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.
The molecular properties of substance P (SP) (Arg-Pro-Lys-Pro-Gln-Gln-Phe-Phe-Gly-Leu-Met amide) and three of its antagonists were derived by measuring the Gibbs adsorption isotherm, providing information on the surface activity, the molecular shape, and the pK values of the different molecules. The following three antagonists were investigated: [D-Arg1,D-Pro2,D-Trp7,9,Leu11]SP, ANT I; [D-Arg1,D-Trp7,9,Leu11]SP, ANT II and [D-Pro2,D-Trp7,9]SP, ANT III. SP is only moderately surface active. The amino acid substitutions lead, however, to an increased surface activity of the antagonists. From the concentration dependence of the surface activity it was possible to quantify the packing characteristics of the individual neuropeptides. SP shows cross-sectional areas of 300 +/- 5 A2 to 240 +/- 5 A2 (pH 5 to 8, 154 mM NaCl) at concentrations below 10(-5) M, i.e., in the physiological concentration range, indicating a folded SP conformation. Upon increasing the packing density to concentrations larger than 10(-5) M the surface area was only half as large (148 +/- 5 A2 to 124 +/- 3 A2) suggesting now a relatively extended conformation of the SP molecule with its long molecular axis perpendicular to the air/water interface. In contrast, the three antagonists were characterized by surface areas of 147 +/- 3 A2 to 126 +/- 3 A2 which were almost independent of concentration. The antagonists thus adopt a relatively extended conformation in the whole concentration range measured. This is further supported by computer modelling which shows that the antagonists are motionally restricted and can adopt neither a bent nor a alpha-helical conformation. The surface activity of the neuropeptides was dependent on the pH of the solution. At low peptide concentrations (about 10(-6) M) it was possible to resolve and determine the pK values of all individual charged amino acid side chains. The pK values observed for the neuropeptides were about two pK units lower than those of the free amino acids in solution. The pK shifts of the neuropeptides at the air/water interface are explained in terms of the Gouy-Chapman theory. SP and its antagonists bind to lipid bilayers in the order of their surface activity. While the binding of SP is mainly due to electrostatic interactions, hydrophobic peptide-lipid interactions contribute to the binding of the antagonists.  相似文献   

2.
The dermaseptins S are closely related peptides with broad-spectrum antibacterial activity that are produced by the skin of the South American hylid frog, Phyllomedusa sauvagei. These peptides are polycationic (Lys-rich), alpha-helical, and amphipathic, with their polar/charged and apolar amino acids on opposing faces along the long axis of the helix cylinder. The amphipathic alpha-helical structure is believed to enable the peptides to interact with membrane bilayers, leading to permeation and disruption of the target cell. We have identified new members of the dermaseptin S family that do not resemble any of the naturally occurring antimicrobial peptides characterized to date. One of these peptides, designated dermaseptin S9, GLRSKIWLWVLLMIWQESNKFKKM, has a tripartite structure that includes a hydrophobic core sequence encompassing residues 6-15 (mean hydrophobicity, +4.40, determined by the Liu-Deber scale) flanked at both termini by cationic and polar residues. This structure is reminiscent of that of synthetic peptides originally designed as transmembrane mimetic models and that spontaneously become inserted into membranes [Liu, L., and Deber, C. M. (1998) Biopolymers 47, 41-62]. Dermaseptin S9 is a potent antibacterial, acting on gram-positive and gram-negative bacteria. The structure of dermaseptin S9 in aqueous solution and in TFE/water mixtures was analyzed by circular dichroism and two-dimensional NMR spectroscopy combined with molecular dynamics calculations. Dermaseptin S9 is aggregated in water, but a monomeric nonamphipathic alpha-helical conformation, mostly in residues 6-21, is stabilized by the addition of TFE. These results, combined with membrane permeabilization assays and surface plasmon resonance analysis of the peptide binding to zwitterionic and anionic phospholipid bilayers, demonstrate that spatial segregation of hydrophobic and hydrophilic/charged residues on opposing faces along the long axis of a helix is not essential for the antimicrobial activity of cationic alpha-helical peptides.  相似文献   

3.
The pi-helix is a secondary structure with 4.4 amino acids per helical turn. Although it was proposed in 1952, no experimental support for its existence was obtained until the mid-1980s. While short peptides are unlikely to assume a marginally stable secondary structure spontaneously, they might do so in the presence of appropriate structural constraints. In this paper, we describe a peptide that is designed to assume a pi-helical conformation when stabilized by cetyltrimethylammonium bromide (CTAB) micelles and Zn(2+). In the designed peptide, lipophilic amino acids are placed such that it would be amphiphilic in the pi-helical, but not in the alpha-helical, conformation. Also, two His residues are incorporated with i, i + 5 spacing, designed to allow binding of Zn(2+) in a pi-helical but not an alpha-helical conformation. The peptide was found to form moderately stable monolayers at the air-water interface, with a collapse pressure that almost doubled when there was Zn(2+) in the subphase. Also, CTAB micelles induced a marked increase in the helicity of the peptide. In 50% TFE, the peptide had a CD spectrum consistent with an alpha-helical structure. The addition of 1 mM Zn(2+) to this solvent caused a saturable decline in ellipticity to approximately half of its original value. The peptide also bound Zn(2+) when it was bound to CTAB micelles, with Zn(2+) again inducing a decrease in ellipticity. The peptide had slightly greater affinity for Zn(2+) in the presence of the CTAB than in a 50% TFE solution (K(d) = 3.1 x 10(-4) M in CTAB and 2.3 x 10(-4) M in TFE). van't Hoff analysis indicated that thermal denaturation of the peptide in 50% TFE containing 1 mM Zn(2+) was associated with both enthalpic and entropic changes that were greater than those in the absence of Zn(2+). These observations are all consistent with the proposal that the peptide assumed a pi-helical conformation in the presence of Zn(2+) and CTAB micelles, and has allowed the stability of this rare conformation to be assessed.  相似文献   

4.
The first 25 amino acids of the coat protein of cowpea chlorotic mottle virus are essential for binding the encapsidated RNA. Although an alpha-helical conformation has been predicted for this highly positively charged N-terminal region [Argos, P. (1981) Virology 110, 55-62; Vriend, G., Verduin, B. J. M., & Hemminga, M. A. (1986) J. Mol. Biol. 191, 453-460], no experimental evidence for this conformation has been presented so far. In this study, two-dimensional proton NMR experiments were performed on a chemically synthesized pentacosapeptide containing the first 25 amino acids of this coat protein [Ten Kortenaar, P. B. W., Krüse, J., Hemminga, M. A., & Tesser, G. I. (1986) Int. J. Pept. Protein Res. 27, 401-413]. All resonances could be assigned by a combined use of two-dimensional correlated spectroscopy and nuclear Overhauser enhancement spectroscopy carried out at four different temperatures. Various NMR parameters indicate the presence of a conformational ensemble consisting of helical structures rapidly converting into more extended states. Differences in chemical shifts and nuclear Overhauser effects indicate that lowering the temperature induces a shift of the dynamic equilibrium toward more helical structures. At 10 degrees C, a perceptible fraction of the conformational ensemble consists of structures with an alpha-helical conformation between residues 9 and 17, likely starting with a turnlike structure around Thr9 and Arg10. Both the conformation and the position of this helical region agree well with the secondary structure predictions mentioned above.  相似文献   

5.
Poly(Leu-Lys-Lys-Leu) and poly(Leu-Lys) are sequential amphiphilic peptide isomers that adopt respectively an alpha-helical conformation and a beta-sheet structure in saline solutions and at the air/water interface. The surface active properties of LKKL and LK sequential isopeptides containing 16, 20, and n residues have been compared in order to evaluate the contributions of the alpha-helical and beta-sheet conformations. Both have a natural tendency to spread at the surface of a saline solution and the values of the equilibrium spreading pressure pi(e) lie in the same range. When dissolved in a saline solution, alpha-helical peptides diffuse faster and adsorb faster at the interface than the beta-sheet isomers. From the compression isotherms of LKKL and LK peptide monolayers it is possible to extract parameters that characterize the behavior of alpha-helical and beta-sheet conformations: beta-sheet peptide monolayers are more stable and less compressible than the monolayers formed with the alpha-helical isomers. The LK peptides differ also by their high degree of self-association at the air/water interface. Copyright 1999 John Wiley & Sons, Inc.  相似文献   

6.
We have studied the conformation of the peptide Ac-EPKRSVAFKKTKKEVKKVATPKK (CH-1), free in solution and bound to the DNA, by Fourier-transform infrared spectroscopy. The peptide belongs to the COOH-terminal domain of histone H1(0) (residues 99-121) and is adjacent to the central globular domain of the protein. In aqueous (D(2)O) solution the amide I' is dominated by component bands at 1643 cm(-1) and 1662 cm(-1), which have been assigned to random coil conformations and turns, respectively. In accordance with previous NMR results, the latter component has been interpreted as arising in turn-like conformations in rapid equilibrium with unfolded states. The peptide becomes fully structured either in 90% trifluoroethanol (TFE) solution or upon interaction with the DNA. In these conditions, the contributions of turn (1662 cm(-1)) and random coil components virtually disappear. In TFE, the spectrum is dominated by the alpha-helical component (1654 cm(-1)). The band at 1662 cm(-1) shifts to 1670 cm(-1), and has been assigned to the COOH-terminal TPKK motif in a more stable turn conformation. A band at 1637 cm(-1), also present in TFE, has been assigned to 3(10) helical structure. The amide I' band of the complexes with the DNA retains the components that were attributed to 3(10) helix and the TPKK turn. In the complexes with the DNA, the alpha-helical component observed in TFE splits into two components at 1657 cm(-1) and 1647 cm(-1). Both components are inside the spectral region of alpha-helical structures. Our results support the presence of inducible helical and turn elements, both sharing the character of DNA-binding motifs.  相似文献   

7.
An NMR study has been made of porcine pancreatic phospholipase A2 (PLA) in three environments: free in solution, in a binary complex with dodecylphosphocholine micelles, and in a ternary complex with a micelle and the substrate-like inhibitor (R)-1-octyl-2-(N-dodecanoylamino)-2-deoxyglycero-3-phosph oglycol. 1H and 15N chemical shifts, amide exchange rates, and NOE intensities are compared for the enzyme in different environments. From these data, structural differences are found for the N-terminal part, the end of the surface loop at residues Tyr69 and Thr70, and the active site residue His48, and also for the Ca-binding loop (residues 28-32). Specifically, when binding to a micelle, the side chains of residues Ala1, Trp3, and Tyr69, as well as all protons of Thr70, are found to be closer together. After subsequent introduction of the competitive inhibitor, further changes are found for these residues. The N-terminus is flexible in PLA free in solution, in contrast with the crystal structures where it adopts an alpha-helical conformation. According to the NMR data, this helix is rigidly formed only in the ternary complex. Furthermore, in the ternary complex, the N-terminal amino group and the exchangeable hydrogen at N3 of the ring of His48 are observed. We propose that PLA is activated in two steps. An initial conformational change occurs upon binding to a micellar interface. The catalytically active conformation of the enzyme, which has an extensive network of hydrogen bonds, is formed only when binding a substrate or competitive inhibitor at a lipid-water interface.  相似文献   

8.
Pulmonary surfactant protein SP-B is known to facilitate adsorption and spreading of surfactant components across the air/water interface. This property appears essential for in vivo function in the alveolar subphase and at the air/alveolar surface. Three peptides with amino acid sequences based on SP-B containing predicted alpha-helical regions (SP-B(1--20), SP-B(9--36A), SP-B(40--60A)) have been synthesized to probe structure-function relationships and protein-lipid interaction in bulk phase and monolayer environments. IR and CD studies are reported along with traditional surface pressure-molecular area (pi-A) isotherms and IR reflection-absorption spectroscopy (IRRAS) investigations conducted at the air/water interface. In bulk phase, helix-promoting environments (methanol and aqueous dispersions of lipid vesicles), SP-B(1--20) and SP-B(9--36A) contained significant amounts of alpha-helical structure, whereas varying degrees of alpha-helix, random coil, and beta-sheet were observed in aqueous solutions and monolayers. The most striking behavior was observed for SP-B(9--36A), which displayed reversible surface pressure-induced beta-sheet formation. Bulk phase lipid melting curves and monolayer experiments with peptide-lipid mixtures showed subtle differences in the degree of bulk phase interaction and substantial differences in peptide surface activity. The uniqueness of IRRAS is emphasized as the importance of evaluating secondary structure in both bulk phase and monolayer environments for lung surfactant peptide mimics is demonstrated.  相似文献   

9.
The interaction of human apolipoprotein (apo-) E3 with heparin was examined using heparin-Sepharose as a model system. The approach taken to determine the region of apo-E that is responsible for binding to heparin was to identify apo-E monoclonal antibodies that inhibited heparin binding, to determine the epitopes of the inhibiting antibodies, and finally to examine the heparin binding of fragments containing the inhibiting antibody epitopes. Three antibodies, designated 1D7, 6C5, and 3H1, were found to inhibit binding, suggesting that multiple heparin binding sites were present on apo-E. The epitopes of the inhibiting antibodies were determined by immunoblot analysis of synthetic or proteolytic fragments of apo-E. Measurement of the heparin binding activity of fragments containing epitopes of the inhibiting antibodies demonstrated that apo-E3 contains two heparin binding sites. The first site is located in the vicinity of residues 142-147 and coincides with the 1D7 epitope. The second binding site is contained in the carboxyl-terminal region of apo-E and is inhibited by 3H1, the epitope of which is located between residues 243 and 272. The epitope of the third inhibiting antibody, 6C5, is located at the amino terminus of apo-E; however, this antibody inhibits the second heparin binding site located in the carboxyl-terminal region. A head-to-tail association of apo-E, in which the 6C5 epitope and the second heparin binding site would be in close proximity, is proposed to account for this observation. In the lipid-free state both heparin binding sites on apo-E are expressed; however, when apo-E is complexed to phospholipid or on the surface of a lipoprotein particle, only the first binding site (residues 142-147) is expressed.  相似文献   

10.
Bacteriophage M13 major coat protein has been isolated with cholate and reconstituted in dimyristoyl- and dioleoylphosphatidylcholine (DMPC and DOPC, respectively) bilayers by dialysis. Fourier transform infrared spectra of DMPC/coat protein recombinants confirmed that, whereas the protein isolated by phenol extraction was predominantly in a beta-sheet conformation, the cholate-isolated coat protein contained a higher proportion of the alpha-helical conformation [cf. Spruijt, R. B., Wolfs, C. J. A. M., & Hemminga, M. A. (1989) Biochemistry 28, 9158-9165]. The cholate-isolated coat protein/lipid recombinants gave different electron spin resonance (ESR) spectral line shapes of incorporated lipid spin labels, as compared with those from recombinants with the phenol-extracted protein that were studied previously [Wolfs, C. J. A. M., Horváth, L. I., Marsh, D., Watts, A., & Hemminga, M. A. (1989) Biochemistry 28, 9995-10001]. Plots of the ratio of the fluid/motionally restricted components in the ESR spectra of spin-labeled phosphatidylglycerol were linear with respect to the lipid/protein ratio in the recombinants up to 20 mol/mol. The corresponding values of the relative association constants, Kr, and number of association sites, N1, on the protein were Kr approximately 1 and N1 approximately 4 for DMPC recombinants and Kr approximately 1 and N1 approximately 5 for DOPC recombinants. Simulation of the two-component lipid spin label ESR spectra with the exchange-coupled Bloch equations gave values for the off-rate of the lipids leaving the protein surface of 2.0 x 10(7) s-1 at 27 degrees C in DMPC recombinants and 3.0 x 10(7) s-1 at 24 degrees C in DOPC recombinants.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The conformation and orientation of synthetic monomeric human sequence SP-B(1-25) (mSP-B(1-25)) was studied in films with phospholipids at the air-water (A/W) interface by polarization modulation infrared reflectance absorption spectroscopy (PM-IRRAS). Modified two-dimensional infrared (2D IR) correlation analysis was applied to PM-IRRAS spectra to define changes in the secondary structure and rates of reorientation of mSP-B(1-25) in the monolayer during compression. PM-IRRAS spectra and 2D IR correlation analysis showed that, in pure films, mSP-B(1-25) had a major alpha-helical conformation plus regions of beta-sheet structure. These alpha-helical regions reoriented later during film compression than beta structural regions, and became oriented normal to the A/W interface as surface pressure increased. In mixed films with 4:1 mol:mol acyl chain perdeuterated 1,2-dipalmitoyl-sn-glycero-3-phosphocholine/1,2-dioleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (sodium salt) (DPPC-d(62):DOPG), the IR spectra of mSP-B(1-25) showed that a significant, concentration-dependent conformational change occurred when mSP-B(1-25) was incorporated into a DPPC-d(62):DOPG monolayer. At an mSP-B(1-25) concentration of 10 wt.%, the peptide assumed a predominantly beta-sheet conformation with no contribution from alpha-helical structures. At lower, more physiological peptide concentrations, 2D IR correlation analysis showed that the propensity of mSP-B(1-25) to form alpha-helical structures was increased. In phospholipid films containing 5 wt.% mSP-B(1-25), a substantial alpha-helical peptide structural component was observed, but regions of alpha and beta structure reoriented together rather than independently during compression. In films containing 1 wt.% mSP-B(1-25), peptide conformation was predominantly alpha-helical and the helical regions reoriented later during compression than the remaining beta structural components. The increased alpha-helical structure of mSP-B(1-25) demonstrated here by PM-IRRAS and 2D IR correlation analysis in monolayers of 4:1 DPPC:DOPG containing 1 wt.% (and, to a lesser extent, 5 wt.%) peptide may be relevant for the formation of the intermediate order 'dendritic' surface phase observed in similar surface films by epi-fluorescence.  相似文献   

12.
The complete nucleotide sequences of three cloned cDNAs corresponding to human liver apolipoprotein E (apo-E) mRNA were determined. Analysis of the longest cDNA showed that it contained 1157 nucleotides of mRNA sequence with a 5'-terminal nontranslated region of 61 nucleotides, a signal peptide region corresponding to 18 amino acids, a mature protein region corresponding to 299 amino acids, and a 3'-terminal nontranslated region of 142 nucleotides. The inferred amino acid sequences from two cDNAs were identical and corresponded to the amino acid sequence for plasma apo-E3 that has been reported previously ( Rall , S. C., Jr., Weisgraber , K. H., and Mahley , R. W. (1982) J. Biol. Chem. 257, 4171-4178). The third cDNA differed from the other two cDNAs in five nucleotide positions. Three of these differences occurred in the third nucleotide position of amino acid codons, resulting in no change in the corresponding amino acids at residues Val-85, Ser-223, and Gln-248. The other two altered nucleotides occurred in the first nucleotide position of codons, leading to changes in the amino acids encoded. In the variant sequence, a threonine replaced the normal alanine at residue 99 and a proline replaced the normal alanine at residue 152. We have concluded that the human liver donor was heterozygous for the epsilon 3 genotype. The variant cDNA corresponds to a new, previously undescribed variant form of apo-E in which the amino acid substitutions of the protein are electrophoretically silent; it would probably be undetectable by standard apo-E phenotyping methods. The amino acid substitution at position 152 occurs in a region of apo-E that appears to be important for receptor binding, and it may have clinical significance.  相似文献   

13.
The conformational properties of two hexapeptides, Ac-LWRILW-NH(2) and its D-amino acid counterpart Ac-lwrilw-NH(2), identified as calmodulin inhibitors using mixture-based synthetic combinatorial library approaches, have been characterised by NMR and CD spectroscopy. The peptides fold into an alpha-helical conformation in aqueous solution. The observed short- and medium-range nuclear Overhauser effects were consistent with the formation of an alpha-helical structure and a reasonably well-defined set of structures was obtained by using restraints from the NMR data in simulated annealing calculations. Analysis of glycine-substitution analogues demonstrated that all the amino acids that make up the peptide sequence are important for the stabilization of the alpha-helical conformation. The results suggest that a well-defined set of interactions is indispensable to allow alpha-helix formation in this short hexapeptide.  相似文献   

14.
We have determined the crystal structure of the ligand binding segment of the Enterococcus faecalis collagen binding MSCRAMM ACE (microbial surface components recognizing adhesive matrix molecules adhesin of collagen from enterococci). This segment is composed of two subdomains, N(1) and N(2), each adopting an IgG-like fold and forming a putative collagen binding surface at the interface between the two subdomains. This structure is very similar to that recently reported for CNA, the collagen binding MSCRAMM of Staphylococcus aureus, for which a unique ligand binding mechanism called the Collagen Hug was proposed. We suggest that ACE binds collagen by a similar mechanism and present the first biochemical evidence for this binding model. Replacing residues in the putative collagen binding trench of ACE N(2) with Ala residues affected collagen binding. A closed conformation of ACE stabilized by an engineered disulfide bond is unable to bind collagen. Finally, the importance of the residues in the N(2) extension in stabilizing the MSCRAMM-ligand complex is demonstrated by selected point and truncation mutations.  相似文献   

15.
Using thrombin and trypsin as probes, we determined: first, that low-density lipoprotein (LDL) receptor binding determinants switch from apolipoprotein (apo) E to apo-B within the very-low-density lipoprotein (VLDL) Sf 20-60 region of the metabolic cascade from VLDL1 (Sf 100-400) of hypertriglyceridemic (HTG) human subjects to LDL. Second, two different conformations of apo-E exist in HTG-VLDL Sf greater than 60, one accessible (greater than or equal to 1 mol/mol of particle) and one inaccessible (1-2 mol/mol) to both thrombin and the LDL receptor; normal VLDL (Sf greater than 60) have only the inaccessible conformation and therefore do not bind to the LDL receptor. Third, thrombin degrades apo-B into large fragments, three of which have electrophoretic mobilities similar to B-48, B-74, and B-26; this, however, has no effect on apo-B-mediated receptor binding. Fibroblast studies showed that thrombin could abolish receptor uptake of HTG-VLDL1 and HTG-VLDL2 (Sf 60-100), had little or no effect on HTG-VLDL3 (Sf 20-60), and no effect on uptake of intermediate-density lipoprotein (IDL) or LDL. Trypsin abolished the binding of HTG-VLDL1 and HTG-VLDL2, reduced that of HTG-VLDL3, but had little to no effect on IDL or LDL binding. Immunochemical techniques revealed that thrombin cleaved some apo-E into the E-22 and E-12 fragments; after trypsin treatment no apo-E was detected in any HTG-lipoprotein. Normal VLDL subclasses contained less apo-E than the corresponding HTG-VLDL subclasses and it was not cleaved by thrombin. Apo-B immunoreactivities of VLDL subclasses were not significantly changed after treatment with thrombin, although thrombin cleaved some of the B-100 of each VLDL subclass, and all apo-B in IDL and LDL, into 4-6 major large fragments. Trypsin converted all of the apo-B of each lipoprotein into smaller fragments (Mr less than 100,000). We conclude that apo-E of the thrombin-accessible conformation mediates uptake of HTG-VLDL1 and HTG-VLDL2 but that apo-B alone is sufficient to mediate receptor binding of IDL and LDL; the switch from apo-E to apo-B as the primary or sufficient binding determinant occurs within the VLDL3 (Sf 20-60) region of the metabolic cascade, where receptor binding first appears in VLDL subclasses from normal subjects.  相似文献   

16.
The purposes of this study were to establish the role of disulfide linkages in the secondary structure of apolipoprotein B, to investigate the effects of sulfhydryl blocking agents, denaturing agents, pH and storage on the conformation of apolipoprotein B and lipoprotein B, and to compare the conformation of water-soluble apolipoprotein B in the presence and absence of its lipids by using circular dichroism. Fresh lipoprotein B examined in Tris/EDTA at pH 9.0, 7.3 and 2.7 exhibited alpha-helical content of 24.4, 26.7 and 26.9%, and beta-pleated sheet 25.1, 15.4 and 18.0%, respectively. The carboxymethylated (CM-) lipoprotein B had similar alpha-helical contents, and lower contents of beta-sheets. Storage of lipoprotein B resulted in marked change of beta-sheets and gradual decrease in alpha-helical structure, in spite of the preventive measures taken for lipid peroxidation and proteolytic degradation. Exposure of apolipoprotein B to 6 M guanidine X HCl led to a complete disappearance of the alpha-helix with an increase in the beta-sheets to 35-40%, irrespective of the use of disulfide-reducing agents. By substituting 6 M urea for guanidine X HCl, the alpha-helical contents for both CM- and reduced CM-apolipoprotein B increased up to 7-9% with a concomitant decrease in beta-structure. When urea was replaced with aqueous buffers, these apolipoprotein B preparations regained their alpha-helical contents (25-27%) to the full extent originally present in the parent lipoprotein samples. No difference was observed between the secondary structure of CM- and reduced CM-apolipoprotein B. Furthermore, the conformation of apolipoprotein B did not vary with pH when pH was changed from 2.7 to 9.0. These results suggest that (1) the conformation of apolipoprotein B is more stable with respect to pH in the absence of lipids than in their presence, (2) intramolecular disulfide linkages play an insignificant role in the conformation of apolipoprotein B, and (3) the changes in alpha-helix structure of lipoprotein B or CM-lipoprotein B due to delipidization and denaturation are reversible.  相似文献   

17.
Alpha-helical coiled coils play a crucial role in mediating specific protein-protein interactions. However, the rules and mechanisms that govern helix-helix association in coiled coils remain incompletely understood. Here we have engineered a seven heptad "Phe-zipper" protein (Phe-14) with phenylalanine residues at all 14 hydrophobic a and d positions, and generated a further variant (Phe-14(M)) in which a single core Phe residue is substituted with Met. Phe-14 forms a discrete alpha-helical pentamer in aqueous solution, while Phe-14(M) folds into a tetrameric helical structure. X-ray crystal structures reveal that in both the tetramer and the pentamer the a and d side-chains interlock in a classical knobs-into-holes packing to produce parallel coiled-coil structures enclosing large tubular cavities. However, the presence of the Met residue in the apolar interface of the tetramer markedly alters its local coiled-coil conformation and superhelical geometry. Thus, short-range interactions involving the Met side-chain serve to preferentially select for tetramer formation, either by inhibiting a nucleation step essential for pentamer folding or by abrogating an intermediate required to form the pentamer. Although specific trigger sequences have not been clearly identified in dimeric coiled coils, higher-order coiled coils, as well as other oligomeric multi-protein complexes, may require such sequences to nucleate and direct their assembly.  相似文献   

18.
Pulmonary surfactant proteins, SP-B and SP-C, if present in preformed monolayers can induce lipid insertion from lipid vesicles into the monolayer after the addition of (divalent) cations [Oosterlaken-Dijksterhuis, M. A., Haagsman, H. P., van Golde, L. M. G., & Demel, R. A. (1991) Biochemistry 30, 8276-8287]. This model system was used to study the mechanisms by which SP-B and SP-C induce monolayer formation from vesicles. Lipid insertion proceeds irrespectively of the molecular class, and PG is not required for this process. In addition to lipids that are immediately inserted from vesicles into the monolayer, large amounts of vesicles are bound to the monolayer and their lipids eventually inserted when the surface area is expanded. SP-B and SP-C are directly responsible for the binding of vesicles to the monolayer. By weight, the vesicle binding capacity of SP-B is approximately 4 times that of SP-C. For vesicle binding and insertion, the formation of close contacts between monolayer and vesicles is essential. SP-B and SP-C show very similar surface properties. Both proteins form extremely stable monolayers (collapse pressures 36-37 mN/m) of alpha-helical structures oriented parallel to the interface. In monolayers consisting of DPPC and SP-B or SP-C, an increase in mean molecular area is observed, which is mainly attributed to the phospholipid. This will greatly enhance the insertion of new lipid material into the monolayer. The results of this study suggest that the surface properties and the hydrophobic nature of SP-B and SP-C are important for the protein-mediated monolayer formation.  相似文献   

19.
We increased drastically the heat stability of Lac repressor (LacR) of Escherichia coli. Wild-type tetrameric LacR denatures irreversibly at 53 degrees C. Improving hydrophobic packing at the dimerisation interface by a single substitution increases LacR heat-resistance by 40 deg. C without abolishing inducer binding at high and low temperatures. Tetrameric LacR mutants carrying substitutions of the positively charged amino acid Lys84 by each of the hydrophobic amino acids Leu, Ile and Met resist heating to temperatures up to 93 degrees C. We performed IPTG binding assays at 80 degrees C and found the mutant Lac repressors active and, thus, the core intact. Furthermore, the activity of LacR following heating is shown at room temperature by a gel retardation assay, which demonstrates normal oligomerisation state and function of the headpiece. The same mutations (K84L/I/M) in the dimer LacR331stop, carrying a stop codon in amino acid 331, increase thermostability of the dimer from 47 degrees C to 87 degrees C. LacRK84M represses beta-galactosidase activity in vivo as well as the wild-type and is sufficiently induced to allow growth on lactose. The results with both tetramer and dimer variants of LacR indicate mutual stabilisation of the tetramerisation region and the stable core.  相似文献   

20.
Conformational states of ribulosebisphosphate carboxylase (Rubisco) from Rhodospirillum rubrum were examined by far-UV circular dichroism (CD), tryptophan fluorescence, and 1-anilino-naphthalenesulfonate (ANS) binding. At pH 2 and low ionic strength (I = 0.01), Rubisco adopts an unfolded, monomeric conformation (UA1 state) as judged by far-UV CD and tryptophan fluorescence. As with other acid-unfolded proteins [Goto, Y., Calciano, L. J., & Fink, A. L. (1990) Proc. Natl. Acad. Sci. U.S.A. 87, 573-577], an intermediate conformation (A1 state) is observed at pH 2 and high ionic strength. The A1 state has an alpha-helical content equivalent to 64% of that present in the native dimer (N2 state). However, fluorescence measurements indicate that the tertiary structure of the A1 state is largely disordered. A site-directed mutant, K168E, which exists as a stable monomer [Mural, R. J., Soper, T. S., Larimer, F. W., & Hartman, F. C. (1990) J. Biol. Chem. 265, 6501-6505] was used to characterize the "native" monomer (N1 state). The far-UV CD spectra of the N1 and N2 states are almost identical, indicating a similar secondary structure content. However, the tertiary structure of the N1 state is less ordered than that of the N2 state. Nevertheless, when appropriately complemented in vitro, K168E forms an active heterodimer. Upon neutralization of acid-denatured Rubisco or dilution of guanidine hydrochloride-denatured Rubisco, unstable folding intermediates (I1 state) are rapidly formed. At concentrations at or below the "critical aggregation concentration" (CAC), the I1 state reverts spontaneously but slowly to the native states with high yield (greater than 65%). The CAC is temperature-dependent. At concentrations above the CAC, the I1 and the A1 states undergo irreversible aggregation. The commitment to aggregation is rapid [ef. Goldberg, M. E., Rudolph, R., & Jaenicke, R. (1991) Biochemistry 30, 2790-2797] and proceeds until the concentration of folding intermediate(s) has fallen to the CAC. In the presence of a molar excess of chaperonin 60 oligomers, the I1 state forms a stable binary complex. No stable binary complex between chaperonin 60 and the N1 state could be detected. Formation of the chaperonin 60-I1 binary complex arrests the spontaneous folding process. The I1 state becomes resistant to interaction with chaperonin 60 with kinetics indistinguishable from those associated with the appearance of the native states. In vitro complementation analysis indicated that the product of the chaperonin-facilitated process is monomeric.(ABSTRACT TRUNCATED AT 400 WORDS)  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