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 0 毫秒
1.
Conformational and dynamic properties of the anticodon loop of yeast tRNAPhe were investigated by analyzing the time resolved fluorescence of wybutine serving as a local structural probe adjacent to the anticodon GmAA on its 3 side. The influence of Mg2+, important for stabilizing the tertiary structure of tRNA, and of the complementary anticodon s2UUC of E. coli tRNA 2 Glu were investigated.Fluorescence lifetimes and anisotropies were measured with ps time resolution using time correlated single photon counting and a mode locked synchronously pumped and frequency doubled dye laser as excitation source. From the analysis of lifetimes () and rotational relaxation times ( R ) we conclude that wybutine occurs in various structural states: (i) one stacked conformation where the base has no free mobility and the only rotational motion reflects the mobility of the whole tRNA molecule (=6 ns, R =19 ns), (ii) an unstacked conformation where the base can freely rotate (=100 ps, R = 370 ps) and (iii) an intermediary state (=2 ns, R = 1.6 ns).Under biological conditions, i. e. in the presence of Mg2+ and neutral salts, wybutine is found in a stacked and immobile state which is consistent with the crystallographic picture. In the presence of the complementary codon however, as exemplified by the E. coli-tRNA 2 Glu anticodon, our analysis indicates that the codon-anticodon complex exists in an equilibrium of structural states with different rotational mobility of wybutine. The conformation with wybutine freely mobile is the predominant one and suggests that this conformation of the codon-anticodon structure differs from the canonical 3–5 stack.  相似文献   

2.
Myosin rod was prepared from hen myosin by chymotryptic digestion. The indigested myosin was successfully removed by ultracentrifugation following alcohol treatment. No significant difference in UV absorption and CD spectra was observed between pH 7.0 and pH 10.5 for both myosin rod and myosin. When pH was raised to 11.7, the phenolic groups of the tyrosyl residues were ionized, and the helical configuration of the myosin rod and myosin could not withstand the electrostatic repulsion. When pH was further raised to 13.6, “abnormal” tyrosyl residues were ionized, resulting in decreased helix content. However, the myosin rod was stabler and less flexible against pH change than myosin, because of the lower content of tyrosyl residues in myosin rod.  相似文献   

3.
4.
The anticodon sequence is a major recognition element for most aminoacyl-tRNA synthetases. We investigated the in vivo effects of changing the anticodon on the aminoacylation specificity in the example of E. coli tRNAPhe. Constructing different anticodon mutants of E. coli tRNAPhe by site-directed mutagenesis, we isolated 22 anticodon mutant tRNAPhe; the anticodons corresponded to 16 amino acids and an opal stop codon. To examine whether the mutant tRNAs had changed their amino acid acceptor specificity in vivo, we tested the viability of E. coli strains containing these tRNAPhe genes in a medium which permitted tRNA induction. Fourteen mutant tRNA genes did not affect host viability. However, eight mutant tRNA genes were toxic to the host and prevented growth, presumably because the anticodon mutants led to translational errors. Many mutant tRNAs which did not affect host viability were not aminoacylated in vivo. Three mutant tRNAs containing anticodon sequences corresponding to lysine (UUU), methionine (CAU) and threonine (UGU) were charged with the amino acid corresponding to their anticodon, but not with phenylalanine. These three tRNAs and tRNAPhe are located in the same cluster in a sequence similarity dendrogram of total E. coli tRNAs. The results support the idea that such tRNAs arising from in vivo evolution are derived by anticodon change from the same ancestor tRNA.  相似文献   

5.
The mucin obtained from a natto sample was found to be composed of 58 % of γ-polyglutamic acid and 40% of polysaccharide. The ratio of l- and d-glutamic acid was determined to be 58:42 using l-glutamic acid decarboxylase. The weight- and z-average molecular weight were 2.08 × 105 and 2.22 × 105, respectively. The distribution curve of the sedimentation coefficient showed a small heterogeneity. The mucin molecule was considered to be randomly coiled at pH 5.0 ~ 8.8 and to be a rod-like molecule in the lower pH region.  相似文献   

6.
7.
8.
9.
Group II chaperonin captures an unfolded protein while in its open conformation and then mediates the folding of the protein during ATP-driven conformational change cycle. In this study, we performed kinetic analyses of the group II chaperonin from a hyperthermophilic archaeon, Thermococcus sp. KS-1 (TKS1-Cpn), by stopped-flow fluorometry and stopped-flow small-angle X-ray scattering to reveal the reaction cycle. Two TKS1-Cpn variants containing a Trp residue at position 265 or position 56 exhibit nearly the same fluorescence kinetics induced by rapid mixing with ATP. Fluorescence started to increase immediately after the start of mixing and reached a maximum at 1–2 s after mixing. Only in the presence of K+ that a gradual decrease in fluorescence was observed after the initial peak. Similar results were obtained by stopped-flow small-angle X-ray scattering. A rapid fluorescence increase, which reflects nucleotide binding, was observed for the mutant containing a Trp residue near the ATP binding site (K485W), irrespective of the presence or absence of K+. Without K+, a small, rapid fluorescence decrease followed the initial increase, and then a gradual decrease was observed. In contrast, with K+, a large, rapid fluorescence decrease occurred just after the initial increase, and then the fluorescence gradually increased. Finally, we observed ATP binding signal and also subtle conformational change in an ATPase-deficient mutant with K485W mutation. Based on these results, we propose a reaction cycle model for group II chaperonins.  相似文献   

10.
The susceptibility of yeast tRNAPhe and Escherichia coli tRNA2Glu to digestion by nucleases Tl and Sl are examined in a variety of environments, and the results are interpreted in view of the available three-dimensional structural information. Significant differences are found in the digestion pattern of the two tRNAs using the guanosine-specific Tl nuclease. In particular, differences are seen due to varying the type of salts in the environment. However, the Sl nuclease results on the two tRNAs do not differ greatly. E. coli tRNA2Glu is known to exist in two different conformations. Nuclease digestion results are presented revealing differences which make it possible to draw some inferences about the structural differences in these two conformations. In carrying out these analyses, the tRNA molecules are labeled either by putting 32P at the 5'-end of the molecular or by adding 32P-labeled pCp at the 3'-end. It is found that both yeast tRNAPhe and E. coli tRNA2Glu have modified Tl nuclease digestion patterns when pCp is added at the 3'-end of the molecule.  相似文献   

11.
12.
Origin of the nucleoside Y in yeast tRNAPhe   总被引:3,自引:0,他引:3  
  相似文献   

13.
The photochemical reaction dynamics of a BLUF (sensors of blue light using FAD) protein, PixD, from a thermophilic cyanobacterium Thermosynechococcus elongatus BP-1 (TePixD, Tll0078) were studied by pulsed laser-induced transient grating method. After the formation of an intermediate species with a red-shifted absorption spectrum, two new reaction phases reflecting protein conformational changes were discovered; one reaction phase manifested itself as expansion of partial molar volume with a time constant of 40 μs, whereas the other reaction phase represented a change in the diffusion coefficient D [i.e., the diffusion-sensitive conformational change (DSCC)]. D decreased from 4.9 × 10− 11 to 4.4 × 10− 11 m2 s− 1 upon the formation of the first intermediate, and subsequently showed a more pronounced decrease to 3.2 × 10− 11 m2 s− 1 upon formation of the second intermediate. From a global analysis of signals at various grating wavenumbers, the time constant of D-change was determined to be 4 ms. Although the magnitude and rate constant of the faster volume change were independent of protein concentration, the amplitude of the signal that reflects the later DSCC significantly decreased as the protein concentration decreased. This concentration dependence suggests that two species exist in solution: a reactive species exhibiting the DSCC, and a second species that is nonreactive. The fraction of these species was found to be dependent on the concentration. The difference in reactivity was attributed to the different oligomeric states of TePixD (i.e., pentamer and decamer). The equilibrium of these states in the dark was confirmed by size-exclusion chromatography at various concentrations. These results demonstrated that only the decamer state is responsible for the conformational change. The results may suggest that the oligomeric state is functionally important in the signal transduction of this photosensory protein.  相似文献   

14.
Properties of tRNAPhe from Drosophila   总被引:3,自引:0,他引:3  
  相似文献   

15.
Current homology-modelling methods do not consider small molecules in their automated processes. Therefore, the development of a reliable tool for protein-ligand homology modelling is an important next step in generating plausible models for molecular interactions. Two automated protein-ligand homology-modelling strategies, requiring no expert knowledge from the user, are investigated here. Both employ the “induced fit” concept with flexibility in side chains and ligand. The most successful strategy superimposes the new ligand over the original ligand before homology modelling, allowing the new ligand to be taken into consideration during protein modelling (rather than after), facilitating conformational change in the local backbone if necessary. We show that this approach results in successful modelling of the ligand and key binding-site residues of angiotensin-converting enzyme 2 (ACE2) from its homologue ACE, which is not possible via conventional homology modelling or by homology modelling followed by docking. Several other difficult target complexes are also successfully modelled, reproducing native protein-ligand contacts with significantly different biological substrates and different binding-site conformations. These include the modelling of Cdk5 (cyclin-dependent kinase 5) from Cdk2, thymidine phosphorylase from a bacterial homologue, and dihydrofolate reductase from a recombinant variant with a markedly different inhibitor. In terms of average modelling quality across 82 targets, the ligand RMSD with respect to the experimental structure is 1.4 Å (and 2.0 Å for the protein binding site) for “easy” cases and 2.9 Å for the ligand (and 2.7 Å for the protein binding site) in “hard” cases. This demonstrates the importance of selecting an optimal template. Ligand-modelling accuracy is strongly dependent on target-template ligand structural similarity, rather than target-template sequence identity. However, protein-modelling accuracy is dependent on both. Our automated protein-ligand homology-modelling strategy generates a higher degree of accuracy than homology modelling followed by docking, generating an average ligand RMSD that is 1-2 Å better than docking with homology models.  相似文献   

16.
Dependence of B-A Conformational Change in DNA on Base Composition   总被引:10,自引:0,他引:10  
IT is generally accepted that all native DNAs, irrespective of their origin, give essentially identical X-ray diffraction patterns characteristic of the A form at high humidities and of the A form at lower humidities1, 2. The purpose of the work described here was an accurate characterization of B-A conformational changes in a range of DNA by infrared dichroism, which seems to be particularly appropriate for the structural characterization of nucleic acids3–6 as a complement of X-ray diffraction. In previous infrared studies the A form and B-A conformational changes were not observed7.  相似文献   

17.
The binding of the codon UUC to the isolated anticodon loop of tRNAPhe (yeast) has been studied as a model of codon recognition by a simple adaptor. Fluorescence titrations demonstrate that UUC binds to the isolated anticodon loop with an equilibrium constant of 1.4 X 10(3) M-1 (at 7.2 degrees C). Equilibrium sedimentation curves reveal that UUC binding induces association of anticodon loops beyond the dimer stage. A set of complete sedimentation curves obtained for various reactant concentrations was analyzed according to a model with an infinite number of subsequent association steps for UUC-anticodon loop complexes and with equal affinity for each step. The coupling of association and sedimentation was considered quantitatively, and the information resulting from conservation of mass was used by integration. According to this procedure, the experimental data can be described by an isodesmic association constant of 8 X 10(3) M-1 with satisfactory accuracy. Temperature-jump relaxation detected by fluorescence measurements provides independent evidence for codon-induced association of the anticodon loop. The data are consistent with the following mechanism: UUC preferentially binds to one of two loop conformations with a rate constant of 4.5 X 10(6) M-1 s-1; the UUC-anticodon loop complex undergoes association with a rate constant of 6.5 X 10(6) M-1 s-1. The reactions observed for the isolated anticodon loop are surprisingly similar to those observed previously for the complete tRNA, suggesting that simple hairpin loops are appropriate adaptors for a translation process at an early stage of evolution; the codon-induced association of the hairpin loop should be very useful to facilitate the transfer of cognate amino acids during translation.  相似文献   

18.
Change point detectors (CPDs) are used to segment recordings of single molecules for the purpose of kinetic analysis. The assessment of the accuracy of CPD algorithms has usually been based on testing them with simulated data. However, there have not been methods to assess the output of CPDs from real data independent of simulation. Here we present one method to do this based on the assumption that the elementary kinetic unit is a stationary period (SP) with a normal distribution of samples, separated from other SPs by change points (CPs). Statistical metrics of normality can then be used to assess SPs detected by a CPD algorithm (detected SPs, DSPs). Two statistics in particular were found to be useful, the z-transformed skew (S Z) and z-transformed kurtosis (K Z). K Z(S Z) plots of DSP from noise, simulated data and single ion channel recordings showed that DSPs with false negative CP could be distinguished. Also they showed that filtering had a significant effect on the normality of data and so filtering should be taken into account when calculating statistics. This method should be useful for analyzing single molecule recordings where there is no simple model for the data.  相似文献   

19.
The nucleotide sequence of tRNAPhe of yellow lupin seeds (Lupinus luteus) is deduced from the composition of pancreatic and T1 ribonuclease digestion products and compared with tRNAPhe of wheat germ. Major lupin tRNAPhe, unlike pea tRNAPhe, differs from wheat germ tRNAPhe in the first base pair of stem TpsiC ("e").  相似文献   

20.
Lead-catalyzed cleavage of yeast tRNAPhe mutants   总被引:23,自引:0,他引:23  
Yeast tRNA(Phe) lacking modified nucleotides undergoes lead-catalyzed cleavage between nucleotides U17 and G18 at a rate very similar to that of its fully modified counterpart. The rates of cleavage for 28 tRNA(Phe) mutants were determined to define the structural requirements of this reaction. The cleavage rate was found to be very dependent on the identity and correct positioning of the two lead-coordinating pyrimidines defined by X-ray crystallography. Nucleotide changes that disrupted the tertiary interactions of tRNAPhe reduced the rate of cleavage even when they were distant from the lead binding pocket. However, nucleotide changes designed to maintain tertiary interactions showed normal rates of cleavage, thereby making the reaction of a useful probe for tRNA(Phe) structure. Certain mutants resulted in the enhancement of cleavage at a "cryptic" site at C48. The sequences of Escherichia coli tRNA(Phe) and yeast tRNA(Arg) were altered such that they acquired the ability to cleave at U17, confirming our understanding of the structural requirements for cleavage. This mutagenic analysis of the lead cleavage domain provides a useful guide for similar analysis of autocatalytic self-cleavage reactions.  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