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 15 毫秒
1.
Alamethicin is a helical 20-amino acid voltage-gated channel-forming peptide, which is known to exhibit segmental flexibility in solution along its backbone near alpha-methylalanine (MeA)-10 and Gly-11. In an alpha-helical configuration, MeA at position 10 would normally hydrogen-bond with position 14, but the presence of proline at this position prevents the formation of this interhelical hydrogen bond. To determine whether the presence of proline at position 14 contributes to the flexibility of this helix, two analogs of alamethicin were synthesized, one with proline 14 replaced by alanine and another with both proline 14 and glycine 11 replaced by alanine. The C-termini of these peptides were derivatized with a proxyl nitroxide, and paramagnetic enhancements produced by the nitroxide on the Calpha protons were used to estimate r-6 weighted distances between the nitroxide and the backbone protons. When compared to native alamethicin, the analog lacking proline 14 exhibited similar C-terminal to Calpha proton distances, indicating that substitution of proline alone does not alter the flexibility of this helix; however, the subsequent removal of glycine 11 resulted in a significant increase in the averaged distances between the C- and N-termini. Thus, the G-X-X-P motif found in alamethicin appears to be largely responsible for mediating high-amplitude bending motions that have been observed in the central helical domain of alamethicin in methanol. To determine whether these substitutions alter the channel behavior of alamethicin, the macroscopic and single-channel currents produced by these analogs were compared. Although the substitution of the G-X-X-P motif produces channels with altered characteristics, this motif is not essential to achieve voltage-dependent gating or alamethicin-like behavior.  相似文献   

2.
The GTP-binding p21 protein encoded by the ras-oncogene can be activated to cause malignant transformation of cells by substitution of a single amino acid at critical positions along the polypeptide chain. Substitution of any non-cyclic L-amino acid for Gly 12 in the normal protein results in a transforming protein. This substitution occurs in a hydrophobic sequence (residues 6-15) which is known to be involved in binding the phosphate moities of GTP (and GDP). We find, using conformational energy calculations, that the 6-15 segment of the normal protein (with Gly 12) adopts structures that contain a bend at residues 11 and 12 with the Gly in the D* conformation, not allowed energetically for L-amino acids. Substitution of non-cyclic L-amino acids for Gly 12 results in shifting this bend to residues 12 and 13. We show that many computed structures for the Gly 12-containing phosphate binding loop, segment 9-15, are superimposable on the corresponding segment of the recently determined X-ray crystallographic structure for residues 1-171 of the p21 protein. All such structures contain bends at residues 11 and 12 and most of these contain Gly 12 in the C* or D* conformational state. Other computed conformations for the 9-15 segment were superimposable on the structure of the corresponding 18-23 segment of EFtu, the bacterial chain elongation factor having structural similarities to the p21 protein in the phosphate-binding regions. This segment contains a Val residue where a Gly occurs in the p21 protein. As previously predicted, all of these superimposable conformations contain a bend at positions 12 and 13, not 11 and 12. If these structures that are superimposable on EFtu are introduced into the p21 protein structure, bad contacts occur between the sidechain of the residue (here Val) at position 12 and another phosphate binding loop region around position 61. These bad contacts between the two segments can be removed by changing the conformation of the 61 region in the p21 protein to the corresponding position of the homologous region in EFtu. In this new conformation, a large site becomes available for the binding of phosphate residues. In addition, such phenomena as autophosphorylation of the p21 protein by GTP can be explained with this new model structure for the activated protein which cannot be explained by the structure for the non-activated protein.  相似文献   

3.
F Ni  H A Scheraga  S T Lord 《Biochemistry》1988,27(12):4481-4491
The proton resonances of the following synthetic linear human fibrinogen-like peptides were completely assigned with two-dimensional NMR techniques in solution: Ala(1)-Asp-Ser-Gly-Glu-Gly-Asp(7)-Phe-Leu-Ala-Glu-Gly(12)-Gly(13)-Gly(14)- Val(15)-Arg(16)-Gly-Pro-Arg-Val-Val-Glu-Arg (F10), Ala-Asp-Ser-Gly-Glu-Gly-Asp-Phe-Leu-Ala-Glu-Gly-Gly(13)-Gly(14)-Val-Arg (F11), and Gly-Pro-Arg-Val-Val-Glu-Arg (F12). No predominant structure was found in the chain segment from Ala(1) to Gly(6) for F10 in both H2O and dimethyl sulfoxide. The previous suggestion that there is a hairpin loop involving residues Gly(12) to Val(15) in the A alpha chain of human fibrinogen is supported by the slow backbone NH exchange rates of Gly(14) and Val(15), by an unusually small NH chemical shift of Val(15), and by strong sequential NOE's involving this region in F10. This local chain fold within residues Asp(7) to Val(20) may place the distant Phe residue near the Arg(16)-Gly(17) peptide bond which is cleaved by thrombin.  相似文献   

4.
Dermaseptins are antimicrobial peptides from frog skin that have high membrane-lytic activity against a broad spectrum of microorganisms. The structure of dermaseptin B2 in aqueous solution, in TFE/water mixtures, and in micellar and nonmicellar SDS was analyzed by CD, FTIR, fluorescence, and NMR spectroscopy combined with molecular dynamics calculations. Dermaseptin B2 is unstructured in water, but helical conformations, mostly in segment 3-18, are stabilized by addition of TFE. SDS titration showed that dermaseptin B2 assumes nonhelical structures at SDS concentrations far below the critical micellar concentration and helical structures at micellar concentrations. Dermaseptin B2 bound to SDS micelles (0.4 mM peptide, 80 mM SDS) adopts a well-defined amphipathic helix between residues 11-31 connected to a more flexible helical segment spanning residues 1-8 by a flexible hinge region around Val9 and Gly10. Experiments using paramagnetic probes showed that dermaseptin B2 lies near the surface of SDS micelles and that residue Trp3 is buried in the SDS micelle, but close to the surface. A slow exchange equilibrium occurs at higher peptide/SDS ratios (2 mM peptide, 80 mM SDS) between forms having distinct sets of resonances in the N-terminal 1-11 segment. This equilibrium could reflect different oligomeric states of dermaseptin B2 interacting with SDS micelles. Structure-activity studies on dermaseptin B2 analogues showed that the N-terminal 1-11 segment is an absolute requirement for antibacterial activity, while the C-terminal 10-33 region is also important for full antibiotic activity.  相似文献   

5.
Disulfide-rich cyclic peptides have generated great interest in the development of peptide-based therapeutics due to their exceptional stability toward chemical, enzymatic, or thermal attack. In particular, they have been used as scaffolds onto which bioactive epitopes can be grafted to take advantage of the favorable biophysical properties of disulfide-rich cyclic peptides. To date, the most commonly used method for the head-to-tail cyclization of peptides has been native chemical ligation. In recent years, however, enzyme-mediated cyclization has become a promising new technology due to its efficiency, safety, and cost-effectiveness. Sortase A (SrtA) is a bacterial enzyme with transpeptidase activity. It recognizes a C-terminal penta-amino acid motif, LPXTG, and cleaves the amide bond between Thr and Gly to form a thioacyl-linked intermediate. This intermediate undergoes nucleophilic attack by an N-terminal poly-Gly sequence to form an amide bond between the Thr and N-terminal Gly. Here, we demonstrate that sortase A can successfully be used to cyclize a variety of small disulfide-rich peptides, including the cyclotide kalata B1, α-conotoxin Vc1.1, and sunflower trypsin inhibitor 1. These peptides range in size from 14 to 29 amino acids and contain three, two, or one disulfide bond, respectively, within their head-to-tail cyclic backbones. Our findings provide proof of concept for the potential broad applicability of enzymatic cyclization of disulfide-rich peptides with therapeutic potential.  相似文献   

6.
Previous studies using chemical modification approach have shown the importance of arginine residues in bile salt activation of carboxyl ester lipase (CEL) activity. However, the x-ray crystal structure of CEL failed to show the involvement of arginine residues in CEL-bile salt interaction. The current study used a site-specific mutagenesis approach to determine the role of arginine residues 63 and 423 in bile salt-dependent and bile salt-independent hydrolytic activities of rat CEL. Mutations of Arg(63) to Ala(63) (R63A) and Arg(423) to Gly(423) (R423G) resulted in enzymes with increased bile salt-independent hydrolytic activity against lysophosphatidylcholine, having 6.5- and 2-fold higher k(cat) values, respectively, in comparison to wild type CEL. In contrast, the R63A and R423A mutant enzymes displayed 5- and 11-fold decreases in k(cat), in comparison with wild type CEL, for bile salt-dependent cholesteryl ester hydrolysis. Although taurocholate induced similar changes in circular dichroism spectra for wild type, R63A, and R423G proteins, this bile salt was less efficient in protecting the mutant enzymes against thermal inactivation in comparison with control CEL. Lipid binding studies revealed less R63A and R423G mutant CEL were bound to 1,2-diolein monolayer at saturation compared with wild type CEL. These results, along with computer modeling of the CEL protein, indicated that Arg(63) and Arg(423) are not involved directly with monomeric bile salt binding. However, these residues participate in micellar bile salt modulation of CEL enzymatic activity through intramolecular hydrogen bonding with the C-terminal domain. These residues are also important, probably through similar intramolecular hydrogen bond formation, in stabilizing the enzyme in solution and at the lipid-water interface.  相似文献   

7.
C-terminal biotin-tagged alamethicin, which has several alpha-aminoisobutyric acid (Aib) residues in its sequence, was synthesized by the preparation of the protected peptide segment using the 2-chlorotrityl resin, followed by conjugation with biotin hydrazide. Suppression of the channel current of the biotin-tagged alamethicin by the addition of streptavidin to the electrolyte was monitorable in real time using the planar lipid-bilayer method. The system was also applicable to the detection of interaction of the biotin-tagged alamethicin with the anti-biotin antibody.  相似文献   

8.
The conformation of the 20-residue antibiotic ionophore alamethicin in macroscopically oriented phospholipid bilayers has been studied using (15)N solid-state nuclear magnetic resonance (NMR) spectroscopy in combination with molecular modeling and molecular dynamics simulations. Differently (15)N-labeled variants of alamethicin and an analog with three of the alpha-amino-isobutyric acid residues replaced by alanines have been investigated to establish experimental structural constraints and determine the orientation of alamethicin in hydrated phospholipid (dimyristoylphosphatidylcholine) bilayers and to investigate the potential for a major kink in the region of the central Pro(14) residue. From the anisotropic (15)N chemical shifts and (1)H-(15)N dipolar couplings determined for alamethicin with (15)N-labeling on the Ala(6), Val(9), and Val(15) residues and incorporated into phospholipid bilayer with a peptide:lipid molar ratio of 1:8, we deduce that alamethicin has a largely linear alpha-helical structure spanning the membrane with the molecular axis tilted by 10-20 degrees relative to the bilayer normal. In particular, we find compatibility with a straight alpha-helix tilted by 17 degrees and a slightly kinked molecular dynamics structure tilted by 11 degrees relative to the bilayer normal. In contrast, the structural constraints derived by solid-state NMR appear not to be compatible with any of several model structures crossing the membrane with vanishing tilt angle or the earlier reported x-ray diffraction structure (Fox and Richards, Nature. 300:325-330, 1982). The solid-state NMR-compatible structures may support the formation of a left-handed and parallel multimeric ion channel.  相似文献   

9.
Angiotensin I converting enzyme (ACE) and neutral endopeptidase ("enkephalinase"; NEP), were purified to homogeneity from human kidney. NEP cleaved substance P (SP) at Gln6-Phe7,-Phe8, and Gly9-Leu10 and neurotensin (NT) at Pro10-Tyr11 and Tyr11-Ile12. NEP hydrolyzed 0.1 mM SP, NT and their C-terminal fragments at the following rates (mumol/min/mg): SP1-11 = 7.8, SP4-11 = 11.7, SP5-11 = 15.4, SP6-11 = 15.6, SP8-11 = 6.7, NT1-13 = 2.9, and NT8-13 = 4.0. Purified ACE rapidly inactivated SP as measured in bioassay. HPLC analysis showed that ACE cleaved SP at Phe8-Gly9 and Gly9-Leu10 to release C-terminal tri- and dipeptide (ratio = 4:1). The hydrolysis was Cl- dependent and inhibited by captopril. ACE released mainly C-terminal tripeptide from SP methyl ester, but only dipeptide from SP free acid. Modification of arginine residues in ACE with cyclohexanedione or butanedione similarly inhibited hydrolysis of SP, bradykinin and Bz-Gly-Phe-Arg (80-93%) indicating an active site arginine is required for hydrolysis of SP. ACE hydrolyzed NT at Tyr11-Ile12 to release Ile12-Leu13. SP, NT and their derivatives (0.1 mM) were cleaved by ACE at the following rates (mumol/min/mg): SP1-11 = 1.2, SP methyl ester = 0.7, SP free acid = 8.5, SP4-11 = 2.4, SP5-11 = 0.9, SP6-11 = 1.4, SP8-11 = 0, NT1-13 = 0.2, and NT8-13 = 1.3. Peptide substrates were used as inhibitors of ACE (substrate = FA-Phe-Gly-Gly) and NEP (substrate = Leu5-enkephalin).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Molecular dynamics simulations of ion channel peptides alamethicin and melittin, solvated in methanol at 27 degrees C, were run with either regular alpha-helical starting structures (alamethicin, 1 ns; melittin 500 ps either with or without chloride counterions), or with the x-ray crystal coordinates of alamethicin as a starting structure (1 ns). The hydrogen bond patterns and stabilities were characterized by analysis of the dynamics trajectories with specified hydrogen bond angle and distance criteria, and were compared with hydrogen bond patterns and stabilities previously determined from high-resolution NMR structural analysis and amide hydrogen exchange measurements in methanol. The two alamethicin simulations rapidly converged to a persistent hydrogen bond pattern with a high level of 3(10) hydrogen bonding involving the amide NH's of residues 3, 4, 9, 15, and 18. The 3(10) hydrogen bonds stabilizing amide NH's of residues C-terminal to P2 and P14 were previously proposed to explain their high amide exchange stabilities. The absence, or low levels of 3(10) hydrogen bonds at the N-terminus or for A15 NH, respectively, in the melittin simulations, is also consistent with interpretations from amide exchange analysis. Perturbation of helical hydrogen bonding in the residues before P14 (Aib10-P14, alamethicin; T11-P14, melittin) was characterized in both peptides by variable hydrogen bond patterns that included pi and gamma hydrogen bonds. The general agreement in hydrogen bond patterns determined in the simulations and from spectroscopic analysis indicates that with suitable conditions (including solvent composition and counterions where required), local hydrogen-bonded secondary structure in helical peptides may be predicted from dynamics simulations from alpha-helical starting structures. Each peptide, particularly alamethicin, underwent some large amplitude structural fluctuations in which several hydrogen bonds were cooperatively broken. The recovery of the persistent hydrogen bonding patterns after these fluctuations demonstrates the stability of intramolecular hydrogen-bonded secondary structure in methanol (consistent with spectroscopic observations), and is promising for simulations on extended timescales to characterize the nature of the backbone fluctuations that underlie amide exchange from isolated helical polypeptides.  相似文献   

11.
A protease secreted by the parasitic helminth Fasciola hepatica, a 37-kDa procathepsin L1 (FheproCL1), autocatalytically processes and activates to its mature enzyme (FheCL1) over a wide pH range of 7.3 to 4.0, although activation is more rapid at low pH. Maturation initiates with cleavages of a small proportion of molecules within the central region of the prosegment, possibly by intramolecular events. However, activation to fully mature enzymes is achieved by a precise intermolecular cleavage at a Leu-12-Ser-11 downward arrowHis-10 sequence within the nonconserved C-terminal region of the prosegment. The importance of this cleavage site in enzyme activation was demonstrated using an active site variant FheproCL1Gly26 (Cys26 to Gly26) and a double variant FheproCL1Pro-12/Gly26 (Leu-12 to Pro-12), and although both of these variants cannot autocatalytically process, the former is susceptible to trans-processing at a Leu-12-Ser-11 downward arrowHis-10 sequence by pre-activated FheCL1, but the latter is not. Another F. hepatica secreted protease FheCL2, which, unlike FheCL1, can readily accept proline in the S2 subsite of its active site, can trans-process the double variant FheproCL1Pro-12/Gly26 by cleavage at the Pro-12-Ser-11 downward arrowHis-10 sequence. Furthermore, the autoactivation of a variant enzyme with a single replacement, FheproCL1Pro-12, was very slow but was increased 40-fold in the presence of FheCL2. These studies provide a molecular insight into the regulation of FheproCL1 autocatalysis.  相似文献   

12.
Understanding the binding and insertion of peptides in lipid bilayers is a prerequisite for understanding phenomena such as antimicrobial activity and membrane-protein folding. We describe molecular dynamics simulations of the antimicrobial peptide alamethicin in lipid/water and octane/water environments, taking into account an external electric field to mimic the membrane potential. At cis-positive potentials, alamethicin does not insert into a phospholipid bilayer in 10 ns of simulation, due to the slow dynamics of the peptide and lipids. However, in octane N-terminal insertion occurs at field strengths from 0.33 V/nm and higher, in simulations of up to 100 ns duration. Insertion of alamethicin occurs in two steps, corresponding to desolvation of the Gln7 side chain, and the backbone of Aib10 and Gly11. The proline induced helix kink angle does not change significantly during insertion. Polyalanine and alamethicin form stable helices both when inserted in octane and at the water/octane interface, where they partition in the same location. In water, both polyalanine and alamethicin partially unfold in multiple simulations. We present a detailed analysis of the insertion of alamethicin into the octane slab and the influence of the external field on the peptide structure. Our findings give new insight into the mechanism of channel formation by alamethicin and the structure and dynamics of membrane-associated helices.  相似文献   

13.
Linear and cyclic analogues of the specific active center of the ACTH molecule have been synthesized, viz. [Lis5]ACTH-(5-10)-, [Lys5, cyclo (Gly10----epsilon Lys5)]-ACTH-(5-10)-hexapeptides, [Lys5 (Gly)]ACTH-(5-10)- and [Lys5, Gly11, cyclo (Gly11----epsilon Lys5)]-ACTH-(5-11)-heptapeptides. The cyclic structures are fixed by covalent bond between the COOH-group of the C-terminal glycine and epsilon-amino group of a lysine residue. Azide method, DCC/HOBT or pentafluorophenyl esters are used for fragment coupling, while cyclization is achieved by means of diphenylphosphoryl azide or pentafluorophenyl esters. Cyclic compounds are 2-3 orders of magnitude more active than their linear counterparts as revealed by assaying their melanocyte-stimulating activity in vitro on frog skin. Only the title heptapeptide possesses a steroidogenic activity similar to that of ACTH-(5-10)-hexapeptide. The results obtained are in accord with the idea implying the formation in the hormone-receptor complexes of quasi-cyclic structures in the region of the specific active center of ACTH molecule.  相似文献   

14.
Ubiquitination-the attachment of ubiquitin to a protein target-is involved in a wide range of cellular processes in eukaryotes. This dynamic posttranslational modification utilizes three enzymes to link, through an isopeptide bond, the C-terminal Gly of ubiquitin to the lysine side chain from a protein target. Progress in the field aiming at deciphering the role of ubiquitination in biological processes has been very dependent on the discovery of the enzymatic machinery, which is known to be very specific to each protein target. Chemical approaches offer a complementary route to the biochemical methods to construct these conjugates in vitro in order to assist in unraveling the role of ubiquitination on protein function. Herein is presented a novel method for the rapid synthesis of ubiquitinated peptides employing solid-phase peptide to generate the critical isopeptide linkage. Using these tools, several ubiquitinated peptides derived from known ubiquitinated proteins were prepared. Among them is the ubiquitinated C-terminal fragment of H2B, which can be used in the synthesis of monoubiquitinated H2B. For the first time, we systematically assessed the effect of the length of the ubiquitinated peptides on the UCH-L3 activity and found that peptides of up to ~20 residues are preferred substrates.  相似文献   

15.
The combined use of several nuclear magnetic resonance and restrained molecular dynamics techniques allowed the formulation of a molecular model for the preferred solution conformation of a synthetic peptide reproducing the [1-20] processing domain of the pro-ocytocin-neurophysin precursor. In the model, the conformation of the 20-membered tocin ring, with the two Cys1 and Cys6 residues bridged by a disulphide bond, is very close to that observed for isolated ocytocin in the solid state; in addition, a type II β-turn is postulated for the 7-10 segment of the acyclic tail containing the Lys11-Arg12 processing site, and connecting ocytocin to the neurophysin domain, while the C-terminal 13-20 segment of the molecule is believed to assume a helical structure. This particular structural organization could be important in participating as the favorable conformation for optimal substrate-enzyme active site recognition and processing by specific endoproteases. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
Isotope labeling of recombinant normal cardiac troponin C (cTnC3) with 15N-enriched amino acids and multidimensional NMR were used to assign the downfield-shifted amide protons of Gly residues at position 6 in Ca(2+)-binding loops II, III, and IV, as well as tightly hydrogen-bonded amides within the short antiparallel beta-sheets between pairs of Ca(2+)-binding loops. The amide protons of Gly70, Gly110, and Gly146 were found to be shifted significantly downfield from the remaining amide proton resonances in Ca(2+)-saturated cTnC3. No downfield-shifted Gly resonance was observed from the naturally inactive site I. Comparison of downfield-shifted amide protons in the Ca(2+)-saturated forms of cTnC3 and CBM-IIA, a mutant having Asp65 replaced by Ala, demonstrated that Gly70 is hydrogen bonded to the carboxylate side chain of Asp65. Thus, the hydrogen bond between Gly and Asp in positions 6 and 1, respectively, of the Ca(2+)-binding loop appears crucial for maintaining the integrity of the helix-loop-helix Ca(2+)-binding sites. In the apo- form of cTnC3, only Gly70 was found to be shifted significantly downfield with respect to the remaining amide proton resonances. Thus, even in the absence of Ca2+ at binding site II, the amide proton of Gly70 is strongly hydrogen bonded to the side-chain carboxylate of Asp65. The amide protons of Ile112 and Ile148 in the C-terminal domain and Ile36 in the N-terminal domain data-sheets exhibit chemical shifts consistent with hydrogen-bond formation between the pair of Ca(2+)-binding loops in each domain of Ca(2+)-saturated cTnC3.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Yao H  Stuart RA  Cai S  Sem DS 《Biochemistry》2008,47(7):1910-1917
F1Fo-ATP synthase is a large multiprotein complex, including at least 10 subunits in the membrane-bound Fo-sector. One of these Fo proteins is subunit e (Su e), involved in the stable dimerization of F1Fo-ATP synthase, and required for the establishment of normal cristae membrane architecture. As a step toward enabling structure-function studies of the Fo-sector, the Su e transmembrane region was structurally characterized in micelles. Based on a series of NMR and CD (circular dichroism) studies, a structural model of the Su e/micelle complex was constructed, indicating Su e is largely helical, and emerges from the micelle with Arg20 near the phosphate head groups. Su e only adopts this folded conformation in the context of the micelle, and is essentially disordered in DMSO, water or trifluoroethanol/water. Within the micelle the C-terminal Ala10-Arg20 stretch is helical, while the region N-terminal may be transiently helical, based on negative CSI (chemical shift index) values. The Ala10-Arg20 helix contains the G14XXXG18 motif, which has been proposed to play an important role in dimer formation with another protein from the Fo-sector. The Gly on the C-terminal end of this motif (Gly18) is slightly more mobile than the more buried Gly14, based on NMR order parameter measurements (Gly14 S2 = 0.950; Gly18 S2 = 0.895). Only one Su e transmembrane peptide is bound per micelle, and micelles are 22-23 A in diameter, composed of 51 +/- 4 dodecylphosphocholine detergent molecules. Although there is no evidence for Su e homodimerization via the transmembrane domain, potentially synergistic roles for N-terminal (membrane) and C-terminal (soluble) domain interactions may still occur. Furthermore, the presence of a buried charged residue (Arg7) suggests there may be interactions with other Fo-sector protein(s) that stabilize this charge, and possibly drive the folding of the N-terminal 9 residues of the transmembrane domain.  相似文献   

18.
The clinical severity of Osteogenesis Imperfecta (OI), also known as the brittle bone disease, relates to the extent of conformational changes in the collagen triple helix induced by Gly substitution mutations. The lingering question is why Gly substitutions at different locations of collagen cause different disruptions of the triple helix. Here, we describe markedly different conformational changes of the triple helix induced by two Gly substitution mutations placed only 12 residues apart. The effects of the Gly substitutions were characterized using a recombinant collagen fragment modeling the 63-residue segment of the alpha1 chain of type I collagen containing no Hyp (residues 877-939) obtained from Escherichia coli. Two Gly --> Ser substitutions at Gly-901 and Gly-913 associated with, respectively, mild and severe OI variants were introduced by site-directed mutagenesis. Biophysical characterization and limited protease digestion experiments revealed that while the substitution at Gly-901 causes relatively minor destabilization of the triple helix, the substitution at Gly-913 induces large scale unfolding of an unstable region C-terminal to the mutation site. This extensive unfolding is caused by the intrinsic low stability of the C-terminal region of the helix and the mutation induced disruption of a set of salt bridges, which functions to lock this unstable region into the triple helical conformation. The extensive conformational changes associated with the loss of the salt bridges highlight the long range impact of the local interactions of triple helix and suggest a new mechanism by which OI mutations cause severe conformational damages in collagen.  相似文献   

19.
Peptidyl-prolyl cis/trans isomerization, observed in the native state of an increasing number of proteins, is of considerable biological significance. The first evidence for an asymmetric transmission along the polypeptide chain of the structural effects of prolyl isomerization is now derived from the statistics of the C(alpha)/C(alpha)-atom distance distributions in the crystal structures of 848 non-homologous proteins. More detailed information on how isomerization affects segments adjacent to proline is obtained from crystal structures of proteins, that are more than 95% homologous, and that exhibit two different states of isomerization at a particular prolyl bond. The resulting 64 cases, which represent 3.8% of the database used, form pairs of coordinates which were analyzed for the existence of isomer-specific intramolecular nonbonded C(alpha)/C(alpha)-atom distances around the critical proline, and for the positional preferences for particular amino acids in the isomeric sequence segment. The probability that a native protein exhibits both prolyl isomers in the crystalline state increases in particular with a Pro at the third position N-terminal to the isomeric bond (-3 position), and with Ser, Gly and Asp at the position preceding the isomeric bond (-1 position). Structural alignment of matched pairs of isomeric proteins generates three classes with respect to position-specific distribution of C(alpha)-atom displacements around an isomeric proline imide bond. In the majority of cases the distribution of these intermolecular isomer-specific C(alpha)-atom distances shows a symmetric behavior for the N-terminal and C-terminal segment flanking the proline residue, and the magnitude did not exceed 1.3+/-0.6 A including the C(alpha) atoms in proximity to the prolyl bond. However, in the remaining 12 protein pairs the structural changes are unidirectional relative to the isomerizing bond whereby the magnitude of the isomer-specific effect exceeds 3.0+/-2.0 A even at positions remote to proline. Interestingly, the magnitude of the intramolecular isomer-specific C(alpha) atom displacements reveals a lever-arm amplification of the isomerization-mediated structural changes in a protein backbone. The observed backbone effects provide a structural basis for isomer-specific reactions of proline-containing polypeptides, and thus may play a role in biological recognition and regulation.  相似文献   

20.
The bee venom constituent, melittin, is structurally and functionally related to alamethicin. By forming solvent-free planar bilayers of small area (approx. 100 microns 2) on the tip of fire-polished glass pipettes we could observe single melittin pores in these membranes. An increase in the applied voltage induced further non-integral conductance levels. This indicates that melittin forms multi-level pores similar to those formed by alamethicin. Trichotoxin A40, an antibiotic analogue of alamethicin, also induces a voltage-dependent bilayer conductivity, but no stable pore states are resolved. However, chemical modification of the C-terminal molecule part by introduction of a dansyl group leads to a steeper voltage-dependence and pore state stabilization. Comparing structure and activity of several natural and synthetic amphiphilic polypeptides, we conclude that a lipophilic, N-terminal alpha-helical part of adequate length (dipole moment) and a large enough hydrophilic, C-terminal region are sufficient prerequisites for voltage-dependent formation of multi-state pores.  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