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.
Kinetic role of helix caps in protein folding is context-dependent   总被引:2,自引:0,他引:2  
Kapp GT  Richardson JS  Oas TG 《Biochemistry》2004,43(13):3814-3823
Secondary structure punctuation through specific backbone and side chain interactions at the beginning and end of alpha-helices has been proposed to play a key role in hierarchical protein folding mechanisms [Baldwin, R. L., and Rose, G. D. (1999) Trends Biochem. Sci. 24, 26-33; Presta, L. G., and Rose, G. D. (1988) Science 240, 1632-1641]. We have made site-specific substitutions in the N- and C-cap motifs of the 5-helix protein monomeric lambda repressor (lambda(6-85)) and have measured the rate constants for folding and unfolding of each variant. The consequences of C-cap changes are strongly context-dependent. When the C-cap was located at the chain terminus, changes had little energetic and no kinetic effect. However, substitutions in a C-cap at the boundary between helix 4 and the subsequent interhelical loop resulted in large changes to the stability and rate constants of the variant, showing a substantial kinetic role for this interior C-cap and suggesting a general kinetic role for interior helix C-caps. Statistical preferences tabulated separately for internal and terminal C-caps also show only weak residue preferences in terminal C-caps. This kinetic distinction between interior and terminal C-caps can explain the discrepancy between the near-absence of stability and kinetic effects seen for C-caps of isolated peptides versus the very strong C-cap effects seen for proteins in statistical sequence preferences and mutational energetics. Introduction of consensus, in-register N-capping motifs resulted in increased stability, accelerated folding, and slower unfolding. The kinetic measurements indicate that some of the new native-state capping interactions remain unformed in the transition state. The accelerated folding rates could result from helix stabilization without invoking a specific role for N-caps in the folding reaction.  相似文献   

2.
The charge-containing hydrophilic functionalities of encoded charged amino acids are linked to the backbone via different numbers of hydrophobic methylenes, despite the apparent electrostatic nature of protein ion pairing interactions. To investigate the effect of side chain length of guanidinium- and carboxylate-containing residues on ion pairing interactions, α-helical peptides containing Zbb–Xaa (i, i + 3), (i, i + 4) and (i, i + 5) (Zbb = carboxylate-containing residues Aad, Glu, Asp in decreasing length; Xaa = guanidinium residues Agh, Arg, Agb, Agp in decreasing length) sequence patterns were studied by circular dichroism spectroscopy (CD). The helicity of Aad- and Glu-containing peptides was similar and mostly pH independent, whereas the helicity of Asp-containing peptides was mostly pH dependent. Furthermore, the Arg-containing peptides consistently exhibited higher helicity compared to the corresponding Agp-, Agb-, and Agh-containing peptides. Side chain conformational analysis by molecular mechanics calculations showed that the Zbb–Xaa (i, i + 3) and (i, i + 4) interactions mainly involved the χ 1 dihedral combinations (g+, g+) and (g?, g+), respectively. These low energy conformations were also observed in intrahelical Asp–Arg and Glu–Arg salt bridges of natural proteins. Accordingly, Asp and Glu provides variation in helix characteristics associated with Arg, but Aad does not provide features beyond those already delivered by Glu. Importantly, nature may have chosen the side chain length of Arg to support helical conformations through inherent high helix propensity coupled with stabilizing intrahelical ion pairing interactions with the carboxylate-containing residues.  相似文献   

3.
Much effort has been invested in seeking to understand the thermodynamic basis of helix stability in both peptides and proteins. Recently, several groups have measured the helix-forming propensities of individual residues (Lyu, P. C., Liff, M. I., Marky, L. A., Kallenbach, N. R. Science 250:669–673, 1990; O'Neil, K. T., DeGrado, W. F. Science 250:646–651, 1990; Padmanabhan, S., Marqusee, S., Ridgeway, T., Laue, T. M., Baldwin, R. L. Nature (London) 344:268–270, 1990). Using Monte Carlo computer simulations, we tested the hypothesis that these differences in measured helix-forming propensity are due primarily to loss of side chain conformational entropy upon helix formation (Creamer, T. P., Rose, G. D. Proc. Natl. Acad. Sci. U.S.A. 89:5937–5941, 1992). Our previous study employed a rigid helix backbone, which is here generalized to a completely flexible helix model in order to ensure that earlier results were not a methodological artifact. Using this flexible model, side chain rotamer distributions and entropy losses are calculated and shown to agree with those obtained earlier. We note that the side chain conformational entropy calculated for Trp in our previous study was in error; a corrected value is presented. Extending earlier work, calculated entropy losses are found to correlate strongly with recent helix propensity scales derived from substitutions made within protein helices (Horovitz, A., Matthews, J. M., Fersht, A. R. J. Mol. Biol. 227:560–568, 1992; Blaber, M., Zhang, X.-J., Matthews, B. M. Science 260:1637–1640, 1993). In contrast, little correlation is found between these helix propensity scales and the accessible surface area buried upon formation of a model polyalanyl α-helix. Taken in sum, our results indicate that loss of side chain entropy is a major determinant of the helix-forming tendency of residues in both peptide and protein helices. © 1994 Wiley-Liss, Inc.  相似文献   

4.
The alpha-helix is a fundamental protein structural motif and is frequently terminated by a glycine residue. Explanations for the predominance of glycine at the C-cap terminal portions of alpha-helices have invoked uniquely favorable energetics of this residue in a left-handed conformation or enhanced solvation of the peptide backbone because of the absence of a side chain. Attempts to quantify the contributions of these two effects have been made previously, but the issue remains unresolved. Here we have used chemical protein synthesis to dissect the energetic basis of alpha-helix termination by comparing a series of ubiquitin variants containing an L-amino acid or the corresponding D-amino acid at the C-cap Gly35 position. D-Amino acids can adopt a left-handed conformation without energetic penalty, so the contributions of conformational strain and backbone solvation can thus be separated. Analysis of the thermodynamic data revealed that the preference for glycine at the C' position of a helix is predominantly a conformational effect.  相似文献   

5.
Understanding the secondary structure of peptides is important in protein folding, enzyme function, and peptide‐based drug design. Previous studies of synthetic Ala‐based peptides (>12 a.a.) have demonstrated the role for charged side chain interactions involving Glu/Lys or Glu/Arg spaced three (i, i + 3) or four (i, i + 4) residues apart. The secondary structure of short peptides (<9 a.a.), however, has not been investigated. In this study, the effect of repetitive Glu/Lys or Glu/Arg side chain interactions, giving rise to E‐R/K helices, on the helicity of short peptides was examined using circular dichroism. Short E‐R/K–based peptides show significant helix content. Peptides containing one or more E‐R interactions display greater helicity than those with similar E‐K interactions. Significant helicity is achieved in Arg‐based E‐R/K peptides eight, six, and five amino acids long. In these short peptides, each additional i + 3 and i + 4 salt bridge has substantial contribution to fractional helix content. The E‐R/K peptides exhibit a strongly linear melt curve indicative of noncooperative folding. The significant helicity of these short peptides with predictable dependence on number, position, and type of side chain interactions makes them an important consideration in peptide design.  相似文献   

6.
It was established previously that helical propensities of different amino acid residues in the middle of α‐helix in peptides and in proteins are very similar. The statistical analysis of the protein helices from the known three‐dimensional structures shows no difference in the frequency of noncharged residues in the middle and at the C terminus. Yet, experimental studies show distinctive differences for the helical propensities of noncharged residues in the middle and in the C terminus in model peptides. Is this a general effect, and is it applicable to protein helices or is it specific to the model alanine‐based peptides? To answer this question, the effects of substitutions at positions 28 (middle residue) and 32 (C2 position at the C terminus) of the α‐helix of ubiquitin on the stability of this protein are measured by using differential scanning calorimetry. The two data sets produce similar values for intrinsic helix propensity, leading to a conclusion that noncharged amino acid residues at the solvent‐exposed positions in the middle and at the C terminus of the α‐helix have the same helical propensity. This conclusion is further supported with an excellent correlation between the helix propensity scale obtained for the two positions in ubiquitin with the experimental helix propensity scale established previously and with the statistical distribution of the residues in protein helices.  相似文献   

7.
A significant fraction of the amino acids in proteins are alpha helical in conformation. Alpha helices in globular proteins are short, with an average length of about twelve residues, so that residues at the ends of helices make up an important fraction of all helical residues. In the middle of a helix, H-bonds connect the NH and CO groups of each residue to partners four residues along the chain. At the ends of a helix, the H-bond potential of the main chain remains unfulfilled, and helix capping interactions involving bonds from polar side chains to the NH or CO of the backbone have been proposed and detected. In a study of synthetic helical peptides, we have found that the sequence Ser-Glu-Asp-Glu stabilizes the alpha helix in a series of helical peptides with consensus sequences. Following the report by Harper and Rose, which identifies SerXaaXaaGlu as a member of a class of common motifs at the N termini of alpha helices in proteins that they refer to as “capping boxes,” we have reexamined the side chain–main chain interactions in a varient sequence using 1H NMR, and find that the postulated reciprocal side chain-backbone bonding between the first Ser and last Glu side chains and their peptide NH partners can be resolved: Deletion of two residues N terminal to the Ser-Glu-Asp-Glu sequence in these peptides has no effect on the initiation of helical structure, as defined by two-dimensional (2D) NMR experiments on this variant. Thus the capping box sequence Ser-Glu-Asp-Glu inhibits N terminal fraying of the N terminus of alpha helix in these peptides, and shows the side chain–main chain interactions proposed by Harper and Rose. It thus acts as a helix initiating signal. Since normal a helix cannot propagate beyond the N terminus of this structure, the box acts as a termination signal in this direction as well. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
A series of artificial peptides bearing cationic functional groups with different side chain lengths were designed, and their ability to increase the thermal stability of nucleic acid duplexes was investigated. The peptides with amino groups selectively increased the stability of RNA/RNA duplexes, and a relationship between the side chain length and the melting temperature (Tm) of the peptide–RNA complexes was observed. On the other hand, while peptides with guanidino groups exhibited a similar tendency with respect to the peptide structure and thermal stability of RNA/RNA duplexes, those with longer side chain lengths, such as l-2-amino-4-guanidinobutyric acid (Agb) or l-arginine (Arg) oligomers, stabilized both RNA/RNA and DNA/DNA duplexes, and those with shorter side chain lengths exhibited a higher ability to selectively stabilize RNA/RNA duplexes. In addition, peptides were designed with different levels of flexibility by introducing glycine (Gly) residues into the l-2-amino-3-guanidinopropionic acid (Agp) oligomers. It was found that insertion of Gly did not affect the thermal stability of the peptide–RNA complexes, but an alternate arrangement of Gly and Agp apparently decreased the thermal stability. Therefore, in the Agp oligomer, consecutive Agp sequences are essential for increasing the stability of RNA/RNA duplexes.  相似文献   

9.
10.
Lactose/H(+) symport by lactose permease of Escherichia coli involves interactions between four irreplaceable charged residues in transmembrane helices that play essential roles in H(+) translocation and coupling [Glu269 (helix VIII) with His322 (helix X) and Arg302 (helix IX) with Glu325 (helix X)], as well as Glu126 (helix IV) and Arg144 (helix V) which are obligatory for substrate binding. The conservative mutation Glu325-->Asp causes a 10-fold reduction in the V(max) for active lactose transport and markedly decreased lactose-induced H(+) influx with no effect on exchange or counterflow, neither of which involves H(+) symport. Thus, shortening the side chain may weaken the interaction of the carboxyl group at position 325 with the guanidino group of Arg302. Therefore, Gly-scanning mutagenesis of helices IX and X and the intervening loop was employed systematically with mutant Glu325-->Asp in an effort to rescue function by introducing conformational flexibility between the two helices. Five Gly replacement mutants in the Glu325-->Asp background are identified that exhibit significantly higher transport activity. Furthermore, mutant Val316-->Gly/Glu325-->Asp catalyzes active transport, efflux, and lactose-induced H(+) influx with kinetic properties approaching those of wild-type permease. It is proposed that introduction of conformational flexibility at the interface between helices IX and X improves juxtapositioning between Arg302 and Asp325 during turnover, thereby allowing more effective deprotonation of the permease on the inner surface of the membrane [Sahin-Tóth, M., Karlin, A., and Kaback, H. R. (2000) Proc. Natl. Acad. Sci. U.S.A. 97, 10729-10732.  相似文献   

11.
The positively charged side chains of cationic antimicrobial peptides are generally thought to provide the initial long-range electrostatic attractive forces that guide them towards the negatively charged bacterial membranes. Peptide analogs were designed to examine the role of the four Arg side chains in the cathelicidin peptide tritrpticin (VRRFPWWWPFLRR). The analogs include several noncoded Arg and Lys derivatives that offer small variations in side chain length and methylation state. The peptides were tested for bactericidal and hemolytic activities, and their membrane insertion and permeabilization properties were characterized by leakage assays and fluorescence spectroscopy. A net charge of +5 for most of the analogs maintains their high antimicrobial activity and directs them towards preferential insertion into model bacterial membrane systems with a similar extent of burial of the Trp side chains. However the peptides exhibit significant functional differences. Analogs with methylated cationic side chains cause lower levels of membrane leakage and are associated with lower hemolytic activities, making them potentially attractive pharmaceutical candidates. Analogs containing the Arg guanidinium groups cause more membrane disruption than those containing the Lys amino groups. Peptides in the latter group with shorter side chains have increased membrane activity and conversely, elongating the Arg residue causes slightly higher membrane activity. Altogether, the potential for strong hydrogen bonding between the four positive Arg side chains with the phospholipid head groups seems to be a determinant for the membrane disruptive properties of tritrpticin and many related cationic antimicrobial peptides.  相似文献   

12.
An hypothesis is tested that individual peptides corresponding to the transmembrane helices of the membrane protein, rhodopsin, would form helices in solution similar to those in the native protein. Peptides containing the sequences of helices 1, 4 and 5 of rhodopsin were synthesized. Two peptides, with overlapping sequences at their termini, were synthesized to cover each of the helices. The peptides from helix 1 and helix 4 were helical throughout most of their length. The N- and C-termini of all the peptides were disordered and proline caused opening of the helical structure in both helix 1 and helix 4. The peptides from helix 5 were helical in the middle segment of each peptide, with larger disordered regions in the N- and C-termini than for helices 1 and 4. These observations show that there is a strong helical propensity in the amino acid sequences corresponding to the transmembrane domain of this G-protein coupled receptor. In the case of the peptides from helix 4, it was possible to superimpose the structures of the overlapping sequences to produce a construct covering the whole of the sequence of helix 4 of rhodopsin. As similar superposition for the peptides from helix 1 also produced a construct, but somewhat less successfully because of the disordering in the region of sequence overlap. This latter problem was more severe for helix 5 and therefore a single peptide was synthesized for the entire sequence of this helix, and its structure determined. It proved to be helical throughout. Comparison of all these structures with the recent crystal structure of rhodopsin revealed that the peptide structures mimicked the structures seen in the whole protein. Thus similar studies of peptides may provide useful information on the secondary structure of other transmembrane proteins built around helical bundles.  相似文献   

13.
The Bcl-2 family of proteins plays a central role in the regulation of mitochondrial outer-membrane permeabilization, a critical step in apoptosis. Heterodimerization between the pro- and anti-apoptotic members of Bcl-2 family is a key event in this process. Anti-apoptotic proteins have high levels of expression in many cancers and they have different affinities for different pro-apoptotic proteins. Experimentally determined structures of all members of Bcl-2 proteins have remarkably similar helical fold despite poor amino acid sequence identity. Peptides representing BH3 region of pro-apoptotic proteins have been shown to bind the hydrophobic cleft of anti-apoptotic proteins and this segment is responsible in modulating the apoptotic pathways in living cells. Understanding the molecular basis of protein-protein recognition is required to develop inhibitors specific to a particular anti-apoptotic protein. We have carried out molecular dynamics simulations on the anti-apoptotic Bcl-X(L) protein in complex with three different BH3 peptides derived from pro-apoptotic Bak, Bad and Bim proteins. Each complex structure was simulated for a period of 50 ns after 2.5 ns equilibration. Analysis of the simulation results showed that in the Bcl-X(L) protein, the helix containing the BH3 region is more flexible than other helices in all three simulations. A network of strong hydrophobic interactions exists between four of the six helices and they contribute significantly to the stability of this helix bundle protein. Analysis of Bcl-X(L)-BH3 peptide interactions reveals the role of loop residues in the protein-peptide interactions in all three simulations. Bad and Bim peptides maintain strong hydrophobic and hydrophilic interactions with the helix preceding the central hydrophobic helix. Residues from this helix interact with an Arg residue in Bad and Bim peptides. This Arg residue is next to the conserved Leu residue and is replaced by Ala in Bak. Absence of these interactions and the helix propensity are likely to be the cause for Bak peptide's weaker binding affinity with the Bcl-X(L) protein. The results of this study have implications in the design of Bcl-X(L)-specific inhibitors.  相似文献   

14.
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   

15.
A peptide fragment corresponding to the third helix of Staphylococcus Aureus protein A, domain B, was chosen to study the effect of the main‒chain direction upon secondary structure formation and stability, applying the retro‒enantio concept. For this purpose, two peptides consisting of the native (Ln) and reversed (Lr) sequences were synthesized and their conformational preferences analysed by CD and NMR spectroscopy. A combination of CD and NMR data, such as molar ellipcitity, NOE connectivities, Hα and NH chemical shifts, 3JαN coupling constants and amide temperature coefficients indicated the presence of nascent helices for both Ln and Lr in water, stabilized upon addition of the fluorinated solvents TFE and HFIP. Helix formation and stabilization appeared to be very similar in both normal and retro peptides, despite the unfavourable charge–macrodipole interactions and bad N-capping in the retro peptide. Thus, these helix stabilization factors are not a secondary structure as determined for this specific peptide. In general, the synthesis and confirmational analysis of peptide pairs with opposite main‒chain directions, normal and retro peptides, could be useful in the determination of secondary structure stabilization factors dependent on the direction. © 1997 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Probable role of amphiphilicity in the binding of mastoparan to calmodulin   总被引:6,自引:0,他引:6  
Two-dimensional helical wheel diagrams and calculations of mean hydrophobic moments show mastoparan, mastoparan X, and Polistes mastoparan to have all the properties expected for amphiphilic helices. Circular dichroic properties are consistent with a random form for these peptides in dilute aqueous solution, but greater than 50% helix is apparent when the peptides are dissolved in 70% trifluoroethanol/water mixtures (v/v) or when the peptides are bound to calmodulin. Changes in the fluorescence spectra, anisotropy, and accessibility of tryptophan whose indole side chain is on the apolar surface of the amphiphilic helix imply a significant role for the apolar surface in the binding of the mastoparans and another amphiphilic peptide, melittin, to calmodulin. These data provide a useful model for designing high-affinity synthetic peptide inhibitors of calmodulin.  相似文献   

17.
R D Sheardy  E J Gabbay 《Biochemistry》1983,22(9):2061-2067
A series of diastereomeric dipeptide amides, containing an N-terminal L-lysyl residue and a C-terminal L- or D-amino acid with a derivatized aromatic ring on the side chain, was synthesized to determine the dependence of (1) the chirality of the N-terminal amino acid alpha-carbon and (2) the length of the N-terminal amino acid side chain for intercalation of the aromatic ring. The nature of the complex between the peptide and DNA (i.e., electrostatic, intercalative, or a combination of these) was determined by UV and CD studies, viscometric titrations, and 1H NMR studies. The results of these studies reveal distinct differences in the binding site of the aromatic rings of the various peptides. In particular, the results suggest that the alpha- and epsilon-amino groups of the lysyl residue bind electrostatically to adjacent phosphates on the DNA backbone in a stereospecific manner. As a result of this stereospecificity, the aromatic rings of the peptides with the L-L designation point toward the DNA helix, while those of the peptides of the L-D designation point away from the helix. This is completely consistent with previously reported work [Gabbay, E.J., Adawadkar, P. D., & Wilson, W. D. (1976) Biochemistry 15, 146; Gabbay, E. J., Adawadkar, P. D., Kapicak, L., Pearce, S., & Wilson, W. D. (1976) Biochemistry 15, 152]. The results also indicate a great dependence on the length of the side chain for intercalation of the aromatic ring. Specifically, if the side chain is long enough, and flexible enough, the aromatic ring can fully or partially intercalate, regardless of the chirality of the N-terminal amino acid alpha-carbon. However, if the side chain is too short, only partial intercalation is observed for peptides of the L-D designation, and no intercalation is observed for peptides of the L-D designation.  相似文献   

18.
On consideration that intrinsic structural weakness could affect the segment spanning the α2‐helical residues 173–195 of the PrP, we have investigated the conformational stabilities of some synthetic Ala‐scanned analogs of the peptide derived from the 180–195 C‐terminal sequence, using a novel approach whose theoretical basis originates from protein thermodynamics. Even though a quantitative comparison among peptides could not be assessed to rank them according to the effect caused by single amino acid substitution, as a general trend, all peptides invariably showed an appreciable preference for an α‐type organization, consistently with the fact that the wild‐type sequence is organized as an α‐helix in the native protein. Moreover, the substitution of whatever single amino acid in the wild‐type sequence reduced the gap between the α‐ and the β‐propensity, invariably enhancing the latter, but in any case this gap was larger than that evaluated for the full‐length α2‐helix‐derived peptide. It appears that the low β‐conformation propensity of the 180–195 region depends on the simultaneous presence of all of the Ala‐scanned residues, indirectly confirming that the N‐terminal 173–179 segment could play a major role in determining the chameleon conformational behavior of the entire 173–195 region in the PrP. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
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   

20.
We surveyed 299 high resolution, non-homologous protein crystal structures for alpha-helix lengths and capping preferences. We find that helices show a preference to have close to an integral number of turns. Helices can be usefully subdivided into either "favoured length" with 6, 7, 10, 11, 13, 14, 17, 18, 21, 22, 24, 25, 28, 29 or 31 residues, or "disfavoured length" with 8, 9, 12, 15, 16, 19, 20, 23, 26, 27 or 30 residues. Favoured length helices have their N and C-caps on the same side of the helix so they can lie on the protein surface. There is no significant difference in amino acid preferences at the N terminus between favoured and disfavoured length helices. At the C terminus, favoured length helices prefer non-polar side-chains at C4 and polar amino acid residues at C2, while disfavoured length helices prefer non-polar amino acid residues at C2. There are strong periodic trends in the likelihood of terminating a helix with a Schellman or alphaL C-capping motif. These can be rationalised by the preference for a non-polar side-chain at C3 with these motifs, favouring placing C3 on the buried side of the helix. We suggest that algorithms aiming to predict helices or C-capping in proteins should include a weight for the helix length.  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