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.
2.
The caerin 1 peptides are among the most powerful of the broad-spectrum antibiotic amphibian peptides. Caerin 1.1 has previously been shown to form an amphipathic helix-bend-helix structure in aqueous trifluoroethanol (H. Wong, J. H. Bowie, and J. A. Carver European Journal Biochemistry, 1997, Vol. 247, pp. 545-557) and structure-activity relationship studies indicate that both helices are required for activity, as well as flexibility in the bend region connecting the two. The structure of caerin 1.1 in dodecylphosphocholine micelles was investigated and shown to be very similar to that determined in aqueous trifluoroethanol. Caerin 1.4, which is identical to caerin 1.1, but with serine residues replacing Val5 and Gly7, is less active than caerin 1.1 against most bacterial species but has improved activity against Escherichia coli and Micrococcus luteus. The solution NMR structure of caerin 1.4 was determined in both aqueous trifluoroethanol and dodecylphosphocholine micelles, and was shown to be similar to caerin 1.1. It was concluded that differences in the hydrophobicity and hydrophilic angle of the first helix are probably responsible for the different spectra of antibacterial activity. The similarity of the structures calculated in aqueous trifluoroethanol and dodecylphosphocholine micelles suggests that, for caerin 1.1 and 1.4, these solvent systems are equally as good at representing a membrane environment.  相似文献   

3.
Summary The 13C chemical shifts for all of the protonated carbons of the 20 common amino acid residues in the protected linear pentapeptide Gly-Gly-X-Gly-Gly have been obtained in water at low pH as well as in aqueous solution containing 10, 20 and 30% acetonitrile or trifluoroethanol. Dioxane was used as an internal reference and its carbon chemical shift value was found to be 66.6 ppm relative to external TMS in water. Comparison of the different referencing methods for 13C chemical shifts in organic cosolvent mixtures showed that an external standard (either TMS or TSP capillary) was the most appropriate. In the present study, external TSP was chosen to define the 0 ppm of the 13C chemical shift scale. When the difference in referencing the dioxane carbon resonance is taken into account, the carbon chemical shift values of the amino acids in aqueous solution are similar to those previously reported (Richarz and Wüthrich (1978) Biopolymers, 17, 2133–2141; Howarth and Lilley (1979) Prog. NMR Spectrosc., 12, 1–40). The pentapeptides studied were assumed to be in a random coil conformation and the measured 13C chemical shifts were used as reference values to correlate carbon chemical shifts with the secondary structure of two well-characterized peptides, bombesin and the 1–29 amino acid fragment of Nle27 human growth hormone-releasing factor. In both cases, the C chemical shifts exhibited a characteristic positive deviation from the random coil values, which indicates the presence of -helices.  相似文献   

4.
Divalent metal‐ion transporter 1 (DMT1) belongs to a large class of metal‐ion transporters that drive the translocation of a wide range of divalent metal substrates across membranes toward the cytosol with couple of protons. Two highly conserved histidines in the sixth transmembrane domain (TMD6) are essential for metal transport activity in DMT1. In the present study, we determine the high‐resolution structures of three 25‐residue peptides, corresponding to TMD6 of the wildtype DMT1 (the segment 255–279) and its H267A and H272A mutants, in 30% TFE‐d2 aqueous solution by the combined use of circular dichroism (CD) and NMR spectroscopies. The wildtype peptide forms an ‘α‐helix‐extended segment‐α‐helix’ structure with two helices spanning over Gly258–Ala262 and Met265–Lys277 linked by a hinge at residues Val263–Ile264. The H267A mutation reduces the hinge to one residue (Ile264), while the H272A mutation extends the flexible region of the central part from Val263 to His267. Diffusion‐ordered spectroscopy (DOSY) study demonstrates that all the peptides are self‐assembly as trimer in 30% TFE‐d2 aqueous solution. The H272A substitution decreases the intermolecular interaction whereas the H267A substitution may enhance the intermolecular interaction. The specific structure of the discontinuous helix and the self‐assembly feature of DMT1–TMD6 may be crucial for its biological function. The changes in conformation and intermolecular interaction induced by histidine substitution may be correlated with the deficiency of DMT1 in metal‐ion permeation. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
Defining the self-association state of a molecule in solution can be an important step in NMR-based structure determination. This is particularly true of peptides, where there can be a relatively small number of long-range interactions and misinterpretation of an intermolecular NOE as an intramolecular contact can have a dramatic influence on the final calculated structure. In this paper, we have investigated the use of translational self-diffusion coefficient measurements to detect self-association in aqueous trifluoroethanol of three peptides which are analogues of the C-terminal region of human neuropeptide Y. Experimentally measured diffusion coefficients were extrapolated to D0, the limiting value as the peptide concentration approaches zero, and then converted to D20,w, the diffusion coefficient after correction for temperature and the viscosity of the solvent. A decrease in D20,w of about 16% was found for all three peptides in aqueous TFE (30% by volume) compared with water, which is in reasonable agreement with the expected decrease upon dimerisation, the presence of which was indicated by sedimentation equilibrium measurements. Apparent molecular masses of these peptides in both solutions were also calculated from their diffusion coefficients and similar results were obtained. Several potential internal standards, including acetone, acetonitrile, dimethylsulfoxide and dioxane, were assessed as monitors of solution viscosity over a range of trifluoroethanol concentrations. Compared with independent measurements of viscosity, acetonitrile was the most accurate standard among these four. The practical limitations of a quantitative assessment of peptide self-association from translational diffusion coefficients measured by PFGNMR, including the calculation of apparent molecular mass, are also discussed.  相似文献   

6.
Cell membrane proteome analysis is limited by inherent membrane hydrophobicity. Conventional membrane protein extraction techniques use detergents, chaotropes and organic acids that require sample clean-up or pH adjustment, and are associated with significant sample loss. We extracted membrane proteins from red blood cells (RBCs) using methanol (MeOH), trifluoroethanol (TFE) and urea, and identified membrane proteins using 2-D LC coupled with MALDI-TOF/TOF-MS. We show that organic solvents MeOH- and TFE-based methods have membrane protein analysis efficiencies comparable to urea, and are complementary for the recovery of both hydrophilic and hydrophobic peptides. The mean grand average of hydropathicity (GRAVY) value of identified peptides from the TFE-based method (-0.107) was significantly higher than that of the MeOH-based method (-0.465) (p<0.001). Sequential and adjunctive use of the organic solvents MeOH and TFE increases the number of proteins identified, and the confidence of their identification. We show that this strategy is effective for shotgun membrane proteome analysis.  相似文献   

7.
Amyloid fibril formation by peptide LYS (11-36) in aqueous trifluoroethanol   总被引:1,自引:0,他引:1  
Peptide LYS (11-36), derived from the beta-sheet region of T4 lysozyme, forms an amyloid fibril in aqueous trifluoroethanol (TFE) at elevated temperature. The peptide has a moderate alpha-helix content in 20 and 50% (v/v) TFE solution; large quantities of fibrils were formed after incubation at 55 degrees C for 2 weeks as monitored by a thioflavin T fluorescence assay. No fibrils were observed when the peptide initially existed predominantly as a random coil or as a complete alpha helix. Our results suggest that a moderate amount of alpha helix and random coil present in the peptide initially facilitates the fibril-formation process, but a high alpha-helix content inhibits fibril formation. Transmission electron microscopy revealed several types of fibril morphologies at different TFE concentrations. The fibrils were highly twisted and consisted of interleaved protofilaments in 50% TFE, while smooth and flat ribbonlike fibrils were found in 20% TFE. In 50% TFE, the fibril growth rate of LYS (11-36) was found to depend strongly on peptide concentration and seeding but was insensitive to solution pH and ionic strength.  相似文献   

8.
9.
The structure of water and the nature of ionic hydration is explored in aqueous solutions of potassium fluoride, chloride, bromide and iodide over a range of concentrations up to 4.8 ion pairs per 100 water molecules, using the combined techniques of neutron diffraction with hydrogen isotope substitution. The diffraction data are interpreted using the method of empirical potential structure refinement, which attempts to build a three-dimensional model of the scattering system consistent with the diffraction data. The water structure is strongly perturbed in the first hydration shells of both anion and cation, but is found to be only mildly perturbed outside of this region, with the largest effects occurring with the smallest anion and highest concentrations. For the potassium ion there are strong orientational correlations in the first hydration shell, with the water molecules lying with their dipole moments pointing almost directly away from the cation on average, but with an angular spread of approximately +/-60 degrees which is mildly dependent on the anion type present. For all the anions the water molecules in the first shell are strongly oriented with one O-H vector pointing directly towards the anion on average, with an angular spread of approximately +/-10 degrees for F(-), increasing to approximately +/-22 degrees for I(-). For both anions and cations the second hydration shell is much more disordered than the first, but there is a weak pattern of orientational correlation which becomes more pronounced with the larger anions. There is some evidence that the fluoride ion structures water significantly in its first hydration shell, but not beyond. The findings throw further light on recent findings that the orientational relaxation time for water outside the first shell of dissolved ions is the same as in the bulk liquid.  相似文献   

10.
CD and nmr studies have been carried out on aqueous trifluoroethanol (TFE) solutions of bradykinin (BK) and a bradykinin antagonist. The CD results exhibit a striking effect of TFE on the spectra of BK, with sequence Arg-Pro-Pro-Gly-Phe-Ser-Pro-Phe-Arg, and the BK antagonist, with sequence D -Arg-Arg-Pro-Hyp-Gly-Thi-D -Ser-D -Cpg-Cpg-Arg [where Hyp is 4-hydroxy-L -proline; Thi refers to β-(2-thienyl)-L -alanine and Cpg refers to α-cyclopentylglycine]. The effect of increasing concentration of TFE in water on the difference ellipticity at 222 nm was examined and showed that BK may be a mixture of at least two different conformers, one of which largely forms when the TFE concentration is increased beyond 80%. The linear extrapolation of 100% of the difference ellipticity of BK at low TFE concentrations yields a value in agreement with that shown by the BK antagonist, indicating that the conformation of BK at the lower TFE concentrations is similar to that of the BK antagonist. The conformational analysis was carried out using both one-dimensional and two-dimensional 1H-nmr techniques. The total correlation spectroscopy (TOCSY) spectrum of BK in a 60/40% (v/v) TFE/H2O solution at 10°C and a nuclear Overhauser effect spectroscopy (NOESY) spectrum that shows only sequential Hα(i) – NH(i + 1) or the Hα(i) – Hδδ′(i + 1) NOEs indicate that the majority of the molecules adopt an all-trans extended conformation. The TOCSY for BK in the 95/5% (v/v) TFE/H2O solution shows that there are two major conformations in the solution with about equal population. The NOESY experiment shows two new important cross peaks for one conformation, namely Pro2(α)-Pro3 (α) and the Pro2(α)-Gly4(NH), indicating a cis Pro2-Pro3 bond and a type VI β-turn between residues Arg1 and Gly4 involving cis proline at position 3, respectively. The low temperature coefficient of Gly4 for this conformation suggests the presence of an intramolecular hydrogen bond, therefore a type VIa β-turn is present. The other conformation is all trans and extended. The BK antafonist shows difference CD spectra in TFE solutions referred to H2O that are superficially indicative of a β-bend. However, nmr speaks against this possibility, as only one set of peaks were observed in the TOCSY and NOESY experiments, indicating an all-trans extended confirmation over the range of TFE concentrations. The BK-antagonist CD data suggest that solvent perturbation of the CD of an extended confirmation perturbation of the optical activity of the thienyl moiety of the peptide since the CD spectrum of N-acetyl-β-thienyl-L -alanine N-methylamide is strongly perturbed by TFE. The present results again demonstrate the complementary relationship between CD and nmr. © 1994 John Wiley & Sons, Inc.  相似文献   

11.
We have developed a solvation function that combines a Generalized Born model for polarization of protein charge by the high dielectric solvent, with a hydrophobic potential of mean force (HPMF) as a model for hydrophobic interaction, to aid in the discrimination of native structures from other misfolded states in protein structure prediction. We find that our energy function outperforms other reported scoring functions in terms of correct native ranking for 91% of proteins and low Z scores for a variety of decoy sets, including the challenging Rosetta decoys. This work shows that the stabilizing effect of hydrophobic exposure to aqueous solvent that defines the HPMF hydration physics is an apparent improvement over solvent-accessible surface area models that penalize hydrophobic exposure. Decoys generated by thermal sampling around the native-state basin reveal a potentially important role for side-chain entropy in the future development of even more accurate free energy surfaces.  相似文献   

12.
The surface properties of spinach chloroplasts, both of intact chloroplasts with surrounding envelope and broken chloroplasts consisting of the inner lamellar system, have been studied by partitioning them between two aqueous phases, especially using counter-current distribution technique. The two-phase system consists of poly(ethyleneglycol), dextran and water. The two polymers are enriched in opposite phases and by binding deoxycholate or palmitate to one of the polymers the affinity of chloroplasts for the corresponding phase is strongly enhanced. The partition of the two classes of chloroplasts, however, is not affected to the same degree and the affinity of the chloroplast envelope for deoxycholate and palmitate is stronger than that of the lamellar system. This has been correlated to the chemical composition of the two types of membranes. By studying the effect of salts on the partition it has been found that the lamellar system bears a larger number of negative charges as compared to the envelope of the intact chloroplast.  相似文献   

13.
A variety of sugars are known to enhance the stability of biomaterials. Trehalose, a nonreducing disaccharide composed of two alpha, alpha(1 --> 1)-linked D-glucopyranose units, appears to be one of the most effective protectants. Both in vivo and in vitro, trehalose protects biostructures such as proteins and membranes from damage due to dehydration, heat, or cold. However, despite the significant amount of experimental data on this disaccharide, no clear picture of the molecular mechanism responsible for its stabilizing properties has emerged yet. Three major hypotheses (water-trehalose hydrogen-bond replacement, coating by a trapped water layer, and mechanical inhibition of the conformational fluctuations) have been proposed to explain the stabilizing effect of trehalose on proteins. To investigate the nature of protein-trehalose-water interactions in solution at the molecular level, two molecular dynamics simulations of the protein lysozyme in solution at room temperature have been carried out, one in the presence (about 0.5 M) and one in the absence of trehalose. The results show that the trehalose molecules cluster and move toward the protein, but neither completely expel water from the protein surface nor form hydrogen bonds with the protein. Furthermore, the coating by trehalose does not significantly reduce the conformational fluctuations of the protein compared to the trehalose-free system. Based on these observations, a model is proposed for the interaction of trehalose molecules with a protein in moderately concentrated solutions, at room temperature and on the nanosecond timescale.  相似文献   

14.
15.
Circular dichroism spectroscopy has been used to study the interactions between heparin and cationic polypeptides in dilute aqueous solution at neutral pH. The results indicate that poly(l-lysine), poly(l-arginine), and poly(l-ornithine) adopt the α-helical conformation in the presence of heparin, rather than the “charged coil” form observed for the polypeptide alone under the same conditions. Maximum interaction for the poly(l-lysine) and poly(l-ornithine) systems occur at an amino acid: disaccharide residue ratio of 2.3 ± 0.1:1, which correlates with the analytical data of 2.3 sulfates per heparin disaccharide. For poly(l-arginine), maximum interaction occurs at a residue ratio of 3.3 ± 0.1:1, and indicates that all the anionic groups (sulfate and carboxyl) of the heparin are involved in this case.The interactions of heparin are analogous to those observed previously for six connective tissue mucopolysaccharides, except that none of the latter had any effect on the conformation of poly(l-ornithine). The poly(l-ornithine)-heparin system shows a thermal “melting” transition at Tm = 56.0 ± 1.0 °C, at which point the polypeptide reverts to the “charged coil” form; the interactions with poly(l-lysine) and poly(l-arginine) are stable up to temperatures > 90 °C. The high thermal stability of these conformation-directing effects indicate a stronger interaction for heparin than the other mucopolysaccharides, which is probably due to the high sulfate content.  相似文献   

16.
Protein-protein interactions have been measured for a mutant (D101F) lysozyme and for native lysozyme in concentrated solutions of ammonium sulfate at pH 7 and sodium chloride at pH 4.5. In the mutant lysozyme, a surface aspartate residue has been replaced with a hydrophobic phenylalanine residue. The protein-protein interactions of D101F lysozyme are more attractive than those of native lysozyme for all conditions studied. The salt-induced attraction is correlated with a solvation potential of mean force given by the work required to desolvate the part of the protein surfaces that is buried by the protein-protein interaction. This work is proportional to the aqueous surface-tension increment of the salt and the fractional non-polar surface coverage of the protein. Experimental measurements of osmotic second virial coefficients validate a proposed potential of mean force that ascribes the salt-induced attraction between protein molecules to an enhancement of the hydrophobic attraction. This model provides a first approximation for predicting the protein-protein potential of mean force in concentrated aqueous electrolyte solutions; this potential is useful for determining solution conditions favorable for protein crystallization.  相似文献   

17.
Dennis S  Camacho CJ  Vajda S 《Proteins》2000,38(2):176-188
To understand water-protein interactions in solution, the electrostatic field is calculated by solving the Poisson-Boltzmann equation, and the free energy surface of water is mapped by translating and rotating an explicit water molecule around the protein. The calculation is applied to T4 lysozyme with data available on the conservation of solvent binding sites in 18 crystallographically independent molecules. The free energy maps around the ordered water sites provide information on the relationship between water positions in crystal structure and in solution. Results show that almost all conserved sites and the majority of nonconserved sites are within 1.3 A of local free energy minima. This finding is in sharp contrast to the behavior of randomly placed water molecules in the boundary layer, which, on the average, must travel more than 3 A to the nearest free energy minimum. Thus, the solvation sites are at least partially determined by protein-water interactions rather than by crystal packing alone. The characteristic water residence times, obtained from the free energies at the local minima, are in good agreement with nuclear magnetic resonance experiments. Only about half of the potential sites show up as ordered water in the 1.7 A resolution X-ray structure. Crystal packing interactions can stabilize weak or mobile potential sites (in fact, some ordered water positions are not close to free energy minima) or can prevent water from occupying certain sites. Apart from a few buried water molecules that are strong binders, the free energies are not very different for conserved and nonconserved sites. We show that conservation of a water site between two crystals occurs if the positions of protein atoms, primarily contributing to the free energy at the local minimum, do not substantially change from one structure to the other. This requirement can be correlated with the nature of the side chain contacting the water molecule in the site.  相似文献   

18.
Aqueous solutions of vanadium(V) have been prepared at different OH vs. vanadium concentrations from both decavanadic acid and metavanadate solutions. These solutions have been extracted with an excess of tetraheptylammonium bromide in benzene. The bromide displaced has been determined in order to gain information on the average charge of the extracting species already known to be variously protonated decavanadate anions. The addition of a constant excess of supporting electrolytes has been avoided, as it is known it could affect the position of equilibria.Besides the already well known H2V10O284−, HV10 O285− and V10O286− species, strong evidence has also been obtained for the existence of H3V10O283−.Moreover, a slow disproportionation of V10O286− into HV10O285− and V2O74− (or HVO42−) during a first stage, and then into metavanadate, resulted.Therefore, H6V10O28 proves a strong acid as triprotic instead of tetraprotic, as reported in the literature. Inconsistent results are almost certainly due to the high ionic media usually employed.  相似文献   

19.
The conformational transition of poly-L -tyrosine in 0.1M KCl was investigated by ORD and infrared spectroscopy, potentiometric titration, and sedimentation velocity experiments. It is shown that the fully ordered conformer is obtained by slow titration of the random coil with 0.1N HCl at 25°C. The charge-induced transition, at variance with other poly-α-amino acids, is completed in a narrow range of α. An aggregation process was detected both by potentiometric titration and sedimentation velocity. The polyamino acid aggregates around α = 0.7 at 25°C when the conformational transition is almost complete. Infrared spectra, in the region of the amide I band (1650 cm?1) showed that the transition is a random coil → antiparallel β one. Evidence exists that the form is of the intramolecular type. The foregoing interpretations of ORD and CD spectra in terms of the α-helix conformation are discussed.  相似文献   

20.
1. The separation of 0.9-S and 10.8-S allantoicase with the aid of a 2H2O-H2O gradient was described. The resulting preparations were subjected to sedimentation equilibrium, optical rotatory dispersion (ORD), circular dichroism (CD) and infrared studies. 2. The molecular weight of 0.9-S allantoicase was determined to be about 1.1 x 10(4) g/mole in studies on the sedimentation behavior, the metal content and amino acid composition. The molecular weight of 10.8-S allantoicase was about 15.4 x 10(4) g/mole. 3. Optical rotatory dispersion, circular dichroism and infrared studies indicated that both molecules contain alpha-helix, beta conformation and random coil. A Cotton effect at 418 nm was ascribed to the asymmetric binding of Mn2+ to the enzyme. Competitive inhibitors decreased the absorption and circular dichroism bands at about 280 nm and 418 nm. These phenomena suggested that the aromatic groups may play an essential role in the binding of substrates and inhibitors by the Mn(2+)-enzyme complex. 4. Comparison of alpha-helical contents of metalloallantoicases showed that the enzymes with low helical contents exhibited high enzymic activities. 5. The nearly identical physicochemical behavior and specific enzymic activity of 0.9-S and 10.8-S allantoicase indicated that they are very similar in structure and conformation.  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