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 0 毫秒
1.
Silva RA  Nguyen JY  Decatur SM 《Biochemistry》2002,41(51):15296-15303
The mechanism of helix stabilization or destabilization by different amino acids has been the subject of several experimental and theoretical studies; these studies suggest that large or bulky side chains may modulate helix stability by altering the hydration of the helix backbone. In this paper, we report a spectroscopic study to determine the effect of alanine to leucine substitutions on the conformation and solvation of specific segments of a model helical peptide. A 25-residue, alanine-rich, helical peptide [Ac-(AAAAK)(4)-AAAAY-NH(2) (AKA)] and its two leucine variants [Ac-LLLLK-(AAAAK)(3)-AAAAY-NH(2) (LKA) and Ac-(AAAAK)(4)-LLLLY-NH(2) (AKL)] were characterized by infrared (IR) and electronic circular dichroism (ECD) spectroscopies. Introduction of (13)C isotopes into specific, consecutive, backbone carbonyls for certain blocks of each of the peptides mentioned above allows the IR spectra to be interpreted in terms of the conformation and solvation of specific residues within the helix. These isotope-edited IR spectra of the leucine peptides do not show evidence of a decrease in the degree of backbone solvation compared to the alanines, but suggest that the peptide may adopt a distorted conformation to accommodate the larger leucine side chains at the N-terminus. These experiments demonstrate the power of isotope-edited IR in dissecting subtle changes in helix conformation at the residue level.  相似文献   

2.
Isotope-edited infrared spectroscopy has the ability to probe the segmental properties of long biopolymers. In this work, we have compared the infrared spectra of a model helical peptide ((12)C) Ac-W-(E-A-A-A-R)(6)-A-NH(2), described originally by Merutka et al. (Biochemistry 1991;30:4245-4248) and three derivatives that are (13)C labeled at the backbone carbonyl of alanines. The locations of six isotopically labeled alanines are at the N-terminal, C-terminal, and the middle two repeating units of the peptide. Variation in temperature from 1 degrees to 91 degrees C transformed the peptides from predominantly helical to predominantly disordered state. Amplitude and position of the infrared amide I' absorption bands from (12)C- and (13)C-labeled segments provided information about the helical content. Temperature dependence of infrared spectra was used to estimate segmental stability. As a control measure of overall peptide stability and helicity (independent of labeling), the temperature dependence of circular dichroism spectra in the far-UV range at identical conditions (temperature and solvent) as infrared spectra was measured. The results indicate that the central quarter of the 32 amino acids helix has the maximal helicity and stability. The midpoint of the melting curve of the central quarter of the helix is 5.4 +/- 0.8 degrees C higher than that of the termini. The N-terminal third of the helix is more helical and is 2.0 +/- 1.4 degrees C more stable than the C-terminus.  相似文献   

3.
4.
Position effect on apparent helical propensities in the C-peptide helix   总被引:1,自引:0,他引:1  
A search has been made for position effects on apparent helix propensities when another amino acid is substituted for alanine in the C-peptide helix of ribonuclease A. Three internal alanine residues (Ala4, Ala5, Ala6) are used as sites for substitution. Five amino acids, Glu, His, Arg, Lys and Phe, are substituted singly in individual peptides at each of these three positions, and the pH profiles of helix content for the substituted peptides have been determined. The effect of using an acetyl or a succinyl amino-terminal-blocking group has also been determined for each substitution. A strong position effect is found at Ala5: the helix content of the substituted peptide is significantly higher for substitution at position 5 than at positions 4 or 6 in almost all cases. The reason for the position 5 effect is unknown. The results also show that electrostatic interactions often influence substitution experiments, and they provide data on the variability of substitution experiments made with a natural sequence peptide.  相似文献   

5.
The ability of several naturally occurring substances known as osmolytes to induce helix formation in an alanine-based peptide have been investigated. As predicted by the osmophobic effect hypothesis, the osmolytes studies here do induce helix formation. Trimethylamine-N-oxide (TMAO) is the best structure-inducing osmolytes investigated here, but it is not as effective in promoting helix formation as the common cosolvent trifluoroethanol (TFE). We also provide a semiquantitative study of the ability of TMAO to induce helix formation and urea, which acts as a helix (and protein) denaturant. We find that on a molar basis, these agents are exactly counteractive as structure inducing and unfolding agents. Finally, we extend the investigations to the effects of urea and TMAO on the stability of a dimeric coiled-coil peptide and find identical results. Together these results support the tenets of the osmophobic hypothesis and highlight the importance of the polypeptide backbone in protein folding and stability.  相似文献   

6.
Repeat proteins are tandem arrays of a small structural motif, in which tertiary structure is stabilized by interactions within a repeat and between neighboring repeats. Several studies have shown that this modular structure is manifest in modular thermodynamic properties. Specifically, the global stability of a repeat protein can be described by simple linear models, considering only two parameters: the stability of the individual repeated units (H) and the coupling interaction between the units (J). If the repeat units are identical, single values of H and J, together with the number of repeated units, is sufficient to completely describe the thermodynamic behavior of any protein within a series. In this work, we demonstrate how the global stability of a repeat protein can be changed, in a predictable fashion, by modifying only the H parameter. Taking a previously characterized series of consensus tetratricopeptide repeats (TPR) (CTPRa) proteins, we introduced mutations into the basic repeating unit, such that the stability of the individual repeat unit was increased, but its interaction with neighboring units was unchanged. In other words, we increased H but kept J constant. We demonstrated that the denaturation curves for a series of such repeat proteins can be fit and additional curves can be predicted by the one-dimensional Ising model in which only H has changed from the original fit for the CTPRa series. Our results show that we can significantly increase the stability of a repeat protein by rationally increasing the stability of the units (H), whereas the interaction between repeats (J) remains unchanged.  相似文献   

7.
Fourier transform infrared spectroscopy (FTIR) can be used for conformational analysis of peptides in a wide range of environments. Measurements can be performed in aqueous solution, organic solvents, detergent micelles as well as in phospholipid membranes. Information on the secondary structure of peptides can be derived from the analysis of the strong amide I band. Orientation of secondary structural elements within a lipid bilayer matrix can be determined by means of polarized attenuated total reflectance–FTIR spectroscopy. Hydrogen–deuterium exchange can be monitored by the analysis of the, amide II band. This review gives some example of peptide systems studied by FTIR spectroscopy. Studies on alamethicin and α-aminoisobutyric acid containing peptides have shown that FTIR spectroscopy is a sensitive tool for identifying 310-helical structures. Changes in the structure of the magainins upon interaction with charged lipids were detected using FTIR spectroscopy. Tachyplesin is an example of a β-sheet containing membrane active peptide. Polarized ir spectroscopy reveals that the antiparallel β-sheet structures of tachyplesin are oriented parallel to the membrane surface. Synthesis of peptides corresponding to functionally/structurally important regions of large proteins is becoming increasingly popular. FTIR spectroscopy has been used to analyze the structure of synthetic peptides corresponding to the ion-selective pore of the voltage-gated potassium channel. In biomembrane systems these peptides adopt a highly helical structure. Under conditions, where these peptides are aggregated the presence of some intermolecular β-sheet structure can also be detected. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
Fourier transform ir spectra have been recorded for three 310‐helical and one α‐helical pentapeptides containing dehydrophenylalanine, in a thin solid film, in order to find marker bands for various secondary structures encountered in peptides containing dehydroaminoacids. The peptide solutions were deposited and dried as thin film on zinc selenide crystal surface. This convenient sampling method has provided reliable estimates of peptide secondary structure in solid state. Detailed vibrational assignments in the spectral region between 1200–1700 cm−1 are reported. In this region, peptide amide I, II, and III vibrations occur. Spectra–structure correlation has been presented based on the amide modes. Comparison of the ir spectra with available crystal structure data provides qualitative support for assignments of ir bands to 310‐helical structure and α‐helical structure in dehydrophenylalanine containing pentapeptides. Band frequency assignments for 310‐helical conformation are consistent for all three peptides. All the assignments agree closely with the theoretical predictions. The spectral differences between 310‐helical peptides and the α‐helical peptide have been highlighted. These findings demonstrate that a method based on ir spectroscopy can be developed for a useful approximation of three‐dimensional structure of dehydropeptides in solid state. © 1999 John Wiley & Sons, Inc. Biopoly 50: 595–601, 1999  相似文献   

9.
The 17-residue peptide acetylAEAAAKEAAAKEAAAKAamide, described as an autonomous folding unit (Marqusee & Baldwin, 1987), has been used to examine the effect of amino acid replacements on helix stability. Alanine residues(s) at positions 4, 9, and 14 in the peptide sequence were replaced either singly or multiply by either serine or methionine residues with solid-phase peptide synthesis. The thermal dependence of the helix/coil transition of each peptide was observed by far-ultraviolet circular dichroism. Within experimental variation, all three single replacements exhibit a common thermal transition, and all three double replacements exhibit a different common thermal transition. These results suggest that replacement of the central alanine residue in the repeat EAAAK located in the N-terminus, in the middle, or in the C-terminus of the peptide helix has the same effect on helix stability. The melting temperature of each thermal transition was estimated by assuming a linear van't Hoff plot and a change in molar ellipticity of 33,500 deg cm2 dmol-1. Such analysis indicates that each replacement of an alanine residue by a serine residue diminishes the melting temperature by 11 +/- 1 degrees C and that each replacement of an alanine residue by a methionine residue diminishes the melting temperature by 6 +/- 1 degrees C. These results suggest that the effect of these replacements on helix stability is additive.  相似文献   

10.
To understand the terminal effect of chiral residue for determining a helical screw sense, we adopted five kinds of peptides IV containing N‐ and/or C‐terminal chiral Leu residue(s): Boc–L ‐Leu–(Aib–ΔPhe)2–Aib–OMe ( I ), Boc–(Aib–ΔPhe)2–L ‐Leu–OMe ( II ), Boc–L ‐Leu–(Aib–ΔPhe)2–L ‐Leu–OMe ( III ), Boc–D ‐Leu–(Aib–ΔPhe)2–L ‐Leu–OMe ( IV ), and Boc–D ‐Leu–(Aib–ΔPhe)2–Aib–OMe ( V ). The segment –(Aib–ΔPhe)2– was used for a backbone composed of two “enantiomeric” (left‐/right‐handed) helices. Actually, this could be confirmed by 1H‐nmr [nuclear Overhauser effect (NOE) and solvent accessibility of NH resonances] and CD spectroscopy on Boc–(Aib–ΔPhe)2–Aib–OMe, which took a left‐/right‐handed 310‐helix. Peptides IV were also found to take 310‐type helical conformations in CDCl3, from difference NOE measurement and solvent accessibility of NH resonances. Chloroform, acetonitrile, methanol, and tetrahydrofuran were used for CD measurement. The CD spectra of peptides IIII in all solvents showed marked exciton couplets with a positive peak at longer wavelengths, indicating that their main chains prefer a left‐handed screw sense over a right‐handed one. Peptide V in all solvents showed exciton couplets with a negative peak at longer wavelengths, indicating it prefers a right‐handed screw sense. Peptide IV in chloroform showed a nonsplit type CD pattern having only a small negative signal around 280 nm, meaning that left‐ and right‐handed helices should exist with almost the same content. In the other solvents, peptide IV showed exciton couplets with a negative peak at longer wavelengths, corresponding to a right‐handed screw sense. From conformational energy calculation and the above 1H‐nmr studies, an N‐ or C‐terminal L ‐Leu residue in the lowest energy left‐handed 310‐helical conformation was found to take an irregular conformation that deviates from a left‐handed helix. The positional effect of the L ‐residue on helical screw sense was discussed based on CD data of peptides IV and of Boc–(L ‐Leu–ΔPhe)n–L ‐Leu–OMe (n = 2 and 3). © 1999 John Wiley & Sons, Inc. Biopoly 49: 551–564, 1999  相似文献   

11.
Summary Solid-state NMR spectroscopy was used to determine the orientations of two amphipathic helical peptides associated with lipid bilayers. A single spectral parameter provides sufficient orientational information for these peptides, which are known, from other methods, to be helical. The orientations of the peptides were determined using the15N chemical shift observed for specifically labeled peptide sites. Magainin, an antibiotic peptide from frog skin, was found to lie in the plane of the bilayer. M2, a helical segment of the nicotinic acetylcholine receptor, was found to span the membrane, perpendicular to the plane of the bilayer. These findings have important implications for the mechanisms of biological functions of these peptides.  相似文献   

12.
13.
Addition of Na(2)CO(3) to almost salt-free DNA solution (5.10(-5)M EDTA, pH=5.7, T(m)=26.5 degrees C) elevates both pH and the DNA melting temperature (T(m)) if Na(2)CO(3) concentration is less than 0.004 M. For 0.004 M Na(2)CO(3), T(m)=58 degrees C is maximal and pH=10.56. Further increase in concentration gives rise to a monotonous decrease in T(m) to 37 degrees C for 1M Na(2)CO(3) (pH=10.57). Increase in pH is also not monotonous. The highest pH=10.87 is reached at 0.04 M Na(2)CO(3) (T(m)=48.3 degrees C). To reveal the cause of this DNA destabilization, which happens in a narrow pH interval (10.56/10.87) and a wide Na(2)CO(3) concentration interval (0.004/1M), a procedure has been developed for determining the separate influences on T(m) of Na(+), pH, and anions formed by Na(2)CO(3) (HCO(3)(-) and CO(3)(2-)). Comparison of influence of anions formed by Na(2)CO(3) on DNA stability with Cl(-) (anion inert to DNA stability), ClO(4)(-) (strong DNA destabilizing "chaotropic" anion) and OH(-) has been carried out. It has been shown that only Na(+) and pH influence T(m) in Na(2)CO(3) solution at concentrations lower than 0.001 M. However, the T(m) decrease with concentration for [Na(2)CO(3)]>/=0.004 M is only partly caused by high pH=10.7. Na(2)CO(3) anions also exert a strong destabilizing influence at these concentrations. For 0.1M Na(2)CO(3) (pH=10.84, [Na(+)]=0.2M, T(m)=42.7 degrees C), the anion destabilizing effect is higher 20 degrees C. For NaClO(4) (ClO(4)(-) is a strong "chaotropic" anion), an equal anion effect occurs at much higher concentrations approximately 3M. This means that Na(2)CO(3) gives rise to a much stronger anion effect than other salts. The effect is pH dependent. It decreases fivefold at neutral pH after addition of HCl to 0.1M Na(2)CO(3) as well as after addition of NaOH for pH greater than 11.2.  相似文献   

14.
A microprobe system has been developed that can record Raman spectra from as little as 2 microL of solution containing only micrograms of biological pigments. The apparatus consists of a liquid nitrogen (l-N2)-cooled cold stage, an epi-illumination microscope, and a substractive-dispersion, double spectrograph coupled to a l-N2-cooled CCD detector. Experiments were performed on native bovine rhodopsin, rhodopsin expressed in COS cells, and four rhodopsin mutants: Glu134 replaced by Gln (E134Q), Glu122 replaced by Gln (E122Q), and Glu113 replaced by Gln (E113Q) or Ala (E113A). Resonance Raman spectra of photostationary steady-state mixtures of 11-cis-rhodopsin, 9-cis-isorhodopsin, and all-trans-bathorhodopsin at 77 K were recorded. The Raman spectra of E134Q and the wild-type are the same, indicating that Glu134 is not located near the chromophore. Substitution at Glu122 also does not affect the C = NH stretching vibration of the chromophore. The fingerprint and Schiff base regions of the Raman spectra of the 380-nm, pH 7 forms of E113Q and E113A are characteristic of unprotonated retinal Schiff bases. The C = NH modes of the approximately 500-nm, pH 5 forms of E113Q and E113A in H2O (D2O) are found at 1648 (1629) and 1645 (1630) cm-1, respectively. These frequencies indicate that the protonated Schiff base interacts more weakly with its protein counterion in the Glu113 mutants than it does in the native pigment. Furthermore, perturbations of the unique bathorhodopsin hydrogen out-of-plane (HOOP) vibrations in E113Q and E113A indicate that the strength of the protein perturbation near C12 is weakened compared to that in native bathorhodopsin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
16.
The rates at which a peptide hexamer and a peptide octamer interconvert between left- and right-handed helical forms in CD2Cl2 solution have been characterized by 13C dynamic NMR (DNMR) spectroscopy. The peptide esters studied are Fmoc-(Aib)n-OtBu (n = 6 and 8), where Fmoc is 9-fluorenylmethyoxycarbonyl and Aib is the strongly helix-forming residue alpha-aminoisobutyric acid. Because the Aib residue is itself achiral, homooligomers of this residue form a 50/50 mixture of enantiomeric 3(10)-helices in solution. It has been demonstrated (R.-P. Hummel, C. Toniolo, and G. Jung, Angewandte Chemie International Edition, 1987, Vol. 26, pp. 1150-1152) that oligomers of Aib interconvert on the millisecond timescale. We have performed lineshape analysis of 13C-NMR spectra collected for our peptides enriched with 13C at a single residue. Rate constants for the octamer range from 6 s(-1) at 196 K to about 56,500 s(-1) at 320 K. At all temperatures, the hexamer interconverts about three times faster than the octamer. Eyring plots of the data reveal experimentally indistinguishable DeltaH++ values for the hexamer and octamer of 37.8 +/- 0.6 and 37.6 +/- 0.4 kJ mol(-1) respectively. The difference in the rates of interconversion is dictated by entropic factors. The hexamer and octamer exhibit negative DeltaS++ values of -29.0(-1) +/- 2.5 and -37.3 +/- 1.7 J K(-1) mol(-1), respectively. A mechanism for the helix-helix interconversion is proposed. and calculated DeltaG++ values are compared to the estimate for a decamer undergoing a helix-helix interconversion.  相似文献   

17.
DNA binding by a 29-residue, monomeric, GCN4 basic region peptide, GCN4br, as well as by peptide br-C, a monomeric basic-region analogue that is helix stabilized at its C-terminal end by a Lys25. Asp29 side-chain lactam-bridged alanine-rich sequence, was studied at 25 C in an aqueous buffer containing 100 mm NaCl. Mixing of both peptides with duplex DNA containing the cAMP-responsive element (CRE) was accompanied by significant helix stabilization in the peptides, whereas mixing of the peptides with duplex DNA containing a scrambled CRE site was not. Peptide NBD-br-C was synthesized as a fluorescent probe to evaluate these peptide-DNA interactions further. Quantitative analysis of the fluorescence quenching of peptide NBD-br-C by CRE half-site DNA indicated the formation of a 1:1 complex with a dissociation constant of 1.41 +/- 0.22 microm. Competitive displacement fluorescence assays of CRE half-site binding gave dissociation constants of 0.65 +/- 0.09 microm for peptide br-C and 3.9 +/- 0.5 microM for GCN4br, which corresponds to a free energy difference of 1.1 kcal/mol that is attributed to the helix stabilization achieved in peptide br-C. This result indicates that helix initiation by the alpha-helical leucine zipper dimerization motif in native bzip proteins, such as GCN4, contributes significantly to the affinity of basic region peptides for their recognition sites on DNA. Our fluorescence assay should also prove useful for determining dissociation constants for CRE binding by other GCN4 basic region analogues under equilibrium conditions and physiological salt concentrations.  相似文献   

18.
Here, we show that the energetics of transmembrane helix heterodimer formation can be characterized in liposomes using F?rster resonance energy transfer (FRET). We present the theory and the protocol for measuring the free energy of heterodimerization, and the total (hetero and homo-dimeric) dimer fraction. We use the presented methodology to determine the propensity for heterodimer formation between wild-type fibroblast growth factor receptor 3 (FGFR3) transmembrane domain and the Ala391Glu mutant, linked to Crouzon syndrome with acanthosis nigricans.  相似文献   

19.
P Pramanik  L S Kan 《Biochemistry》1987,26(13):3807-3812
Temperature variation studies of the imido proton NMR and 31P NMR resonances of the self-associated d(C-C-A-A-G-A-T-T-G-G) and d[C-C-A-A-G-p(Et)-A-T-T-G-G] duplexes (both the R and S diastereoisomers) and the heteroduplexes formed with their complementary strand d(C-C-A-A-T-C-T-T-G-G) were carried out in aqueous solution. Results demonstrate that phosphate backbone ethylation did not disrupt the interstrand hydrogen bonding involved in double-helix formation but perturbed the helix. The S isomer perturbed the duplex more than the R isomer. The line broadening patterns and faster fraying motion in the alkylated duplexes compared to those in the nonalkylated duplexes indicate that the perturbation introduced in the middle propagates along the backbone to the end of the duplex.  相似文献   

20.
Reducing a CO to a CH2 moiety in a peptide bond destroys the ability of the peptide link to act as a proton acceptor in a hydrogen-bonded structure. Here, this modification is introduced into different positions of the helical peptide, acetyl-WGG(RAAAA)4R-amide, and the melting of these peptides is followed using CD. Effects of this modification on helical peptides are compared to our previous N-methylation studies [C. F. Chang and M.H. Zehfus (1996) Biopolymers, Vol. 40, pp. 609–616]. While the experiments were designed to remove the same hydrogen bond from the peptide, no consistent results are obtained between these two modifications. This result suggests that these modifications not only break the backbone hydrogen bonds, but also involve other destabilizing effects. When our data is analyzed using different helix-coil transition models, the results show that as the models increase in complexity the energy associated with a single residue modification increases. Unfortunately, the most detailed dichroic model, which should best describe this system, works for only one peptide. Apparently, the models need to be further improved to better mimic our system. © 1998 John Wiley & Sons, Inc. Biopoly 46: 181–193, 1998  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