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.
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   

2.
The B-domain of protein A has one of the simplest protein topologies, a three-helix bundle. Its folding has been studied as a model for elementary steps in the folding of larger proteins. Earlier studies suggested that folding might occur by way of a helical hairpin intermediate. Equilibrium hydrogen exchange measurements indicate that the C-terminal helical hairpin could be a potential folding intermediate. Kinetic refolding experiments were performed using stopped-flow circular dichroism and NMR hydrogen-deuterium exchange pulse labeling. Folding of the entire molecule is essentially complete within the 6 ms dead time of the quench-flow apparatus, indicating that the intermediate, if formed, progresses rapidly to the final folded state. Site-directed mutagenesis of the isoleucine residue at position 16 was used to generate a variant protein containing tryptophan (the 116 W mutant). The formation of the putative folding intermediate was expected to be favored in this mutant at the expense of the native folded form, due to predicted unfavorable steric interactions of the bulky tryptophan side chain in the folded state. The 116 W mutant refolds completely within the dead time of a stopped-flow fluorescence experiment. No partly folded intermediate could be detected by either kinetic or equilibrium measurements. Studies of peptide fragments suggest that the protein A sequence has an intrinsic propensity to form a helix II/helix III hairpin. However, its stability appears to be marginal (of the order of 1/2 kT) and it could not be an obligatory intermediate on a defined folding pathway. These results explicitly demonstrate that the protein A B domain folds extremely rapidly by an apparent two-state mechanism without formation of stable partly folded intermediates. Similar mechanisms may also be involved in the rapid folding of subdomains of larger proteins to form the compact molten globule intermediates that often accumulate during the folding process.  相似文献   

3.
Folding propensities of peptide fragments of myoglobin.   总被引:8,自引:5,他引:3       下载免费PDF全文
Myoglobin has been studied extensively as a paradigm for protein folding. As part of an ongoing study of potential folding initiation sites in myoglobin, we have synthetized a series of peptides covering the entire sequence of sperm whale myoglobin. We report here on the conformation preferences of a series of peptides that cover the region from the A helix to the FG turn. Structural propensities were determined using circular dichroism and nuclear magnetic resonance spectroscopy in aqueous solution, trifluoroethanol, and methanol. Peptides corresponding to helical regions in the native protein, namely the B, C, D, and E helices, populate the alpha region of (phi, psi) space in water solution but show no measurable helix formation except in the presence of trifluoroethanol. The F-helix sequence has a much lower propensity to populate helical conformations even in TFE. Despite several attempts, we were not successful in synthesizing a peptide corresponding to the A-helix region that was soluble in water. A peptide termed the AB domain was constructed spanning the A- and B-helix sequences. The AB domain is not soluble in water, but shows extensive helix formation throughout the peptide when dissolved in methanol, with a break in the helix at a site close to the A-B helix junction in the intact folded myoglobin protein. With the exception of one local preference for a turn conformation stabilized by hydrophobic interactions, the peptides corresponding to turns in the folded protein do not measurably populate beta-turn conformations in water, and the addition of trifluoroethanol does not enhance the formation of either helical or turn structure. In contrast to the series of peptides described here, either studies of peptides from the GH region of myoglobin show a marked tendency to populate helical structures (H), nascent helical structures (G), or turn conformations (GH peptide) in water solution. This region, together with the A-helix and part of the B-helix, has been shown to participate in an early folding intermediate. The complete analysis of conformational properties of isolated myoglobin peptides supports the hypothesis that spontaneous secondary structure formation in local regions of the polypeptide may play an important role in the initiation of protein folding.  相似文献   

4.
The effect of an aqueous/organic interface on the folding and aggregation of amphipathic peptides is examined by applying discontinuous molecular dynamics (DMD) simulations combined with an intermediate resolution protein model, PRIME20, to a peptide/interface system. The systems contain 48 (KLLK)4 peptides in random coil or α-helical conformations interacting with both strong and weak interfaces. In the absence of an interface, most of the oligomers form helical bundles, a small fraction of which convert to β-sheets when the temperature is above the folding transition. Adding a weak interface decreases oligomer formation above the folding temperature and increases it below. Little monolayer formation is observed at the weak interface; instead reversible adsorption increases the local peptide concentration near the interface, promoting helical bundle formation in the aqueous phase below the folding temperature and β-sheet formation above the folding temperature. Introducing a strong interface leads to irreversible adsorption, promoting formation of helical monolayers below the folding temperature and mixed β-sheet/amorphous monolayers above the folding temperature. The (KLLK)4 peptide is more likely to adsorb to the interface when it is in an α-helical conformation, as opposed to a random coil, because of its larger hydrophobic moment.  相似文献   

5.
CD and nmr spectroscopy were used to compare the conformational properties of two related peptides. One of the peptides, Model AB, was designed to adopt a helix-turn-extended strand (αβ) tertiary structure in water that might be stabilized by hydrophobic interactions between two leucine residues in the amino-terminal segment and two methionine residues in the carboxyl terminal segment. The other peptide, AB Helix, has the same amino acid sequence as Model AB except that it lacks the-Pro-Met-Thr-Met-Thr-Gly segment at the carboxyl-terminus. Although the carboxyl-terminal segment of Model AB was found to be unstructured, its presence increases the number of residues in a helical conformation, shifts the pKas of three ionizable side chains by 1 pH unit or more compared to an unstructured peptide, stabilizes the peptide as a monomer in high concentrations of ammonium sulfate, increases the conformational stability of residues at the terminal ends of the helix, and results in many slowly exchanging amide protons throughout the entire backbone of the peptide. These results suggest that interactions between adjacent segments in a small peptide can have significant structure organizing effects. Similar kinds of interactions may be important in determining the structure of early intermediates in protein folding and may be useful in the de novo design of independently folding peptides. © 1995 John Wiley & Sons, Inc.  相似文献   

6.
Du D  Bunagan MR  Gai F 《Biophysical journal》2007,93(11):4076-4082
The formation of the monomeric α-helix represents one of the simplest scenarios in protein folding; however, our current understanding of the folding dynamics of the α-helix motif is mainly based on studies of alanine-rich model peptides. To examine the effect of peptide sequence on the folding kinetics of α-helices, we studied the relaxation kinetics of a 21-residue helical peptide, Conantokin-T (Con-T), using time-resolved infrared spectroscopy in conjunction with a laser-induced temperature jump technique. Con-T is a neuroactive peptide containing a large number of charged residues that is found in the venom of the piscivorous cone snail Conus tulipa. The temperature-jump relaxation kinetics of Con-T is distinctly slower than that of previously studied alanine-based peptides, suggesting that the folding time of α-helices is sequence-dependent. Furthermore, it appears that the slower folding of Con-T can be attributed to the fact that its helical conformation is stabilized by charge-charge interactions or salt bridges. Although this finding contradicts an earlier molecular dynamics simulation, it also has implications for existing models of protein folding.  相似文献   

7.
A 14 amino acid residue peptide from the helical region of Scorpion neurotoxin has been structurally characterized using CD and NMR spectroscopy in different solvent conditions. 2,2,2-Trifluoroethanol (TFE) titration has been carried out in 11 steps from 0 to 90% TFE and the gradual stabilization of the conformation to form predominantly alpha-helix covering all of the 14 residues has been studied by 1H and 13C NMR spectroscopy. Detailed information such as coupling constants, chemical shift indices, NOESY peak intensities and amide proton temperature coefficients at each TFE concentration has been extracted and analysed to derive the step-wise preferential stabilization of the helical segments along the length of the peptide. It was found that there is a finite amount of the helical conformation in the middle residues 5-11 even at low TFE concentrations. It was also observed that > 75% TFE (v/v) is required for the propagation of the helix to the N and C termini and for correct packing of the side chains of all of the residues. These observations are significant to understanding the folding of this segment in the protein and may throw light on the inherent preferences and side chain interactions in the formation of the helix in the peptide.  相似文献   

8.
The peptide backbones in folded native proteins contain distinctive secondary structures, alpha-helices, beta-sheets, and turns, with significant frequency. One question that arises in folding is how the stability of this secondary structure relates to that of the protein as a whole. To address this question, we substituted the alpha-helix-stabilizing alanine side chain at 16 selected sites in the sequence of sperm whale myoglobin, 12 at helical sites on the surface of the protein, and 4 at obviously internal sites. Substitution of alanine for bulky side chains at internal sites destabilizes the protein, as expected if packing interactions are disrupted. Alanine substitutions do not uniformly stabilize the protein, either in capping positions near the ends of helices or at mid-helical sites near the surface of myoglobin. When corrected for the extent of exposure of each side chain replaced by alanine at a mid-helix position, alanine replacement still has no clear effect in stabilizing the native structure. Thus linkage between the stabilization of secondary structure and tertiary structure in myoglobin cannot be demonstrated, probably because of the relatively small free energy differences between side chains in stabilizing isolated helix. By contrast, about 80% of the variance in free energy observed can be accounted for by the loss in buried surface area of the native residue substituted by alanine. The differential free energy of helix stabilization does not account for any additional variation.  相似文献   

9.
Environment and exposure to solvent of protein atoms. Lysozyme and insulin   总被引:32,自引:0,他引:32  
A computer program is described for calculating the environment and the exposure to solvent of atoms of a protein. The computation is based on the atomic co-ordinates of the protein and on assumptions like those of Lee &; Richards (1971). Results for lysozyme and insulin are presented. Changes in exposure to solvent and in the nature of contacts that develop through folding, association reactions and crystallization are described numerically. The computations suggest several generalizations. (a) Lattice contacts within the protein crystal are characterized by a significantly smaller involvement of non-polar side chains and a proportionately greater involvement of ionizable side chains than is found for protein folding or for protein association reactions important for biological function, (b) In helical regions the carbonyl oxygen of the first residue in the helix has high probability of being shielded from solvent, (c) Glycine is among the residues having exposure least affected by folding; this accords with the expectation that it lies at bends of the peptide chain on the surface of the molecule.  相似文献   

10.
Helix propensities of the amino acids have been measured in alanine-based peptides in the absence of helix-stabilizing side-chain interactions. Fifty-eight peptides have been studied. A modified form of the Lifson-Roig theory for the helix-coil transition, which includes helix capping (Doig AJ, Chakrabartty A, Klingler TM, Baldwin RL, 1994, Biochemistry 33:3396-3403), was used to analyze the results. Substitutions were made at various positions of homologous helical peptides. Helix-capping interactions were found to contribute to helix stability, even when the substitution site was not at the end of the peptide. Analysis of our data with the original Lifson-Roig theory, which neglects capping effects, does not produce as good a fit to the experimental data as does analysis with the modified Lifson-Roig theory. At 0 degrees C, Ala is a strong helix former, Leu and Arg are helix-indifferent, and all other amino acids are helix breakers of varying severity. Because Ala has a small side chain that cannot interact significantly with other side chains, helix formation by Ala is stabilized predominantly by the backbone ("peptide H-bonds"). The implication for protein folding is that formation of peptide H-bonds can largely offset the unfavorable entropy change caused by fixing the peptide backbone. The helix propensities of most amino acids oppose folding; consequently, the majority of isolated helices derived from proteins are unstable, unless specific side-chain interactions stabilize them.  相似文献   

11.
Anderson MW  Gorski J 《Biochemistry》2005,44(15):5617-5624
To generate an effective immune response, class II major histocompatibility complex molecules (MHCII) must present a diverse array of peptide ligands for recognition by T lymphocytes. Peptide/MHCII complexes are stabilized by hydrophobic anchoring of peptide side chains to pockets in the MHCII protein and the formation of hydrogen bonds to the peptide backbone. Many current models of peptide/MHCII association assume an additive and independent contribution of the interactions between major MHCII pockets and corresponding side chains in the peptide. However, significant conformational rearrangements occur in both the peptide and MHCII during binding. Therefore, we hypothesize that peptide binding to MHCII could be viewed as a folding process in which both molecules cooperate to produce the final conformation. To directly test this hypothesis, we adapt a serial mutagenesis strategy to study cooperativity in the interaction of the human MHCII HLA-DR1 and a peptide derived from influenza hemagglutinin. Substitutions in either the peptide or HLA-DR1 that are predicted to interfere with hydrogen bond formation show cooperative effects on complex stability and affinity. Substitution of a peptide side chain that provides a hydrophobic contact also contributes to the cooperative effect, suggesting a role for all energetic sources in the folding process. We propose that cooperativity throughout the peptide-binding groove reflects the folding of segments of the MHCII molecule into helices around the peptide with a concomitant folding of the peptide into a polyproline helix. The implications of cooperativity for peptide/MHCII structure and epitope selection are discussed.  相似文献   

12.
In this work, we studied how an amphipathic peptide of the surface of the globular protein thioredoxin, TRX94‐108, acquires a native‐like structure when it becomes involved in an apolar interaction network. We designed peptide variants where the tendency to form α‐helical conformation is modulated by replacing each of the leucine amino acid residues by an alanine. The induction of structure caused by sodium dodecyl sulfate (SDS) binding was studied by capillary zone electrophoresis, circular dichroism, DOSY‐NMR, and molecular dynamics simulations (MDS). In addition, we analyzed the strength of the interaction between a C18 RP‐HPLC matrix and the peptides. The results presented here reveal that (a) critical elements in the sequence of the wild‐type peptide stabilize a SDS/peptide supramolecular cluster; (b) the hydrophobic nature of the interaction between SDS molecules and the peptide constrains the ensemble of conformations; (c) nonspecific apolar surfaces are sufficient to stabilize peptide secondary structure. Remarkably, MDS shed light on a contact network formed by a limited number of SDS molecules that serves as a structural scaffold preserving the helical conformation of this module. This mechanism might prevail when a peptide with low helical propensity is involved in structure consolidation. We suggest that folding of peptides sharing this feature does not require a preformed tightly‐packed protein core. Thus, the formation of specific tertiary interactions would be the consequence of peptide folding and not its cause. In this scenario, folding might be thought of as a process that includes unspecific rounds of structure stabilization guiding the protein to the native state. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
In an attempt to understand the earliest events in the protein folding pathway, the complete sequence of French bean plastocyanin has been synthesized as a series of short peptide fragments, and the conformational preferences of each peptide examined in aqueous solution using proton n.m.r. methods. Plastocyanin consists largely of beta-sheet, with reverse turns and loops between the strands of the sheet, and one short helix. The n.m.r. experiments indicate that most of the peptides derived from the plastocyanin sequence have remarkably little propensity to adopt folded conformations in aqueous solution, in marked contrast to the peptides derived from the helical protein, myohemerythrin (accompanying paper). For most plastocyanin peptides, the backbone dihedral angles are predominantly in the beta-region of conformational space. Some of the peptides show weak NOE connectivities between adjacent amide protons, indicative of small local populations of backbone conformations in the a region of (phi,psi) space. A conformational preference for a reverse turn is seen in the sequence Ala65-Pro-Gly-Glu68, where a turn structure is found in the folded protein. Significantly, the peptide sequences that populate the alpha-region of (phi,psi) space are mostly derived from turn and loop regions in the protein. The addition of trifluoroethanol does not drive the peptides into helical conformations. In one region of the sequence, the n.m.r. spectra provide evidence of the formation of a hydrophobic cluster involving aromatic and aliphatic side-chains. These results have significance for understanding the initiation of protein folding. From these studies of the fragments of plastocyanin (this paper) and myohemerythrin (accompanying paper), it appears that there is a pre-partitioning of the conformational space sampled by the polypeptide backbone that is related to the secondary structure in the final folded state.  相似文献   

14.
Determining the minimal sequence necessary to induce protein folding is beneficial in understanding the role of protein–protein interactions in biological systems, as their three-dimensional structures often dictate their activity. Proteins are generally comprised of discrete secondary structures, from α-helices to β-turns and larger β-sheets, each of which is influenced by its primary structure. Manipulating the sequence of short, moderately helical peptides can help elucidate the influences on folding. We created two new scaffolds based on a modestly helical eight-residue peptide, PT3, we previously published. Using circular dichroism (CD) spectroscopy and changing the possible salt-bridging residues to new combinations of Lys, Arg, Glu, and Asp, we found that our most helical improvements came from the Arg–Glu combination, whereas the Lys–Asp was not significantly different from the Lys–Glu of the parent scaffold, PT3. The marked 310-helical contributions in PT3 were lessened in the Arg–Glu-containing peptide with the beginning of cooperative unfolding seen through a thermal denaturation. However, a unique and unexpected signature was seen for the denaturation of the Lys–Asp peptide which could help elucidate the stages of folding between the 310 and α-helix. In addition, we developed a short six-residue peptide with β-turn/sheet CD signature, again to help study minimal sequences needed for folding. Overall, the results indicate that improvements made to short peptide scaffolds by fine-tuning the salt-bridging residues can enhance scaffold structure. Likewise, with the results from the new, short β-turn motif, these can help impact future peptidomimetic designs in creating biologically useful, short, structured β-sheet-forming peptides.  相似文献   

15.
S J Landry  L M Gierasch 《Biochemistry》1991,30(30):7359-7362
Chaperones facilitate folding and assembly of nascent polypeptides in vivo and prevent aggregation in refolding assays in vitro. A given chaperone acts on a number of different proteins. Thus, chaperones must recognize features present in incompletely folded polypeptide chains and not strictly dependent on primary structural information. We have used transferred nuclear Overhauser effects to demonstrate that the Escherichia coli chaperonin GroEL binds to a peptide corresponding to the N-terminal alpha-helix in rhodanese, a mitochondrial protein whose in vitro refolding is facilitated by addition of GroEL, GroES, and ATP. Furthermore, the peptide, which is unstructured when free in aqueous solution, adopts an alpha-helical conformation upon binding to GroEL. Modification of the peptide to reduce its intrinsic propensity to take up alpha-helical structure lowered its affinity for GroEL, but, nonetheless, it could be bound and took up a helical conformation when bound. We propose that GroEL interacts with sequences in an incompletely folded chain that have the potential to adopt an amphipathic alpha-helix and that the chaperonin binding site promotes formation of a helix.  相似文献   

16.
We study the effects of confinement between planar walls on the folding thermodynamics of a β-hairpin, using large-scale replica-exchange molecular-dynamics simulations with an all-atom model and explicit solvent. We find that the folding free-energy landscape of this peptide observed in bulk is significantly modified when the peptide is confined between the walls. Most notably, the propensity of the peptide to form a misfolded state observed in the bulk solution becomes negligible under confinement. The absence of the misfolded state under confinement can be explained by an increased tendency of hydrophobic aromatic side chains to stay near the walls, because the misfolded state is characterized by a nonnative arrangement of aromatic side chains. These results from a simple confinement model may provide clues about the role of chaperonin confinement in smoothing folding landscapes by avoiding trapped intermediates.  相似文献   

17.
A seven-residue peptide that is highly conserved in SH3 domains despite being far from the active site has been shown by NMR to be stable in solution. This peptide, biologically important because it is a likely folding nucleus for SH3 domains, provides a challenging subject for molecular dynamics because it is highly charged. We present stable, 10-ns simulations of both the native-like diverging turn structure and a helical model. Free energies of these two conformations, estimated through MM-PBSA analysis using several force fields, suggest a comparable free energy (DeltaDeltaG < or =6 kcal/mol) for native and helix conformations. NOE intensities calculated from the native trajectory reproduce experimental data quite well, suggesting that the conformations sampled by the trajectory reasonably represent those observed in the NMR experiment. The molecular dynamics results, as well as sequence analysis of a diverse 690-member family of SH3 domain proteins, suggest that the presence of two elements is essential for formation of the diverging turn structure: a pair of residues with low helical propensity in the turn region and, as previously recognized, two hydrophobic residues to close the end of the diverging turn. Thus, these two sequence features may form the structural basis for the function of this peptide as a folding nucleus in this family of proteins.  相似文献   

18.
Antimicrobial peptides (AMPs) are found in various classes of organisms as part of the innate immune system. Despite high sequence variability, they share common features such as net positive charge and an amphipathic fold when interacting with biologic membranes. Esculentin-1b is a 46-mer frog-skin peptide, which shows an outstanding antimicrobial activity. Experimental studies revealed that the N-terminal fragment encompassing the first 18 residues, Esc(1-18), is responsible for the antimicrobial activity of the whole peptide, with a negligible toxicity toward eukaryotic cells, thus representing an excellent candidate for future pharmaceutical applications. Similarly to most of the known AMPs, Esc(1-18) is expected to act by destroying/permeating the bacterial plasma-membrane but, to date, its 3D structure and the detailed mode of action remains unexplored. Before an in-depth investigation on peptide/membranes interactions could be undertaken, it is necessary to characterize peptide's folding propensity in solution, to understand what is intrinsically due to the peptide sequence, and what is actually driven by the membrane interaction. Circular dichroism and nuclear magnetic resonance spectroscopy were used to determine the structure adopted by the peptide, moving from water to increasing amounts of trifluoroethanol. The results showed that Esc(1-18) has a clear tendency to fold in a helical conformation as hydrophobicity of the environment increases, revealing an intriguing amphipathic structure. The helical folding is adopted only by the N-terminal portion of the peptide, while the rest is unstructured. The presence of a hydrophobic cluster of residues in the C-terminal portion suggests its possible membrane-anchoring role.  相似文献   

19.
We have examined the kinetics of the adsorption of melittin, a secondary amphipathic peptide extracted from bee venom, on lipid membranes using three independent and complementary approaches. We probed (i) the change in the polarity of the 19Trp of the peptide upon binding, (ii) the insertion of this residue in the apolar core of the membrane, measuring the 19Trp-fluorescence quenching by bromine atoms attached on lipid acyl chains, and (iii) the folding of the peptide, by circular dichroism (CD). We report a tight coupling of the insertion of the peptide with its folding as an α-helix. For all the investigated membrane systems (cholesterol-containing, phosphoglycerol-containing, and pure phosphocholine bilayers), the decrease in the polarity of 19Trp was found to be significantly faster than the increase in the helical content of melittin. Therefore, from a kinetics point of view, the formation of the α-helix is a consequence of the insertion of melittin. The rate of melittin folding was found to be influenced by the lipid composition of the bilayer and we propose that this was achieved by the modulation of the kinetics of insertion. The study reports a clear example of the coupling existing between protein penetration and folding, an interconnection that must be considered in the general scheme of membrane protein folding.  相似文献   

20.
Kinetic intermediates in protein folding are short-lived and therefore difficult to detect and to characterize. In the folding of polypeptide chains with incorrect isomers of Xaa-Pro peptide bonds the final rate-limiting transition to the native state is slow, since it is coupled to prolyl isomerization. Incorrect prolyl isomers thus act as effective traps for folding intermediates and allow their properties to be studied more easily. We employed this strategy to investigate the mechanism of slow folding of ribonuclease T1. In our experiments we use a mutant form of this protein with a single cis peptide bond at proline 39. During refolding, protein chains with an incorrect trans proline 39 can rapidly form extensive secondary structure. The CD signal in the amide region is regained within the dead-time of stopped-flow mixing (15 ms), indicating a fast formation of the single alpha-helix of ribonuclease T1. This step is correlated with partial formation of a hydrophobic core, because the fluorescence emission maximum of tryptophan 59 is shifted from 349 nm to 325 nm within less than a second. After about 20 s of refolding an intermediate is present that shows about 40% enzymatic activity compared to the completely refolded protein. In addition, the solvent accessibility of tryptophan 59 is drastically reduced in this intermediate and comparable to that of the native state as determined by acrylamide quenching of the tryptophan fluorescence. Activity and quenching measurements have long dead-times and therefore we do not know whether enzymatic activity and solvent accessibility also change in the time range of milliseconds. At this stage of folding at least part of the beta-sheet structure is already present, since it hosts the active site of the enzyme. The trans to cis isomerization of the tyrosine 38-proline 39 peptide bond in the intermediate and consequently the formation of native protein is very slow (tau = 6,500 s at pH 5.0 and 10 degrees C). It is accompanied by an additional increase in tryptophan fluorescence, by the development of the fine structure of the tryptophan emission spectrum, and by the regain of the full enzymatic activity. This indicates that the packing of the hydrophobic core, which involves both tryptophan 59 and proline 39, is optimized in this step. Apparently, refolding polypeptide chains with an incorrect prolyl isomer can very rapidly form partially folded intermediates with native-like properties.  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