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.
An increasing number of experimental and theoretical studies have demonstrated the importance of the 3(10)-helix/ alpha-helix/coil equilibrium for the structure and folding of peptides and proteins. One way to perturb this equilibrium is to introduce side-chain interactions that stabilize or destabilize one helix. For example, an attractive i, i + 4 interaction, present only in the alpha-helix, will favor the alpha-helix over 3(10), while an i, i + 4 repulsion will favor the 3(10)-helix over alpha. To quantify the 3(10)/alpha/coil equilibrium, it is essential to use a helix/coil theory that considers the stability of every possible conformation of a peptide. We have previously developed models for the 3(10)-helix/coil and 3(10)-helix/alpha-helix/ coil equilibria. Here we extend this work by adding i, i + 3 and i, i + 4 side-chain interaction energies to the models. The theory is based on classifying residues into alpha-helical, 3(10)-helical, or nonhelical (coil) conformations. Statistical weights are assigned to residues in a helical conformation with an associated helical hydrogen bond, a helical conformation with no hydrogen bond, an N-cap position, a C-cap position, or the reference coil conformation plus i, i + 3 and i, i + 4 side-chain interactions. This work may provide a framework for quantitatively rationalizing experimental work on isolated 3(10)-helices and mixed 3(10)-/alpha-helices and for predicting the locations and stabilities of these structures in peptides and proteins. We conclude that strong i, i + 4 side-chain interactions favor alpha-helix formation, while the 3(10)-helix population is maximized when weaker i, i + 4 side-chain interactions are present.  相似文献   

2.
As an additional step toward the dissection of the factors responsible for the onset of 3(10)-helix vs alpha-helix in peptides, in this paper we describe the results of a three-dimensional (3D) structural analysis by x-ray diffraction of the N(alpha)-acylated heptapeptide alkylamide mBrBz-L-Iva-L-(alphaMe)Val-L-Abu-L-(alphaMe)Val-L-(alphaMe)Phe-L-(alphaMe)Val-L-Iva-NHMe characterized by a single (L-Abu3) C(alpha)-trisubstituted and six C(alpha)-tetrasubstituted alpha-amino acids. We find that in the crystal state this peptide is folded in a mixed helical structure with short elements of 3(10)-helix at either terminus and a central region of alpha-helix. This finding, taken together with the published NMR and x-ray diffraction data on the all C(alpha)-methylated parent sequence and its L-Val2 analog (also the latter heptapeptide has a single C(alpha)-trisubstituted alpha-amino acid) strongly supports the view that one C(alpha)-trisubstituted alpha-amino acid inserted near the N-terminus of an N(alpha)-acylated heptapeptide alkylamide sequence may be enough to switch a regular 3(10)-helix into an essentially alpha-helical conformation. As a corollary of this work, the x-ray diffraction structure of the N(alpha)-protected, C-terminal tetrapeptide alkylamide Z-L-(alphaMe)Val-L-(alphaMe)Phe-L-(alphaMe)Val-L-Iva-NHMe, also reported here, is clearly indicative of the preference of this fully C(alpha)-methylated, short peptide for the 3(10)-helix. As the same terminally blocked sequence is mixed 3(10)/alpha-helical in the L-Abu3 heptapeptide amide but regular 3(10)-helical in the tetrapeptide amide and in the parent heptapeptide amide, these results point to an evident plasticity even of a fully C(alpha)-methylated short peptide.  相似文献   

3.
Structures of N-termini of helices in proteins.   总被引:8,自引:7,他引:1       下载免费PDF全文
We have surveyed 393 N-termini of alpha-helices and 156 N-termini of 3(10)-helices in 85 high resolution, non-homologous protein crystal structures for N-cap side-chain rotamer preferences, hydrogen bonding patterns, and solvent accessibilities. We find very strong rotamer preferences that are unique to N-cap sites. The following rules are generally observed for N-capping in alpha-helices: Thr and Ser N-cap side chains adopt the gauche - rotamer, hydrogen bond to the N3 NH and have psi restricted to 164 +/- 8 degrees. Asp and Asn N-cap side chains either adopt the gauche - rotamer and hydrogen bond to the N3 NH with psi = 172 +/- 10 degrees, or adopt the trans rotamer and hydrogen bond to both the N2 and N3 NH groups with psi = 1-7 +/- 19 degrees. With all other N-caps, the side chain is found in the gauche + rotamer so that the side chain does not interact unfavorably with the N-terminus by blocking solvation and psi is unrestricted. An i, i + 3 hydrogen bond from N3 NH to the N-cap backbone C = O in more likely to form at the N-terminus when an unfavorable N-cap is present. In the 3(10)-helix Asn and Asp remain favorable N-caps as they can hydrogen bond to the N2 NH while in the trans rotamer; in contrast, Ser and Thr are disfavored as their preferred hydrogen bonding partner (N3 NH) is inaccessible. This suggests that Ser is the optimum choice of N-cap when alpha-helix formation is to be encouraged while 3(10)-helix formation discouraged. The strong energetic and structural preferences found for N-caps, which differ greatly from positions within helix interiors, suggest that N-caps should be treated explicitly in any consideration of helical structure in peptides or proteins.  相似文献   

4.
Schievano E  Pagano K  Mammi S  Peggion E 《Biopolymers》2005,80(2-3):294-302
Aib-rich side-chain lactam-bridged oligomers Ac-(Glu-Aib-Aib-Lys)n-Ala-OH with n = 1,2,3 were designed and synthesized as putative models of the 3(10)-helix. The lactam bridge between the side chains of L-Glu and L-Lys in (i)--(i + 3) positions was introduced in order to enhance the structural preference toward the right-handed 3(10)-helix. The conformational properties of the three peptides were studied in trifluoroethanol (TFE) solution by CD, NMR, and computer simulations. The structural information was derived mainly from the analysis of nuclear Overhauser effect spectroscopy spectra. The presence of alpha H(i)-HN(i + 2) and of alpha H(i)-HN(i + 3) connectivities and the absence of alpha H(i)-HN(i + 4) connectivities indicate that these peptides fold into a 3(10)-helix rather than into an alpha-helix. Based on these conformational features, stereospecific assignment of the Aib methyl groups was possible. The results of such experiments and of the subsequent distance geometry and restrained molecular dynamics simulations reveal a marked preference of these peptides for 3(10)-helix. The CD spectra of these peptides indicate that the helix content increases upon chain elongation. The CD spectrum of the trimer is characterized by a negative band at 200 nm and by a weak positive band around 220 nm. The CD spectrum in TFE is different from that observed in aqueous solution in the presence of SDS micelles, reported in our previous work, and from those reported by a different research group for 3(10)-helical peptides. A possible reason for these differences could rest in the presence of different equilibria of the conformer populations of the various peptides in different solvent systems.  相似文献   

5.
A series of short, amphipathic peptides incorporating 80% C(alpha),C(alpha)-disubstituted glycines has been prepared to investigate amphipathicity as a helix-stabilizing effect. The peptides were designed to adopt 3(10)- or alpha-helices based on amphipathic design of the primary sequence. Characterization by circular dichroism spectroscopy in various media (1 : 1 acetonitrile/water; 9 : 1 acetonitrile/water; 9 : 1 acetonitrile/TFE; 25 mM SDS micelles in water) indicates that the peptides selectively adopt their designed conformation in micellar environments. We speculate that steric effects from ith and ith + 3 residues interactions may destabilize the 3(10)-helix in peptides containing amino acids with large side-chains, as with 1-aminocyclohexane-1-carboxylic acid (Ac(6)c). This problem may be overcome by alternating large and small amino acids in the ith and ith + 3 residues, which are staggered in the 3(10)-helix.  相似文献   

6.
A structural transition from a 3(10)-helix to an alpha-helix has been characterized at high resolution for an octapeptide segment located in 3 different sequences. Three synthetic peptides, decapeptide (A) Boc-Aib-Trp-(Leu-Aib-Ala)2-Phe-Aib-OMe, nonapeptide (B) Boc-Trp-(Leu-Aib-Ala)2-Phe-Aib-OMe, and octapeptide (C) Boc-(Leu-Aib-Ala)2-Phe-Aib-OMe, are completely helical in their respective crystals. At 0.9 A resolution, R factors for A, B, and C are 8.3%, 5.4%, and 7.3%, respectively. The octapeptide and nonapeptide form ideal 3(10)-helices with average torsional angles phi(N-C alpha) and psi(C alpha-C') of -57 degrees, -26 degrees C and -60 degrees, -27 degrees for B. The 10-residue peptide (A) begins as a 3(10)-helix and abruptly changes to an alpha-helix at carbonyl O(3), which is the acceptor for both a 4-->1 hydrogen bond with N(6)H and a 5-->1 hydrogen with N(7)H, even though the last 8 residues have the same sequence in all 3 peptides. The average phi, psi angles in the decapeptide are -58 degrees, -28 degrees for residues 1-3 and -63 degrees, -41 degrees for residues 4-10. The packing of helices in the crystals does not provide any obvious reason for the transition in helix type. Fourier transform infrared studies in the solid state also provide evidence for a 3(10)- to alpha-helix transition with the amide I band appearing at 1,656-1,657 cm-1 in the 9- and 10-residue peptides, whereas in shorter sequences the band is observed at 1,667 cm-1.  相似文献   

7.
M Hatano  M Yoneyama  Y Sato 《Biopolymers》1973,12(4):895-903
Interaction of poly-α,L -glutamic acid (PGLA) with acridine orange (AO) was studied with circular dichroism and absorption spectra measurements. The following results were observed: (1) the addition of a comparable amount of AO with the glutamyl residue to the PLGA solution at pH of 4.5 reduced the fraction of helix of the polymer; (2) when AO was added to the PGLA solution, the pH range of the helix-coil transition of the polymer shifted toward higher pH regions; and (3) when the mixture of the same amount of AO and the glutamyl residue was brought to the neutral and alkaline pH region, some induced circular dichroism bands were observed. In this case, it was assumed that PLGA in the system takes a helical form due to the neutralization of the anionized side chains by the cationic species of AO. We concluded that AO molecules bound to the carboxylate groups of the side chains of PLGA arrange to from a righthanded super-helix which surrounds the core of the righthanded α-helix of PLGA.  相似文献   

8.
The difference in length between the more elongated peptide 3(10)-helix and the more compact alpha-helix is about 0.4 A/residue. This property makes the 3(10)-/alpha-helix reversible conversion very promising as a molecular switching tool between the N- and C-terminal functions of a peptide backbone. In this work, using homo-peptides of various main-chain length, all based on the strongly helicogenic, Calpha-tetrasubstituted alpha-amino acid Calpha-methyl-L-valine, we show that a well defined, solvent controlled, reversible 3(10)-/alpha-helix transition takes place even in a homo-oligomer as short as a terminally blocked hexapeptide. Homo-peptide sequences blocked as a urethane or an acetamide at the N-terminus and as a methyl ester or an N-alkyl amide at the C-terminus are all appropriate. The nature of the occurring helical species in the various solvents tested was assessed by electronic or vibrational circular dichroism.  相似文献   

9.
T Becker  K Weber    N Johnsson 《The EMBO journal》1990,9(13):4207-4213
Annexin II (p36) interacts with its ligand p11 via the short stretch of 12 amino acids (Ac-S-T-V-H-E-I-L-C-K-L-S-L) situated at the N-terminus. We have now synthesized some 37 tetradecapeptides, which differ from the original p11 binding sequence (Ac1-14) by single amino acid substitutions. The relative affinity of each peptide for p11 was determined by fluorescence spectroscopy using a competitive binding assay. The binding behaviour of the different peptides confirms the model of an amphiphilic alpha-helix induced upon binding to p11. The apparent affinities delta delta Gbind of the mutant peptides revealed that the N-acetyl group of serine 1 and the hydrophobic side chains at positions 3, 6, 7 and 10 contribute most to the binding. The observed destabilization of the complex upon removal of signal methyl groups from the hydrophobic side of the helix is comparable with the destabilization of proteins in which methyl groups have been removed from the inner core. We conclude that upon binding to p11 the hydrophobic side of the amphiphatic alpha-helix becomes fully buried.  相似文献   

10.
W C Johnson 《Proteins》1999,35(3):307-312
We have developed an algorithm to analyze the circular dichroism of proteins for secondary structure. Its hallmark is tremendous flexibility in creating the basis set, and it also combines the ideas of many previous workers. We also present a new basis set containing the CD spectra of 22 proteins with secondary structures from high quality X-ray diffraction data. High flexibility is obtained by doing the analysis with a variable selection basis set of only eight proteins. Many variable selection basis sets fail to give a good analysis, but good analyses can be selected without any a priori knowledge by using the following criteria: (1) the sum of secondary structures should be close to 1.0, (2) no fraction of secondary structure should be less than -0.03, (3) the reconstructed CD spectrum should fit the original CD spectrum with only a small error, and (4) the fraction of alpha-helix should be similar to that obtained using all the proteins in the basis set. This algorithm gives a root mean square error for the predicted secondary structure for the proteins in the basis set of 3.3% for alpha-helix, 2.6% for 3(10)-helix, 4.2% for beta-strand, 4.2% for beta-turn, 2.7% for poly(L-proline) II type 3(1)-helix, and 5.1% for other structures when compared with the X-ray structure.  相似文献   

11.
Models for the 3(10)-helix/coil and pi-helix/coil equilibria have been derived. The theory is based on classifying residues into helical or nonhelical (coil) conformations. Statistical weights are assigned to residues in a helical conformation with an associated helical hydrogen bond, a helical conformation with no hydrogen bond, an N-cap position, a C-cap position, or the reference coil conformation. The models for alpha-helix formation and 3(10)-helix formation have also been combined to describe a three-state equilibrium in which alpha-helical, 3(10)-helical, and coil conformations are populated. The results are compared with the modified Lifson-Roig theory for the alpha-helix/coil equilibrium. The comparison accounts for the experimental observations that 3(10)-helices tend to be short and pi-helices are not favored for any length. This work may provide a framework for quantitatively rationalizing experimental work on isolated 3(10)-helices and mixed 3(10)-/alpha-helices.  相似文献   

12.
Structural properties of signal peptides and their membrane insertion   总被引:5,自引:0,他引:5  
Garnier J  Gaye P  Mercier JC  Robson B 《Biochimie》1980,62(4):231-239
Structural properties of the amino acid sequences from 22 signal peptides have been analyzed and compared with peptides known to interact with biological membranes and liposomes, melittin, a lytic peptide of bee venom, and the non-polar C-terminal segment of cytochrome b5. All these peptides evidence a double amphipatic structure with an hydrophobic core of 9 to 24 amino acid residues and two charged polar ends. They all exhibit a high potential for making alpha-helix and, to a lesser degree, extended or beta-sheet conformation with low or negative potentials for making reverse turns or aperiodic conformation. A model of spontaneous insertion of these peptides into the lipid bilayer without specific surface receptor protein is proposed, where the two polar ends interact with each polar face of the lipid bilayer and the hydrophobic core inserts into the non-hydrogen bonding environment of the fatty acid side chains. This insertion could be the molecular trigger for ribophorin assembly around the signal peptide and subsequent attachment to the ribosome prior to the transfer of the polypeptide chain through the endoplasmic reticulum membrane.  相似文献   

13.
Xiong K  Punihaole D  Asher SA 《Biochemistry》2012,51(29):5822-5830
We utilize 198 and 204 nm excited UV resonance Raman spectroscopy (UVRR) and circular dichroism spectroscopy (CD) to monitor the backbone conformation and the Gln side chain hydrogen bonding (HB) of a short, mainly polyGln peptide with a D(2)Q(10)K(2) sequence (Q10). We measured the UVRR spectra of valeramide to determine the dependence of the primary amide vibrations on amide HB. We observe that a nondisaggregated Q10 (NDQ10) solution (prepared by directly dissolving the original synthesized peptide in pure water) exists in a β-sheet conformation, where the Gln side chains form hydrogen bonds to either the backbone or other Gln side chains. At 60 °C, these solutions readily form amyloid fibrils. We used the polyGln disaggregation protocol of Wetzel et al. [Wetzel, R., et al. (2006) Methods Enzymol.413, 34-74] to dissolve the Q10 β-sheet aggregates. We observe that the disaggregated Q10 (DQ10) solutions adopt PPII-like and 2.5(1)-helix conformations where the Gln side chains form hydrogen bonds with water. In contrast, these samples do not form fibrils. The NDQ10 β-sheet solution structure is essentially identical to that found in the NDQ10 solid formed upon evaporation of the solution. The DQ10 PPII and 2.5(1)-helix solution structure is essentially identical to that in the DQ10 solid. Although the NDQ10 solution readily forms fibrils when heated, the DQ10 solution does not form fibrils unless seeded with the NDQ10 solution. This result demonstrates very high activation barriers between these solution conformations. The NDQ10 fibril secondary structure is essentially identical to that of the NDQ10 solution, except that the NDQ10 fibril backbone conformational distribution is narrower than in the dissolved species. The NDQ10 fibril Gln side chain geometry is more constrained than when NDQ10 is in solution. The NDQ10 fibril structure is identical to that of the DQ10 fibril seeded by the NDQ10 solution.  相似文献   

14.
A new approach for site-directed placement of nitroxide spin labels in chemically synthesized peptides and proteins is described. The scheme takes advantage of a novel diaminopropionic acid scaffold to independently control backbone and side chain elongation. The result is a spin-labeled side chain, referred to as Dap-SL, in which an amide bond forms a linker between the nitroxide and the peptide backbone. The method was demonstrated in a series of helical peptides. Circular dichroism and nuclear magnetic resonance showed that Dap-SL introduces only a minor perturbation in the helical structure. The electron paramagnetic resonance spectrum of the singly labeled species allowed for determination of the spin label rotational correlation time and suggests that the Dap-SL side chain is more flexible than the modified Cys side chain frequently used in site-directed spin label studies. Spectra of the doubly labeled peptides indicate a mixture of 3(10)-helix and alpha-helix, which parallels findings from previous studies. The scheme demonstrated here offers a fundamentally new approach for introducing spin labels into proteins and promises to significantly extend biophysical investigations of large proteins and receptors. In addition, the technique is readily modified for incorporation of any biophysical probe.  相似文献   

15.
We test molecular level hypotheses for the high thermal stability of alpha-helical conformations of alanine-based peptides by performing detailed atomistic simulations of a 20-amino-acid peptide with explicit treatment of water. To assess the contribution of large side chains to alpha-helix stability through backbone desolvation and salt-bridge formation, we simulate the alanine-rich peptide, Ac-YAEAAKAAEAAKAAEAAKAF-Nme, referred to as the EK peptide, that has three pairs of "i, i + 3" glutamic acid(-) and lysine(+) substitutions. Efficient configurational sampling of the EK peptide over a wide temperature range enabled by the replica exchange molecular dynamics technique allows characterization of the stability of alpha-helix with respect to heat-induced unfolding. We find that near ambient temperatures, the EK peptide predominately samples alpha-helical configurations with 80% fractional helicity at 300 K. The helix melts over a broad range of temperatures with melting temperature, T(m), equal to 350 K, that is significantly higher than the T(m) of a 21-residue polyalanine peptide, A(21). Salt-bridges between oppositely charged Glu(-) and Lys(+) side chains can, in principle, provide thermal stability to alpha-helical conformers. For the specific EK peptide sequence, we observe infrequent formation of Glu-Lys salt-bridges (with approximately 10-20% probability) and therefore we conclude that salt-bridge formation does not contribute significantly to the EK peptide's helical stability. However, lysine side chains are found to shield specific "i, i + 4" backbone hydrogen bonds from water, indicating that large side-chain substituents can play an important role in stabilizing alpha-helical configurations of short peptides in aqueous solution through mediation of water access to backbone hydrogen bonds. These observations have implications on molecular engineering of peptides and biomolecules in the design of their thermostable variants where the shielding mechanism can act in concert with other factors such as salt-bridge formation, thereby increasing thermal stability considerably.  相似文献   

16.
The secondary structure of the catalytic domain from protein kinase C zeta was studied using IR spectroscopy. In the presence of the substrate MgATP, there was a significant change in the secondary structure. After heating to 80 degrees C, a 14% decrease in the alpha-helix component was observed, accompanied by a 6% decrease in the beta-pleated sheet; no change was observed in the large loops or in 3(10)-helix plus associated loops. The maximum increase with heating was observed in the aggregated beta-sheet component, with an increase of 14%. In the presence of MgATP, and compared with the sample heated in its absence, there was a substantial decrease in the 3(10)-helix plus associated loops and an increase in alpha-helix. Synchronous 2D-IR correlation showed that the main changes occurred at 1617 cm(-1), which was assigned to changes in the intermolecular aggregated beta-sheet of the denaturated protein. This increase was mainly correlated with the change in alpha-helix. In the presence of MgATP, the main correlation was between aggregated beta-sheet and the large loops component. The asynchronous 2D-correlation spectrum indicated that a number of components are transformed in intermolecularly aggregated beta-sheet, especially the alpha-helix and beta-sheet components. It is interesting that changes in 3(10)-helix plus associated loops and in alpha-helix preceded changes in large loops, which suggests that the open loops structure exists as an intermediate state during denaturation. In summary, IR spectroscopy revealed an important effect of MgATP on the secondary structure and on the thermal unfolding process when this was induced, whereas 2D-IR correlation spectroscopy allowed us to show the establishment of the denaturation pathway of this protein.  相似文献   

17.
All known naturally occurring linear cationic peptides adopt an amphipathic alpha-helical conformation upon binding to lipids as an initial step in the induction of cell leakage. We designed an 18-residue peptide, (KIGAKI)3-NH2, that has no amphipathic character as an alpha-helix but can form a highly amphipathic beta-sheet. When bound to lipids, (KIGAKI)3-NH2 did indeed form a beta-sheet structure as evidenced by Fourier transform infrared and circular dichroism spectroscopy. The antimicrobial activity of this peptide was compared with that of (KIAGKIA)3-NH2, and it was better than that of GMASKAGAIAGKIAKVALKAL-NH2 (PGLa) and (KLAGLAK)3-NH2, all of which form amphipathic alpha-helices when bound to membranes. (KIGAKI)3-NH2 was much less effective at inducing leakage in lipid vesicles composed of mixtures of the acidic lipid, phosphatidylglycerol, and the neutral lipid, phosphatidylcholine, as compared with the other peptides. However, when phosphatidylethanolamine replaced phosphatidylcholine, the lytic potency of PGLa and the alpha-helical model peptides was reduced, whereas that of (KIGAKI)3-NH2 was improved. Fluorescence experiments using analogs containing a single tryptophan residue showed significant differences between (KIGAKI)3-NH2 and the alpha-helical peptides in their interactions with lipid vesicles. Because the data suggest enhanced selectivity between bacterial and mammalian lipids, linear amphipathic beta-sheet peptides such as (KIGAKI)3-NH2 warrant further investigation as potential antimicrobial agents.  相似文献   

18.
Relevant parameters and stereochemical consequences of helices [alpha-helix, 3(10)-helix, beta-bend ribbon spiral, gamma-helix, 2.0(5)-helix, poly(Pro)(n) type-I and -II helices, and collagen triple helix] of peptides based on alpha-amino acids for use as templates in various branches of chemistry are briefly discussed.  相似文献   

19.
Kimura S  Naito A  Tuzi S  Saitô H 《Biopolymers》2001,58(1):78-88
We have recorded (13)C NMR spectra of selectively [3-(13)C]Ala-, [1-(13)C]Ala-, or [1-(13)C]Val-labeled synthetic transmembrane peptides of bacteriorhodopsin (bR) and enzymatically cleaved C-2 fragment in the solid and dimyristoylphosphatidylcholine bilayer. It turned out that these transmembrane peptides either in hexafluoroisopropanol or cast from it take an ordinary alpha-helix (alpha(I)-helix) irrespective of their amino acid sequences with reference to the conformation-dependent (13)C chemical shifts of (Ala)(n) taking the alpha-helix form. These transmembrane peptides are not always static in the lipid bilayer as in the solid state but undergo rigid-body motions with various frequencies as estimated from suppressed peaks either by fast isotropic or large-amplitude motions (>10(8) Hz) or intermediate frequencies (10(5) or 10(3) Hz). Further, (13)C chemical shifts of the [3-(13)C]Ala-labeled peptides in the bilayer were displaced downfield by 0.3-1.1 ppm depending upon amino acid sequence with respect to those in the solid state, which were explained in terms of local conformational fluctuation (10(2) Hz) deviated from the torsion angles (alpha(II)-helix) from those of standard alpha-helix, under anisotropic environment in lipid bilayer, in addition to the above-mentioned rigid-body motions. The carbonyl (13)C peaks, on the other hand, are not sensitively displaced by such local anisotropic fluctuations, because they are more sensitive to the manner of hydrogen-bond interactions. The amino acid sequences of these peptides inserted within the bilayer were not always the same as those of intact bR, causing disposition of the transmembrane alpha-helical segment from that of intact bR. Finally, we confirmed that the (13)C NMR peak positions of the random coil form are located at the boundary between the alpha-helix and a turned structure in loop regions.  相似文献   

20.
Chugh JK  Brückner H  Wallace BA 《Biochemistry》2002,41(43):12934-12941
Trichotoxin_A50E is an 18-residue peptaibol antibiotic which forms multimeric transmembrane channels through self-association. The crystal structure of trichotoxin has been determined at a resolution of 0.9 A. The trichotoxin sequence contains nine helix-promoting Aib residues, which contribute to the formation of an entirely helical structure that has a central bend of 8-10 degrees located between residues 10-13. Trichotoxin is the first solved structure of the peptaibol family that is all alpha-helix as opposed to containing part or all 3(10)-helix. Gln residues in positions 6 and 17 produce a polar face, and are proposed to form the channel lumen. An octameric model channel has been constructed from the crystal structure. It has a central pore of approximately 4-5 A radius, a size sufficient to enable transport of ions, with a constricted region at one end, formed by a ring of Gln6 residues. Electrostatic calculations are consistent with it being a cationic channel.  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