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 78 毫秒
1.
We used frequency-domain measurements of fluorescence resonance energy transfer to measure the distribution of distances between Trp-19 of melittin and a 1-dimethylamino-5-sulfonylnaphthalene (dansyl) residue on the N-terminal-alpha-amino group. Distance distributions were obtained for melittin free in solution and when complexed with calmodulin (CaM), troponin C (TnC), or palmitoyloleoyl-L-alpha-phosphatidylcholine (POPC) vesicles. A wide range of donor (Trp-19)-to-acceptor (dansyl) distances was found for free melittin, which is consistent with that expected for the random coil state, characterized by a Gaussian width (full width at half maxima) of 28.2 A. In contrast, narrow distance distributions were found for melittin complexed with CaM, 8.2 A, or with POPC vesicles, 4.9 A. A somewhat wider distribution was found for the melittin complex with TnC, 12.8 A, suggesting the presence of heterogeneity in the mode of binding between melittin and TnC. For all the complexes the mean Trp-19 to dansyl distance was near 20 A. This value is somewhat smaller than expected for the free alpha-helical state of melittin, suggesting that binding with CaM or TnC results in a modest decrease in the length of the melittin molecule.  相似文献   

2.
Delta-Hemolysin forms a 1:1 complex with Ca2+ -liganded calmodulin. Probably because of the pronounced tendency of delta-hemolysin to self-associate, the apparent binding affinity is much less than that for melittin. Complex formation is reflected by an increase in quantum yield of Trp-15 of delta-hemolysin and by increased shielding from acrylamide quenching. There is, however, no indication of a change in peptide molecular ellipticity. The binding of 2-toluidinyl-naphthalene-6-sulfonate is reduced by complex formation, suggesting the involvement of a hydrophobic region. Complex formation also blocks the proteolysis by trypsin of the bond between residues 77 and 78. The time decays of fluorescence intensity and anisotropy for tryptophan are multiexponential for both free and complexed delta-hemolysin; the average decay time for intensity is substantially increased for the complex. The localized mobility of tryptophan is greatly reduced in the complex. Complex formation appears to involve both the C-terminal lobe and the connecting strand of calmodulin.  相似文献   

3.
The dynamics in isotopic solvents of selectively 13C labeled synthetic melittin and three analogues have been investigated by using NMR and fluorescence techniques both separately and in combination. In conjunction with the "model-free" approach to interpretation of NMR relaxation data [Lipari, G., & Szabo, A. (1982) J. Am. Chem. Soc. 104, 4546-4570], the availability of steady-state fluorescence anisotropy and lifetime data augment T1, T2, and NOE data to provide quantitative information about fluorophore dynamics in these peptides. A method is presented for using combined fluorescence and NMR data to obtain technique- and model-independent values for parameters describing local motion of 13C-labeled fluorophores in peptides and proteins. The dynamics of melittin and melittin analogues are found to be consistent with structural characteristics inferred from CD, fluorescence, and NMR spectral information presented in the preceding paper (Weaver et al., 1989). In particular, the mobility of the random coil peptide monomers is shown to be quite similar, while side-chain as well as peptide backbone motion in the aggregated or oligomeric species differs markedly among the analogues. For melittin itself, experimentally determined overall rotational correlation times for the monomer and tetramer agree very well with values predicted on the basis of solvent-accessible protein surface area. The local dynamics of selectively 13C-labeled Trp-19 and Gly-12 residues of melittin are also found to be consistent with peptide structure. In random coil melittin monomer, a specific model for the motion indicates that the Trp side chain moves through an approximate angle of +/- 71 degrees about the beta-gamma bond with a correlation time of 159 +/- 24 ps. In melittin tetramer, the indole moiety is spatially more confined with a flip angle of +/- 37 degrees, yet demonstrates an increased rate of motion with a correlation time of 56 +/- 8 ps. The constrained mobility of the Trp-19 side chain is consistent with motional constraints inferred from the X-ray structure of melittin tetramer. These results show that protein side-chain motion, even of moieties as large as indole, can occur on the picosecond time scale and that these motions are reasonably similar to those inferred from molecular dynamics simulations.  相似文献   

4.
The shielding of lysine groups from acetylation by acetic anhydride has been used to identify the regions of calmodulin in contact with melittin in the 1:1 complex. The estimation of the degree of acetylation was done by examining cyanogen bromide and cyanogen bromide/trypsin digests by mass spectrometry. Evidence was obtained that lysines-21, -75, and -148 are protected to some extent, with the implication that both the N- and C-terminal lobes and the connecting strand are involved in the interaction.  相似文献   

5.
The interaction of calmodulin with melittin   总被引:1,自引:0,他引:1  
Studies utilizing the interaction of melittin with the 1-106 fragment of calmodulin, the protection of calmodulin from tryptic digestion by melittin, and the interaction of the carbocyanine dye Stains-all with the calmodulin-melittin complex have indicated that complex formation of calmodulin with melittin involves the alpha-helical connecting bridge joining the N- and C-terminal lobes of calmodulin.  相似文献   

6.
To design melittin (ME) analogues that are not cytotoxic against mammalian cells but which possessing potent antimicrobial activity, we synthesized a ME analogue (ME-w) in which the Trp-19 residue of ME was replaced by a Trp-peptoid residue (Nhtrp). ME-w exhibited similar antimicrobial activity compared to ME against the tested six bacteria and C. albicans. However, it was much less cytotoxic against the hRBCs and HeLa and NIH-3T3 cells than ME. Tryptophan fluorescence and CD spectra revealed that the Trp-19 --> Nhtrp substitution in ME contributed to a much lower helical assembly in an aqueous environment and structural flexibility and exterior localization to zwitterionic membrane which modulates its selectivity toward bacterial cells.  相似文献   

7.
Lipopolysaccharide (LPS), the major constituent of the outer membrane of Gram-negative bacteria, is the very first site of interactions with the antimicrobial peptides. In this work, we have determined a solution conformation of melittin, a well-known membrane active amphiphilic peptide from honey bee venom, by transferred nuclear Overhauser effect (Tr-NOE) spectroscopy in its bound state with lipopolysaccharide. The LPS bound conformation of melittin is characterized by a helical structure restricted only to the C-terminus region (residues A15-R24) of the molecule. Saturation transfer difference (STD) NMR studies reveal that several C-terminal residues of melittin including Trp19 are in close proximity with LPS. Isothermal titration calorimetry (ITC) data demonstrates that melittin binding to LPS or lipid A is an endothermic process. The interaction between melittin and lipid A is further characterized by an equilibrium association constant (Ka) of 2.85 × 106 M− 1 and a stoichiometry of 0.80, melittin/lipid A. The estimated free energy of binding (ΔG0), − 8.8 kcal mol− 1, obtained from ITC experiments correlates well with a partial helical structure of melittin in complex with LPS. Moreover, a synthetic peptide fragment, residues L13-Q26 or mel-C, derived from the C-terminus of melittin has been found to contain comparable outer membrane permeabilizing activity against Escherichia coli cells. Intrinsic tryptophan fluorescence experiments of melittin and mel-C demonstrate very similar emission maxima and quenching in presence of LPS micelles. The Red Edge Excitation Shift (REES) studies of tryptophan residue indicate that both peptides are located in very similar environment in complex with LPS. Collectively, these results suggest that a helical conformation of melittin, at its C-terminus, could be an important element in recognition of LPS in the outer membrane.  相似文献   

8.
Micellar complexes of melittin with fully deuterated detergents have been studied by high resolution 1H nuclear magnetic resonance (NMR). The synthesis of deuterated micelles is described and it is shown that the 1H NMR spectrum of micelle-bound melittin is well resolved and suitable for detailed analysis by conventional high-resolution NMR methods. A preliminary characterization of micelle-bound melittin shows that interaction with the micelle results in different conformational and dynamic features for the hydrophobic and hydrophilic regions of the melittin amino acid sequence. The present experiments on melittin and preliminary results with other polypeptides and proteins demonstrate that in favourable cases high-resolution 1H NMR studies of the complexes formed between membrane proteins and deuterated micelles provides a viable method for conformational studies of membrane-bound proteins.  相似文献   

9.
We have previously shown that melittin, a bee venom peptide, potently inhibited the catalytic and transport functions of rabbit gastric (H+ + K+)ATPase. A radioactive photoaffinity analog of melittin, ([125I]azidosalicylyl melittin), labeled the (H+ + K+)ATPase. These results suggested that melittin exerted inhibitory effects through direct interaction with the (H+ + K+)ATPase. In this study we attempt to define the melittin-binding domain of the (H+ + K+)ATPase using conformation-dependent proteolytic fragmentation of [125I]azidosalicylyl melittin-labeled hog gastric (H+ + K+)ATPase. In the presence of KCl (E2 form) the 95,000-Da [125I]-azidosalicylyl melittin-labeled (H+ + K+)ATPase was cleaved by trypsin to a 40,000-Da NH2-terminal tryptic fragment and a 56,000-Da COOH-terminal fragment through cleavage at Arg 454 of the (H+ + K+)ATPase. The 40,000-Da fragment was labeled by [125I]-azidosalicylyl melittin. The 56,000-Da fragment was not labeled. When unmodified (H+ + K+)ATPase was trypsinized in the presence of KCl, and the fragments were then reacted with [125I]azidosalicylyl melittin, similar tryptic fragmentation results were obtained. In the absence of KCl (E1 form), the 56,000- and 40,000-Da fragments did not accumulate. Chymotryptic hydrolysis of [125I]azidosalicylyl melittin-labeled (H+ + K+)-ATPase was very slow in the presence of KCl (E2 form). In the absence of KCl (E1 form), chymotryptic hydrolysis was more rapid, with accumulation of a major 42,000-Da fragment which was radiolabeled. The melittin-binding region on the (H+ + K+)ATPase is N-terminal to Arg 454 of the (H+ + K+)ATPase. This region is known to contain the aspartyl phosphate residue (Asp 385), the site of phosphoenzyme formation on the (H+ + K+)ATPase. Melittin is also known to bind to calmodulin and other proteins. Another known calmodulin-binding peptide with a different sequence but similar structure, Trp-3, (Leu-Lys-Trp-Lys-Lys-Leu-Leu-Lys-Leu-Leu-Lys-Lys-Leu-Leu-Lys-Leu-Gly) also inhibited the (H+ + K+)ATPase and label incorporation by [125I]azidosalicylyl melittin. These Trp-3 results suggested that the (H+ + K+)ATPase contains a peptide-binding domain which is similar to the peptide-binding domains found on other melittin-binding proteins.  相似文献   

10.
Assignments have been obtained for most of the 1H-NMR lines of melittin bound to fully deuterated dodecylphosphocholine micelles by combined use of two-dimensional spin echo correlated spectroscopy and one-dimensional NMR methods. Nuclear Overhauser enhancement measurements showed that the mobility of the entire polypeptide chain is reduced by binding of melittin to the detergent micelle and that the amino-terminal and carboxy-terminal halves of the primary structure constitute separate, compact domains within the conformation of micelle-bound melittin. p2H titration experiments showed that the presence of positive charges on the four amino groups of melittin had little influence on the conformation of the micelle-bound polypeptide. Titration of tetrameric melittin with detergent provided evidence that melittin assumes similar conformations as a self-aggregated tetramer and as a monomer bound to micelles.  相似文献   

11.
The conformation of a melittin—inhibitor complex was studied by solution NMR, solid-state NMR, and circular dichroism. In solution, binding was studied by titrating inhibitor against melittin in dimethyl sulfoxide, methanol, aqueous buffer, and dodecylphosphocholine micelles. The change in chemical shift of Trp19 resonances and the formation of a precipitate at 1:1 molar ratio indicated that the inhibitor was bound to melittin. Solid-state NMR also showed a change in chemical shift of two labeled carbons of melittin near Pro14 and a change in 1H T 1 relaxation times when complexed with inhibitor. Rotational resonance experiments of melittin labeled in the proline region indicated a change in conformation for melittin complexed with inhibitor. This observation was also supported by circular dichroism measurements, indicating a reduction in α-helical structure for increasing ratios of inhibitor bound to melittin.  相似文献   

12.
The N-terminal α-amino groups of β1-bungarotoxin (β1-Bgt) fromBungarus multicinctus venom were modified with trinitrobenzene sulfonic acid and the modified derivative was separated by high performance liquid chromatography. The trinitrophenylated (TNP) derivative contained two TNP groups at the α-amino groups of A chain and B chain and showed a marked decrease in enzymatic activity. Methionine residues at positions 6 and 8 of the A chain were oxidized with chloramine T or cleaved with cyanogen bromide to remove the N-terminal octapeptide. Oxidation of methionine residues and removal of the N-terminal octapeptide caused a precipitous decrease in enzymatic activity, whereas antigenicity remained unchanged. The presence of dihexanoyllecithin influenced the interaction between β1-Bgt and 8-antilinonaphthalene sulfonate (ANS) and revealed that β1-Bgt consists of two types of ANS-binding sites, one at the substrate binding site of the A chain and the other might be at the B chain. The modified derivatives still retained their affinity for Ca2+ and ANS, indicating that the N-terminal region is not involved in Ca2+ and substrate binding. A fluorescence study revealed that the α-amino group of the A chain was in the vicinity of substrate binding site and that the TNP α-amino groups were in proximity to Trp-19 of the A chain. In addition, the study showed that the N-terminal region is important for stabilizing the architectural environment of Trp-19. The results, together with the proposal that Trp-19 of the A chain is involved in substrate binding, suggest that the N-terminal region of the A chain plays a crucial role in maintaining a functional active site for β1-Bgt.  相似文献   

13.
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   

14.
In the prokaryotic potassium channel KcsA activation gating at the inner bundle gate is followed by C-type inactivation at the selectivity filter. Entry into the C-type inactivated state has been directly linked to the strength of the H-bond interaction between residues Glu-71 and Asp-80 behind the filter, and is allosterically triggered by the rearrangement of the inner bundle gate. Here, we show that H-bond pairing between residues Trp-67 and Asp-80, conserved in most K+ channels, constitutes another critical interaction that determines the rate and extent of KcsA C-type inactivation. Disruption of the equivalent interaction in Shaker (Trp-434-Asp-447) and Kv1.2 (Trp-366-Asp-379) leads also to modulation of the inactivation process, suggesting that these residues also play an analogous role in the inactivation gating of Kv channels. The present results show that in KcsA C-type inactivation gating is governed by a multipoint hydrogen-bond network formed by the triad Trp-67-Glu71-Asp-80. This triad exerts a critical role in the dynamics and conformational stability of the selectivity filter and might serve as a general modulator of selectivity filter gating in other members of the K+ channel family.  相似文献   

15.
The filamentous bacteriophage Pf3 consists of a covalently closed DNA single strand of 5833 nucleotides sheathed by approximately 2500 copies of a 44-residue capsid subunit. The capsid subunit contains a single tryptophan residue (Trp-38), which is located within the basic C-terminal sequence (-RWIKAQFF) and is essential for virion assembly in vivo. Polarized Raman microspectroscopy has been employed to determine the orientation of the Trp-38 side chain in the native virus structure. The polarized Raman measurements show that the plane of the indolyl ring is tilted by 17 degrees from the virion axis and that the indolyl pseudo-twofold axis is inclined at 46 degrees to the virion axis. Using the presently determined orientation of the indolyl ring and side-chain torsion angles, chi(1) (N-C(alpha)-C(beta)-C(gamma)) and chi(2,1) (C(alpha)-C(beta)-C(gamma)-C(delta1)), we propose a detailed molecular model for the local structure of Trp-38 in the Pf3 virion. The present Pf3 model is consistent with previously reported Raman, ultraviolet-resonance Raman and fluorescence results suggesting an unusual environment for Trp-38 in the virion assembly, probably involving an intrasubunit cation-pi interaction between the guanidinium moiety of Arg-37 and the indolyl moiety of Trp-38. Such a C-terminal Trp-38/Arg-37 interaction may be important for the stabilization of a subunit conformation that is required for binding to the single-stranded DNA genome during virion assembly.  相似文献   

16.
The interaction of melittin with calmodulin and its tryptic fragments   总被引:4,自引:0,他引:4  
Melittin has been found to interact with both the N- and C-terminal half-molecules of calmodulin, as well as the intact molecule, in the presence of Ca2+. The interaction results in a major change in the microenvironment of Trp-19, which is in a more nonpolar, solvent-shielded, and immobilized microenvironment in the complex. The properties of Tyr-99 and Tyr-138 of calmodulin are altered by complex formation. From measurements of the efficiencies of radiationless energy transfer from Trp-19 to the nitro derivatives of Tyr-99 and/or Tyr-138, it is concluded that Trp-19 is located in proximity to the C-terminal lobe of calmodulin in the complex.  相似文献   

17.
The conformation of a melittin—inhibitor complex was studied by solution NMR, solid-state NMR, and circular dichroism. In solution, binding was studied by titrating inhibitor against melittin in dimethyl sulfoxide, methanol, aqueous buffer, and dodecylphosphocholine micelles. The change in chemical shift of Trp19 resonances and the formation of a precipitate at 1:1 molar ratio indicated that the inhibitor was bound to melittin. Solid-state NMR also showed a change in chemical shift of two labeled carbons of melittin near Pro14 and a change in 1H T 1 relaxation times when complexed with inhibitor. Rotational resonance experiments of melittin labeled in the proline region indicated a change in conformation for melittin complexed with inhibitor. This observation was also supported by circular dichroism measurements, indicating a reduction in -helical structure for increasing ratios of inhibitor bound to melittin.  相似文献   

18.
Structure and activity of D-Pro14 melittin   总被引:1,自引:0,他引:1  
D-Pro14 melittin was synthesized to investigate the effect of increasing the angle of the bend in the hinge region between the helical segments of the molecule. Structural analysis by nuclear magnetic resonance indicated that, in methanol, the molecule consisted of two helices separated at Pro14, as in melittin. However, the two helices in D-Pro14 melittin were laterally displaced relative to each other by approximately 7 , and in addition, there was a small rotation of the carboxyl-terminal helix relative to the amino-terminal helix around the long axis of the molecule. The peptide had less than 5% of the cytolytic activity of melittin. Modification of Arg22 with the 2,2,5,7,8-pentamethyl-chroman-6-sulphonyl (pmc) group restored hemolytic activity to close to that of unmodified melittin. Replacement of Arg22 with Phe was less effective in restoring hemolytic activity. Electron-paramagnetic resonance studies suggest that there is a positive correlation between hemolytic activity of the peptides and interaction with phospholipid bilayers.Donald E. Rivett: Deceased, 5 April 1998  相似文献   

19.
Summary This paper describes experiments designed to explore interactions between human red blood cell membranes and melittin, the main component of bee venom. We found that melittin binds to human red cell membranes suspended in isotonic NaCl at room temperature, with an apparent dissociation constant of 3×10–8 m and maximum binding capacity of 1.8×107 molecules/cell. When about 1% of the melittin binding sites are occupied, cell lysis can be observed, and progressive, further increases in the fraction of the total sites occupied lead to progressively greater lysis in a graded manner. 50% lysis occurs when there are about 2×106 molecules bound to the cell membrane. For any particular extent of melittin binding, lysis proceeds rapidly during the first few minutes but then slows and stops so that no further lysis occurs after one hour of exposure of cells to melittin. The graded lysis of erythrocytes by melittin is due to complete lysis of some of the cells, since both the density and the hemoglobin content of surviving, intact cells in a suspension that has undergone graded melittin lysis are similar to the values observed in the same cells prior to the addition of melittin. The cells surviving graded melittin lysis have an increased Na and reduced K, proportional to the extent of occupation of the melittin binding sites. Like lysis, Na accumulation and K loss proceed rapidly during the first few minutes of exposure to melittin but then stops so that Na, K and hemoglobin content of the cells remain constant after the first hour. These kinetic characteristics of both lysis and cation movements suggest that melittin modifies the permeability of the red cell membrane only for the first few minutes after the start of the interaction. Direct observation of cells by Nomarsky optics revealed that they crenate, become swollen and lyse within 10 to 30 sec after these changes in morphology are first seen. Taken together, these results are consistent with the idea that melittin produces lysis of human red cells at room temperature by a colloid osmotic mechanism.  相似文献   

20.
When studying bee venom melittin in an ordered tetrameric form we found a shift of the fluorescence spectrum to a longer wavelength with a rise in temperature above 25 degrees C. The application of the methods of circular dichroism, temperature-perturbation difference spectrophotometry, gel filtration, ionic quenching and polarization of Trp-19 fluorescence argues against the possibility of dissociation and change in conformation with the rise in temperature. The spectral shifts are, probably, caused by dipole-orientational structural relaxation of the tryptophanyl environment in the excited state at nanosecond times. The dependence of the fluorescence spectrum on the excitation wavelength was found to be a function of temperature. This function was applied to determine the dipole-orientational relaxation times.  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