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.
Melittin exhibits a transition from random-coil monomer to helical tetramer as a function of peptide concentration [ J. Bello, H. R. Bello, and E. Granados (1982) Biochemistry, Vol. 21, pp. 461–465]. When permethylated on each of the four amino groups (Gly-1 Nα and Lys-7, 21, and 23 Nε) to yield trimethylammonium groups, melittin exists as a random coil and does not show any concentration-dependent conformational transition (up to 290 μM). Acylation of the amino groups of melittin with glycine or 5-aminopentanoic acid followed by permethylation increases helix formation, but to a lesser extent than for the unmethylated aminoacylmelittin derivatives. The results are discussed in relation to hydrophobicity, charge repulsions, and ion binding. Melittin, and more weakly, permethylated melittin (MLT-Me) form helical hybrids with an anionic random-coil melittin analogue (E-MLT), in which all the lysine and arginine residues of melittin were replaced by glutamate residues. The hybrid between MLT-Me and E-MLT shows a concentration-dependent increase in helicity. E-MLT, when succinylated at the N-terminal glycine (E-MLT-suc), forms a stronger hybrid with MLT-Me, possibly as a result of increased electrostatic interaction between equal but opposite charges in E-MLT-suc (net charge –6) and MLT-Me (net charge +6). The hybrids exhibit both cold- and heat-induced denaturation, similar to the phenomenon exhibited by proteins. The hybrids also exhibit significant residual structures in the temperature range of 80–100°C, which may be similar to the molten globular states that have been suggested for proteins. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
Whereas melittin at micromolar concentrations is unfolded under conditions of low salt at neutral pH, it transforms to a tetrameric alpha-helical structure under several conditions, such as high peptide concentration, high anion concentration, or alkaline pH. The anion- and pH-dependent stabilization of the tetrameric structure is similar to that of the molten globule state of several acid-denatured proteins, suggesting that tetrameric melittin might be a state similar to the molten globule state. To test this possibility, we studied the thermal unfolding of tetrameric melittin using far-UV CD and differential scanning calorimetry. The latter technique revealed a broad but distinct heat absorption peak. The heat absorption curves were consistent with the unfolding transition observed by CD and were explainable by a 2-state transition mechanism between the tetrameric alpha-helical state and the monomeric unfolded state. From the peptide or salt-concentration dependence of unfolding, the heat capacity change upon unfolding was estimated to be 5 kJ (mol of tetramer)-1 K-1 at 30 degrees C and decreased with increasing temperature. The observed change in heat capacity was much smaller than that predicted from the crystallographic structure (9.2 kJ (mol of tetramer)-1 K-1), suggesting that the hydrophobic residues of tetrameric melittin in solution are exposed in comparison with the crystallographic structure. However, the results also indicate that the structure is more ordered than that of a typical molten globule state. We consider that the conformation is intermediate between the molten globule state and the native state of globular proteins.  相似文献   

3.
The intracellular C-terminal domain (CTD) of KcsA, a bacterial homotetrameric potassium channel, is a 40-residue-long segment that natively adopts a helical bundle conformation with 4-fold symmetry. A hallmark of KcsA behavior is pH-induced conformational change, which leads to the opening of the channel at acidic pH. Previous studies have reached conflicting conclusions as to the role of the CTD in this transition. Here, we investigate the involvement of this domain in pH-mediated channel opening by NMR using a soluble peptide corresponding to residues 128-160 of the CTD (CTD34). At neutral pH, CTD34 exhibits concentration-dependent spectral changes consistent with oligomer formation. We prove this slowly tumbling species to be a tetramer with a dissociation constant of (2.0±0.5)×10(-)(11)?M(3) by NMR and sedimentation equilibrium experiments. Whereas monomeric CTD34 is only mildly helical, secondary chemical shifts prove that the tetrameric species adopts a tight native-like helical bundle conformation. The tetrameric species undergoes pH-dependent dissociation, and CTD34 is fully monomeric below pH?5.0. The structural basis for this phenomenon is the destabilization of the tetrameric CTD34 by protonation of residue H145 in the monomeric form of the peptide. We conclude that (i) the CTD34 peptide is independently capable of forming a tetrameric helical bundle, and (ii) this structurally significant conformational shift is modulated by the effects of solution pH on residue H145. Therefore, the involvement of this domain in the pH gating of the channel is strongly suggested.  相似文献   

4.
Guanidine x HCl (GdnHCl)-induced unfolding of tetrameric N(5)-(L-1-carboxyethyl)-L-ornithine synthase (CEOS; 141,300 M(r)) from Lactococcus lactis at pH 7.2 and 25 degrees C occurred in several phases. The enzyme was inactivated at approximately 1 M GdnHCl. A time-, temperature-, and concentration-dependent formation of soluble protein aggregates occurred at 0.5-1.5 M GdnHCl due to an increased exposure of apolar surfaces. A transition from tetramer to unfolded monomer was observed between 2 and 3.5 M GdnHCl (without observable dimer or trimer intermediates), as evidenced by tyrosyl and tryptophanyl fluorescence changes, sulfhydryl group exposure, loss of secondary structure, size-exclusion chromatography, and sedimentation equilibrium data. GdnHCl-induced dissociation and unfolding of tetrameric CEOS was concerted, and yields of reactivated CEOS by dilution from 5 M GdnHCl were improved when unfolding took place on ice rather than at 25 degrees C. Refolding and reconstitution of the enzyme were optimal at 相似文献   

5.
SecB is a cytosolic tetrameric chaperone in Escherichia coli, which maintains polypeptides, destined for export in a translocation competent state. The thermodynamics of unfolding of SecB was studied as a function of protein concentration, by using high sensitivity-differential scanning calorimetry and spectroscopic methods. The thermal unfolding of tetrameric SecB is reversible and can be well described as a two-state transition in which the folded tetramer is converted directly to unfolded monomers. Increasing the pH decreases the stability of the tetramer significantly, the T(m) changing from 341.3 K at pH 6.5 to 332.6 K at pH 9.5. The value of DeltaC(p) obtained from measurements of DeltaH(m) as a function of T(m) was 10.7 +/- 0.7 kcal mol(-1) K(-1). The value of DeltaC(p) is among the highest measured for a multimeric protein. At 298 K, pH 7.4, the DeltaG degrees (u) for the SecB tetramer is 27.9 +/- 2 kcal mol(-1). Denaturant-mediated unfolding of SecB was found to be irreversible. The reactivity of the four solvent-exposed free thiols in tetrameric SecB is salt dependent. The kinetics of reactivity suggests that these four cysteines are in close proximity to each other and that these residues on each monomer are in chemically identical environments. The thermodynamic data suggest that SecB is a stable, well-folded, and tightly packed tetramer and that substrate binding occurs at a surface site rather than at an interior cavity.  相似文献   

6.
Mixed micelles of the 26-residue, lytic peptide melittin (MLT) and 1-myristoyl-2-hydroxyl-sn-glycero-3-phosphocholine (MMPC) in aqueous solution at 25 degrees C were investigated by (13)C- and (31)P-NMR spectroscopy. (13)C alpha chemical shifts of isotopically labeled synthetic MLT revealed that MLT in the micelle is predominantly alpha-helical and that the peptide secondary structure is stable from pH 4 to pH 11. Although the helical transformation of MLT as determined from NMR is evident at lipid:peptide molar ratios as low as 1:2, tryptophan fluorescence measurements demonstrate that well-defined micellar complexes do not predominate until lipid:peptide ratios exceed 30:1. (31)P linewidth measurements indicate that the interaction between phosphate ions in solution and cationic groups on MLT is pH dependent, and that the phosphoryl group of MMPC senses a constant charge, most likely +2, on MLT from pH 4 to pH 10. (13)C-NMR relaxation data, analyzed using the model-free formalism, show that the peptide backbone of MLT is partially, but not completely, immobilized in the mixed micelles. Specifically, order parameters (S(2)) of C alpha-H vectors averaged 0.7 and were somewhat larger for residues in the N-terminal half of the molecule. The amino terminal glycine had essentially the same range of motion as the backbone carbons. Likewise, order parameters for the trp side chain were similar to those found for the peptide C alpha moieties, as was verified by trp fluorescence anisotropy decay data. In contrast, the motion of the lysine side chains was less restricted, the average S(2) values for the C epsilon-H vectors being 0.19, 0.30, and 0.44 for lys-7, 21, and 23, respectively, for MLT in the mixed micelles. Values of the effective correlation time of the local motion tau e were in the motional narrowing limit and usually longer for side-chain atoms than for those in the backbone. The dynamics were independent of pH from pH 4 to pH 9, but at pH 11 the correlation time for the rotational motion of the mixed micelles as a whole increased from 10 ns to 16 ns, and S(2) for the lys side chains increased. Overall it appears that the MLT helix lies near the surface of the micelle at low to neutral pH, but at higher pH its orientation changes, accompanied by deeper penetration of the lysine side chains into the micelle interior. It is apparent, however, that the MLT-lipid interaction is not dependent on deprotonation of any of the titratable cationic groups in the peptide in the pH 4-10 range, and that there is substantial backbone and side-chain mobility in micelle-bound MLT.  相似文献   

7.
Y Hagihara  M Kataoka  S Aimoto  Y Goto 《Biochemistry》1992,31(47):11908-11914
Electrostatic repulsion between positively charged groups has been suggested to be critical in determining the conformation of melittin. To clarify the role of repulsive forces, we prepared a series of succinylated melittins, an acetylated melittin, and a synthetic melittin mutant, with various degrees of charge repulsion. The conformation of the melittin derivatives was examined by far-UV circular dichroism under various conditions of pH and salt at 20 degrees C. The stability of the tetrameric helical state was found to be dependent on the net charge of the peptides. The charge repulsive forces destabilized the helical state of intact melittin by 600 cal/(charge.mol of tetramer). This value was close to the corresponding one (450 cal/(charge.mol)) obtained for the acidic molten globule of horse cytochrome c [Goto, Y., & Nishikiori, S. (1991) J. Mol. Biol. 222, 679-686], which has a molecular weight and a net charge comparable to those of the tetrameric melittin. Small-angle X-ray scattering of the tetrameric melittin and the molten globule of cytochrome c showed that the two states are also comparable to each other in the radius of gyration. These results suggest that the contribution of electrostatic repulsion to the conformational stability of melittin is similar to that of the molten globule.  相似文献   

8.
Anionic phospholipids induce a marked conformational change in a synthetic peptide corresponding to residues 1-27 of pre-ornithine carbamyltransferase. The peptide designated, pO-(1-27)-peptide amide, becomes more alpha-helical in the presence of cardiolipin or dimyristoylphosphatidylglycerol but not in the presence of dimyristoylphosphatidylcholine. The greater helix-promoting action of anionic versus zwitterionic lipids is predicted by helix-coil transition theory. This statistical mechanical theory also predicts that a shorter peptide, N-acetyl-pO-(16-27)-peptide amide, has less helix-forming tendency, even in the presence of sodium dodecyl sulfate, despite the fact that it has a comparable number of positive charges. The N-acetyl-pO-(16-27)-peptide amide has no helical structure in buffer with or without dimyristoylphosphatidylglycerol but it has a small (5%) helical content in methanol. Thus, the ability of anionic lipids to promote helix formation requires more than the presence of cationic groups on the peptide. The angular dependence of the hydrophobic moment of the putative helical segment of pO-(1-27)-peptide amide demonstrates that any helical structure which is formed would have some amphiphilic character. The pO-(1-27)-peptide amide disrupts large lipid aggregates to form discoid micelles about 30 to 50 nm in diameter. The ability to lyse membranes into disc-shaped micelles is characteristic of peptides containing an amphiphathic helix. In the case of the mitochondrial signal peptide, this membrane-lytic behavior may contribute to the translocation of the protein into the organelle.  相似文献   

9.
D D Haas  B R Ware 《Biochemistry》1978,17(23):4946-4950
Diffusion studies by photon correlation of scattered laser light confirm the dissociation of the tetrameric form of human carboxyhemoglobin to dimers above pH 10 and provide new estimates of the subunit dissociation equilibrium constants in this pH range. Electrophoretic light-scattering experiments under the same conditions reveal that the electrophoretic mobilities of tetramers and dimers are indistinguishable to within instrumental resolution (ca. 7% in these experiments). The data imply an increase of the electrical charge on the dimer of at least 2.8 to 4.4 net negative charges upon dissociation. Mechanisms for the accumulation of negative charge by the dimer upon dissociation of the tetramer are proposed.  相似文献   

10.
Conformations of isolated fragments of pancreatic polypeptide   总被引:2,自引:0,他引:2  
K Tonan  Y Kawata  K Hamaguchi 《Biochemistry》1990,29(18):4424-4429
In spite of its short polypeptide chain, the pancreatic polypeptide molecule consists of a polyproline II type helix and alpha-helix. To understand the stability and formation of the alpha-helical region, we prepared some peptide fragments including the helical segment of chicken pancreatic polypeptide and studied their conformations by circular dichroism (CD). PP7-36 (a peptide fragment corresponding to residues 7-36 of chicken pancreatic polypeptide) showed a CD spectrum characteristic of the helix at pH 4.6 and at peptide concentrations as low as 1 microM. PP11-36 was able to form a helical conformation only at high peptide concentrations and not at concentrations lower than 10 microM. However, acetyl PP11-36 (in which the alpha-amino group is acetylated so that no positive charge exists at the N terminus) was able to form the helical conformation at pH 4.6 and at the peptide concentrations where PP11-36 could not. Succinyl PP12-36 (in which the alpha-amino group is succinylated to introduce a negative charge) was also able to form the helical conformation. The CD spectra of PP12-36 and PP13-36 were not characteristic of the helical conformation at all the pH values and peptide concentrations studied. Acetyl PP13-36, which has no charge at the N terminus, did not form the helix. On the other hand, succinyl PP13-36, which has a negative charge at the N-terminal end, did form the helix at pH 4.6. These findings indicate that the presence of the negative charge of carboxylate at the N-terminal region of a peptide fragment is important for helix formation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Zhu HL  Atkinson D 《Biochemistry》2007,46(6):1624-1634
Human apolipoprotein A-I (apoA-I) is the principle apolipoprotein of high-density lipoproteins that are critically involved in reverse cholesterol transport. The intrinsically flexibility of apoA-I has hindered studies of the structural and functional details of the protein. Our strategy is to study peptide models representing different regions of apoA-I. Our previous report on [1-44]apoA-I demonstrated that this N-terminal region is unstructured and folds into approximately 60% alpha-helix with a moderate lipid binding affinity. We now present details of the conformation and lipid interaction of a C-terminal 46-residue peptide, [198-243]apoA-I, encompassing putative helix repeats 10 and 9 and the second half of repeat 8 from the C-terminus of apoA-I. Far-ultraviolet circular dichroism spectra show that [198-243]apoA-I is also unfolded in aqueous solution. However, self-association induces approximately 50% alpha-helix in the peptide. The self-associated peptide exists mainly as a tetramer, as determined by native electrophoresis, cross-linking with glutaraldehyde, and unfolding data from circular dichroism (CD) and differential scanning calorimetry (DSC). In the presence of a number of lipid-mimicking detergents, above their CMC, approximately 60% alpha-helix was induced in the peptide. In contrast, SDS, an anionic lipid-mimicking detergent, induced helical folding in the peptide at a concentration of approximately 0.003% (approximately 100 microM), approximately 70-fold below its typical CMC (0.17-0.23% or 6-8 mM). Both monomeric and tetrameric peptide can solubilize dimyristoylphosphatidylcholine (DMPC) liposomes and fold into approximately 60% alpha-helix. Fractionation by density gradient ultracentrifugation and visualization by negative staining electromicroscopy demonstrated that the peptide binds to DMPC with a high affinity to form at least two sizes of relatively homogeneous discoidal HDL-like particles depending on the initial lipid:peptide ratio. The characteristics (lipid:peptide weight ratio, diameter, and density) of both complexes are similar to those of plasma A-I/DMPC complexes formed under similar conditions: small discoidal complexes (approximately 3:1 weight ratio, approximately 110 A, and approximately 1.10 g/cm3) formed at an initial 1:1 weight ratio and larger discoidal complexes (approximately 4.6:1 weight ratio, approximately 165 A, and approximately 1.085 g/cm3) formed at initial 4:1 weight ratio. The cross-linking data for the peptide on the complexes of two sizes is consistent with the calculated peptide numbers per particle. Compared to the approximately 100 A disk-like complex formed by the N-terminal peptide in which helical structure was insufficient to cover the disk edge by a single belt, the compositions of these two types of complexes formed by the C-terminal peptide are more consistent with a "double belt" model, similar to that proposed for full-length apoA-I. Thus, our data provide direct evidence that this C-terminal region of apoA-I is responsible for the self-association of apoA-I, and this C-terminal peptide model can mimic the interaction with the phospholipid of plasma apoA-I to form two sizes of homogeneous discoidal complexes and thus may be responsible for apoA-I function in the formation and maintenance of HDL subspecies in plasma.  相似文献   

12.
The urea and guanidine hydrochloride (GdnHCl)-induced denaturation of tetrameric concanavalin A (ConA) at pH 7.2 has been studied by using intrinsic fluorescence, 8-anilino-1-naphthalenesulfonate (ANS) binding, far-UV circular dichroism (CD), and size-exclusion chromatography. The equilibrium denaturation pathway of ConA, as monitored by steady state fluorescence, exhibits a three-state mechanism involving an intermediate state, which has been characterized as a structured monomer of the protein by ANS binding, far-UV CD and gel filtration size analysis. The three-state equilibrium is analyzed in terms of two distinct and separate dissociation (native tetramer<-->structured monomer) and unfolding (structured monomer<-->unfolded monomer) reaction steps, with the apparent transition midpoints (C(m)), respectively, at 1.4 and 4.5 M in urea, and at 0.8 and 2.4 M in GdnHCl. The results show that the free energy of stabilization of structured monomer relative to the unfolded state (-DeltaG(unf, aq)), is 4.4-5.5 kcal mol(-1), and that of native tetramer relative to structured monomer (-DeltaG(dis, aq)) is 7.2-7.4 kcal mol(-1), giving an overall free energy of stabilization (-DeltaG(dis&unf, aq)) of 11.6-12.9 kcal mol(-1) (monomer mass) for the native protein. However, the free energy preference at the level of quaternary tetrameric structure is found to be far greater than that at the tertiary monomeric level, which reveals that the structural stability of ConA is maintained mostly by subunit association.  相似文献   

13.
The effect of peptide hydrophobicity and charge on peptide interaction with model lipid bilayers was investigated for the C3a-derived peptide CNY21 by fluorescence spectroscopy, circular dichroism, ellipsometry, z-potential, and photon correlation spectroscopy measurements. For both zwitterionic and anionic liposomes, the membrane-disruptive potency for CNY21 variants increased with increasing net positive charge and mean hydrophobicity and was completely lost on elimination of all peptide positive charges. Analogous effects of elimination of the peptide positive net charge in particular were found regarding bacteria killing for both Pseudomonas aeruginosa and Bacillus subtilis. The peptides, characterized by moderate helix content both in buffer and when attached to the liposomes, displayed high adsorption for the net positively charged peptide variants, whereas adsorption was non-measurable for the uncharged peptide. That electrostatically driven adsorption represents the main driving force for membrane disruption in lipid systems was also demonstrated by a drastic reduction in both liposome leakage and peptide adsorption with increasing ionic strength, and this salt inactivation can be partly avoided by increasing the peptide hydrophobicity. This increased electrolyte resistance translates also to a higher antibacterial effect for the hydrophobically modified variant at high salt concentration. Overall, our findings demonstrate the importance of the peptide adsorption and resulting peptide interfacial density for membrane-disruptive effects of these peptides.  相似文献   

14.
J F Collawn  Y Paterson 《Biopolymers》1990,29(8-9):1289-1296
The conformations of two 17-residue peptide analogues derived from the C-terminal sequence of pigeon cytochrome c (native sequence = KAERADLIAYLKQATAK) were examined in aqueous and lipid environments by CD spectroscopy. The two analogues, KKLLKKLIAYLKQATAK (K peptide) and EELLEELIAYLKQATAK (E peptide), were made amphipathic with respect to helical segregation by substituting a 6-residue sequence at the N-terminus of the native peptide. Their structures were compared to the native peptide under aqueous conditions of varying pH and temperature, and in the presence of liposomes composed of phosphatidylcholine and phosphatidylserine in the ratio of 9:1. The results indicated that the native peptide remains unstructured under all the conditions examined even though this region of the native molecule is surface exposed and helical. The E peptide, however, was helical under aqueous conditions at 25 degrees C from pH 2-10 with a maximum helicity at pH 4 (54% helix from analysis of CD data). The ellipticity of the E peptide at pH 4 and 8 was concentration dependent, indicating an aggregation phenomenon. In studies in which the CD spectrum was measured at different temperatures, the E peptide became more helical at lower temperatures at pH 4 but not at pH 8. Upon interaction with a lipid membrane in the form of liposomes, there appeared to be a slight destabilization in the structure of the E peptide. The K peptide in an aqueous environment behaved like the native peptide in that it was structureless at all pHs and temperatures examined. In the presence of liposomes, however, this peptide had a high helical content (75% helix from analysis of CD data). These findings suggest that while stabilization of the helix dipole with negative charges at the N-terminus are important in inducing helical conformation in the E peptide, hydrophobic interactions created during aggregation appear to provide the principal stabilizing force. The results with the K peptide demonstrate that the positive N-terminal sequence of this peptide is able to interact with the negatively charged head groups in the phospholipid membrane in such a fashion as to stabilize a helical structure that is not apparent in an aqueous environment alone.  相似文献   

15.
Chitosan, a naturally occurring cationic polyelectrolyte, restores the adsorption of the clinical lung surfactant Survanta to the air-water interface in the presence of albumin at much lower concentrations than uncharged polymers such as polyethylene glycol. This is consistent with the positively charged chitosan forming ion pairs with negative charges on the albumin and lung surfactant particles, reducing the net charge in the double-layer, and decreasing the electrostatic energy barrier to adsorption to the air-water interface. However, chitosan, like other polyelectrolytes, cannot perfectly match the charge distribution on the surfactant, which leads to patches of positive and negative charge at net neutrality. Increasing the chitosan concentration further leads to a reduction in the rate of surfactant adsorption consistent with an over-compensation of the negative charge on the surfactant and albumin surfaces, which creates a new repulsive electrostatic potential between the now cationic surfaces. This charge neutralization followed by charge inversion explains the window of polyelectrolyte concentration that enhances surfactant adsorption; the same physical mechanism is observed in flocculation and re-stabilization of anionic colloids by chitosan and in alternate layer deposition of anionic and cationic polyelectrolytes on charged colloids.  相似文献   

16.
Salom D  Hill BR  Lear JD  DeGrado WF 《Biochemistry》2000,39(46):14160-14170
The M2 proton channel from the influenza A virus is a small protein with a single transmembrane helix that associates to form a tetramer in vivo. This protein forms proton-selective ion channels, which are the target of the drug amantadine. Here, we propose a mechanism for the pH-dependent association, and amantadine binding of M2, based on studies of a peptide representing the M2 transmembrane segment in dodecylphosphocholine micelles. Using analytical ultracentrifugation, we find that the sedimentation curves for the peptide depend on its concentration in the micellar phase. The data are well-described by a monomer-tetramer equilibrium, and the binding of amantadine shifts the monomer-tetramer equilibrium toward tetrameric species. Both tetramerization and the binding of amantadine lead to increases in the magnitude of the ellipticity at 223 nm in the circular dichroism spectrum of the peptide. The tetramerization and binding of amantadine are more favorable at elevated pH, with a pK(a) that is assigned to a His side chain, the only ionizable residue within the transmembrane helix. Our results, interpreted quantitatively in terms of a reversible monomer and tetramer protonation equilibrium model, suggest that amantadine competes with protons for binding to the deprotonated tetramer, thereby stabilizing the tetramer in a slightly altered conformation. This model accounts for the observed inhibition of proton flux by amantadine. Additionally, our measurements suggest that the M2 tetramer is substantially protonated at neutral pH and that both singly and doubly protonated states could be involved in M2's proton conduction at more acidic pHs.  相似文献   

17.
Yuko Okamoto 《Biopolymers》1994,34(4):529-539
Monte Carlo simulated annealing is applied to the tertiary structure prediction of a 17-residue synthetic peptide, which is known by experiment to exhibit high helical content at low pH. Two dielectric models are considered: sigmoidal distance-dependent dielectric function and a constant dielectric function (? = 2). Starting from completely random initial conformations, our simulations for both dielectric models at low pH gave many helical conformations. The obtained low-energy conformations are compared with the nuclear Overhauser effect spectroscopy cross-peak data for both main chain and side chains, and it is shown that the results for the sigmoidal dielectric function are in remarkable agreement with the experimental data. The results predict the existence of two disjoint helices around residues 5–9 and 11–16, while nmr experiments imply significant α-helix content between residues 5 and 14. Simulations with high pH, on the other hand, hardly gave a helical conformation, which is also in accord with the experiment. These findings indicate that when side chains are charged, electrostatic interactions due to these charges play a major role in the helix stability. Our results are compared with the previous 500 ps molecular dynamics simulations of the same peptide. It is argued that simulated annealing is superior to molecular dynamics in two respects: (1) direct folding of α-helix from completely random initial conformations is possible for the former, whereas only unfolding of an α-helix can be studied by the latter; (2) while both methods predict high helix content for low pH, the results for high pH agree with experiment (low helix content) only for the former method. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
Exendin-4 is a natural, 39-residue peptide first isolated from the salivary secretions of a Gila Monster (Heloderma suspectum) that has some pharmacological properties similar to glucagon-like-peptide-1 (GLP-1). This paper reports differences in the structural preferences of these two peptides. For GLP-1 in aqueous buffer (pH 3.5 or 5.9), the concentration dependence of circular dichroism spectra suggests that substantial helicity results only as a consequence of helix bundle formation. In contrast, exendin-4 is significantly helical in aqueous buffer even at the lowest concentration examined (2.3 microM). The pH dependence of the helical signal for exendin-4 indicates that helicity is enhanced by a more favorable sequence alignment of oppositely charged sidechains. Both peptides become more helical upon addition of either lipid micelles or fluoroalcohols. The stabilities of the helices were assessed from the thermal gradient of ellipticity (partial differential theta(221)/partial differential T values); on this basis, the exendin helix does not melt appreciably until temperatures significantly above ambient. The extent of helix formation for exendin-4 in aqueous buffer (and the thermal stability of the resulting helix) suggests the presence of a stable helix-capping interaction which was localized to the C-terminal segment by NMR studies of NH exchange protection. Solvent effects on the thermal stability of the helix indicate that the C-terminal capping interaction is hydrophobic in nature. The absence of this C-capping interaction and the presence of a flexible, helix-destabilizing glycine at residue 16 in GLP-1 are the likely causes of the greater fragility of the monomeric helical state of GLP-1. The intramolecular hydrophobic clustering in exendin-4 also appears to decrease the extent of helical aggregate formation.  相似文献   

19.
Amyloid fibril formation and deposition is a common feature of a wide range of fatal diseases including spongiform encephalopathies, Alzheimer's disease, and familial amyloidotic polyneuropathies (FAP), among many others. In certain forms of FAP, the amyloid fibrils are mostly constituted by variants of transthyretin (TTR), a homotetrameric plasma protein. Recently, we showed that transthyretin in solution may undergo dissociation to a non-native monomer, even under close to physiological conditions of temperature, pH, ionic strength, and protein concentration. We also showed that this non-native monomer is a compact structure, does not behave as a molten globule, and may lead to the formation of partially unfolded monomeric species and high molecular mass soluble aggregates (Quintas, A., Saraiva, M. J. M., and Brito, R. M. M. (1999) J. Biol. Chem. 274, 32943-32949). Here, based on aging experiments of tetrameric TTR and chemically induced protein unfolding experiments of the non-native monomeric forms, we show that tetramer dissociation and partial unfolding of the monomer precedes amyloid fibril formation. We also show that TTR variants with the least thermodynamically stable non-native monomer produce the largest amount of partially unfolded monomeric species and soluble aggregates under conditions that are close to physiological. Additionally, the soluble aggregates formed by the amyloidogenic TTR variants showed morphological and thioflavin-T fluorescence properties characteristic of amyloid. These results allowed us to conclude that amyloid fibril formation by some TTR variants might be triggered by tetramer dissociation to a compact non-native monomer with low conformational stability, which originates partially unfolded monomeric species with a high tendency for ordered aggregation into amyloid fibrils. Thus, partial unfolding and conformational fluctuations of molecular species with marginal thermodynamic stability may play a crucial role on amyloid formation in vivo.  相似文献   

20.
C S Wu  A Hachimori  J T Yang 《Biochemistry》1982,21(19):4556-4562
The conformation of several naturally occurring peptide hormones and bioactive oligopeptides in phospholipid solutions was studied by circular dichroism. Phosphatidylcholine induced a partial helix in human gastrin I at neutral pH, but phosphatidylserine did not unless the five consecutive glutamic acid residues in gastrin were protonated. Reduced somatostatin with two lysines and substance P with one arginine and one lysine were partially helical in phosphatidylserine, but not phosphatidylcholine, solution. Both lipids induced a helical conformation in glucagon and its COOH-terminal fragment (19-29) probably because the helical segment is primarily located at the uncharged COOH terminus. Thus, polypeptides with a helix-forming potential can have the helical conformation only when the peptides carry no charge or charges opposite to those on the polar head of the lipid. Renin substrate, which has potentials for the beta form and beta turn, seemed to form a mixture of the two conformations in phosphatidylserine solution. Angiotensin I with a strong probability for the beta form adopted the beta form in phosphatidylserine solution and sleep peptide with no structure-forming potential remained unordered in lipid solutions. The helix usually predominated over the beta form in lipid solutions if the peptide has potentials for both conformations. This could account for the preponderance of helices in bacteriorhodopsin of the purple membrane, which according to its amino acid sequence would have favored the beta form.  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