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.
The refined crystal structure of deoxyhemoglobin S (Padlan, E. A., and Love, W. E. (1985) J. Biol. Chem. 260, 8272-8279) was used to analyze in detail the molecular interactions between hemoglobin tetramers in the crystal. The analysis confirms the close similarity and also the nonequivalence of the molecular interactions involving the two independent tetramers in the asymmetric unit of the crystal. The residue at the site of the hemoglobin S mutation, beta 6, is intimately involved in the lateral contacts between adjacent molecules. The molecular contacts in the crystals of deoxyhemoglobin S, deoxyhemoglobin A, and deoxyhemoglobin F were compared; some contacts involve the same regions of the molecule although the details of the interactions are very different. The effect of introducing an R state tetramer into the deoxyhemoglobin S strands was investigated using the known structure of carbon monoxyhemoglobin A. It was found that substituting a molecule of carbon monoxyhemoglobin A for one of the deoxyhemoglobin S tetramers results in extensive molecular interpenetration.  相似文献   

2.
Fibers of deoxyhemoglobin S obtained directly from lysed sickled red blood cells have been compared with fibers from chromatographically pure deoxyhemoglobin S solutions of known chemical composition. Electron micrographs of negatively stained specimens reveal that the molecular packing within the fibers remains largely invariant with changes in pH, ionic strength, Mg2+ concentration, 2,3-diphosphoglycerate concentration, temperature or the method of deoxygenation.When solutions of chromatographically pure deoxyhemoglobin S are stirred, the fibers align into well defined fascicles. After several hours of stirring, long needles and twisted ribbons develop and in a relatively short time replace the fascicles in solution. With continued stirring all forms are replaced by small crystals. By use of electron microscopy and low-angle X-ray diffraction we have found these crystals to have cell parameters indistinguishable from those of crystals grown in polyethylene glycol and citrate/phosphate buffer at pH 5 to 6 (Wishner et al., 1975a).Our evidence indicates that crystal formation in stirred solutions of deoxyhemoglobin S is the result of a progressive alignment and fusion of the fibers, and that the molecular arrangement within the fibers is closely related to that within the crystal. The remarkable pH invariance of the molecular packing within the fiber and crystal structures is consistent with the dominance of hydrophobic bonding between molecules. The β6-valine contact observed by Wishner et al. (1975b) is apparently the pathological contact responsible for the polymerization of deoxyhemoglobin S in vivo. On the basis of our observations and knowledge of the crystal structure we propose that the deoxyhemoglobin S fiber consists of eight molecular double strands, four of which run in each direction along the length of the fiber.  相似文献   

3.
The mutations in hemoglobin Nancy beta145(HC2) Tyr leads to Asp and hemoglobin Cochin-Portal-Royal beta146(HC3) His leads to Arg involve residues which are thought to be essential for the full expression of allosteric action in hemoglobin. Relative to the structure of deoxyhemoglobin A, our x-ray study of deoxyhemoglobin Nancy shows severe disordering of the beta chain COOH-terminal tetrapeptide and a possible movement of the beta heme iron atom toward the plane of the porphyrin ring. These structural perturbations result in a high oxygen affinity, reduced Bohr effect, and lack of cooperatively in hemoglobin Nancy. In the presence of inositol hexaphosphate (IHP), the Hill constant for hemoglobin Nancy increases from 1.1 to 2.0. But relative to its action on hemoglobin A, IHP is much less effective in reducing the oxygen affinity and in increasing the Bohr effect of hemoglobin Nancy. This indicates that IHP does not influence the R in equilibrium T equilibrium as much in hemoglobin Nancy as in hemoglobin A, and this probably is due to the disordering of His 143beta which is known to be part of the IHP binding site. IHP is also known to produce large changes in the absorption spectrum of methemoglobin A, but we find that it has no effect on the spectrum of methemoglobin Nancy. In contrast to the large structural changes in deoxyhemoglobin Nancy, the structure of deoxyhemoglobin Cochin-Port-Royal differs from deoxyhemoglobin A only in the position of the side chain of residue 146beta. The intrasubunit salt bridge between His 146beta and Asp 94beta in deoxyhemoglobin A is lost in deoxyhemoglobin Cochin-Portal-Royal with the guanidinium ion of Arg 146beta floating freely in solution. This small difference in structure results in a reduced Bohr effect, but does not cause a change in the Hill coefficient, the response to 2,3-diphosphoglycerate, or the oxygen affinity at physiological pH.  相似文献   

4.
Closing in on the resting state of the Shaker K(+) channel   总被引:4,自引:0,他引:4  
Membrane depolarization causes voltage-gated ion channels to transition from a resting/closed conformation to an activated/open conformation. We used voltage-clamp fluorometry to measure protein motion at specific regions of the Shaker Kv channel. This enabled us to construct new structural models of the resting/closed and activated/open states based on the Kv1.2 crystal structure using the Rosetta-Membrane method and molecular dynamics simulations. Our models account for the measured gating charge displacement and suggest a molecular mechanism of activation in which the primary voltage sensors, S4s, rotate by approximately 180 degrees as they move "outward" by 6-8 A. A subsequent tilting motion of the S4s and the pore domain helices, S5s, of all four subunits induces a concerted movement of the channel's S4-S5 linkers and S6 helices, allowing ion conduction. Our models are compatible with a wide body of data and resolve apparent contradictions that previously led to several distinct models of voltage sensing.  相似文献   

5.
J H Kim  A G Marshall 《Biopolymers》1992,32(9):1263-1270
The structures of the helices II-III region and the helix IV region of B. megaterium 5S rRNA have been examined by means of energy minimization and molecular dynamics calculations. Calculated distances between neighboring hydrogen-bonded imino protons in helices II, III, and IV were between 3.5 and 4.5 A. The overall axis for the helices II-III region is warped rather than straight. Formation of additional Watson-Crick base pairs in loop B and loop C was not evident from the atomic positions calculated by molecular dynamics. Bases in loop C are well stacked, showing no significant change during dynamics. Bulge migration in helix III does not seem to be possible; the helices II-III region prefers one conformation. Helix II is more stable than helix III. Five base pairs in helix IV were sufficiently stable to establish that helix IV is terminated by a hairpin loop of three nucleotides. U87 protrudes from loop D. Structures of the helices II-III segment and the helix IV segment of B. megaterium 5S rRNA obtained by molecular dynamics were generally consistent with the solution structure inferred from high-field proton nmr spectroscopy.  相似文献   

6.
Correlation times for the tumbling motion of the spin probe 2,2,6,6,-tetramethyl piperidone-N-oxyl (Tempone) were obtained in the presence of different concentrations of oxyhemoglobin A, oxyhemoglobin S, and deoxyhemoglobin S and compared to the viscosity of non-gelling hemoglobin solutions. Reorientational motion (or tumbling) of Tempone in gelled solutions of deoxyhemoglobin S is as great as that in non-gelled hemoglobins of the same total concentration. It is concluded that the gel does not exclusively partition Tempone into an aqueous phase of lower solute concentration after gel formation. The gel at room temperature is a highly mobile and dynamic structure on the microscopic level.  相似文献   

7.
Secondary structure model for 23S ribosomal RNA.   总被引:31,自引:32,他引:31       下载免费PDF全文
A secondary structure model for 23S ribosomal RNA has been constructed on the basis of comparative sequence data, including the complete sequences from E. coli. Bacillus stearothermophilis, human and mouse mitochondria and several partial sequences. The model has been tested extensively with single strand-specific chemical and enzymatic probes. Long range base-paired interactions organize the molecule into six major structural domains containing over 100 individual helices in all. Regions containing the sites of interaction with several ribosomal proteins and 5S RNA have been located. Segments of the 23S RNA structure corresponding to eucaryotic 5.8S and 25 RNA have been identified, and base paired interactions in the model suggest how they are attached to 28S RNA. Functionally important regions, including possible sites of contact with 30S ribosomal subunits, the peptidyl transferase center and locations of intervening sequences in various organisms are discussed. Models for molecular 'switching' of RNA molecules based on coaxial stacking of helices are presented, including a scheme for tRNA-23S RNA interaction.  相似文献   

8.
Secondary structure predictions for membrane proteins are relatively reliable and permit the construction of model structures that may serve as initial conformations for molecular dynamics simulations. This might provide a scheme to predict the three-dimensional structures of membrane proteins. The feasibility of such an approach is tested for bacteriorhodopsin. We were not able to fully predict the kidney-shaped structure of bacteriorhodopsin. However, features compatible with this structure developed in a simulation starting from a circular arrangement of the seven predicted helices. When instead we started from the kidney shape, assigning the seven predicted helices in different ways to those on the structure, we could distinguish between the different assignments on the basis of energy and tilt of the helices. In this way we could select the correct assignment from a few others. For the correct assignment, the helices spontaneously adopted a tilt that agrees remarkably well with the experimental model structure derived by others. The root-mean-square deviation between our best molecular dynamics structure and the experimental model structure is 3.8 A, caused mainly by deviations in the internal degrees of freedom of the helices.  相似文献   

9.
Electron micrographs of deoxyhemoglobin S fiber cross sections provide an end-on view of the fiber whose appearance is sensitive to small changes in orientation. We have developed a procedure to exploit this sensitivity in order to determine the hand of these particles. In a sickle hemoglobin fiber the hemoglobin molecules form long pitch helical strands which twist about the particle axis with a pitch of about 3000 A. Tilting a 400-A-thick cross section by a few degrees aligns one of the long pitch helices so that it is nearly parallel to the direction of view. When a strand of hemoglobin molecules in a fiber is aligned in this manner it appears as a strongly contrasted bright spot. It is this spot, rather than the fiber axis, which appears to be the apparent center of rotation of the cross section. The direction of the displacement of the spot from the particle axis depends upon the particle hand and tilt direction. We have used this property to determine that sickle hemoglobin fibers are right-handed particles. This method may be applicable to other particles with long pitch helices as well.  相似文献   

10.
T B Woolf 《Biophysical journal》1997,73(5):2376-2392
Understanding the role of the lipid bilayer in membrane protein structure and dynamics is needed for tertiary structure determination methods. However, the molecular details are not well understood. Molecular dynamics computer calculations can provide insight into these molecular details of protein:lipid interactions. This paper reports on 10 simulations of individual alpha-helices in explicit lipid bilayers. The 10 helices were selected from the bacteriorhodopsin structure as representative alpha-helical membrane folding components. The bilayer is constructed of dimyristoyl phosphatidylcholine molecules. The only major difference between simulations is the primary sequence of the alpha-helix. The results show dramatic differences in motional behavior between alpha-helices. For example, helix A has much smaller root-mean-squared deviations than does helix D. This can be understood in terms of the presence of aromatic residues at the interface for helix A that are not present in helix D. Additional motions are possible for the helices that contain proline side chains relative to other amino acids. The results thus provide insight into the types of motion and the average structures possible for helices within the bilayer setting and demonstrate the strength of molecular simulations in providing molecular details that are not directly visualized in experiments.  相似文献   

11.
Several types of lipid-associating helices exist: transmembrane helices such as in receptor proteins, pore-forming helices in ion channel proteins, fusion-inducing peptides in viral proteins, and amphipathic helices such as in plasma apolipoproteins. In order to propose a classification of these helices according to their molecular properties, we introduce the concept of molecular hydrophobicity potential for such helical segments. The calculation of this parameter for alpha-helices enables the visualization of the hydrophobic and hydrophilic envelopes around the peptide and their three-dimensional representation by molecular graphics. We have used this parameter to differentiate between pore-forming helices with a hydrophobic envelope larger than the hydrophilic component, membrane-spanning helices surrounded almost entirely by an hydrophobic envelope, fusiogenic peptides with an hydrophobicity gradient both around the helix and along the axis, and finally, amphipathic helices with a predominantly hydrophilic envelope. The structure of the lipid-protein complexes is determined by a number of different interactions: the hydrophobic interaction of the apolar faces of the helices with lipids, the polar interaction of the hydrophilic sides of different helices with each other, and the interaction of hydrophilic residues with the aqueous solvent. The relative magnitude of the hydrophobic and hydrophilic envelopes accounts for the differences in the structure of the lipid-protein complexes. Purely hydrophobic interactions stabilize transmembrane helical segments, while hydrophobic interactions with the lipid phase and with each other are involved in the stabilization of the pore-forming helices. In contrast, both hydrophobic interactions with the lipids and hydrophilic interactions with the aqueous phase contribute to the arrangement of amphipathic helices around the edges of the discoidal lipid-apoprotein complexes.  相似文献   

12.
We report the results of thermodynamic and kinetic studies on the gelation of mixtures of sickle cell (S) deoxyhemoglobin with normal human adult (A) and fetal (F) deoxyhemoglobins. The delay time of thermally induced gelation was monitored by the increase in turbidity. At the completion of gelation the solubility was determined by sedimenting the polymers and measuring the supernatant concentration spectrophotometrically. Addition of hemoglobins A or F, at mole fractions from 0 to 0.6, resulted in large increases in both the solubility and the delay time. For a 50:50 mixture of deoxyhemoglobin F with deoxyhemoglobin S, the solubility increased by a factor of 1.8 and the delay time by a factor of 107 relative to pure deoxyhemoglobin S at the same total concentration, while for a 50:50 mixture of deoxyhemoglobins A and S the solubility increased by a factor of 1.4 and the delay time by a factor of 104. The relative delay times were independent of both temperature and total hemoglobin concentration. The data have been analyzed according to theoretical models which treat the effects of temperature, concentration, non-ideality and solution composition on the thermodynamics and kinetics of gelation. The increased solubility in mixtures with deoxyhemoglobin F is fully explained by a model in which only deoxyhemoglobin S molecules polymerize. The effect of fetal hemoglobin (α2γ2) and hybrid α2γβS molecules is to increase the solution non-ideality through the contribution of their excluded volume. The smaller increase in the solubility observed in comparable mixtures with deoxyhemoglobin A requires that the hybrid α2βAβS molecules copolymerize with the deoxyhemoglobin S. The kinetic results for the mixtures can be quantitatively accounted for using a nucleation model in which the equilibrium properties of the polymer are used to describe the critical nucleus. The very large increases in delay time observed for the SF mixtures can be explained by assuming that only α2β2S molecules participate in the formation of a nucleus containing about 25 monomers. As in the thermodynamic analysis, the smaller effect of adding deoxyhemoglobin A can be attributed to the contribution of the hybrid molecules in forming the critical nucleus. Thus the difference between the polymerization properties of mixtures of deoxyhemoglobin S with deoxyhemoglobins A and F can be attributed solely to the copolymerization of the α2βAβS hybrid molecule and the absence of any significant copolymerization of the α2γβS hybrid.  相似文献   

13.
From a consideration of the primary sequence of hemoglobin A2 and the reported 5 å molecular contacts between deoxyhemoglobin S molecules in a crystal, it is predicted that hemoglobin A2 might act as an inhibitor of the polymerization of deoxyhemoglobin S in a manner similar to hemoglobin. F. This has been tested experimentally by measuring the rate of change of the transverse water proton relaxation times (T2) in equimolar mixtures of hemoglobin S and one of the non-gelling hemoglobins A, F or A2. Hemoglobins A2 and F have far more pronounced inhibitory effects on the rate of polymerization than does hemoglobin A. These molecules contain several amino acid differences from hemoglobin A beta chains which are located in the 5 Å molecular crystal contacts and these altered crystal contacts result in a much stronger inhibition of the rate of polymerization. Since hemoglobin A2 is a normal hemoglobin found in small amounts in all adult red cells, increased delta chain synthesis may have potential importance in therapy for sickle cell disease.  相似文献   

14.
The self-association of deoxyhemoglobin S was measured in dilute solutions (0 to 5 g/dl) by Rayleigh light scattering at 630 nm and osmometry in 0.05 M potassium phosphate buffer (pH 7.35). Weight and number average molecular weights (Mw and Mn, respectively) and the second or higher virial coefficients, B' were determined. No experimentally significant differences were observed between oxy- and deoxy-Hb S up to the concentration of 2 g/dl; their apparent average molecular weights were within experimental error. Above that concentration, both Mn and Mw of deoxy-Hb S were significantly different from that of oxy-Hb S. The negative second viral coefficent of deoxy-Hb S, observed by both techniques, is consistent with the self-association of this protein. The lack of effect of 0.4 M propylurea on the state of aggregation and the significant influence of 0.1 M NaCl suggests that polar interactions are involved in formation of these aggregates.  相似文献   

15.
The structure of ribosomal 5 S RNA has been examined using several physical biochemical techniques. Hydrodynamic measurements yield a s020,omega and [eta] of 5.5 x 10(-13) x and 6.9 ml/g, respectively. Other parameters calculated from these values indicate the shape of 5 S RNA is consistent with that of a prolate ellipsoid 160 A in length and 32 A wide. Sedimentation equilibrium results show that 5 S RNA exists as a monomer in the reconstitution buffer with an apparent molecular weight of 44,000. Ultraviolet absorption difference spectra show that approximately 75% of the bases in 5 S RNA are involved in base pairing, and of these base pairs 70% are G-C and 30% are A-U. These results on the overall shape and secondary structure of 5 S RNA have been incorporated with the results of other investigators as to the possible location of single-stranded and double-stranded helical regions, and a molecular model for 5 S RNA is proposed. The molecular model consists of three double helices in the shape of a prolate ellipsoid, with two of the double helical regions at one end of the molecule. The structure is consistent with the available data on the structure and function of 5 S RNA and bears similarity to the molecular model proposed by Osterberg et al. ((1976) Eur. J. Biochem. 68, 481-487) based on small angle x-ray scattering results and the secondary structure proposed by Madison ((1968) Annu. Rev. Biochem. 37, 131-148).  相似文献   

16.
In normal deoxyhemoglobin A, the beta chain COOH-terminal peptide adopts a well ordered structure which is needed for the full expression of allosteric action. Our crystallographic studies of deoxyhemoglobin Creteil (beta 89 Ser replaced by Asn), a variant hemoglobin characterized by high oxygen affinity and a very low level of allosteric function, show that replacement of Ser 89 beta by asparagine causes severe disordering of the beta chain COOH-terminal tetrapeptide. This results, as shown by our spectroscopic studies, in the destabilization of the quaternary structure of deoxyhemoglobin Creteil. We find, furthermore, that the changes in tertiary structure observed in deoxyhemoglobin Creteil are common to other variant hemoglobins having similar functional abnormalities but very different changes in primary structure. In particular, direct comparison of the difference electron density map of deoxyhemoglobin Creteil with that of deoxyhemoglobin Nancy (beta 145 Tyr replaced by Asp) suggests that these two abnormal hemoglobins may have the same mechanism of dysfunction despite the very different nature of their respective sequence changes.  相似文献   

17.
Secondary structure models of the 5.8S rRNA and both internal transcribed spacers (ITS1 and ITS2) are proposed for Calciodinelloideae (Peridiniaceae) and are also plausible for other dinoflagellates. The secondary structure of the 5.8S rRNA corresponds to previously developed models, with two internal paired regions and at least one 5.8S rRNA–28S rRNA interaction. A general secondary structure model of ITS1 for Calciodinelloideae (and other dinoflagellates), consisting of an open multibranch loop with three major helices, is proposed. The homology of these paired regions with those found in other taxa, published in previous studies (e.g. yeast, green algae and Platyhelmithes) remains to be determined. Finally, a general secondary structure model of ITS2 for Calciodinelloideae (and other dinoflagellates) is reconstructed. Based on the 5.8S rRNA–28S rRNA interaction, it consists of a closed multibranch loop, with four major helices. At least helix III and IV have homology with paired regions found in other eukaryotic taxa (e.g. yeast, green algae and vertebrates). Since the secondary structures of both ITS regions are more conserved than the nucleotide sequences, their analysis helps in understanding molecular evolution and increases the number of structural characters. Thus, the structure models developed in this study may be generally useful for future phylogenetic analyses.  相似文献   

18.
We have used the "osmotic stress" method to determine the phase diagram of deoxyhemoglobin S polymerization. This method involves equilibration, through a semipermeable membrane, of the protein with solutions of inert polymers of known osmotic pressure. With deoxyhemoglobin A and S solutions, in which we have demonstrated achievement of equilibrium, plots of osmotic pressure versus concentration initially agree closely with the results of other methods of measurement of colligative properties. However, once the known solubility value is exceeded for the deoxyhemoglobin S solutions at various temperatures, there is a rapid rise in hemoglobin concentration over a narrow osmotic pressure range and then a more gradual increase in concentration. We believe that these two regions correspond, respectively, to the onset of the polymerization process, and of subsequent continuing growth and compression or alignment of polymer. We derive the thermodynamic values for these processes and show that the behavior of the deoxyhemoglobin S system is analogous to the phase transition for a simple chemical system. These results are relevant to understanding the intracellular polymerization of deoxyhemoglobin S in sickle cell disease, and these concepts are applicable to other protein assembly systems.  相似文献   

19.
Posson DJ  Selvin PR 《Neuron》2008,59(1):98-109
Voltage-driven activation of Kv channels results from conformational changes of four voltage sensor domains (VSDs) that surround the K(+) selective pore domain. How the VSD helices rearrange during gating is an area of active research. Luminescence resonance energy transfer (LRET) is a powerful spectroscopic ruler uniquely suitable for addressing the conformational trajectory of these helices. Using a geometric analysis of numerous LRET measurements, we were able to estimate LRET probe positions relative to existing structural models. The experimental movement of helix S4 does not support a large 15-20 A transmembrane "paddle-type" movement or a near-zero A vertical "transporter-type" model. Rather, our measurements demonstrate a moderate S4 displacement of 10 +/- 5 A, with a vertical component of 5 +/- 2 A. The S3 segment moves 2 +/- 1 A in the opposite direction and is therefore not moving as an S3-S4 rigid body.  相似文献   

20.
BACKGROUND: Ribosomes are complex macromolecular machines that perform the translation of the genetic message. Cryo-electron microscopic (cryo-EM) maps of the Escherichia coli 70S ribosome are approaching a resolution of 10 A and X-ray maps of the 30S and 50S subunits are now available at 5 A. These maps show a lot of details about the inner architecture of the ribosome and ribosomal RNA helices are clearly visible. However, in the absence of further biological information, even at the higher resolution of the X-ray maps many rRNA helices can be placed only tentatively. Here we show that genetic tagging in combination with cryo-EM can place and orient double-stranded RNA helices with high accuracy. RESULTS: A tRNA sequence inserted into the E. coli 23S ribosomal RNA gene, at one of the points of sequence expansion in eukaryotic ribosomes, is visible in the cryo-EM map as a peripheral 'foot' structure. By tracing its acceptor-stem end, the location of helix 63 in domain IV and helix 98 in domain VI of the 50S subunit could be precisely determined. CONCLUSIONS: Our study demonstrates for the first time that features of a three-dimensional cryo-EM map of an asymmetric macromolecular complex can be interpreted in terms of secondary and primary structure. Using the identified helices as a starting point, it is possible to model and interpret, in molecular terms, a larger portion of the ribosome. Our results might be also useful in interpreting and refining the current X-ray maps.  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