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 protein environment of mRNA 3′ of the A-site codon (the decoding site) in the human 80S ribosome was studied using a set of oligoribonucleotide derivatives bearing a UUU triplet at the 5′-end and a perfluoroarylazide group at one of the nucleotide residues 3′ of this triplet. Analogues of mRNA were phased into the ribosome using binding at the tRNAPhe P-site, which recognizes the UUU codon. Mild UV irradiation of ribosome complexes with tRNAPhe and mRNA analogues resulted in the predominant crosslinking of the analogues with the 40S subunit components, mainly with proteins and, to a lesser extent, with rRNA. Among the 40S subunit ribosomal proteins, the S3 protein was the main target for modification in all cases. In addition, minor crosslinking with the S2 protein was observed. The crosslinking with the S3 and S2 proteins occurred both in ternary complexes and in the absence of tRNA. Within ternary complexes, crosslinking with S15 protein was also found, its efficiency considerably falling when the modified nucleotide was moved from positions +5 to +12 relative to the first codon nucleotide in the P-site. In some cases, crosslinking with the S30 protein was observed; it was most efficient for the derivative containing a photoreactive group at the +7 adenosine residue. The results indicate that the S3 protein in the human ribosome plays a key role in the formation of the mRNA binding site 3′ of the codon in the decoding site.  相似文献   

2.
There is a special emphasis today on integrating traditional healing within health services. However, most areas in which there is a system of traditional healing have undergone colonization and a number of pressures suppressing tradition for hundreds of years. The question arises as to how one can understand today’s tradition in light of earlier traditions. This article is based on material collected in Sámi areas of Finnmark and Nord-Troms Norway; it compares local healing traditions with what is known of earlier shamanic traditions in the area. The study is based on 27 interviews among healers and their patients. The findings suggest that although local healing traditions among the Sámi in northern Norway have undergone major transformations during the last several hundred years, they may be considered an extension of a long-standing tradition with deep roots in the region. Of special interest are also the new forms tradition may take in today’s changing global society.  相似文献   

3.
The 2201-bp spacer between the chloroplast ribosomal 16S and 23S genes ofSpinacia oleracea was sequenced. It contains the genes of the tRNAIle (GAU) and tRNAAla (UGC) which are both interrupted by introns of respectively 728 and 816 bp. These introns belong to the class II according to the classfication of Michel and Dujon [17]. Comparison of the rDNA spacer sequence of maize, tobacco and spinach indicates that no conserved polypeptide is encoded within the introns of the two tRNA genes and that the two main insertions/deletions between the three plants are located within two loops of the class II introns secondary structure, which is therefore conserved. Based on the sequence complementarity observed between the upstream and downstream parts, of the 16S and 23S rRNA genes, RNase III-like secondary structures involved in the processing of the rRNA precursor are proposed.  相似文献   

4.
5.
The arrangement of the template sequence 3′ of the A-site codon on the 80S ribosome was studied using mRNA analogs containing Phe codon UUU at the 5′ end and a photoreactive perfluoroarylazido group linked to C5 of U or N7 of G. The analogs were positioned on the ribosome with the use of tRNAPhe, which directed the UUU codon to the P site, bringing a modified nucleotide to position +9 or +12 relative to the first nucleotide of the P-site codon. Upon mild UV irradiation of ribosome complexes, the analogs of both types crosslinked to the 18S rRNA and proteins of the 40S subunit. Comparisons were made with the crosslinking patterns of complexes in which an mRNA analog contained a modified nucleotide in position +7 (the crosslinking to 18S rRNA in such complexes has been studied previously). The efficiency of crosslinking to ribosomal components depended on the nature of the modified nucleotide of an mRNA analog and its position on the ribosome. The extent of crosslinking to the 18S rRNA drastically decreased as the modified nucleotide was transferred from position +7 to position +12. The 18S rRNA nucleotides involved in crosslinking were identified. A modified nucleotide in position +9 crosslinked to the invariant dinucleotide A1824/A1825 and variable A1823 in the 3′ minidomain of the 18S rRNA and to S15. The same ribosomal components have earlier been shown to crosslink to modified nucleotides in positions +4 to +7. In addition, all mRNA analogs crosslinked to invariant C1698 in the 3′ minidomain and to conserved region 605–620, which closes helix 18 in the 5′ domain.  相似文献   

6.
Summary The nucleotide sequence of a spacer region between rice 17S and 25S rRNA genes (rDNAs) has been determined. The coding regions for the mature 17S, 5.8S and 25S rRNAs were identified by sequencing terminal regions of these rRNAs. The first internal transcribed spacer (ITS1), between 17S and 5.8S rDNAs, is 194–195 bp long. The second internal transcribed spacer (ITS2), between 5.8S and 25S rDNAs, is 233 bp long. Both spacers are very rich in G+C, 72.7% for ITS1 and 77.3% for ITS2. The 5.8S rDNA is 163–164 bp long and similar in primary and secondary structures to other eukaryotic 5.8S rDNAs. The 5.8S rDNA is capable of interacting with the 5′ terminal region of 25S rDNA.  相似文献   

7.
16S–23S rRNA internally transcribed spacer (ITS) sequences from 53 Frankia strains were sequenced and sized from polymerase chain reaction amplification products and compiled with 14 selected 16S–23S ITS sequences from public database. Frankia genomes included two to three ITS copies lacking length polymorphism except for nine strains. No tRNA gene was encountered in this region. Frankia strains exhibited various lengths (369 to 452 nt) and a wide range of sequence similarity (35–100%) in the ITS region. The average pairwise distance varied from 0.368 (clusters 1 and 2) to 0.964 (clusters 3 and 4) and was 0.397, 0.138, 0.129, and 0.016, respectively, for cluster 4 (saprophytic non-infective/non-effective), clusters 1 and 3 (facultative symbiotic), and cluster 2 (obligate symbiotic). This suggests a gradual erosion of Frankia diversity concomitantly with a shift from saprophytic non-infective/non-effective to facultative and symbiotic lifestyle. Comparative sequence analyses of the 16S–23S rRNA intergenic spacer region of Frankia strains are not useful to assign them to their respective cluster or host infection group. Accurate assignment required the inclusion of the adjacent 16S and 23S rRNA gene fragments.  相似文献   

8.
The sequence of the ribosomal spacer region of soybean chloroplast DNA including the 3 end of the 16S rRNA gene, the tRNAAla and tRNAIle genes (but not their introns), the three intergenic regions and the 5 end of the 23S rRNA gene, has been determined. This sequence has been compared to corresponding regions of other angiosperm chloroplast DNAs. Secondary structure models are proposed for the entirety of the intergenic regions a, b and c and for the flanking rRNA regions. A model for a common secondary structure of the ribosomal spacer intergenic regions from chloroplasts of higher plants is proposed, which is supported by comparative evidence.  相似文献   

9.
10.
The aim of this study was to investigate the feasibility of using short tandem repeats (STRs) to diagnose Edwards’ syndrome (ES). Quantitative fluorescence polymerase chain reaction (QF-PCR) was performed to amplify STR loci on chromosome 18, specifically D18S53, D18S59, and D18S488. The amplified products were subjected to a fluorescence signal analysis and their application to ES diagnosis was examined. Among the 807 cases that showed normal results in the karyotype analysis, 793 showed one or two fluorescence bands with a fluorescence intensity ratio of 1:1, and 14 cases showed 3 bands, which were false-positive results. ES was diagnosed in 9 samples. The sensitivities of D18S53, D18S59, and D18S488 for the diagnosis of ES were 77.78, 44.44, and 55.56 % and the specificities were 96.16, 96.03, and 96.28 %, respectively. The combined sensitivity of the three loci for diagnosing DS was 100 % (9/9), with a specificity of 98.27 % (793/807). QF-PCR amplification of STR loci had high sensitivity, strong specificity, and was simple and rapid. Thus, it might have wide clinical applications, and could be an ideal tool for large-scale genetic and prenatal diagnosis of ES.  相似文献   

11.
12.
Jason S. King 《Autophagy》2012,8(7):1159-1162
Autophagy is conserved throughout the eukaryotes and for many years, work in Saccharomyces cerevisiae has been at the forefront of autophagy research. However as our knowledge of the autophagic machinery has increased, differences between S. cerevisiae and mammalian cells have become apparent. Recent work in other organisms, such as the amoeba Dictyostelium discoideum, indicate an autophagic pathway much more similar to mammalian cells than S. cerevisiae, despite its earlier evolutionary divergence. S. cerevisiae therefore appear to have significantly specialized, and the autophagic pathway in mammals is much more ancient than previously appreciated, which has implications for how we interpret data from organisms throughout the eukaryotic tree.  相似文献   

13.
The ribosomal RNA multigene family in Escherichia coli comprises seven rrn operons of similar, but not identical, sequence. Four operons (rrnC, B, G, and E) contain genes in the 16S–23S intergenic spacer region (ISR) for tRNAGlu-2 and three (rrnA, D, and H) contain genes for tRNAIle-1 and tRNAAla-1B. To increase our understanding of their molecular evolution, we have determined the ISR sequence of the seven operons in a set of 12 strains from the ECOR collection. Each operon was specifically amplified using polymerase chain reaction primers designed from genes or open reading frames located upstream of the 16S rRNA genes in E. coli K12. With a single exception (ECOR 40), ISRs containing one or two tRNA genes were found at the same respective loci as those of strain K12. Intercistronic heterogeneity already found in K12 was representative of most variation among the strains studied and the location of polymorphic sites was the same. Dispersed nucleotide substitutions were very few but 21 variable sites were found grouped in a stem-loop, although the secondary structure was conserved. Some regions were found in which a stretch of nucleotides was substituted in block by one alternative, apparently unrelated, sequence (as illustrated by the known putative insertion of rsl in K12). Except for substitutions of different sizes and insertions/deletions found in the ISR, the pattern of nucleotide variation is very similar to that found for the 16S rRNA gene in E. coli. Strains K12 and ECOR 40 showed the highest intercistronic heterogeneity. Most strains showed a strong tendency to homogenization. Concerted evolution could explain the notorious conservation of this region that is supposed to have low functional restrictions. Received: 31 July 1997 / Accepted: 17 October 1997  相似文献   

14.
The DEAE-cellulose-purified 4 S form of the rat liver glucocorticoid receptor can associate with cytosolic factors, as evidenced by an alteration of the sedimentation value of the 7–8 S form. On the basis of sedimentation profile, this form is indistinguishable from the activated, low-salt 7–8 S form isolated from rat liver cytosol. In addition, both the endogenous and reconstituted 7–8 S receptor can bind DNA as the 7–8 S form. In keeping with our reports that the endogenous form of the 7–8 S receptor is sensitive to RNAase digestion, treatment of the cytosol with RNAase prior to mixing with the 4 S receptor prevents the formation of the 7–8 S material. Moreover, warming the cytosol to 50°C prior to mixing with the 4 S receptor also eliminates the ability to form the heavier material. Since RNA is heat-stable, this suggests that other factors may be involved. Treatment of the cytosol with N-ethylmaleimide, a reagent reported to be specific for sulfhydryl groups, also eliminates 7–8 S generating ability. These observations suggest that a protein may be a component of the 7–8 S generating material. This is substantiated by the observation that trypsin or chymotrypsin treatment of the cytosol mitigates the ability of the cytosol to form the 7–8 S material and results in the appearance of a form of the receptor that sediments at approximately 6 S. Protease treatment of partially purified material eliminates the 7–8 S generating activity entirely. We conclude that the 7–8 S form of the receptor can be reconstituted from the 4 S receptor via association with at least two other cytosolic factors, a protein and an RNA.  相似文献   

15.
Both structural and thermodynamic studies are necessary to understand the ribosome assembly. An initial step was made in studying the interaction between a 16S rRNA fragment and S7, a key protein in assembling the prokaryotic ribosome small subunit. The apparent dissociation constant was obtained for complexes of recombinant Escherichia coliandThermus thermophilusS7 with a fragment of the 3" domain of the E. coli16S rRNA. Both proteins showed high rRNA-binding activity, which was not observed earlier. Since RNA and proteins are conformationally labile, their folding must be considered to correctly describe the RNA–protein interactions.  相似文献   

16.
As initial steps to define how the 26S proteasome degrades ubiquitinated proteins in plants, we have characterized many of the subunits that comprise the proteolytic complex from Arabidopsis thaliana. A set of 23 Arabidopsis genes encoding the full complement of core particle (CP) subunits and a collection encoding 12 out of 18 known eukaryotic regulatory particle (RP) subunits, including six AAA-ATPase subunits, were identified. Several of these 26S proteasome genes could complement yeast strains missing the corresponding orthologs. Using this ability of plant subunits to functionally replace yeast counterparts, a parallel structure/function analysis was performed with the RP subunit RPN10/MCB1, a putative receptor for ubiquitin conjugates. RPN10 is not essential for yeast viability but is required for amino acid analog tolerance and degradation of proteins via the ubiquitin-fusion degradation pathway, a subpathway within the ubiquitin system. Surprisingly, we found that the C-terminal motif required for conjugate recognition by RPN10 is not essential for in vivo functions. Instead, a domain near the N-terminus is required. We have begun to exploit the moss Physcomitrella patens as a model to characterize the plant 26S proteasome using reverse genetics. By homologous recombination, we have successfully disrupted the RPN10 gene. Unlike yeast rpn10 strains which grow normally, Physcomitrella rpn10 strains are developmentally arrested, being unable to initiate gametophorogenesis. Further analysis of these mutants revealed that RPN10 is likely required for a developmental program triggered by plant hormones.  相似文献   

17.
18.
Normal activation–inactivation coupling in sodium channels insures that inactivation is slow at small but rapid at large depolarizations. M1651Q/M1652Q substitutions in the cytoplasmic loop connecting the fourth and fifth transmembrane segments of Domain 4 (S4–S5/D4) of the human heart sodium channel subtype 1 (hH1) affect the kinetics and voltage dependence of inactivation (Tang, L., R.G. Kallen, and R. Horn. 1996. J. Gen. Physiol. 108:89–104.). We now show that glutamine substitutions NH2-terminal to the methionines (L1646, L1647, F1648, A1649, L1650) also influence the kinetics and voltage dependence of inactivation compared with the wild-type channel. In contrast, mutations at the COOH-terminal end of the S4–S5/D4 segment (L1654, P1655, A1656) are without significant effect. Strikingly, the A1649Q mutation renders the current decay time constants virtually voltage independent and decreases the voltage dependences of steady state inactivation and the time constants for the recovery from inactivation. Single-channel measurements show that at negative voltages latency times to first opening are shorter and less voltage dependent in A1649Q than in wild-type channels; peak open probabilities are significantly smaller and the mean open times are shorter. This indicates that the rate constants for inactivation and, probably, activation are increased at negative voltages by the A1649Q mutation reminiscent of Y1494Q/ Y1495Q mutations in the cytoplasmic loop between the third and fourth domains (O''Leary, M.E., L.Q. Chen, R.G. Kallen, and R. Horn. 1995. J. Gen. Physiol. 106:641–658.). Other substitutions, A1649S and A1649V, decrease but fail to eliminate the voltage dependence of time constants for inactivation, suggesting that the decreased hydrophobicity of glutamine at either residues A1649 or Y1494Y1495 may disrupt a linkage between S4–S5/D4 and the interdomain 3–4 loop interfering with normal activation–inactivation coupling.  相似文献   

19.
p34cdc2: the S and M kinase?   总被引:14,自引:0,他引:14  
In the yeast cell cycle, the induction of two very different processes, DNA synthesis (S-phase) and mitosis (M-phase), requires the same serine/threonine-specific protein kinase p34cdc2, which has been highly conserved through evolution. On the basis of work conducted largely in multicellular eukaryotes, it has recently been suggested that p34cdc2 is able to perform these two mutually exclusive roles by phosphorylating different sets of substrates through a cell cycle-dependent association with other proteins that dictate the substrate specificity of the protein kinase. To recognize its mitotic substrates, p34cdc2 associates with one of the cyclins--a family of proteins of two distinct but related types (A and B) characterized by their periodic destruction at each mitosis. In interphase, the formation of a complex between p34cdc2 and another protein (or proteins) would allow the phosphorylation of a different set of proteins involved in the G1 to S transition. This review focuses on the evidence for this appealing simple model and the nature of the putative substrates proposed.  相似文献   

20.
S100β is a soluble calcium binding protein released by glial cells. It has been reported as a neurotrophic factor that promotes neurite maturation and outgrowth during development. This protein also plays a role in axonal stability and in long term potentiation in the adult brain. The ability of S100β to modulate neuronal morphology raises the important question whether there is an age-related difference in the expression of S100β in the cerebral and cerebellar cortices of AKR strain mice and is this change is region specific. Our RT–PCR and Western blotting experiments show that the expression of S100β gene in the cerebral and cerebellar cortices starts from 0 day, peaks at about 45 days. However, in 70-week old mice its expression is significantly up-regulated as compared to that of 20-week old mice. S100β follows the same age-related pattern in both cerebral and cerebellar cortices. These results suggest that S100β is important for brain development and establishment of proper brain functions. Up-regulation of S100β in old age may have some role in development of age-related pathological systems in the brain.  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