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.
The conformation and dynamic structure of single-stranded poly(inosinic acid), poly(I), in aqueous solution at neutral pH have been investigated by nmr of four nuclei at different frequencies: 1H (90 and 250 MHz), 2H (13.8 MHz), 13C (75.4 MHz), and 31P (36.4 and 111.6 MHz). Measurements of the proton-proton coupling constants and of the 1H and 13C chemical shifts versus temperature show that the ribose is flexible and that base-base stacking is not very significant for concentrations varying from 0.04 to 0.10M in the monomer unit. On the other hand, the proton T1 ratios between the sugar protons, T1 (H1′)/T1 (H3′), indicate a predominance of the anti orientation of the base around the glycosidic bond. The local motions of the ribose and the base were studied at different temperatures by measurements of nuclear Overhauser enhancement (NOE) of protonated carbons, the ratio of the proton relaxation times measured at two frequencies (90 and 250 MHz), and the deuterium quadrupolar transverse relaxation time T2. For a given temperature between 22 and 62°C, the 13C-{1H} NOE value is practically the same for seven protonated carbons (C2, C8, C1′, C2′, C3′, C4′, C5′). This is also true for the T1 ratio of the corresponding protons. Thus, the motion of the ribose–base unit can be considered as isotropic and characterized by a single correlation time, τc, for all protons and carbons. The τc values determined from either the 13C-{1H} NOE or proton T1 ratios, T1(90 MHz)/T1(250 MHz), and/or deuterium transverse relaxation time T2 agree well. The molecular motion of the sugar-phosphate backbone (O-P-O) and the chemical-shift anisotropy (CSA) were deduced from T1 (31P) and 31P-{1H} NOE measurements at two frequencies. The CSA contribution to the phosphorus relaxation is about 12% at 36.4 MHz and 72% at 111.6 MHz, corresponding to a value of 118 ppm for the CSA (σ = σ∥ ? σ?). Activation energies of 2–6 kcal/mol for the motion of the ribose–base unit and the sugarphosphate backbone were evaluated from the proton and phosphorus relaxation data.  相似文献   

3.
4.
A new thionin from barley, ω-hordothionin, has been shown to exist in aqueous solution as a mixture of two different isoforms in a 3:2 ratio, as revealed by a complete analysis of its two-dimensional 1H-nmr spectra. The conformational heterogeneity arises frtm cis–trans isomerism ahout the Phe 12–Pro 13 peptide bond, where the major, form corresponds to the cis conformation. The complete assignment of chemical shifts and nuclear Overhaiiser effects (NOES) of the two isoforms allow a detailed comparative analysis of their conformational properties, even though a complete calculation of their solution structures is not possible because of a somewhat limited number of NOE constraints. Structures for the two isomers could be modeled, however, on the basis of the high structural homology between ω-hordothionin and related γ-thionins, and under the conditions of satisfying all observed experimental data. The two isoforms adopt practically identical global folds and the structural changes imposed by cis–trans isomerization are confined to the region proximal to Pro 13. The cis–trans isomerism occurs in a conserved loop connecting the first β-strand of the triple-stranded antiparallel β-sheet and the α-helix. A comparative analysis of the sequences of this loop in the different thionins suggests that the cis–trans equilibrium about the X-Pro peptide bond depends on the size of the side chain of X (X = Gly in γ-thionins and Phe in ω-thionin). The structural homology of this new thionin with γ-thionins as well as with some scorpion toxins and insect defensins suggests that these proteins may share a common mode of functional activity. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
The physical properties of the DNA oligomer d(CGCGCGTTTTCGCGCG) in solvents containing 4 M NaClO4 and 0.1 M NaCl were investigated by proton NMR, optical melting, and circular dichroism spectroscopy. Results of these investigations are as follows: (i) The DNA hexadecamer exists as a unimolecular hairpin in either high or low salt. (ii) In high salt the stem region of the hairpin is in the left-handed Z conformation. (iii) In either high or low salt, the duplex stem of the hairpin is stabilized against melting by approximately 40 degrees C compared to the linear core duplex. The added stability of the hairpin is entropic in origin. (iv) In high salt, as the temperature is elevated, the equilibrium structure of the duplex stem of the hairpin shifts from the Z to the B conformation before melting. (v) In low salt, when the DNA duplex exists in the B conformation, attachment of a T4 single-strand loop to one end only slightly decreases (by 14%) the correlation time of the CH5-CH6 interproton vector. In high salt, when the DNA duplex exists in the Z conformation, the correlation time of the CH5-CH6 interproton vector decreases by 51%. Since these viscosity-corrected correlation times are taken to be indicators of duplex motions on the nanosecond time scale, this result directly suggests a larger amplitude of these motions is present in the duplex stem of the hairpin when it exists in the Z conformation.  相似文献   

6.
The effect of the number of methylene groups in the side chains on the conformation of polypeptides is assessed for three poly(L -lysine) homologs with R = –(CH2)nNH2. Circular dichroism studies show a pH-induced helix–coil transition in 0.05 M KCl with midpoints at 9.6, 9.0, and 8.7 for n = 5, 6, and 7, respectively, as compared with 10.1 for (Lys)x (n = 4). Homologs with n = 6 and 7 could be partially helical even when the side groups are fully charged (with n = 7, the compound is highly aggregated above pH 9.1). Thus, the longer the number of methylene groups the more stable is the helical conformation of these homologs. Potentiometric titration of the n = 5 homolog gives a ΔG° of ?310 cal/mol (residue) for the uncharged coil-to-helix transition at 25°C. The corresponding ΔH° and ΔS° are ?1740 cal/mol (residue) and ?4.8 e.u./mol (residue). Unlike (Lys)x, the uncharged helix-to-β transition is slow and incomplete even after heating at 80°C for 1 hr. Addition of methanol enhances the helical formation in neutral solution with midpoints at 72, 52, and 27% methanol (v/v) for n = 5, 6, and 7, respectively [cf. 88% for (Lys)x]. Addition of sodium dodecyl sulfate induces a coil-to-helix transition for all three homologs in contrast with the β form of (Lys)x under similar conditions.  相似文献   

7.
E K Achter  G Felsenfeld 《Biopolymers》1971,10(9):1625-1634
To elucidate the role of the bases in single-strand polynucleolide conformation, we have studied apurinic acid, a single-strand polydeoxyribonucleotide from which almost half the bases have been removed. The conformation of apurinic acid in aqueous solution near θ condition has been investigated by sedimentation velocity and sedimentation equilibrium measurements. The unperturbed coil dimensions of apurinic acid are essentially identical to those of poly rU of the same degree of polymerization. The dimensions are also similar to those of poly rA at high temperature, where the adenine residues are not stacked upon one another. We conclude that the considerable rigidity of these polynuclotides is conferred not by residual, undetected base stacking, but by restrictions in rotation about the bonds of the backbone. Furthermore, the rigidity of the ribose-phosphate backbone cannot be attributed to interactions involving the 2'—OH group.  相似文献   

8.
9.
Yen-Yau H. Chao  R. Bersohn 《Biopolymers》1978,17(12):2761-2767
In aqueous solutions, 13C- and 1H-nmr studies show that the percentage of trans conformation of proline oligomers +H2H Pro-(Pro)n-CO increases substantially from n = 1 (65% trans) to n = 2 (90% trans). The relatively low percentage of trans structure for the dimer (n = 1) very likely is caused by the extra stability acquired by the end-to-end intramolecular H-bonding of the cis dimer. As n increase from 2 to 3 (or 5) in +H2N-Pro-(Pro)n-CO, the percentage of trans conformation stays more or less constant (~0.9). A high salt concentration (4M CaCl2) causes a conformation randomization, so that the short-chain oligomer (n = 1, 3, 5) and the long-chain poly (L -proline) all show about the same frantion of trans conformation (0.7-0.8).  相似文献   

10.
Ubiquitin modification of proteins is used as a signal in many cellular processes. Lysine side-chains can be modified by a single ubiquitin or by a polyubiquitin chain, which is defined by an isopeptide bond between the C terminus of one ubiquitin and a specific lysine in a neighboring ubiquitin. Polyubiquitin conformations that result from different lysine linkages presumably differentiate their roles and ability to bind specific targets and enzymes. However, conflicting results have been obtained regarding the precise conformation of Lys48-linked tetraubiquitin. We report the crystal structure of Lys48-linked tetraubiquitin at near-neutral pH. The two tetraubiquitin complexes in the asymmetric unit show the complete connectivity of the chain and the molecular details of the interactions. This tetraubiquitin conformation is consistent with our NMR data as well as with previous studies of diubiquitin and tetraubiquitin in solution at neutral pH. The structure provides a basis for understanding Lys48-linked polyubiquitin recognition under physiological conditions.  相似文献   

11.
A heat-stable endoribonuclease isolated from chicken liver has been purified to homogeneity as evidenced by the presence of a single protein band upon polyacrylamide gel electrophoresis. The enzyme can, in limit digests of 5 S rRNA and 5.8 S rRNA, dinstinguish between cytidylic and uridylic acids bonds at a ratio of 61:1 and, therefore, may be useful in RNA sequence analysis. The means by which the enzyme hydrolyzes substrate is unusual in that kinetic data do not support a simple formation and breakdown of an enzyme . substrate complex. Rather, the existence of a second complex, consisting of 2 mol of substrate and one of enzyme, derived from the initial enzyme . substrate complex, is postulated. In common with the other endonucleases, enzyme activity is inhibited by free poly(A) or tracts of the polypurine present at the 3'-terminus of RNA. Reversal of inhibition and restoration of activity may be achieved by the addition of low concentrations of spermidine to reaction mixtures.  相似文献   

12.
The secondary structure of the human growth hormone releasing factor (GRF 1-29) in a solution mixture of 60% aqueous phosphate buffer:40% 2,2,2-trifluoroethanol-d3 has been investigated by two-dimensional 1H-nmr spectroscopy. Sequential resonance assignments and elements of secondary structure were obtained from phase-sensitive correlation spectroscopy, relayed coherence spectroscopy, and nuclear Overhauser spectroscopy experiments. The observation of a large number of α-amide and amide-amide interresidual nuclear Overhauser effect connectivities as well as the existence of 11 slowly exchanging amide protons indicates that the peptide adopts a well-defined secondary structure most likely constituted of a single long helix. This conclusion is consistent with the CD measurements.  相似文献   

13.
14.
Binding of the recA gene product from Escherichia coli to single-stranded polynucleotides has been investigated using poly(dA) that have been modified by chloroacetaldehyde to yield fluorescent 1,N6-ethenoadenine (epsilon A) bases. A strong enhancement of the fluorescent quantum yield of poly(d epsilon A) is induced upon RecA protein binding. A 4-fold increase is observed in the absence of ATP or ATP gamma S and a 7-fold increase in the presence of either nucleoside triphosphate. RecA protein can bind to poly(d epsilon A) in the absence of both Mg2+ ions and ATP (or ATP gamma S) but Mg2+ ions are required to observe RecA protein binding in the presence of ATP (or ATP gamma S) at pH 7.5. ATP binding to the RecA-poly(d epsilon A) complex induces a dissociation of RecA from the polynucleotide followed by re-binding of [RecA-ATP-Mg2+] ternary complex. Whereas ATP-induced dissociation of RecA-poly(d epsilon A) complexes is a fast process, the subsequent binding reaction of [RecA-ATP-Mg2+] is slow. A model is proposed whereby [RecA-ATP-Mg2+] binding to poly(d epsilon A) involves slow nucleation and elongation processes along the polynucleotide backbone. The nucleation reaction is shown to involve at least a trimer or a tetramer. Polymerization of the [RecA-ATP-Mg2+] ternary complex stops when the polynucleotide is entirely covered with 6 +/- 1 nucleotides per RecA monomer. ATP hydrolysis then induces a release of RecA-ADP complexes from the polynucleotide template.  相似文献   

15.
A single multi-domain viral protein, termed Gag, is sufficient for assembly of retrovirus-like particles in mammalian cells. We have purified the human immunodeficiency virus type 1 (HIV-1) Gag protein (lacking myristate at its N terminus and the p6 domain at its C terminus) from bacteria. This protein is capable of assembly into virus-like particles in a defined in vitro system. We have reported that it is in monomer-dimer equilibrium in solution, and have described a mutant Gag protein that remains monomeric at high concentrations in solution. We report that the mutant protein retains several properties of wild-type Gag. This mutant enabled us to analyze solutions of monomeric protein. Hydrodynamic studies on the mutant protein showed that it is highly asymmetric, with a frictional ratio of 1.66. Small-angle neutron scattering (SANS) experiments confirmed its asymmetry and yielded an R(g) value of 34 A. Atomic-level structures of individual domains within Gag have previously been determined, but these domains are connected in Gag by flexible linkers. We constructed a series of models of the mutant Gag protein based on these domain structures, and tested each model computationally for its agreement with the experimental hydrodynamic and SANS data. The only models consistent with the data were those in which Gag was folded over, with its N-terminal matrix domain near its C-terminal nucleocapsid domain in three-dimensional space. Since Gag is a rod-shaped molecule in the assembled immature virion, these findings imply that Gag undergoes a major conformational change upon virus assembly.  相似文献   

16.
Laser Raman spectra of the trinucleoside diphoshate ApApA and dinucleoside phosphates ApU, UpA, GpC, CpG, and GpU are reported and discussed. Assignments of conformationally sensitive frequencies are-facilitated by comparison with spectra reported here of poly(rA), poly(rC), and poly(rU) in deuterium oxide solutions. The significant spectral differences between ApU and UpA, and between GpC and CpG, reveal that the sequence isomers have nonidentical conformations in aqueous solution. In UpA at low temperature the bases are stacked and the backbone conformation is similar to that found in ordered polynucleotide structures and RNA. In ApU no base stacking can be detected and the backbone conformation differs from that found in UpA, both in the orientation of phosphodiester linkages and in the internal conformation of ribose. At the conditions employed neither ApU nor UpA exhibits base pairing in aqueous solutions. In both GpC and CpG the bases are stacked and the phosphodiester conformations are similar to those encountered for UpA and RNA. However, major differences between spectra of GpC and CpG indicate that the geometries of stacking and ribosyl conformations are different. In GpC the Raman data favor the formation of hydrogen bonded dimers containing GC pairs. Protonation of C in GpC is sufficient to eliminate the ordered conformation detected by Raman spectroscopy. Despite the ordered backbone conformation evident in GpU, this dinucleoside apparently contains neither stacked nor hydrogen bonded bases at the conditions employed here. The Raman data also confirm the stacking interactions in ApApA, poly(rA), and poly(rC) but suggest that the backbone conformation in poly(rC) differs qualitatively from that found in most ordered polynucleotide structures and is thermally more stable. The present results demonstrate the sensitivity of the Raman technique to sequence-related structural differences in oligonucleotides and provide additional spectra–structure correlations for future conformational studies of RNA by laser Raman spectroscopy.  相似文献   

17.
Increasing the pH of a neutral salt solution of sodium hyaluronate to 12.5 produces a rapid drop in viscosity which is reversible upon restoring the pH to neutrality. Light scattering data showing a decrease in radius of gyration with no change in molecular weight and negative results with chondroitin and other acidic glycosaminoglycans suggest that the conformational change is specific for hyaluronate molecules.  相似文献   

18.
Uridylate tracts were released from rat liver mRNA by nuclease digestion and terminally-labeled invitro with 32P using polynucleotide kinase. The pattern of fragments released from A+ and U+ mRNAs were the same as judged by electrophoresis on urea-polyacrylamide gels. The bulk of the fragments were in the size range 20–40 residues, but larger components (up to 70 residues in length) could also be seen. The ability of U+ and A+ mRNAs to direct the synthesis of proteins in a rabbit reticulocyte cell-free system was evaluated. The translation products were resolved by two-dimensional polyacrylamide gel electrophoresis. A comparison of the proteins made by the two classes of RNA showed that the U+ mRNA fraction represents a subset of A+ mRNA species, although the proportions of the protein products were quite different and several proteins were found to be unique to the U+ mRNA class.  相似文献   

19.
The absorption and circular dichroism spectra of the 1:1 copolymer (L -Lys, L -Tyr)n have been investigated in aqueous solutions at pH ranging from 3 to 13. The spectral patterns indicate that the fully charged polympholyte assumes a nonperiodic conformation on the acid and basic sides of the isoelectric point. At pH ranging from 9.2 to 11.6, the polymer is largely ordered and takes mostly an antiparallel β-structure as is shown by the infrared spectra in D2O solutions. Moreover, the rotational strength of the La transition of tyrosyl is independent of the polymer conformation, whereas that of the Lb transition is strongly sensitive to it.  相似文献   

20.
The backbone dynamics of human α-thrombin inhibited at the active site serine were analyzed using R1, R2, and heteronuclear NOE experiments, variable temperature TROSY 2D [1H-15N] correlation spectra, and Rex measurements. The N-terminus of the heavy chain, which is formed upon zymogen activation and inserts into the protein core, is highly ordered, as is much of the double beta-barrel core. Some of the surface loops, by contrast, remain very dynamic with order parameters as low as 0.5 indicating significant motions on the ps-ns timescale. Regions of the protein that were thought to be dynamic in the zymogen and to become rigid upon activation, in particular the γ-loop, the 180s loop, and the Na+ binding site have order parameters below 0.8. Significant Rex was observed in most of the γ-loop, in regions proximal to the light chain, and in the β-sheet core. Accelerated molecular dynamics simulations yielded a molecular ensemble consistent with measured residual dipolar couplings that revealed dynamic motions up to milliseconds. Several regions, including the light chain and two proximal loops, did not appear highly dynamic on the ps-ns timescale, but had significant motions on slower timescales.  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