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 890 毫秒
1.
The complete sequence-specific assignment of resonances in the1H-NMR spectrum of the polypeptide neurotoxin III (Hm III) from the sea anemoneHeteractis macrodactylus is described. Comparison of the chemical shifts and pattern of NOEs for Hm III with those for the related toxin Hp III fromHeteractis paumotensis, which differs only in the substitution of Asn for Tyr at position 11, shows that the overall secondary and tertiary structures are conserved. The largest differences in chemical shift caused by the substitution at position 11 are observed for the NH resonances of Arg-13, Thr-14, Ala-15, Leu-17, and Cys-26. The CαH resonances influenced most are those of ASP-6, Gly-9, Leu-17, and Glu-42, while the most affected CβH resonances are from Leu-17, Glu-28, and Lys-32. The absence of long-range NOEs to the aromatic ring of Tyr-11 as well as the lack of significant chemical shift effects on residues outside the loop comprising residues 7–16 confirm that this part of the loop makes no long-lived contacts with the rest of the molecule. The deviations from random coil shifts of Hm III are compared with those of the related anemone toxins Hp II, Hp III, and toxin I fromStichodactyla helianthus (Sh I). The similarity in deviations in chemical shift as a function of sequence position for these four toxins emphasizes the overall structural homology among these polypeptides.  相似文献   

2.
In this paper, the N-terminus of glycoprotein-41, the HIV-1 fusion peptide, was studied by molecular dynamics simulations in an explicit sodium dodecyl sulfate micelle. The simulation provides a detailed picture of the equilibrium structure and peptide stability as it interacts with the micelle. The equilibrium location of the peptide shows the peptide at the surface of the micelle with hydrophobic residues interacting with the micelle's core. At equilibrium, the peptide adopts an alpha-helical structure from residues 5-16 and a type-1 beta-turn from 17-20 with the other residues exhibiting more flexible conformations. The primary hydrophobic interactions with the micelle are from the leucine and phenylalanine residues (Leu-7, Phe-8, Leu-9, Phe-11, Leu-12) while the alanine and glycine residues (Ala-1, Gly-3, Gly-5, Ala-6, Gly-10, Gly-13, Ala-14, Ala-15, Gly-16, Gly-10, Ala-21) interact favorably with water molecules. The results suggest that Phe-8, part of the highly conserved FLG motif of the fusion peptide, plays a key role in the interaction of the peptide with membranes. Our simulations corroborate experimental investigations of the fusion peptide in SDS micelles, providing a high-resolution picture that explains the experimental findings.  相似文献   

3.
The complete sequence-specific assignment of resonances in the1H-NMR spectrum of the polypeptide neurotoxin III (Hm III) from the sea anemoneHeteractis macrodactylus is described. Comparison of the chemical shifts and pattern of NOEs for Hm III with those for the related toxin Hp III fromHeteractis paumotensis, which differs only in the substitution of Asn for Tyr at position 11, shows that the overall secondary and tertiary structures are conserved. The largest differences in chemical shift caused by the substitution at position 11 are observed for the NH resonances of Arg-13, Thr-14, Ala-15, Leu-17, and Cys-26. The CH resonances influenced most are those of ASP-6, Gly-9, Leu-17, and Glu-42, while the most affected CH resonances are from Leu-17, Glu-28, and Lys-32. The absence of long-range NOEs to the aromatic ring of Tyr-11 as well as the lack of significant chemical shift effects on residues outside the loop comprising residues 7–16 confirm that this part of the loop makes no long-lived contacts with the rest of the molecule. The deviations from random coil shifts of Hm III are compared with those of the related anemone toxins Hp II, Hp III, and toxin I fromStichodactyla helianthus (Sh I). The similarity in deviations in chemical shift as a function of sequence position for these four toxins emphasizes the overall structural homology among these polypeptides.  相似文献   

4.
13C-nmr chemical shifts of backbone carbonyl and side-chain β-carbons in polypeptides provide structural information. Recent utilization of substituent effects on 13C-nmr chemical shifts (principally γ-effects) has permitted the rationalization of their sequence and conformation dependence when observed in linear, flexible polypeptides. In this report, we apply the γ-effect method to study the 13C-nmr chemical shifts observed in solution and in the solid state for the backbone carbonyl and side-chain β-carbons in conformationally rigid polypeptides, which are usually cyclic. As found previously for flexible, linear polypetides, the relative 13C-nmr chemical shifts observed for the backbone carbonyl and side-chain β-carbons in conformationally rigid polypeptides are predictable from knowledge of their peptide residue sequence (primary structure) and conformation (secondary structure) via the γ-effect method.  相似文献   

5.
Gerig JT 《Biophysical journal》2004,86(5):3166-3175
Fluorinated alcohols can induce peptides and proteins to take up helical conformations. Nuclear Overhauser effect (NOE) spectroscopy experiments and analysis of C(alpha)H proton chemical shifts show that the conformation of melittin in 35% hexafluoro-2-propanol/water is alpha-helical from residues Ile-2 to Val-8 and from Leu-13 to Gln-25. As has been found in other solvent systems, the two helical regions are not colinear; the interhelix angle (73 +/- 15 degrees ) in 35% 1,1,1,3,3,3-hexafluoro-2-propanol/water is smaller than the angle found in other fluoroalcohol-water mixtures or in the crystal. Intermolecular (1)H(19)F and (1)H(1)H nuclear Overhauser effects were used to explore interaction of solvent components with melittin dissolved in this solvent mixture. The NOEs observed indicate that fluoroalcohol and water molecules are both tightly bound to the peptide in the vicinity of the interhelix bend. For the remainder of the molecule, solute-solvent NOEs are consistent with preferential solvation of the peptide by the fluoroalcohol component of the solvent mixture.  相似文献   

6.
pH titration by 1H-NMR spectroscopy of the peptide hormone bradykinin was carried out in 2H2O. Assignment of all α-proton signals and of most of the other resonances permitted the extraction of vicinal coupling constants 3Jαβ,β′ from which side chain conformation of all residues could be followed and analyzed as a function of pH. It is shown that the ionization of the terminal COOH group affects simultaneously the Arg-9 and Phe-8 chemical shifts and side chain orientation, and the non-equivalence of the Gly-4 methylene protons. Cooperative effects along the peptide backbone or a folded structure of the C-terminal part of bradykinin could explain this effect.  相似文献   

7.
We present a program, named Promega, to predict the Xaa-Pro peptide bond conformation on the basis of backbone chemical shifts and the amino acid sequence. Using a chemical shift database of proteins of known structure together with the PDB-extracted amino acid preference of cis Xaa-Pro peptide bonds, a cis/trans probability score is calculated from the backbone and 13Cβ chemical shifts of the proline and its neighboring residues. For an arbitrary number of input chemical shifts, which may include Pro-13Cγ, Promega calculates the statistical probability that a Xaa-Pro peptide bond is cis. Besides its potential as a validation tool, Promega is particularly useful for studies of larger proteins where Pro-13Cγ assignments can be challenging, and for on-going efforts to determine protein structures exclusively on the basis of backbone and 13Cβ chemical shifts.  相似文献   

8.
The crystal structure of the triple‐helical peptide (Pro‐Hyp‐Gly)3‐Pro‐Arg‐Gly‐(Pro‐Hyp‐Gly)4 (POG3‐PRG‐POG4) was determined at 1.45 Å resolution. POG3‐PRG‐POG4 was designed to permit investigation of the side‐chain conformation of the Arg residues in a triple‐helical structure. Because of the alternative structure of one of three Arg residues, four side‐chain conformations were observed in an asymmetric unit. Among them, three adopt a ttg?t conformation and the other adopts a tg?g?t conformation. A statistical analysis of 80 Arg residues in various triple‐helical peptides showed that, unlike those in globular proteins, they preferentially adopt a tt conformation for χ1 and χ2, as observed in POG3‐PRG‐POG4. This conformation permits van der Waals contacts between the side‐chain atoms of Arg and the main‐chain atoms of the adjacent strand in the same molecule. Unlike many other host–guest peptides, in which there is a significant difference between the helical twists in the guest and the host peptides, POG3‐PRG‐POG4 shows a marked difference between the helical twists in the N‐terminal peptide and those in the C‐terminal peptide, separated near the Arg residue. This suggested that the unique side‐chain conformation of the Arg residue affects not only the conformation of the guest peptide, but also the conformation of the peptide away from the Arg residue. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1000–1009, 2014.  相似文献   

9.
Variant semisynthetic ribonuclease-S complexes were characterized in which the helical glutamic acid 9 residue was replaced by either leucine or glycine. The Leu-9 and Gly-9 synthetic peptides, corresponding otherwise to residues 1 through 15 of bovine pancreatic ribonuclease, were studied with respect to the ability to bind, and generate enzymic activity, with the complementary native protein fragment containing residues 21 through 124 of ribonuclease (RNAase-S-(21–124)). Both the Leu and Gly peptides bind to the RNAase-S-(21–124) to yield complexes with catalytic properties similar to those obtained with the Glu-9-containing peptide of residues 1 through 20 of ribonuclease (RNAase-S-(1–20)). However, whereas the binding affinity of Leu peptide to RNAase-S-(21–124) is only a factor of three less than that for RNAase-S-(1–20), that for Gly peptide is about 20-fold less than that for RNAase-S-(1–20). The stronger binding of Leu than Gly peptide corresponds to the observed propensity of leucine but not glycine for the α-helical conformation in globular proteins.In spite of the weakened affinity of the Gly peptide for RNAase-S-(21–124), it is essentially fully as capable as the Leu-9 and RNAase-S-(1–20) peptides in directing the re-formation of correct disulfide-containing conformation of RNAase-S-(21–124) after disulfide randomization of the latter.  相似文献   

10.
A 17 residue peptide corresponding to the C-helix of hen lysozyme (residues 86 to 102) has been investigated in detail to assess the factors that determine its conformation in both aqueous and trifluoroethanol (TFE) solutions. A thorough characterisation of the peptide by CD and NMR techniques under both conditions has been performed including the determination of complete NMR proton sequential assignments, and measurement of NOE effects,3JHNαcoupling constants, temperature coefficients and residue-specific hydrogen-exchange rates. In water, the peptide adopts a largely unstructured conformation and NMR data, particularly coupling constants and chemical shift deviations, have been shown to agree closely with predictions from a model for a random coil based on the φ,ψ distributions in a protein database. This indicates that under these conditions the intrinsic conformational preferences of the individual amino acid residues are the dominating factors that determine the population of conformers adopted. With increasing concentrations of TFE a cooperative transition to an extensively helical conformation occurs and the resultant changes in CαH chemical shifts have been shown to correlate with the changes in φ,ψ populations. Using NOE and coupling constant data for this state, an ensemble of structures has been calculated and provides a model for a helix in the absence of tertiary interactions. In this model fluctuations, which increase in amplitude towards the termini, occur about the average helical φ,ψ angles and are responsible for increasing the values of3JHNαcoupling constants above those anticipated for a static helix. The residue-specific rates of hydrogen exchange for the peptide in 50% TFE-d3are consistent with such a model, the maximum protection from exchange being observed for residues in the centre of the helix.  相似文献   

11.
Mastoparan-B, a tetradecapeptide isolated from the venom of the hornet Vespa basalis, belongs to the mastoparan analogs of vespid venom with the lysine residues common for all mastoparan family toxins at positions 4, 11 and 12. Here we use 1H-NMR spectroscopy and hybrid distance geometry-simulated annealing calculation to investigate its three-dimensional structure in trifluoroethanol-containing aqueous solution. The calculated structure shows that residues 3–14 adopt an amphiphilic α-helical structure in which the residues with hydrophilic side chains (i.e. Lys-4, Ser-5, Ser-8, Lys-11, Lys-12) are located on one side and the residues with hydrophobic side chains (i.e. Leu-3, Ile-6, Trp-9, A a-10, Val-13, Leu-14) located on the other side of the molecule. The overall structural features are very similar to the conformation of mastoparan-X reconstituted in vesicles [Wakamatsu et al. (1992) Biochemistry 31, 5654–5660] in spite of the substitutions made for eight residues with distinctly different hydrophobicity. These substitutions lead to a larger hydrophobic moment for the α-helical segment and further mobilized N-terminal. This study will help reveal the conformational significance of mastoparan toxins with respect to their potency and activity in G protein regulation.  相似文献   

12.
The three-dimensional structures of the two peptides, PlnJ and PlnK, that constitutes the two-peptide bacteriocin plantaricin JK have been solved in water/TFE and water/DPC-micellar solutions using nuclear magnetic resonance (NMR) spectroscopy. PlnJ, a 25 residue peptide, has an N-terminal amphiphilic α-helix between Trp-3 and Tyr-15. The 32 residues long PlnK forms a central amphiphilic α-helix between Gly-9 and Leu-24. Measurements of the effect on anti-microbial activity of single glycine replacements in PlnJ and PlnK show that Gly-13 and Gly-17 in both peptides are very sensitive, giving more than a 100-fold reduction in activity when large residues replace glycine. In variants where other glycine residues, Gly-20 in PlnJ and Gly-7, Gly-9, Gly-24 and Gly-25 in PlnK, were replaced, the activity was reduced less than 10-fold. It is proposed that the detrimental effect on activity when exchanging Gly-13 and Gly-17 in PlnJ and PlnK is a result of reduced ability of the two peptides to interact through the GxxxG-motifs constituting Gly-13 and Gly-17.  相似文献   

13.
The membrane-bound conformation of a cell-penetrating peptide, penetratin, is investigated using solid-state NMR spectroscopy. The 13C chemical shifts of 13C, 15N-labeled residues in the peptide indicate a reversible conformational change from β-sheet at low temperature to coil-like at high temperature. This conformational change occurs for all residues examined between positions 3 and 13, at peptide/lipid molar ratios of 1:15 and 1:30, in membranes with 25-50% anionic lipids, and in both saturated DMPC/DMPG (1,2-dimyristoyl-sn-glycero-3-phosphatidylchloline/1,2-dimyristoyl-sn-glycero-3-phosphatidylglycerol) membranes and unsaturated POPC/POPG (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylcholine/1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidylglycerol) membranes. Thus, it is an intrinsic property of penetratin. The coil state of the peptide has C-H order parameters of 0.23-0.52 for Cα and Cβ sites, indicating that the peptide backbone is unstructured. Moreover, chemical shift anisotropy lineshapes are uniaxially averaged, suggesting that the peptide backbone undergoes uniaxial rotation around the bilayer normal. These observations suggest that the dynamic state of penetratin at high temperature is a structured turn instead of an isotropic random coil. The thermodynamic parameters of this sheet-turn transition are extracted and compared to other membrane peptides reported to exhibit conformational changes. We suggest that the function of this turn conformation may be to reduce hydrophobic interactions with the lipid chains and facilitate penetratin translocation across the bilayer without causing permanent membrane damage.  相似文献   

14.
Rhodopsin undergoes rearrangements of its transmembrane helices after photon absorption to transfer a light signal to the G-protein transducin. To investigate the mechanism by which rhodopsin adopts the transducin-activating conformation, the local environmental changes in the transmembrane region were probed using the cysteine S-H group, whose stretching frequency is well isolated from the other protein vibrational modes. The S-H stretching modes of cysteine residues introduced into Helix III, which contains several key residues for the helical movements, and of native cysteine residues were measured by Fourier transform infrared spectroscopy. This method was applied to metarhodopsin IIa, a precursor of the transducin-activating state in which the intramolecular interactions are likely to produce a state ready for helical movements. No environmental change was observed near the ionic lock between Arg-135 in Helix III and Glu-247 in Helix VI that maintains the inactive conformation. Rather, the cysteine residues that showed environmental changes were located around the chromophore, Ala-164, His-211, and Phe-261. These findings imply that the hydrogen bond between Helix III and Helix V involving Glu-122 and His-211 and the hydrophobic packing between Helix III and Helix VI involving Gly-121, Leu-125, Phe-261, and Trp-265 are altered before the helical rearrangement leading toward the active conformation.  相似文献   

15.
Structure and membrane interaction of a 31 amino acid residue fragment of the membrane bound FKBP-like protein twisted dwarf 1 (TWD1) from Arabidopsis thaliana was investigated by solid-state NMR spectroscopy. The studied peptide TWD1(335–365) contained the putative membrane anchor of the protein (residues 339–357) that was previously predicted by sequence hydrophobicity analysis. The TWD1 peptide was synthesized by standard solid phase peptide synthesis and contained three uniformly 13C- and 15N-labelled residues (Phe 340, Val 350, Ala 364). The peptide was incorporated into either multilamellar vesicles or oriented planar membranes composed of an equimolar ternary phospholipid mixture (POPC, POPE, POPG), where the POPC was sn-1 chain-deuterated. 31P NMR spectra of the membrane in the absence and in the presence of the peptide showed axially symmetric powder patterns indicative of a lamellar bilayer phase. Further, the addition of peptide caused a decrease in the lipid hydrocarbon chain order as indicated by reduced quadrupolar splittings in the 2H NMR spectra of the POPC in the membrane. The conformation of TWD1(335–365) was investigated by 13C cross-polarization magic-angle spinning NMR spectroscopy. At a temperature of −30°C all peptide signals were resolved and could be fully assigned in two-dimensional proton-driven 13C spin diffusion and 13C single quantum/double quantum correlation experiments. The isotropic chemical shift values for Phe 340 and Val 350 exhibited the signature of a regular α-helix. Chemical shifts typical for a random coil conformation were observed for Ala 364 located close to the C-terminus of the peptide. Static 15N NMR spectra of TWD1(335–365) in mechanically aligned lipid bilayers demonstrated that the helical segment of TWD1(335–365) adopts an orientation perpendicular to the membrane normal. At 30°C, the peptide undergoes intermediate time scale motions. Dedicated to Prof. K. Arnold on the occasion of his 65th birthday.  相似文献   

16.
1H- and 13C-nmr studies of conformational transitions of random amino acid copolymers containing aromatic residues (Lys50Tyr50)n and (Lys50Phe50)n in the presence of neutral salts were performed to serve as models of the aggregation behavior of polypeptides of biological significance. The 1H and 13C signal intensities of Tyr and Phe residues decreased preferentially with increasing concentration of neutral salts such as NaCl and NaClO4. This behavior contrasts with that of (Lys)n in the presence of similar neutral salts, where the displacement of the 13C signal is clearly seen on transition from the random-coil to the helical conformation. On the basis of the previous conformational studies, the loss of the peak areas is ascribed to the presence of immobilized helical segments by hydrophobic interaction between aromatic side chains. The remaining resonances are due to the residual random-coil regions, since the values of nuclear Overhauser enhancements and chemical shifts are unchanged in the presence and absence of the neutral salts.  相似文献   

17.
The tetradecapeptide Ac-D-F-L-A-E-G-G-G-V-R-G-P-R-V-OMe, which mimics residues 7f-20f of the A alpha-chain of human fibrinogen, has been co-crystallized with bovine thrombin from ammonium sulfate solutions in space group P2(1) with unit cell dimensions of a = 83.0 A, b = 89.4 A, c = 99.3 A, and beta = 106.6 degrees. Three crystallographically independent complexes were located in the asymmetric unit by molecular replacement using the native bovine thrombin structure as a model. The standard crystallographic R-factor is 0.167 at 2.3-A resolution. Excellent electron density could be traced for the decapeptide, beginning with Asp-7f and ending with Arg-16f in the active site of thrombin; the remaining 4 residues, which have been cleaved from the tetradecapeptide at the Arg-16f/Gly-17f bond, are not seen. Residues 7f-11f at the NH2 terminus of the peptide form a single turn of alpha-helix that is connected by Gly-12f, which has a positive phi angle, to an extended chain containing residues 13f-16f. The major specific interactions between the peptide and thrombin are 1) a hydrophobic cage formed by residues Tyr-60A, Trp-60D, Leu-99, Ile-174, Trp-215, Leu-9f, Gly-13f, and Val-15f that surrounds Phe-8f; 2) a hydrogen bond linking Phe-8f NH to Lys-97 O;3) a salt link between Glu-11f and Arg-173; 4) two antiparallel beta-sheet hydrogen bonds between Gly-14f and Gly-216; and 5) the insertion of Arg-16f into the specificity pocket. Binding of the peptide is accompanied by a considerable shift in two of the loops near the active site relative to human D-phenyl-L-prolyl-L-arginyl chloromethyl ketone (PPACK)-thrombin.  相似文献   

18.
The combined use of several nuclear magnetic resonance and restrained molecular dynamics techniques allowed the formulation of a molecular model for the preferred solution conformation of a synthetic peptide reproducing the [1-20] processing domain of the pro-ocytocin-neurophysin precursor. In the model, the conformation of the 20-membered tocin ring, with the two Cys1 and Cys6 residues bridged by a disulphide bond, is very close to that observed for isolated ocytocin in the solid state; in addition, a type II β-turn is postulated for the 7-10 segment of the acyclic tail containing the Lys11-Arg12 processing site, and connecting ocytocin to the neurophysin domain, while the C-terminal 13-20 segment of the molecule is believed to assume a helical structure. This particular structural organization could be important in participating as the favorable conformation for optimal substrate-enzyme active site recognition and processing by specific endoproteases. © 1996 John Wiley & Sons, Inc.  相似文献   

19.
WASp-interacting protein (WIP) is a 503-residue proline-rich polypeptide expressed in human T cells. The WIP C-terminal domain binds to Wiskott-Aldrich syndrome protein (WASp) and regulates its activation and degradation, and the WIP-WASp interaction has been shown to be critical for actin polymerization and implicated in the onset of WAS and X-linked thrombocytopenia. WIP is predicted to be an intrinsically disordered protein, a class of polypeptides that are of great interest because they violate the traditional structure-function paradigm. In this first (to our knowledge) study of WIP in its unbound state, we used NMR to investigate the biophysical behavior of WIPC, a C-terminal domain fragment of WIP that includes residues 407–503 and contains the WASp-binding site. In light of the poor spectral dispersion exhibited by WIPC and the high occurrence (25%) of proline residues, we employed 5D-NMR13C-detected NMR experiments with nonuniform sampling to accomplish full resonance assignment. Secondary chemical-shift analysis, 15N relaxation rates, and protection from solvent exchange all concurred in detecting transient structure located in motifs that span the WASp-binding site. Residues 446–456 exhibited a propensity for helical conformation, and an extended conformation followed by a short, capped helix was observed for residues 468–478. The 13C-detected approach allows chemical-shift assignment in the WIPC polyproline stretches and thus sheds light on their conformation and dynamics. The effects of temperature on chemical shifts referenced to a denatured sample of the polypeptide demonstrate that heating reduces the structural character of WIPC. Thus, we conclude that the disordered WIPC fragment is comprised of regions with latent structure connected by flexible loops, an architecture with implications for binding affinity and function.  相似文献   

20.
The nanoscale peptide YSGVCHTDLHAWHGDWPLPVK exhibits molecular chaperone activity and prevents protein aggregation under chemical and/or thermal stress. Here, His mutations of this peptide and their impact on chaperone activity were evaluated using theoretical techniques. Molecular dynamic (MD) simulations with simulated annealing (SA) of different mutant nanopeptides were employed to determine the contribution of the scaffolding His residues (H45, H49, H52), when mutated to Pro, on chaperone action in vitro. The in silico mutations of His residues to Pro (H45P, H49P, H52P) revealed loss of secondary ordered strand structure. However, a small part of the strand conformation was formed in the middle region of the native chaperone peptide. The His‐to‐Pro mutations resulted in decreased gyration radius (Rg) values and surface accessibility of the mutant peptides under the simulation times. The invariant dihedral angle (ϕ) values and the disrupting effects of the Pro residues indicated the coil conformation of mutant peptides. The failure of the chaperone‐like action in the Pro mutant peptides was consistent with their decreased effective accessible surfaces. The high variation of Φ value for His residues in native chaperone peptide leads to high flexibility, such as a minichaperone acting as a nanomachine at the molecular level. Our findings demonstrate that the peptide strand conformation motif with high flexibility at nanoscale is critical for chaperone activity. Copyright © 2008 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