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.
Equine β-lactoglobulin (ELG) assumes non-native helices during refolding and in partially folded states. Previously, circular dichroism (CD) combined with site-directed mutagenesis identified helical regions in the acid- and cold-denatured states of ELG. It is also known that a fragment of ELG, CHIBL (residues 88-142), has a structure similar to that of the cold-denatured state. For the study reported herein, the structure of a shorter fragment, CHIBLΔF (residues 97-142), was investigated by CD and nuclear magnetic resonance spectroscopy. The secondary chemical shifts clearly showed that non-native α-helices are present in two different regions, residues 98-107 and 114-135, and are connected by a native disulfide bond. The CD spectra of two peptides that correspond to the helical regions are characterized by weak helical signatures, and the sum of their CD spectra is nearly the same as the spectrum of disulfide-reduced CHIBLΔF. Therefore, the non-native helices are stabilized by the disulfide, and non-native helix formation may occur only during the refolding of the disulfide-intact protein. Supporting this conclusion is the observation that tear lipocalin, a homologue of ELG that lacks the disulfide, does not form non-native helices during folding.  相似文献   

2.
Apomyoglobin folds by a sequential mechanism in which the A, G, and H helix regions undergo rapid collapse to form a compact intermediate onto which the central portion of the B helix subsequently docks. To investigate the factors that frustrate folding, we have made mutations in the N-terminus of the B helix to stabilize helical structure (in the mutant G23A/G25A) and to promote native-like hydrophobic packing interactions with helix G (in the mutant H24L/H119F). The kinetic and equilibrium intermediates of G23A/G25A and H24L/H119F were studied by hydrogen exchange pulse labeling and interrupted hydrogen/deuterium exchange combined with NMR. For both mutants, stabilization of helical structure in the N-terminal region of the B helix is confirmed by increased exchange protection in the equilibrium molten globule states near pH 4. Increased protection is also observed in the GH turn region in the G23A/G25A mutant, suggesting that stabilization of the B helix facilitates native-like interactions with the C-terminal region of helix G. These interactions are further enhanced in H24L/H119F. The kinetic burst phase intermediates of both mutants show increased protection, relative to wild-type protein, of amides in the N-terminus of the B helix and in part of the E helix. Stabilization of the E helix in the intermediate is attributed to direct interactions between E helix residues and the newly stabilized N-terminus of helix B. Stabilization of native packing between the B and G helices in H24L/H119F also favors formation of native-like interactions in the GH turn and between the G and H helices in the ensemble of burst phase intermediates. We conclude that instability at the N-terminus of the B helix of apomyoglobin contributes to the energetic frustration of folding by preventing docking and stabilization of the E helix.  相似文献   

3.
The response regulator DosR is essential for promoting long-term survival of Mycobacterium tuberculosis under low oxygen conditions in a dormant state and may be responsible for latent tuberculosis in one-third of the world's population. Here, we report crystal structures of full-length unphosphorylated DosR at 2.2 Å resolution and its C-terminal DNA-binding domain at 1.7 Å resolution. The full-length DosR structure reveals several features never seen before in other response regulators. The N-terminal domain of the full-length DosR structure has an unexpected (βα)4 topology instead of the canonical (βα)5 fold observed in other response regulators. The linker region adopts a unique conformation that contains two helices forming a four-helix bundle with two helices from another subunit, resulting in dimer formation. The C-terminal domain in the full-length DosR structure displays a novel location of helix α10, which allows Gln199 to interact with the catalytic Asp54 residue of the N-terminal domain. In contrast, the structure of the DosR C-terminal domain alone displays a remarkable unstructured conformation for helix α10 residues, different from the well-defined helical conformations in all other known structures, indicating considerable flexibility within the C-terminal domain. Our structures suggest a mode of DosR activation by phosphorylation via a helix rearrangement mechanism.  相似文献   

4.
Kinetic and equilibrium studies of apomyoglobin folding pathways and intermediates have provided important insights into the mechanism of protein folding. To investigate the role of intrinsic helical propensities in the apomyoglobin folding process, a mutant has been prepared in which Asn132 and Glu136 have been substituted with glycine to destabilize the H helix. The structure and dynamics of the equilibrium molten globule state formed at pH 4.1 have been examined using NMR spectroscopy. Deviations of backbone (13)C(alpha) and (13)CO chemical shifts from random coil values reveal high populations of helical structure in the A and G helix regions and in part of the B helix. However, the H helix is significantly destabilized compared to the wild-type molten globule. Heteronuclear [(1)H]-(15)N NOEs show that, although the polypeptide backbone in the H helix region is more flexible than in the wild-type protein, its motions are restricted by transient hydrophobic interactions with the molten globule core. Quench flow hydrogen exchange measurements reveal stable helical structure in the A and G helices and part of the B helix in the burst phase kinetic intermediate and confirm that the H helix is largely unstructured. Stabilization of structure in the H helix occurs during the slow folding phases, in synchrony with the C and E helices and the CD region. The kinetic and equilibrium molten globule intermediates formed by N132G/E136G are similar in structure. Although both the wild-type apomyoglobin and the mutant fold via compact helical intermediates, the structures of the intermediates and consequently the detailed folding pathways differ. Apomyoglobin is therefore capable of compensating for mutations by using alternative folding pathways within a common basic framework. Tertiary hydrophobic interactions appear to play an important role in the formation and stabilization of secondary structure in the H helix of the N132G/E136G mutant. These studies provide important insights into the interplay between secondary and tertiary structure formation in protein folding.  相似文献   

5.
An understanding of the structural transitions that an alpha-helix undergoes will help to elucidate such motions in proteins and their role in protein folding. We present the results of molecular dynamics simulations to investigate these transitions in a short polyalanine peptide (13 residues) both in vacuo and in the presence of solvent. The denaturation of this peptide was monitored as a function of temperature (ranging from 5 to 200 degrees C). In vacuo, the helical state predominated at all temperatures, whereas in solution the helix melted with increasing temperature. The peptide was predominantly helical at low temperature in solution, while at intermediate temperatures the peptide spent the bulk of the time fluctuating between different conformations with intermediate amounts of helix, e.g. not completely helical nor entirely non-helical. Many of these conformations consisted of short helical segments with intervening non-helical residues. At high temperature the peptide unfolded and adopted various collapsed unstructured states. The intrahelical hydrogen bonds that break at high temperature were not fully compensated by hydrogen bonds with water molecules in the partially unfolded forms of the peptide. Increases in temperature disrupted both the helical structure and the peptide-water interactions. Water played a major but indirect role in facilitating unfolding, as opposed to specifically competing for the intrapeptide hydrogen bonds. The implications of our results to protein folding are discussed.  相似文献   

6.
A low resolution model has been proposed for the exofacial conformation of the Glut1 glucose transporter in which eight transmembrane segments form an inner helical bundle stabilized by four outer helices. The role of transmembrane segment 4, predicted to be an inner helix in this structural model, was investigated by cysteine-scanning mutagenesis in conjunction with the substituted cysteine accessibility method using the membrane-impermeant, sulfhydryl-specific reagent, p-chloromercuribenzenesulfonate (pCMBS). A functional, cysteine-less, parental Glut1 molecule was used to produce 21 Glut1 point mutants by individually changing each residue along transmembrane helix 4 to a cysteine. The single cysteine mutants were then expressed in Xenopus oocytes, and their expression levels, transport activities, and sensitivities to pCMBS were determined. In striking contrast to all of the other seven predicted inner helices, none of the 21 helix 4 single-cysteine mutants was demonstrably inhibited by pCMBS. However, cysteine substitution within helix 4 resulted in an unusually high number of severely transport-defective mutants. The low absolute transport activities of two of these mutants (G130C and G134C) were due to their extremely low levels of expression, presumably a result of structural instability and consequent degradation in oocytes, suggesting that these two residues play an important role in maintaining the native structure of Glut1. The other two transport-defective mutants (Y143C and E146C) exhibited low specific transport activities, implying that these two residues play an important role in the transport cycle. Based on these data, we conclude that the exoplasmic end of helix 4 lies outside the inner helical bundle in the exofacial configuration of Glut1.  相似文献   

7.
The structural changes of bacteriorhodopsin during its photochemical cycle, as revealed by crystal structures of trapped intermediates, have provided insights to the proton translocation mechanism. Because accumulation of the last photointermediate, O, appears to be hindered by lattice forces in the crystals, the only information about the structure of this state is from suggested analogies with the determined structures of the non-illuminated D85S mutant and wild-type bacteriorhodopsin at low pH. We used electron paramagnetic resonance spectroscopy of site-directed spin labels at the extracellular protein surface in membranes to test these models. Spin-spin dipolar interactions in the authentic O state compared to the non-illuminated state revealed that the distance between helices C and F increases by ca 4 Angstroms, there is no distance change between helices D and F, and the distance between helix D and helix B of the adjacent monomer increases. Further, the mobility changes of single labels indicate that helices E and F move outward from the proton channel at the center of the protein, and helix D tilts inward. The overall pattern of movements suggests that the model at acid pH is a better representation of the O state than D85S. However, the mobility analysis of spin-labels on the B-C interhelical loop indicates that the antiparallel beta-sheet maintains its ordered secondary structure in O, instead of the predicted disorder in the two structural models. During decay of the O state, the last step of the photocycle, a proton is transferred from Asp85 to proton release complex in the extracellular proton channel. The structural changes in O suggest the need of large conformational changes to drive the Arg82 side-chain back to its initial orientation towards Asp85, and to rearrange the numerous water molecules in this region in order to conduct the proton away from Asp85.  相似文献   

8.
Scott KA  Alonso DO  Pan Y  Daggett V 《Biochemistry》2006,45(13):4153-4163
Molecular dynamics simulations can be used to reveal the detailed conformational behaviors of peptides and proteins. By comparing fragment and full-length protein simulations, we can investigate the role of each peptide segment in the folding process. Here, we take advantage of information regarding the helix formation process from our previous simulations of barnase and protein A as well as new simulations of four helical fragments from these proteins at three different temperatures, starting with both helical and extended structures. Segments with high helical propensity began the folding process by tethering the chain through side chain interactions involving either polar interactions, such as salt bridges, or hydrophobic staples. These tethers were frequently nonnative (i.e., not i --> i + 4 spacing) and provided a scaffold for other residues, thereby limiting the conformational search. The helical structure then propagated on both sides of the tether. Segments with low stability and propensity formed later in the folding process and utilized contacts with other portions of the protein when folding. These helices formed via a tertiary contact-assisted mechanism, primarily via hydrophobic contacts between residues distant in sequence. Thus, segments with different helical propensities appear to play different roles during protein folding. Furthermore, the active role of nonlocal side chains in helix formation highlights why we must move beyond simple hierarchical models of protein folding.  相似文献   

9.
We developed a method of predicting the tertiary structures of seven transmembrane helical proteins in triangle lattice models, assuming that the configuration of helices is stabilized by polar interactions. Triangle lattice models having 12 or 11 nearest neighbor pairs were used as general templates of a seven-helix system, then the orientation angles of all helices were varied at intervals of 15 degrees. The polar interaction energy for all possible positions of each helix was estimated using the calculated polar indices of transmembrane helices. An automated system was constructed and applied to bacteriorhodopsin, a typical membrane protein with seven transmembrane helices. The predicted optimal and actual structures were similar. The top 100 predicted helical configurations indicated that the helix-triangle, CFG, occurred at the highest frequency. In fact, this helix-triangle of bacteriorhodopsin forms an active proton-pumping site, suggesting that the present method can identify functionally important helices in membrane proteins. The possibility of studying the structure change of bacteriorhodopsin during the functional process by this method is discussed, and may serve to explain the experimental structures of photointermediate states.  相似文献   

10.
Eliezer D  Chung J  Dyson HJ  Wright PE 《Biochemistry》2000,39(11):2894-2901
The partly folded state of apomyoglobin at pH 4 represents an excellent model for an obligatory kinetic folding intermediate. The structure and dynamics of this intermediate state have been extensively examined using NMR spectroscopy. Secondary chemical shifts, (1)H-(1)H NOEs, and amide proton temperature coefficients have been used to probe residual structure in the intermediate state, and NMR relaxation parameters T(1) and T(2) and ?(1)H?-(15)N NOE have been analyzed using spectral densities to correlate motion of the polypeptide chain with these structural observations. A significant amount of helical structure remains in the pH 4 state, indicated by the secondary chemical shifts of the (13)C(alpha), (13)CO, (1)H(alpha), and (13)C(beta) nuclei, and the boundaries of this helical structure are confirmed by the locations of (1)H-(1)H NOEs. Hydrogen bonding in the structured regions is predominantly native-like according to the amide proton chemical shifts and their temperature dependence. The locations of the A, G, and H helix segments and the C-terminal part of the B helix are similar to those in native apomyoglobin, consistent with the early, complete protection of the amides of residues in these helices in quench-flow experiments. These results confirm the similarity of the equilibrium form of apoMb at pH 4 and the kinetic intermediate observed at short times in the quench-flow experiment. Flexibility in this structured core is severely curtailed compared with the remainder of the protein, as indicated by the analysis of the NMR relaxation parameters. Regions with relatively high values of J(0) and low values of J(750) correspond well with the A, B, G, and H helices, an indication that nanosecond time scale backbone fluctuations in these regions of the sequence are restricted. Other parts of the protein show much greater flexibility and much reduced secondary chemical shifts. Nevertheless, several regions show evidence of the beginnings of helical structure, including stretches encompassing the C helix-CD loop, the boundary of the D and E helices, and the C-terminal half of the E helix. These regions are clearly not well-structured in the pH 4 state, unlike the A, B, G, and H helices, which form a native-like structured core. However, the proximity of this structured core most likely influences the region between the B and F helices, inducing at least transient helical structure.  相似文献   

11.
RNA junctions are secondary-structure elements formed when three or more helices come together. They are present in diverse RNA molecules with various fundamental functions in the cell. To better understand the intricate architecture of three-dimensional (3D) RNAs, we analyze currently solved 3D RNA junctions in terms of base-pair interactions and 3D configurations. First, we study base-pair interaction diagrams for solved RNA junctions with 5 to 10 helices and discuss common features. Second, we compare these higher-order junctions to those containing 3 or 4 helices and identify global motif patterns such as coaxial stacking and parallel and perpendicular helical configurations. These analyses show that higher-order junctions organize their helical components in parallel and helical configurations similar to lower-order junctions. Their sub-junctions also resemble local helical configurations found in three- and four-way junctions and are stabilized by similar long-range interaction preferences such as A-minor interactions. Furthermore, loop regions within junctions are high in adenine but low in cytosine, and in agreement with previous studies, we suggest that coaxial stacking between helices likely forms when the common single-stranded loop is small in size; however, other factors such as stacking interactions involving noncanonical base pairs and proteins can greatly determine or disrupt coaxial stacking. Finally, we introduce the ribo-base interactions: when combined with the along-groove packing motif, these ribo-base interactions form novel motifs involved in perpendicular helix-helix interactions. Overall, these analyses suggest recurrent tertiary motifs that stabilize junction architecture, pack helices, and help form helical configurations that occur as sub-elements of larger junction networks. The frequent occurrence of similar helical motifs suggest nature's finite and perhaps limited repertoire of RNA helical conformation preferences. More generally, studies of RNA junctions and tertiary building blocks can ultimately help in the difficult task of RNA 3D structure prediction.  相似文献   

12.
Proline residues occur frequently in transmembrane alpha helices, which contrasts with their behaviour as helix-breakers in water-soluble proteins. The three membrane-embedded proline residues of bacteriorhodopsin have been replaced individually by alanine and glycine to give P50A, or P50G on helix B, P91A, or P91G on helix C, and P186A or P186G on helix F, and the effect on the protein folding kinetics has been investigated. The rate-limiting apoprotein folding step, which results in formation of a seven transmembrane, alpha helical state, was slower than wild-type protein for the Pro50 and Pro91 mutants, regardless of whether they were mutated to Ala or Gly. These proline residues give rise to several inter-helix contacts, which are therefore important in folding to the seven transmembrane helix state. No evidence for cis-trans isomerisations of the peptidyl prolyl bonds was found during this rate-limiting apoprotein folding step. Mutations of all three membrane-embedded proline residues affected the subsequent retinal binding and final folding to bacteriorhodopsin, suggesting that these proline residues contribute to formation of the retinal binding pocket within the helix bundle, again via helix/helix interactions. These results point to proline residues in transmembrane alpha helices being important in the folding of integral membrane proteins. The helix/helix interactions and hydrogen bonds that arise from the presence of proline residues in transmembrane alpha helices can affect the formation of transmembrane alpha helix bundles as well as cofactor binding pockets.  相似文献   

13.
The four-helical immunity protein Im7 folds through an on-pathway intermediate that has a specific, but partially misfolded, hydrophobic core. In order to gain further insight into the structure of this species, we have identified the backbone hydrogen bonds formed in the ensemble by measuring the amide exchange rates (under EX2 conditions) of the wild-type protein and a variant, I72V. In this mutant the intermediate is significantly destabilised relative to the unfolded state (deltadeltaG(ui) = 4.4 kJ/mol) but the native state is only slightly destabilised (deltadeltaG(nu) = 1.8 kJ/mol) at 10 degrees C in 2H2O, pH* 7.0 containing 0.4 M Na2SO4, consistent with the view that this residue forms significant non-native stabilising interactions in the intermediate state. Comparison of the hydrogen exchange rates of the two proteins, therefore, enables the state from which hydrogen exchange occurs to be identified. The data show that amides in helices I, II and IV in both proteins exchange slowly with a free energy similar to that associated with global unfolding, suggesting that these helices form highly protected hydrogen-bonded helical structure in the intermediate. By contrast, amides in helix III exchange rapidly in both proteins. Importantly, the rate of exchange of amides in helix III are slowed substantially in the Im7* variant, I72V, compared with the wild-type protein, whilst other amides exchange more rapidly in the mutant protein, in accord with the kinetics of folding/unfolding measured using chevron analysis. These data demonstrate, therefore, that local fluctuations do not dominate the exchange mechanism and confirm that helix III does not form stable secondary structure in the intermediate. By combining these results with previously obtained Phi-values, we show that the on-pathway folding intermediate of Im7 contains extensive, stable hydrogen-bonded structure in helices I, II and IV, and that this structure is stabilised by both native and non-native interactions involving amino acid side-chains in these helices.  相似文献   

14.
Im7 folds via an on‐pathway intermediate that contains three of the four native α‐helices. The missing helix, helix III, is the shortest and its failure to be formed until late in the pathway is related to frustration in the structure. Im7H3M3, a 94‐residue variant of the 87‐residue Im7 in which helix III is the longest of the four native helices, also folds via an intermediate. To investigate the structural basis for this we calculated the frustration in the structure of Im7H3M3 and used NMR to investigate its dynamics. We found that the native state of Im7H3M3 is highly frustrated and in equilibrium with an intermediate state that lacks helix III, similar to Im7. Model‐free analysis identified residues with chemical exchange contributions to their relaxation that aligned with the residues predicted to have highly frustrated interactions, also like Im7. Finally, we determined properties of urea‐denatured Im7H3M3 and identified four clusters of interacting residues that corresponded to the α‐helices of the native protein. In Im7 the cluster sizes were related to the lengths of the α‐helices with cluster III being the smallest but in Im7H3M3 cluster III was also the smallest, despite this region forming the longest helix in the native state. These results suggest that the conformational properties of the urea‐denatured states promote formation of a three‐helix intermediate in which the residues that form helix III remain non‐helical. Thus it appears that features of the native structure are formed early in folding linked to collapse of the unfolded state.  相似文献   

15.
Site-directed mutagenesis has been used to probe the interactions that stabilize the equilibrium and burst phase kinetic intermediates formed by apomyoglobin. Nine bulky hydrophobic residues in the A, E, G and H helices were replaced by alanine, and the effects on protein stability and kinetic folding pathways were determined. Hydrogen exchange pulse-labeling experiments, with NMR detection, were performed for all mutants. All of the alanine substitutions resulted in changes in proton occupancy or an increased rate of hydrogen-deuterium exchange for amides in the immediate vicinity of the mutation. In addition, most mutations affected residues in distant parts of the amino acid sequence, providing insights into the topology of the burst phase intermediate and the interactions that stabilize its structure. Differences between the pH 4 equilibrium molten globule and the kinetic intermediate are evident: the E helix region plays no discernible role in the equilibrium intermediate, but contributes significantly to stabilization of the ensemble of compact intermediates formed during kinetic refolding. Mutations that interfere with docking of the E helix onto the preformed A/B/G/H helix core substantially decrease the folding rate, indicating that docking and folding of the E helix region occurs prior to formation of the apomyoglobin folding transition state. The results of the mutagenesis experiments are consistent with rapid formation of an ensemble of compact burst phase intermediates with an overall native-like topological arrangement of the A, B, E, G, and H helices. However, the experiments also point to disorder in docking of the E helix and to non-native contacts in the kinetic intermediate. In particular, there is evidence for translocation of the H helix by approximately one helical turn towards its N terminus to maximize hydrophobic interactions with helix G. Thus, the burst phase intermediate observed during kinetic refolding of apomyoglobin consists of an ensemble of compact, kinetically trapped states in which the helix docking appears to be topologically correct, but in which there are local non-native interactions that must be resolved before the protein can fold to the native structure.  相似文献   

16.
Amphipathic helices in membrane proteins that interact with the hydrophobic/hydrophilic interface of the lipid bilayer have been difficult to structurally characterize. Here, the backbone structure and orientation of an amphipathic helix in the full-length M2 protein from influenza A virus has been characterized. The protein has been studied in hydrated DMPC/DMPG lipid bilayers above the gel to liquid-crystalline phase transition temperature by solid-state NMR spectroscopy. Characteristic PISA (Polar Index Slant Angle) wheels reflecting helical wheels have been observed in uniformly aligned bilayer preparations of both uniformly 15N labeled and amino acid specific labeled M2 samples. Hydrogen/deuterium exchange studies have shown the very slow exchange of some residues in the amphipathic helix and more rapid exchange for the transmembrane helix. These latter results clearly suggest the presence of an aqueous pore. A variation in exchange rate about the transmembrane helical axis provides additional support for this claim and suggests that motions occur about the helical axes in this tetramer to expose the entire backbone to the pore.  相似文献   

17.
Electrostatic interactions are often critical for determining the specificity of protein-protein complexes. To study the role of electrostatic interactions for assembly of helical bundles, we previously designed a thermostable, heterotrimeric coiled coil, ABC, in which charged residues were employed to drive preferential association of three distinct, 34-residue helices. To investigate the basis for heterotrimer specificity, we have used multiwavelength anomalous diffraction (MAD) analysis to determine the 1.8 A resolution crystal structure of ABC. The structure shows that ABC forms a heterotrimeric coiled coil with the intended arrangement of parallel chains. Over half of the ion pairs engineered to restrict helix associations were apparent in the experimental electron density map. As seen in other trimeric coiled coils, ABC displays acute knobs-into-holes packing and a buried anion coordinated by core polar amino acids. These interactions validate the design strategy and illustrate how packing and polar contacts determine structural uniqueness.  相似文献   

18.
The human copper transporter 1 (hCtr1) mediates cellular uptake of copper and Pt‐based chemotherapeutic anticancer drugs. In this paper, we determined the three‐dimensional structure and oligomerization of the transmembrane domains (TMDs) of hCtr1 in 40% HFIP aqueous solution by using solution‐state NMR spectroscopy. We firstly revealed that TMD1 forms an α‐helical structure from Gly67 to Glu84 and is dimerized by close packing of its C‐terminal helix; TMD2 forms an α‐helical structure from Leu134 to Thr155 and is self‐associated as a trimer by the hydrophobic contact of TMD2 monomers; TMD3 adopts a discontinuous helix structure, known as ‘α‐helix‐coiled segment‐α‐helix’, and is dimerized by the interaction between the N‐terminal helices. The motif GxxxG in TMD3 is not fully involved in the helix, but partially unstructured as a linker between helices. The flexible linker of TMD3 may serve as a gating adapter to mediate pore on and off switch. The differences in the structure and aggregation of the TMD peptides may be related to their different roles in the channel formation and transport function. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Apolipoprotein A‐I (apoA‐I), which constitutes the principal protein component of high‐density lipoprotein, is responsible for its major antiatherogenic functions. Aiming at contributing to the development of potent inhibitors of low‐density lipoprotein (LDL) peptide models of helices 4,6 and 9,10 of apoA‐I were designed and synthesized. Specific amino acid substitutions, resulting in transformation of the original helix class A and Y to G according to the Schiffer and Edmundson helical wheel representation, were introduced in order to validate the contribution of these modifications in the inhibitory activity of the synthesized peptide models against the LDL oxidation. The role of Met at positions 112 (helix 4) and 148 (helix 6) as oxidant scavenger was also investigated. The helical characteristics of all the peptide models were studied by CD in membrane‐mimicking microenvironments and compared with the original helices. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
A 34‐residue α/β peptide [IG(28–61)], derived from the C‐terminal part of the B3 domain of the immunoglobulin binding protein G from Streptoccocus, was studied using CD and NMR spectroscopy at various temperatures and by differential scanning calorimetry. It was found that the C‐terminal part (a 16‐residue‐long fragment) of this peptide, which corresponds to the sequence of the β‐hairpin in the native structure, forms structure similar to the β‐hairpin only at T = 313 K, and the structure is stabilized by non‐native long‐range hydrophobic interactions (Val47–Val59). On the other hand, the N‐terminal part of IG(28–61), which corresponds to the middle α‐helix in the native structure, is unstructured at low temperature (283 K) and forms an α‐helix‐like structure at 305 K, and only one helical turn is observed at 313 K. At all temperatures at which NMR experiments were performed (283, 305, and 313 K), we do not observe any long‐range connectivities which would have supported packing between the C‐terminal (β‐hairpin) and the N‐terminal (α‐helix) parts of the sequence. Such interactions are absent, in contrast to the folding pathway of the B domain of protein G, proposed recently by Kmiecik and Kolinski (Biophys J 2008, 94, 726–736), based on Monte‐Carlo dynamics studies. Alternative folding mechanisms are proposed and discussed. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 469–480, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  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