeta changes from a nonpathogenic to a pathogenic conformation resulting in self-aggregation and deposition of the peptide. It has been established that denaturing factors (such as the interaction with membranes) are involved in the structural transition. This work is aimed at determining the effect of hydrophobic Teflon on the conformation of the Abeta (1-40). Prior to adsorption, the secondary structure and self-aggregation state of the Abeta in solution were established as a function of pH. Three different species coexist: unordered monomers/dimers, small oligomers in mainly a regular beta-sheet structure, and bigger aggregates having a twisted beta-sheet conformation. Transferring the Abeta from the solution to the Teflon surface strongly promotes alpha-helix formation. Furthermore, increasing the degree of coverage of the Teflon by the Alphabeta protein leads to a conformational change toward a more enriched beta-sheet structure.  相似文献   

18.
The fusion peptides of HIV and influenza virus are crucial for viral entry into a host cell. We report the membrane-perturbing and structural properties of fusion peptides from the HA fusion protein of influenza virus and the gp41 fusion protein of HIV. Our goals were to determine: 1), how fusion peptides alter structure within the bilayers of fusogenic and nonfusogenic lipid vesicles and 2), how fusion peptide structure is related to the ability to promote fusion. Fluorescent probes revealed that neither peptide had a significant effect on bilayer packing at the water-membrane interface, but both increased acyl chain order in both fusogenic and nonfusogenic vesicles. Both also reduced free volume within the bilayer as indicated by partitioning of a lipophilic fluorophore into membranes. These membrane ordering effects were smaller for the gp41 peptide than for the HA peptide at low peptide/lipid ratio, suggesting that the two peptides assume different structures on membranes. The influenza peptide was predominantly helical, and the gp41 peptide was predominantly antiparallel beta-sheet when membrane bound, however, the depths of penetration of Trps of both peptides into neutral membranes were similar and independent of membrane composition. We previously demonstrated: 1), the abilities of both peptides to promote fusion but not initial intermediate formation during PEG-mediated fusion and 2), the ability of hexadecane to compete with this effect of the fusion peptides. Taken together, our current and past results suggest a hypothesis for a common mechanism by which these two viral fusion peptides promote fusion.  相似文献   

19.
The precipitation by antibodies to intact myelin basic protein (BP) and to synthetic peptides containing a sequence based on the region 65 to 83 of bovine BP, S82, S81, S79, and S24, of intact BP in solution or bound to lipid vesicles was compared, using 125I-BP or 14C-DPPC-labeled lipid-BP vesicles. The antipeptide antibodies were shown earlier to recognize conformational determinants which are not expressed in the intact protein in solution. Several anti-BP antibodies precipitated more of the BP free in solution than when bound to lipid vesicles, suggesting that some of the determinants recognized by these antibodies were either sequestered in the bilayer or were altered in conformation. In contrast, one anti-peptide antisera, which had a high titer for the conformational determinant in two of these peptides, S82 and S81, precipitated the protein to a significant degree when it was bound to PG vesicles, even though it did not react with the intact protein in solution. These results indicated that PG was able to confer on the protein the unique peptide conformation recognized by this antibody. PS was less effective, and other lipids were ineffective at conferring this conformation on the protein, supporting earlier results which showed that the conformation of the protein is influenced by the lipid composition of its environment. None of the other anti-peptide antibodies studied bound to the protein either in solution or in lipid vesicles. These results indicate that the lipid environment can sequester or alter the conformation of some antigenic determinants, preventing recognition by some anti-BP antibodies, and can expose or generate other conformational determinants, allowing recognition by an anti-peptide antiserum.  相似文献   

20.
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