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.
2.
A novel method is described, which uses changes in NMR chemical shifts to characterise the structural change in a protein with pressure. Melittin in methanol is a small -helical protein, and its chemical shifts change linearly and reversibly with pressure between 1 and 2000 bar. An improved relationship between structure and HN shift has been calculated, and used to drive a molecular dynamics-based calculation of the change in structure. With pressure, the helix is compressed, with the H—O distance of the NH—O=C hydrogen bonds decreased by 0.021 ± 0.039 Å, leading to an overall compression along the entire helix of about 0.4 Å, corresponding to a static compressibility of 6 ×10–6 bar–1. The backbone dihedral angles and are altered by no more than ± 3° for most residues with a negative correlation coefficient of –0.85 between i and i–1, indicating that the local conformation alters to maintain hydrogen bonds in good geometries. The method is shown to be capable of calculating structural change with high precision, and the results agree with structural changes determined using other methodologies.  相似文献   

3.
Protein disulfide isomerase (PDI) is a multifunctional protein of the endoplasmic reticulum, which catalyzes the formation, breakage and rearrangement of disulfide bonds during protein folding. It consists of four domains designated a, b, b and a. Both a and a domains contain an active site with the sequence motif -Cys-Gly-His-Cys- involved directly in thiol-disulfide exchange reactions. As expected these domains have structures very similar to the ubiquitous redox protein thioredoxin. A low-resolution NMR structure of the b domain revealed that this domain adopts a fold similar to the PDI a domain and thioredoxin [Kemmink, J., Darby, N.J., Dijkstra, K., Nilges, M. and Creighton, T.E. (1997) Curr. Biol., 7, 239–245]. A refined ensemble of solution structures based on the input of 1865 structural restraints shows that the structure of PDI b is well defined throughout the complete protein except for about 10 residues at the C-terminus of the sequence. 15N relaxation data show that these residues are disordered and not part of this structural domain. Therefore the domain boundaries of PDI can now be fixed with reasonable precision. Structural comparison of the PDI b domain with thioredoxin and PDI a reveals several features important for thiol-disulfide exchange activity.  相似文献   

4.
The low-energy orientational oscillations of the peptide groups of an -helix are considered and the value of the frequency is estimated to be in agreement with experiments. Approximate formulae are derived for the projection of a dipole moment on the helix axis and for the helix parameters. Within the framework of a three-chain model, the asymptotics of the soliton solution is obtained using a discrete approach.The analysis of -helix geometry exhibits two types of low-frequency oscillations of the -helix. The first one is connected with atom movements along the helix axis with the peptide groups twisting around the helix axis. Accordingly, it changes the hydrogen bond lengths between neighbouring peptide groups. In the second case, the slopes of the peptide groups to the helix axis oscillate without the helix parameters changing. Here, the energy of interactions between peptide-group dipoles is changed and, as a result, the oscillations have an optical nature. The frequency of the optical orientational oscillations is approximately 100 cm-1.  相似文献   

5.
15N–1H residual dipolar couplings (RDC) have been used as additional restraints to refine the solution structure of the ribotoxin -sarcin. The RDC values were obtained by partial alignment of -sarcin in the binary mixture of n-dodecyl hexa(ethylene glycol)/hexanol. A total of 131 RDCs were measured and 106 were introduced in the final steps of the calculation protocol following the main calculation based on nuclear Overhauser enhancements and torsion angle restraints. A homogeneous family of 81 conformers was obtained. The resulting average pairwise root-mean-square deviation corresponding to the superposition of the 20 best structures is 0.69±0.12 Å for the backbone and 1.29±0.14 Å for all heavy atoms. The new structural features derived from the refined structure, compared with the non-refined structure of -sarcin, consist of new hydrogen bonds and a better definition of the backbone conformation. In particular, the loop segment spanning Gly 60 to Lys 70 shows a single conformation, corresponding to the most populated family of conformers observed in the unrefined structure. The information derived from the analysis of the refined structure and the comparison with the homologous protein restrictocin could help in establishing further structure–function relationships concerning -sarcin which can be reasonably extrapolated to other members of the ribotoxin family.  相似文献   

6.
Summary The 15N relaxation rates of the -aminoisobutyric acid (Aib)-rich peptide alamethicin dissolved in methanol at 27°C and 5°C, and dissolved in aqueous sodium dodecylsulfate (SDS) at 27°C, were measured using inverse-detected one-and two-dimensional 1H–15N NMR spectroscopy. Measurements of 15N longitudinal (RN(Nz)) and transverse (RN(Nx,y)) relaxation rates and the {1H} 15N nuclear Overhauser enhancement (NOE) at 11.7 Tesla were used to calculate (quasi-) spectral density values at 0, 50, and 450 MHz for the peptide in methanol and in SDS. Spectral density mapping at 0, 50, 450, 500, and 550 MHz was done using additional measurements of the 1H–15N lingitudinal two-spin order, RNH(2H infZ supN NZ), two-spin antiphase coherence, RNH(2H infN supZ Nx,y), and the proton longitudinal relaxation rate, RH(H infN supZ ), for the peptide dissolved in methanol only. The spectral density of motions was also modeled using the three-parameter Lipari-Szabo function. The overall rotational correlation times were determined to be 1.1, 2.5, and 5.7 ns for alamethicin in methanol at 27°C and 5°C, and in SDS at 27°C, respectively. From the rotational correlation time determined in SDS the number of detergent molecules associated with the peptide was estimated to be about 40. The average order parameter was about 0.7 and the internal correlation times were about 70 ps for the majority of backbone amide 15N sites of alamethicin in methanol and in SDS. The relaxation data, spectral densities, and order parameters suggest that the peptide N-H vectors of alamethicin are not as highly constrained as the core regions of folded globular proteins. However, the peptide backbone is clearly not as mobile as the most unconstrained regions of folded proteins, such as those found in the frayed C-and N-termini of some proteins, or in randomcoil peptides. The data also suggest significant mobility at both ends of the peptide dissolved in methanol. In SDS the mobility in the middle and at the ends of the peptide is reduced. The implications of the results with respect to the sterically hindered Aib residues and the biological activities of the peptide are discussed.To whom correspondence should be addressed.  相似文献   

7.
8.
The fusion of biological membranes is mediated by integral membrane proteins with α-helical transmembrane segments. Additionally, those proteins are often modified by the covalent attachment of hydrocarbon chains. Previously, a series of de novo designed α-helical peptides with mixed Leu/Val sequences was presented, mimicking fusiogenically active transmembrane segments in model membranes (Hofmann et al., Proc. Natl. Acad. Sci. USA 101 (2004) 14776-14781). From this series, we have investigated the peptide LV16 (KKKW LVLV LVLV LVLV LVLV KKK), which was synthesized featuring either a free N-terminus or a saturated N-acylation of 2, 8, 12, or 16 carbons. We used 2H and 31P NMR spectroscopy to investigate the structure and dynamics of those peptide lipid modifications in POPC and DLPC bilayers and compared them to the hydrocarbon chains of the surrounding membrane. Except for the C2 chain, all peptide acyl chains were found to insert well into the membrane. This can be explained by the high local lipid concentrations the N-terminal lipid chains experience. Further, the insertion of these peptides did not influence the membrane structure and dynamics as seen from the 2H and 31P NMR data. In spite of the fact that the longer acyl chains insert into the membrane, they do not adapt their lengths to the thickness of the bilayer. Even the C16 lipid chain on the peptide, which could match the length of the POPC palmitoyl chain, exhibited lower order parameters in the upper chain, which get closer and finally reach similar values in the lower chain region. 2H NMR square law plots reveal motions of slightly larger amplitudes for the peptide lipid chains compared to the surrounding phospholipids. In spite of the significantly different chain lengths of the acylations, the fraction of gauche defects in the inserted chains is constant.  相似文献   

9.
TRPM7 is an unusual bifunctional protein consisting of an α-kinase domain fused to a TRP ion channel. Previously, we have identified annexin A1 as a substrate for TRPM7 kinase and found that TRPM7 phosphorylates annexin A1 at Ser5 within the N-terminal α-helix. Annexin A1 is a Ca(2+)-dependent membrane binding protein, which has been implicated in membrane trafficking and reorganization. The N-terminal tail of annexin A1 can interact with either membranes or S100A11 protein, and it adopts the conformation of an amphipathic α-helix upon these interactions. Moreover, the existing evidence indicates that the formation of an α-helix is essential for these interactions. Here we show that phosphorylation at Ser5 prevents the N-terminal peptide of annexin A1 from adopting an α-helical conformation in the presence of membrane-mimetic micelles as well as phospholipid vesicles. We also show that phosphorylation at Ser5 dramatically weakens the binding of the peptide to S100A11. Our data suggest that phosphorylation at Ser5 regulates the interaction of annexin A1 with membranes as well as S100A11 protein.  相似文献   

10.
Summary α-conotoxin EI is an 18-residue peptide (RDOCCYHPTCNMSNPQIC; 4–10, 5–18) isolated from the venom ofConus ermineus, the only fish-hunting cone snail of the Atlantic Ocean. This peptide targets specifically the nicotinic acetylcholine receptor (nAChR) found in mammalian skeletal muscle and the electric organTorpedo, showing a novel selectivity profile when compared to other α-conotoxins. The 3D structure of EI has been determined by 2D-NMR methods in combination with dynamical simulated annealing protocols. A total of 133 NOE-derived distances were used to produce 13 structures with minimum energy that complied with the NOE restraints. The structure of EI is characterized by a helical loop between THr9 and Met12 that is stabilized by the Cys4-Cys10 disulfide bond and turns involving Cys4-Cys5 and Asn14-Pro15. Other regions of the peptide appear to be flexible. The overall fold of EI is similar to that of other α4/7-conotoxins (PnIA/B, MII, EpI). However, unlike these other α4/7-conotoxins, EI targets the muscular type nAChR. The differences in selectivity can be attributed to differences in the surface charge distribution among these α4/7-conotoxins. The implications for binding of EI to the muscular nAChR are discussed with respect to the current NMR structure of EI. Supplementary material available:1H resonance assignments of α-conotoxin EI.  相似文献   

11.
(1) High-resolution 31P-NMR was used to study the environment of the phosphoserine residues of the phosphoproteins, αs1-casein B, β-casein A2 and β-casein C. For reference purposes 31P-NMR spectra of phosvitin and ovalbumin were also collected. (2) 31P resonances were assigned to specific phosphoserine residues as a result of comparisons of the high-resolution 31P-NMR spectra for αs1- and β-caseins and for peptide fragments of these proteins obtained by cyanogen bromide and trypsin cleavage. (3) Measurements of the enhancement of the relaxation rate for water protons (1H) on addition of Mn2+ to αs1-casein B and to a fragment αs1-CN3, obtained by cyanogen bromide cleavage, gave approximate pK values for the binding groups and suggest the possibility of a conformational change induced by varying the concentration of divalent cation.  相似文献   

12.
α-Conotoxin EI is an 18-residue peptide (RDOCCYHPTCNMSNPQIC; 4–10, 5–18) isolated from the venom of Conus ermineus, the only fish-hunting cone snail of the Atlantic Ocean. This peptide targets specifically the nicotinic acetylcholine receptor (nAChR) found in mammalian skeletal muscle and the electric organ Torpedo, showing a novel selectivity profile when compared to other α-conotoxins. The 3D structure of EI has been determined by 2D-NMR methods in combination with dynamical simulated annealing protocols. A total of 133 NOE-derived distances were used to produce 13 structures with minimum energy that complied with the NOE restraints. The structure of EI is characterized by a helical loop between Thr9 and Met12 that is stabilized by the Cys4-Cys10 disulfide bond and turns involving Cys4-Cys5 and Asn14-Pro15. Other regions of the peptide appear to be flexible. The overall fold of EI is similar to that of other α4/7-conotoxins (PnIA/B, MII, EpI). However, unlike these other α4/7-conotoxins, EI targets the muscular type nAChR. The differences in selectivity can be attributed to differences in the surface charge distribution among these α4/7-conotoxins. The implications for binding of EI to the muscular nAChR are discussed with respect to the current NMR structure of EI.  相似文献   

13.
Hyperphosphorylation of tau protein (tau) causes neurodegenerative diseases such as Alzheimer's disease (AD). Recent studies of the physiological correlation between tau and α-synuclein (α-SN) have demonstrated that: (a) phosphorylated tau is also present in Lewy bodies, which are cytoplasmic inclusions formed by abnormal aggregation of α-SN; and (b) the neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) increases the phosphorylation of tau as well as the protein level of α-SN in cultured neuronal cells, and also in mice. However, the molecular mechanism responsible for the α-SN-mediated hyperphosphorylation of tau remains to be elucidated. In this in vitro study, we found that: (a) α-SN directly stimulates the phosphorylation of tau by glycogen synthase kinase-3β (GSK-3β), (b) α-SN forms a heterotrimeric complex with tau and GSK-3β, and (c) the nonamyloid beta component (NAC) domain and an acidic region of α-SN are responsible for the stimulation of GSK-3β-mediated tau phosphorylation. Thus, it is concluded that α-SN functions as a connecting mediator for tau and GSK-3β, resulting in GSK-3β-mediated tau phosphorylation. Because the expression of α-SN is promoted by oxidative stress, the accumulation of α-SN induced by such stress may directly induce the hyperphosphorylation of tau by GSK-3β. Furthermore, we found that heat shock protein 70 (Hsp70) suppresses the α-SN-induced phosphorylation of tau by GSK-3β through its direct binding to α-SN, suggesting that Hsp70 acts as a physiological suppressor of α-SN-mediated tau hyperphosphorylation. These results suggest that the cellular level of Hsp70 may be a novel therapeutic target to counteract α-SN-mediated tau phosphorylation in the initial stage of neurodegenerative disease.  相似文献   

14.
The backbone dynamics of the bacteriorhodopsin fragment (1–36)BR solubilized in a 1:1 chloroform/methanol mixture were investigated by heteronuclear 1H-15N NMR spectroscopy. The heteronuclear 15N longitudinal and transverse relaxation rates and 15N{1H} steady-state NOEs were measured at three magnetic fields (11.7, 14.1, and 17.6 T). Careful statistical analysis resulted in the selection of the extended model-free form of the spectral density function [Clore et al. (1990) J. Am. Chem. Soc., 112, 4989–4991] for all the backbone amides of (1–36)BR. The peptide exhibits motions on the micro-, nano-, and picosecond time scales. The dynamics of the -helical part of the peptide (residues 9–31) are characterised by nanosecond and picosecond motions with mean order parameters S s 2 = 0.60 and S f 2 = 0.84, respectively. The nanosecond motions were attributed to the peptide's helix-coil transitions in equilibrium. Residues 3–7 and 30–35 also exhibit motions on the pico- and nanosecond time scales, but with lower order parameters. Residue 10 at the beginning of the -helix and residues 30–35 at the C-terminus are involved in conformational exchange processes on the microsecond time scale. The implications of the obtained results for the studies of helix-coil transitions and the dynamics of membrane proteins are discussed.  相似文献   

15.
Understanding how small molecules affect amyloid formation is of major biomedical and pharmaceutical importance due to the association of amyloid with incurable diseases including Alzheimer's, Parkinson's, and type II diabetes. Using solution state (1)H NMR, we demonstrate that curcumin, a planar biphenolic compound found in the Indian spice turmeric, delays the self-assembly of islet amyloid polypeptide to NMR-invisible assemblies. Accompanying circular dichroism studies show that curcumin disassembles α-helix in maturing assemblies of IAPP. The amount of α-helix disassembled correlates with predicted and experimentally determined helical content of IAPP obtained by others. Taken together, these results indicate that curcumin modulates IAPP self-assembly by unfolding α-helix on pathway to amyloid. The implications of this work in the elucidation of the mechanism for amyloid formation by IAPP in the presence of curcumin are discussed.  相似文献   

16.
Summary The backbone dynamics of uniformly 15N-labelled fragments (residues 1–71 and 1–36) of bacterioopsin, solubilized in two media (methanol-chloroform (1:1), 0.1 M 2HCO2NH4, or SDS micelles) have been investigated using 2D proton-detected heteronuclear 1H-15N NMR spectroscopy at two spectrometer frequencies, 600 and 400 MHz. Contributions of the conformational exchange to the transverse relaxation rates of individual nitrogens were elucidated using a set of different rates of the CPMG spin-lock pulse train and were essentially suppressed by the high-frequency CPMG spin-lock. We found that most of the backbone amide groups of (1–71)bacterioopsin in SDS micelles are involved in the conformational exchange process over a rate range of 103 to 104 s-1. This conformational exchange is supposed to be due to an interaction between two -helixes of (1–71)bacterioopsin, since the hydrolysis of the peptide bond in the loop region results in the disappearance of exchange line broadening. 15N relaxation rates and 1H-15N NOE values were interpreted using the model-free approach of Lipari and Szabo [Lipari, G. and Szabo, A. (1982) J. Am. Chem. Soc., 104, 4546–4559]. In addition to overall rotation of the molecule, the backbone N-H vectors of the peptides are involved in two types of internal motions: fast, on a time scale <20 ps, and intermediate, on a time scale close to 1 ns. The intermediate dynamics in the -helical stretches was mostly attributed to bending motions. A decrease in the order parameter of intermediate motions was also observed for residues next to Pro50, indicating an anisotropy of the overall rotational diffusion of the molecule. Distinctly mobile regions are identified by a large decrease in the order parameter of intermediate motions and correspond to the N- and C-termini, and to a loop connecting the -helixes of (1–71)bacterioopsin. The internal dynamics of the -helixes on the millisecond and nanosecond time scales should be taken into account in the development of a model of the functioning bacteriorhodopsin.Abbreviations BO bacterioopsin - 2D two-dimensional - CPMG Carr-Purcell-Meiboom-Gill (Carr and Purcell, 1954) - SDS sodium dodecyl(2H25) sulfate - R(Sx), R(Sz) 15N transverse and longitudinal relaxation rates, respectively  相似文献   

17.
Conformational flexibility of α-helices in glucoamylase of the fungus Aspergillus awamori was studied by molecular dynamics methods. Several amino acid substitutions (G127A, P128A, I136L, G137A, and G139A) optimizing intrinsic interactions in one of the α-helices (D) within the hydrophobic core of this protein were constructed and studied. It was found that these point mutations had different effects on the glucoamylase thermal inactivation constant. Unlike amino acid substitution P128A and substitutions G137A and A246C, I136L and G139A displayed a pronounced additive thermostabilizing effect.  相似文献   

18.
E.F. Elstner  A. Heupel 《BBA》1973,325(1):182-188
The mechanism of the decarboxylation of α-keto acids by isolated chloroplasts has been studied with the aid of superoxide dismutase and catalase. Using photosynthetic and enzymatic systems, which are known to catalyze peroxidic oxidations, we have been able to demonstrate that both the superoxide free radical ion and H2O2 are necessary for maximal rates of decarboxylation. In isolated chloroplasts, an auto-oxidizable electron acceptor as well as an electron donor for Photosystem I are absolute requirements for the decarboxylation. H2O2 seems to be the primary oxidant in the decarboxylation of pyruvate or glyoxylate by isolated chloroplasts. A secondary rate of decarboxylation is superimposed on the primary one, mediated by superoxide free radical ion. Mn2+ stimulates the decarboxylation probably via intermediarily-formed Mn3+ in a reaction, which is neither inhibited by catalase nor by superoxide dismutase. A decarboxylation of pyruvate or glyoxylate by isolated chloroplasts in the presence of NADP+ is initiated, as soon as the available NADP+ is fully reduced. In this case, the open-chain electron transport seems to switch from NADP+ to oxygen as the terminal electron acceptor.  相似文献   

19.
Fibrillar aggregates of the islet amyloid polypeptide (IAPP) and amyloid-β (Aβ) are known to deposit at pancreatic β-cells and neuronal cells and are associated with the cell degenerative diseases type-2 diabetes mellitus (T2DM) and Alzheimer's disease (AD), respectively. Since IAPP is secreted by β-cells and a membrane-damaging effect of IAPP has been discussed as a reason for β-cell dysfunction and the development of T2DM, studies of the interaction of IAPP with the β-cell membrane are of high relevance for gaining a molecular-level understanding of the underlying mechanism. Recently, it has also been shown that patients suffering from T2DM exhibit an increased risk to develop AD and vice versa, and a molecular link between AD and T2DM has been suggested. In this study, membrane lipids from the rat insulinoma-derived INS-1E β-cell line were isolated, and their interaction with the amyloidogenic peptides IAPP and Aβ and a mixture of both peptides has been studied. To yield insight into the associated peptides' conformational changes and their effect on the membrane integrity during aggregation, we have carried out attenuated total reflection Fourier transform infrared spectroscopy, fluorescence microscopy, and atomic force microscopy experiments. The IAPP-Aβ heterocomplexes formed were shown to adsorb, aggregate, and permeabilize the isolated β-cell membrane significantly slower than pure IAPP, however, at a rate that is much faster than that of pure Aβ. In addition, it could be shown that isolated β-cell membranes cause similar effects on the kinetics of IAPP and IAPP-Aβ fibril formation as anionic heterogeneous model membranes.  相似文献   

20.
This report describes the backbone amide dynamics of the uniformly 15N labeled catalytic domain of human stromelysin complexed to PNU-99533, a hydroxamate-containing ligand that binds to the S1-S3 region (right side) of the stromelysin active site, and to PNU-107859 and PNU-142372, both thiadiazole-containing ligands that bind to the S1-S3 region (left side) of the stromelysin active site. 15N R1, R2 and NOE NMR relaxation measurements were recorded and analyzed for each complex. Different dynamic behaviors were observed for stromelysin complexed to the two classes of ligands, indicating that it may be possible to use protein dynamics to distinguish between different binding orientations. In the absence of bound ligand at the S1-S3 subsites, the S1-S3 residues were found to be relatively rigid. In contrast, the S1-S3 subsites were found to be flexible in the absence of interactions with ligand. The relative rigidness of the S1-S3 subsites may be responsible for MMP binding specificity by discriminating between ligands of different shapes. By contrast, the inherent flexibility of the S1-S3 subsites allows structural rearrangement to accommodate a broad range of incoming substrates or inhibitors. Similarities and differences in dynamics observed for each complex provide insights into the interactions responsible for protein–ligand recognition. The relevance of protein dynamics to structure-based drug design is discussed.  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