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 31 毫秒
1.
The existence of specific sites in tRNA for the binding of divalent cations has been seriously questioned by electrostatic considerations [Leroy & Guéron (1979) Biopolymers, 16, 2429-2446]. However, our earlier studies of the binding of Mg2+ and Mn2+ to yeast tRNA(Tyr) have indicated that spermine creates new binding sites for divalent cations [Weygand-Durasevi? et al. (1977) Biochim. Biophys, Acta, 479, 332-344; N?thig-Laslo et al. (1981) Eur. J. Biochem. 117, 263-267]. We have now used yeast tRNA(Tyr), spin labeled at the hypermodified purine (i6A-37) in the anticodon loop, to study the effect of spermine on the binding of manganese ions. The presence of eight spermine molecules per tRNA(Tyr) at high ionic strength (0.2 M NaCl, 0.05 M triethanolamine.HCl) and at low temperature (7 degrees C) enhances the binding of manganese to tRNA(Tyr). This effect could not be explained by electrostatic binding. The initial binding of manganese to tRNA(Tyr) affects the motional properties of the spin label indicating a change of the conformation of the anticodon loop. From the absence of the paramagnetic effect of manganese on the ESR spectra of the spin label one can conclude that the first binding site for manganese is at a distance from i6A-37, influencing the spin label motion through a long-range effect. The enhancement of the binding of manganese to tRNA(Tyr) by spermine is lost upon destruction of its specific macromolecular structure and it does not occur in single stranded or in double-stranded polynucleotides. The observed effect can be explained by the binding of Mn2+ to new sites, created by the binding of spermine, which are specific for the macromolecular structure of tRNA.  相似文献   

2.
Binding of the polyamines spermidine (∼-+3) and spermine (∼-+4) to yeast tRNAphe has been investigated by equilibrium dialysis under the same conditions used to study Mn2+-tRNAphe interactions (Schreier & Schimmel, 1974). The polyamines bind to tRNAphe in a co-operative and a non-co-operative phase, which is analogous to the behavior found with Mn2+. In the co-operative phase, the empirical index of co-operativity is somewhat greater for the polyamines, however. Binding constants for both the co-operative and non-co-operative phases are similar for Mn2+ and spermidine, and are strongest for spermine. Estimates of the total number of ligand binding sites indicate that these numbers are inversely proportional to the charge on the ligand for all three ligands. The interaction of polyamines with four large fragments of tRNAphe shows no evidence for co-operativity. These results, together with recent kinetic studies, collectively suggest that polyamine binding to the co-operative sites is associated with tertiary structure formation and that polyamine and divalent metal ion interactions with tRNA occur by phenomenologically similar mechanisms, in spite of their structural diversity.  相似文献   

3.
Transfer RNATyr (anticodon GA) was isolated from Drosophila melanogaster by means of Sepharose 4B, RPC-5, and polyacrylamide gel electrophoresis. The tRNA was iodinated in vitro with Na125I and hybridized in situ to salivary gland chromosomes from Drosophila. The genes of tRNATyr were localized in eight regions of the genome by autoradiography. Restriction enzyme analysis of genomic DNA indicated that the haploid Drosophila genome codes for about 23 tRNATyr genes. The regions 22F and 85A each contain four to five tRNATyr genes, whereas the regions 28C, 41AB, 42A, 42E, and 56D each contain two to three tRNATyr genes.  相似文献   

4.
Primary structure of Bacillus subtilis tRNAsTyr   总被引:4,自引:0,他引:4  
tRNAITyr and tRNAIITyr have been purified from B.subtilis and their nucleotide sequence determined. tRNAITyr differs from tRNAIITyr only by the extent of modification of the adenosine in 3′ position adjacent to the anticodon, i6A and ms2i6A respectively.  相似文献   

5.
tRNA structure and binding sites for cations   总被引:4,自引:0,他引:4  
A Danchin 《Biopolymers》1972,11(7):1317-1333
Equilibrium dialysis and electronic and nuclear resonance spectroscopy show that tRNA cooperatively binds divalent metal ions at very low concentrations (free metal concentration 3 × 10 ?6 M). The first two methods show that different purified tRNAs have a very similar behavior, including initiator tRNAFmet. tRNAs with an extra arm in the clover-leaf model, however, appear to have a slightly different behavior. The binding can be described in terms of two classes of sites. The cooperative association of divalent ions binding first does not parallel a cooperative change in the hyperchromism of the tRNA, while the non-cooperative association of the second class of divalent ions corresponds to the concentrations needed to obtain a cooperative melting of the tRNA. The temperature dependence of the number of binding sites and of their binding constants is also presented. The nature of the divalent ion gives the following efficiency: for the cooperativity Co++>Mg++>Mn++ for the weak binding sites Mn++>Co++>Mg++  相似文献   

6.
Electron paramagnetic resonance spectroscopy has been used to study the coupling of Mn2+ ions with the tRNAVal1 modified with a spin label at four pseudouridylic residues and with the valyl-tRNAVal1 modified with a spin label at the α-amine group of the valyl residue. A sharp increase of spin-label mobility has been found in these samples, due to the conformational transition induced by the first and second Mn2+ ions. Analysis of dipole–dipole couplings of spin labels with the coordinated ions revealed a definite order in the occupation of ion coordination sites in the tRNA. For some valyl-tRNAVal1 molecules, the second Mn2+ ions were shown to coordinate on the α-amine group of the valyl residue at a distance of 15–25 Å from a spin label. As a result of the conformational transition, a coordination site appeared in the tRNA at one of the pseudouridylic residues, its distance from the spin label being less than 10 Å. It has been suggested that the conformational transition induced by ions excluded some bases from the system of hydrogen bonds at the level of the tRNA tertiary structure. As a result, these bases acquired sufficient sterical freedom to participate in the Mn2+ ion coordination.  相似文献   

7.
The anticodon-anticodon complex   总被引:6,自引:0,他引:6  
Gel electrophoresis has been used to measure the binding between two tRNAs with complementary anticodons, tRNAVal (Escherichia coli) (anticodon X,A,C) and tRNATyr (E. coli) (anticodon Q,U,A). The association constant K at 0 °C was found to be 4 × 105 m?1 which is about three orders of magnitude greater than the association constant for tRNATyr (E. coli) binding its trinucleotide codon UAC. The temperature dependence of K suggests that this results from the rigidity of the anticodon loop. tRNATyr (E. coli) binds an order of magnitude more weakly to tRNAVal (yeast) than to tRNAVal (E. coli), presumably because it contains the wobble base pair A · I. The relationship between the anticodon-anticodon complex and codon recognition is discussed.  相似文献   

8.
Further investigations into the properties of the mercury derivative formed by the reaction of 4-thiouridine-containing tRNAs and pentafluorophenylmercury chloride have been carried out. tRNAfMet (which contains only one 4-thiouridine residue) has been isolated by a one-step column Chromatographic procedure from unfractionated Escherichia coli tRNA and has been shown to react with the mercury compound to give a derivative which has similar properties to those previously reported for the corresponding mercury derivative of tRNATyr which contains two adjacent 4-thiouridine residues. The mercury derivative of tRNATyr appears to be a competitive inhibitor of tRNATyr in the aminoacylation reaction (tRNATyrKm = 0.42 μM, mercury derivative of tRNATyrKi = 0.11 μM). The mercury derivative of Tyr-tRNATyr can be made, but only by the reaction of the mercury compound with the aminoacylated tRNA.  相似文献   

9.
Intron-containing tRNA genes are exceptional within nuclear plant genomes. It appears that merely two tRNA gene families coding for tRNATyr G A and elongator tRNAMet CmAU contain intervening sequences. We have previously investigated the features required by wheat germ splicing endonuclease for efficient and accurate intron excision from Arabidopsis pre-tRNATyr. Here we have studied the expression of an Arabidopsis elongator tRNAMet gene in two plant extracts of different origin. This gene was first transcribed either in HeLa or in tobacco cell nuclear extract and splicing of intron-containing tRNAMet precursors was then examined in wheat germ S23 extract and in the tobacco system. The results show that conversion of pre-tRNAMet to mature tRNA proceeds very efficiently in both plant extracts. In order to elucidate the potential role of specific nucleotides at the 3 and 5 splice sites and of a structured intron for pre-tRNAMet splicing in either extract, we have performed a systematic survey by mutational analyses. The results show that cytidine residues at intron-exon boundaries impair pre-tRNAMet splicing and that a highly structured intron is indispensable for pre-tRNAMet splicing. tRNA precursors with an extended anticodon stem of three to four base pairs are readily accepted as substrates by wheat and tobacco splicing endonuclease, whereas pre-tRNA molecules that can form an extended anticodon stem of only two putative base pairs are not spliced at all. An amber suppressor, generated from the intron-containing elongator tRNAMet gene, is efficiently processed and spliced in both plant extracts.  相似文献   

10.
Dietary cadmium causes the queuine-containing, Q(+), isoacceptors to increase relative to the guanine-containing, Q(?), ones of tRNATyr, tRNAHis and tRNAAsp of Drosophila melanogaster. Of the other divalent cations examined, Sr2+, Ni2+, Cu2+, Zn2+ and Hg2+, only Hg2+ failed to cause an increase in Q(+)tRNATyr. For these results, all pre-adult stages of the organism were spent on media containing the divalent ions. Adult flies that had developed on a normal diet also responsed to divalent ions; Hg2+ as well as Cd2+, Sr2+ and Zn2+ caused an increase in Q(+)tRNATyr in 4 days. Using adult flies, the rate of the response was measured; when placed on a Cd2+-containing diet, they formed significantly more Q(+)tRNATyr within 24 h as compared to adults on a normal diet. Whether the queuine is derived from the diet or from de novo synthesis is yet to be determined. Since the metal ions represent a range of values in the ‘hard-soft’ classification, different sites of reaction are expected, yet for Drosophila a common result is an alteration in the ratio of Q(+) and Q(?) isoacceptors of these tRNAs. The transition to Q(+)tRNA may be an early indication of the metabolic imbalances resulting from the presence of the divalent cation.  相似文献   

11.
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   

12.
tRNA molecules contain 93 chemically unique nucleotide base modifications that expand the chemical and biophysical diversity of RNA and contribute to the overall fitness of the cell. Nucleotide modifications of tRNA confer fidelity and efficiency to translation and are important in tRNA-dependent RNA-mediated regulatory processes. The three-dimensional structure of the anticodon is crucial to tRNA-mRNA specificity, and the diverse modifications of nucleotide bases in the anticodon region modulate this specificity. We have determined the solution structures and thermodynamic properties of Bacillus subtilis tRNATyr anticodon arms containing the natural base modifications N6-dimethylallyl adenine (i6A37) and pseudouridine (ψ39). UV melting and differential scanning calorimetry indicate that the modifications stabilize the stem and may enhance base stacking in the loop. The i6A37 modification disrupts the hydrogen bond network of the unmodified anticodon loop including a C32-A38+ base pair and an A37-U33 base-base interaction. Although the i6A37 modification increases the dynamic nature of the loop nucleotides, metal ion coordination reestablishes conformational homogeneity. Interestingly, the i6A37 modification and Mg2+ are sufficient to promote the U-turn fold of the anticodon loop of Escherichia coli tRNAPhe, but these elements do not result in this signature feature of the anticodon loop in tRNATyr.  相似文献   

13.
14.
We have noticed that during a long storage and handling, the plant methionine initiator tRNA is spontaneously hydrolyzed within the anticodon loop at the C34-A35 phosphodiester bond. A literature search indicated that there is also the case for human initiator tRNAMet but not for yeast tRNAMet i or E. coli tRNAMet f. All these tRNAs have an identical nucleotide sequence of the anticodon stems and loops with only one difference at position 33 within the loop. It means that cytosine 33 (C33) makes the anticodon loop of plant and human tRNAMet i susceptible to the specific cleavage reaction. Using crystallographic data of tRNAMet f of E. coli with U33, we modeled the anticodon loop of this tRNA with C33. We found that C33 within the anticodon loop creates a pocket that can accomodate a hydrogen bonded water molecule that acts as a general base and catalyzes a hydrolysis of C-A bond. We conclude that a single nucleotide change in the primary structure of tRNAMet i made changes in hydration pattern and readjustment in hydrogen bonding which lead to a cleavage of the phosphodiester bond.  相似文献   

15.
Binding of Mn2+ to the whole molecule, fragments and complementary fragment recombinations of yeast tRNAPhe, and to synthetic polynucleotides was studied by equilibrium dialysis. The comparison of the binding patterns of the fragments, fragment recombinations and synthetic polynucleotides with that of intact tRNAPhe permits reasonable conclusions concerning the nature and location of the various classes of sites on tRNAPhe. Binding of Mn2+ to intact tRNAPhe consists of a co-operative and a non-co-operative phase. There are about 17 “strong” sites and several “weak” ones. Five of the 17 strong sites are associated with the co-operative phase. This phase is completely lacking in the binding of Mn2+ to tRNAPhe fragments (5′-12, 3′-12, 5′-35, 3′-25), poly-(A):poly(U) and poly(I):poly(C) helices, and single stranded poly(A) and poly(U). This argues that the co-operative sites arise from the tRNA tertiary structure. This conclusion is further strengthened by the observation that cooperativity is present in a tRNAPhe molecule which has been split in the anticodon loop, but it is absent in one which has been split in the extra loop. It is in the vicinity of the latter loop, but not the former, that tertiary interactions are seen in the crystal structure. The remaining 12 strong sites are “independent” and appear to be associated with cloverleaf helical sections.  相似文献   

16.
Lambda bacteriophage containing yeast tyrosine transfer RNA genes were prepared by molecular recombination. These phage were identified by hybridization of 125I-labeled yeast tRNATyr to plaques from lambda-yeast recombinant phage pools. The cloned yeast EcoRI fragments that hybridize to 125I-labeled tRNATyr were compared in size with the fragments in total yeast DNA that hybridize to the same probe. These comparisons indicate that seven of the eight different tRNATyr genes have been isolated. Unambiguous evidence that these seven fragments contain tRNATyr coding regions was obtained by showing that they hybridize to aminoacylated [3H]Tyr-tRNATyr. Only one of the fragments hybridizes to 32P-labeled total yeast tRNA in the presence of competing unlabeled tRNATyr; the tRNATyr genes, therefore, are not predominantly organized into heteroclusters of tRNA genes.  相似文献   

17.
Import of nucleus-encoded tRNAs into the mitochondria of the kinetoplastid protozoon Leishmania involves recognition of specific import signals by the membrane-bound import machinery. Multiple signals on different tRNA domains may be present, and further, importable RNAs interact positively (Type I) or negatively (Type II) with one another at the inner membrane in vitro. By co-transfection assays, it is shown here that tRNATyr (Type I) transiently stimulates the rate of entry of tRNAIle (Type II) into Leishmania mitochondria in transfected cells, and conversely, is inhibited by tRNAIle. Truncation and mutagenesis experiments led to the co-localization of the effector and import activities of tRNATyr to the D domain, and those of tRNAIle to the variable region–T domain (V-T region), indicating that both activities originate from a single RNA–receptor interaction. A third tRNA, human tRNALys, is imported into Leishmania mitochondria in vitro as well as in vivo. This tRNA has Type I and Type II motifs in the D domain and the V-T region, respectively, and shows both Type I and Type II effector activities. Such dual-type tRNAs may interact simultaneously with the Type I and Type II binding sites of the inner membrane import machinery.  相似文献   

18.
We have constructed a model of the complex between tyrosyl-tRNA synthetase (TyrRS) from Bacillus stearothermophilus and tRNATyr by successive cycles of predictions, mutagenesis of TyrRS and molecular modeling. We confront this model with data obtained independently, compare it to the crystal structures of other complexes and review recent data on the discrimination between tRNAs by TyrRS. Comparison of the crystal structures of TyrRs and GlnRS, both of which are class I synthetases, and comparison of the identity elements of tRNATyr and tRNAGln indicate that the two synthetases bind their cognate tRNAs differently. The mutagenesis data on tRNATyr confirm the model of the TyrRS:tRNATyr complex on the following points. TyrRS approaches tRNATyr on the side of the variable loop. The bases of the first three pairs of the acceptor stem are not recognized. The presence of the NH2 group in position C6 and the absence of a bulky group in position C2 are important for the recognition of the discriminator base A73 by TyrRS, which is fully realized only in the transition state for the acyl transfer. The anticodon is the major identity element of tRNATyr. We have set up an in vivo approach to study the effects of synthetase mutations on the discrimination between tRNAs. Using this approach, we have shown that residue Glul52 of TyrRS acts as a purely negative discriminant towards non-cognate tRNAs, by electrostatic and steric repulsions. The overproductions of the wild type TyrRSs from E coli and B stearothermophilus are toxic to E coli, due to the mischarging or the non-productive binding of tRNAs. The construction of a family of hybrids between the TyrRSs from E coli and B stearothermophilus has shown that their sequences and structures have remained locally compatible through evolution, for holding and function, in particular for the specific recognition and charging of tRNATyr.  相似文献   

19.
Yu H  Li Q 《PloS one》2011,6(1):e16147

Background

Animal mitochondrial genomes typically encode one tRNA for each synonymous codon family, so that each tRNA anticodon essentially has to wobble to recognize two or four synonymous codons. Several factors have been hypothesized to determine the nucleotide at the wobble site of a tRNA anticodon in mitochondrial genomes, such as the codon-anticodon adaptation hypothesis, the wobble versatility hypothesis, the translation initiation and elongation conflict hypothesis, and the wobble cost hypothesis.

Principal Findings

In this study, we analyzed codon usage and tRNA anticodon wobble sites of 29 marine bivalve mitochondrial genomes to evaluate features of the wobble nucleotides in tRNA anticodons. The strand-specific mutation bias favors G and T on the H strand in all the 29 marine bivalve mitochondrial genomes. A bias favoring G and T is also visible in the third codon positions of protein-coding genes and the wobble sites of anticodons, rejecting that codon usage bias drives the wobble sites of tRNA anticodons or tRNA anticodon bias drives the evolution of codon usage. Almost all codon families (98.9%) from marine bivalve mitogenomes support the wobble versatility hypothesis. There are a few interesting exceptions involving tRNATrp with an anticodon CCA fixed in Pectinoida species, tRNASer with a GCU anticodon fixed in Mytiloida mitogenomes, and the uniform anticodon CAU of tRNAMet translating the AUR codon family.

Conclusions/Significance

These results demonstrate that most of the nucleotides at the wobble sites of tRNA anticodons in marine bivalve mitogenomes are determined by wobble versatility. Other factors such as the translation initiation and elongation conflict, and the cost of wobble translation may contribute to the determination of the wobble nucleotide in tRNA anticodons. The finding presented here provides valuable insights into the previous hypotheses of the wobble nucleotide in tRNA anticodons by adding some new evidence.  相似文献   

20.
J L Leroy  M Guéron 《Biopolymers》1977,16(11):2429-2446
The binding of Mn2+ to yeast tRNAPhe at 25°C is measured by esr, and found to depend strongly on the concentration of monovalent cations, showing the importance of electrostatic effects. In low sodium (<15mM/l.), the affinity is high and the Scatchard plots are distinctly curved. In high sodium (>50mM/l.), the affinity and the curvature are reduced. In a limited range of sodium concentrations (15–30mM/l.), the folding of tRNA which is induced by the divalent ions results in cooperative binding, leading to upwards convexity of the Scatchard plot. An electrostatic model is developed, based on a single type of binding site which we take to be the phosphates, with a binding constant for Mn2+ in the range of that found for ApA, 10 l./M. We compute the change in the binding constant due to the electrostatic potential of the distant charges (other phosphates and counterions), using a single set of parameters for all sodium concentrations. The model predicts that the plots in low sodium are curved, and a good fit to the experimental results is obtained: it is therefore not legitimate or necessary to interpret these results in terms of two types of binding sites. In high salt, the model gives plots that are only slightly curved, corresponding to weaker electrostatic effects. This shows that a search for sites with a special binding mode should be done in high salt. The computed plots are in good agreement with the data, except for slight differences concerning the first bound ions, which give a possible indication in favor of special binding. Given the observation of one special site for Mg2+ at 4°C in high sodium [Stein, A. & Crothers, D. M. (1976) Biochemistry 15 , 157–160] in E. coli tRNAfMet, we have measured the binding of Mn2+ at lower temperature. At 12°C, in both yeast tRNAPhe and E. coli tRNAfMet, the plots clearly indicate special binding. A site found in high sodium is on a very different footing from the four to six so-called strong sites unduly derived from low-salt binding plots.  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