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.
Macrolide antibiotics block the entrance of nascent peptides to the peptide exit tunnel of the large ribosomal subunit. Expression of specific cis-acting peptides confers low-level macrolide-resistance. We show that, in the case of josamycin, peptide expression does not eject josamycin from the ribosome, implying a peptide resistance mechanism different from that previously suggested for erythromycin. We find dipeptide formation and dipeptidyl-tRNA drop-off in the presence of josamycin to be much slower during translation of resistance than of control mRNAs. We demonstrate low-level josamycin resistance by over-expression of peptidyl-tRNA hydrolase. These findings suggest dual growth-inhibitory action of josamycin by (i) direct inhibition of peptide-elongation and (ii) indirect inhibition of peptide-elongation through rapid peptidyl-tRNA drop-off, leading to depletion of tRNA isoacceptors available for protein synthesis. We propose that josamycin resistance peptide expression brings ribosomes into a “quarantine” state with small drop-off rate, thereby eliminating the josamycin dependent depletion of tRNA isoacceptors in the protein-synthesis-active state.  相似文献   

2.
There is a single-site interaction of [methylene-14C]thiamphenicol and [methylene-14C]chloramphenicol with run-off ribosomes with dissociation constants Kd = 6.8 micronM and Kd = 4.6 micronM respectively. Similar affinities for the antibiotics are observed in polysomes totally deprived of nascent peptides, or bearing nascent peptides on the A-site. However, two types of interaction are observed in endogenous polysomes with some ribosomes bearing nascent peptides on the P-site and other in the A-site. The lower-affinity bindings (dissociation constants Kd = 6.4 micronM and Kd = 1.5 micronM for thiamphenicol and chloramphenicol respectively) are due to the ribosomes bearing nascent peptides on the A-site. The higher-affinity bindings (dissociation constants Kd = 2.3 micronM and Kd = 1.5 micronM for thiamphenicol and chloramphenicol, respectively) are due to the ribosomes bearing nascent peptides on the P-site. Therefore binding of nascent peptides to the A-site does not affect the affinities of thiamphenicol and chloramphenicol for the ribosome. On the other hand interaction of the nascent peptides with the P-site of the ribosomes increases the affinities of both antibiotics for the ribosome. Thiamphenicol and chloramphenicol are thus good inhibitors of peptide bond formation in ribosomes and polysomes. Their affinities are increased precisely when the peptidyl-tRNA is placed in the P-site preceeding the peptide bond formation step, which is specifically blocked by the antibiotics. There is a single-site interaction per ribosome for [35S]thiostrepton, which does not appear to be affected by the attachment to the ribosomes of mRNA, tRNA and nascent peptides either to the A or the P-site. [N-methyl-14C]Lincomycin, [N-methyl-14C]erythromycin, [G-3H]streptogramin B and [G-3H]-streptogramin A bind to run-off ribosomes and polysomes totally free from nascent peptides. However, these antibiotics do not interact with ribosomes bearing nascent peptides either in the A or the P-site and therefore are not active on preformed polysomes. Thus lincomycin and streptogramin A only interact with free ribosomes and 50-S subunits and block the early rounds of peptide bond formation prior to polysome formation. Erythromycin and streptogramin B do not inhibit either initiation or the first round of peptide bond formation. However, erythromycin and streptogramin B, prebound to the ribosome, block peptide elongation probably by steric hindrance with the growing oligopeptide chain when this reaches a certain critical length.  相似文献   

3.
Erythromycin (a 14-membered macrolide) and virginiamycin S (a type B synergimycin) block protein biosynthesis in bacteria, but are virtually inactive on poly(U)-directed poly(Phe) synthesis. We have recently shown, however, that these antibiotics inhibit the in vitro polypeptide synthesis directed by synthetic copolymers: this effect is analyzed further in the present work. We were unable to find any consistent alteration produced by these antibiotics on coupled and uncoupled EF-G- and EF-Tu-dependent GTPases, on the EF-Tu-directed binding of aminoacyl-tRNA to ribosomes, and on the EF-G- and GTP-mediated translocation of peptidyl-tRNA bound to poly(U,C).ribosome complexes. With these complexes, the peptidyl transfer reaction, as measured by peptidylpuromycin synthesis, was 10-30% inhibited by virginiamycin S and erythromycin. A direct relationship between the virginiamycin S- and erythromycin-promoted inhibition of poly(A,C)-directed polypeptide synthesis, on the one hand, and the EF-G concentration and the rate of the polymerization reaction, on the other hand, was observed, in agreement with a postulated reversible inhibitor action of these antibiotics. The increased inhibitory activity, which was observed during the first 4-6 rounds of elongation, in the presence of virginiamycin S or erythromycin, was suggestive of a specific action of these antibiotics on the correct positioning of peptidyl-tRNA at the P site. The marked stimulation of premature release of peptidyl-tRNA from poly(A,C).ribosome complexes can be referred to an altered interaction of the C-terminal aminoacyl residue of the growing peptidyl chain with the ribosome. We conclude that the action of virginiamycin S and erythromycin entails a template-dependent alteration of the interaction of peptidyl-tRNA with the donor site of peptidyltransferase, which may lead to a transient functional block of the ribosome and in some instances to a premature release of peptidyl-tRNA and termination of the elongation process.  相似文献   

4.
The macrolide-lincosamide-streptogramin B class (MLS) of antibiotics contains structurally different but functionally similar drugs, that all bind to the 50S ribosomal subunit. It has been suggested that these compounds block the path by which nascent peptides exit the ribosome. We have studied the mechanisms of action of four macrolides (erythromycin, josamycin, spiramycin and telithromycin), one lincosamide (clindamycin) and one streptogramin B (pristinamycin IA). All these MLS drugs cause dissociation of peptidyl-tRNA from the ribosome. Josamycin, spiramycin and clindamycin, that extend to the peptidyl transferase center, cause dissociation of peptidyl-tRNAs containing two, three or four amino acid residues. Erythromycin, which does not reach the peptidyl transferase center, induces dissociation of peptidyl-tRNAs containing six, seven or eight amino acid residues. Pristinamycin IA causes dissociation of peptidyl-tRNAs with six amino acid residues and telithromycin allows polymerisation of nine or ten amino acid residues before peptidyl-tRNA dissociates. Our data, in combination with previous structural information, suggest a common mode of action for all MLS antibiotics, which is modulated by the space available between the peptidyl transferase center and the drug.  相似文献   

5.
Macrolide antibiotics, such as erythromycin, bind to the nascent peptide exit tunnel (NPET) of the bacterial ribosome and modulate protein synthesis depending on the nascent peptide sequence. Whereas in vitro biochemical and structural methods have been instrumental in dissecting and explaining the molecular details of macrolide-induced peptidyl-tRNA drop-off and ribosome stalling, the dynamic effects of the drugs on ongoing protein synthesis inside live bacterial cells are far less explored. In the present study, we used single-particle tracking of dye-labeled tRNAs to study the kinetics of mRNA translation in the presence of erythromycin, directly inside live Escherichia coli cells. In erythromycin-treated cells, we find that the dwells of elongator tRNAPhe on ribosomes extend significantly, but they occur much more seldom. In contrast, the drug barely affects the ribosome binding events of the initiator tRNAfMet. By overexpressing specific short peptides, we further find context-specific ribosome binding dynamics of tRNAPhe, underscoring the complexity of erythromycin’s effect on protein synthesis in bacterial cells.  相似文献   

6.
The ribosome catalyzes peptide bond formation between peptidyl-tRNA in the P site and aminoacyl-tRNA in the A site. Here, we show that the nature of the C-terminal amino acid residue in the P-site peptidyl-tRNA strongly affects the rate of peptidyl transfer. Depending on the C-terminal amino acid of the peptidyl-tRNA, the rate of reaction with the small A-site substrate puromycin varied between 100 and 0.14 s(-1), regardless of the tRNA identity. The reactivity decreased in the order Lys = Arg > Ala > Ser > Phe = Val > Asp > Pro, with Pro being by far the slowest. However, when Phe-tRNA(Phe) was used as A-site substrate, the rate of peptide bond formation with any peptidyl-tRNA was approximately 7 s(-1), which corresponds to the rate of binding of Phe-tRNA(Phe) to the A site (accommodation). Because accommodation is rate-limiting for peptide bond formation, the reaction rate is uniform for all peptidyl-tRNAs, regardless of the variations of the intrinsic chemical reactivities. On the other hand, the 50-fold increase in the reaction rate for peptidyl-tRNA ending with Pro suggests that full-length aminoacyl-tRNA in the A site greatly accelerates peptide bond formation.  相似文献   

7.
Peptidyl-tRNA dissociation from ribosomes is an energetically costly but apparently inevitable process that accompanies normal protein synthesis. The drop-off products of these events are hydrolysed by peptidyl-tRNA hydrolase. Mutant selections have been made to identify genes involved in the drop-off of peptidyl-tRNA, using a thermosensitive peptidyl-tRNA hydrolase mutant in Escherichia coli. Transposon insertions upstream of the frr gene, which encodes RF4 (ribosome release or recycling factor), restored growth to this mutant. The insertions impaired expression of the frr gene. Mutations inactivating prfC, encoding RF3 (release factor 3), displayed a similar phenotype. Conversely, production of RF4 from a plasmid increased the thermosensitivity of the peptidyl-tRNA hydrolase mutant. In vitro measurements of peptidyl-tRNA release from ribosomes paused at stop signals or sense codons confirmed that RF3 and RF4 were able to stimulate peptidyl-tRNA release from ribosomes, and showed that this action of RF4 required the presence of translocation factor EF2, known to be needed for the function of RF4 in ribosome recycling. When present together, the three factors were able to stimulate release up to 12-fold. It is suggested that RF4 may displace peptidyl-tRNA from the ribosome in a manner related to its proposed function in removing deacylated tRNA during ribosome recycling.  相似文献   

8.
9.
A stochastic model of protein synthesis was modified by including the process of dissociating peptidyl-tRNA from ribosomes. To simulate ribosome editing, the probability of dissociation was assumed to be high if the peptidyl-tRNA was erroneous; that is, if it resulted from transfer of a peptide to an aminoacyl-tRNA that was inappropriate relative to the mRNA codon. The effects of amino acid starvation on protein synthesis were simulated both by increasing the probability of such erring at and by reducing the conditional probability of elongation at “hungry” codons, those whose correct amino acid was in short supply. These probabilities were varied systematically to simulate tryptophan limitation during synthesis of coat protein from bacteriophage MS2.Significant reduction, during starvation, in the synthesis of complete coat protein required large reductions in the probability of elongation at hungry codons but only small increases in the probability of erring. Enhanced dissociation of peptidyl-tRNA during starvation, followed rapidly by dissociation of ribosomes from mRNA, led to reductions in mean polysome size, a result that had been interpreted by others as due to some effect of starvation on the initiation of protein synthesis.Results from experiments by Goldman (1982) on the cell-free synthesis of MS2 coat protein during tryptophan starvation could be mimicked in detail by the computer simulations. A simple competition between correct and erroneous amino acids was sufficient to explain the tryptophan dependence of complete coat protein and internal peptide syntheses. Values for the Michaelis constants were derived from the computer simulations.  相似文献   

10.
The ribosomal peptidyl transferase center is responsible for two fundamental reactions, peptide bond formation and nascent peptide release, during the elongation and termination phases of protein synthesis, respectively. We used in vitro genetics to investigate the functional importance of conserved 23S rRNA nucleotides located in the peptidyl transferase active site for transpeptidation and peptidyl-tRNA hydrolysis. While mutations at A2451, U2585, and C2063 (E. coli numbering) did not significantly affect either of the reactions, substitution of A2602 with C or its deletion abolished the ribosome ability to promote peptide release but had little effect on transpeptidation. This indicates that the mechanism of peptide release is distinct from that of peptide bond formation, with A2602 playing a critical role in peptide release during translation termination.  相似文献   

11.
O W Odom  W D Picking  B Hardesty 《Biochemistry》1990,29(48):10734-10744
The results from experiments involving nonradiative energy transfer indicate that a fluorescent probe on the 5'-end of tRNA(Phe) moves more than 20 A towards probes on ribosomal protein L1 as a peptide bond is formed during the peptidyl transferase reaction on Escherichia coli ribosomes. The peptide itself moves no more than a few angstroms during peptide bond formation, as judged by the movement of fluorescent probes attached to the phenylalanine amino group of phenylalanyl-tRNA. Other results demonstrate that an analogue of peptidyl-tRNA, deacylated tRNA, and puromycin can be bound simultaneously to the same ribosome, indicating that there are three physically distinct sites to which tRNA is bound during the reaction steps by which peptides are elongated. The results appear to be consistent with the displacement model of peptide elongation.  相似文献   

12.
During translation, ribosomes stall on mRNA when the aminoacyl-tRNA to be read is not readily available. The stalled ribosomes are deleterious to the cell and should be rescued to maintain its viability. To investigate the contribution of some of the cellular translation factors on ribosome rescuing, we provoked stalling at AGA codons in mutants that affected the factors and then analyzed the accumulation of oligopeptidyl (peptides of up to 6 amino acid residues, oligopep-)-tRNA or polypeptidyl (peptides of more than 300 amino acids in length, polypep-)-tRNA associated with ribosomes. Stalling was achieved by starvation for aminoacyl-tRNA(Arg4) upon induced expression of engineered lacZ (β-galactosidase) reporter gene harboring contiguous AGA codons close to the initiation codon or at internal codon positions together with minigene ATGAGATAA accompanied by reduced peptidyl-tRNA hydrolase (Pth). Our results showed accumulations of peptidyl-tRNA associated with ribosomes in mutants for release factors (RF1, RF2, and RF3), ribosome recycling factor (RRF), Pth, and transfer-messenger RNA (tmRNA), implying that each of these factors cooperate in rescuing stalled ribosomes. The role of these factors in ribosome releasing from the stalled complex may vary depending on the length of the peptide in the peptidyl-tRNA. RF3 and RRF rescue stalled ribosomes by "drop-off" of peptidyl-tRNA, while RF1, RF2 (in the absence of termination codon), or Pth may rescue by hydrolyzing the associated peptidyl-tRNA. This is followed by the disassembly of the ribosomal complex of tRNA and mRNA by RRF and elongation factor G.  相似文献   

13.
Expression of minigenes encoding tetra- or pentapeptides MXLX or MXLXV (E peptides), where X is a nonpolar amino acid, renders cells erythromycin resistant whereas expression of minigenes encoding tripeptide MXL does not. By using a 3A′ reporter gene system beginning with an E-peptide-encoding sequence, we asked whether the codons UGG and GGG, which are known to promote peptidyl-tRNA drop-off at early positions in mRNA, would result in a phenotype of erythromycin resistance if located after this sequence. We find that UGG or GGG, at either position +4 or +5, without a following stop codon, is associated with an erythromycin resistance phenotype upon gene induction. Our results suggest that, while a stop codon at +4 gives a tripeptide product (MIL) and erythromycin sensitivity, UGG or GGG codons at the same position give a tetrapeptide product (MILW or MILG) and phenotype of erythromycin resistance. Thus, the drop-off event on GGG or UGG codons occurs after incorporation of the corresponding amino acid into the growing peptide chain. Drop-off gives rise to a peptidyl-tRNA where the peptide moiety functionally mimics a minigene peptide product of the type previously associated with erythromycin resistance. Several genes in Escherichia coli fulfill the requirements of high mRNA expression and an E-peptide sequence followed by UGG or GGG at position +4 or +5 and should potentially be able to give an erythromycin resistance phenotype.  相似文献   

14.
The major enzymatic activity of the ribosome is the catalysis of peptide bond formation. The active site -- the peptidyl transferase center -- is composed of ribosomal RNA (rRNA), and interactions between rRNA and the reactants, peptidyl-tRNA and aminoacyl-tRNA, are crucial for the reaction to proceed rapidly and efficiently. Here, we describe the influence of rRNA interactions with cytidine residues in A-site substrate analogs (C-puromycin or CC-puromycin), mimicking C74 and C75 of tRNA on the reaction. Base-pairing of C75 with G2553 of 23S rRNA accelerates peptide bond formation, presumably by stabilizing the peptidyl transferase center in its productive conformation. When C74 is also present in the substrate analog, the reaction is slowed down considerably, indicating a slow step in substrate binding to the active site, which limits the reaction rate. The tRNA-rRNA interactions lead to a robust reaction that is insensitive to pH changes or base substitutions in 23S rRNA at the active site of the ribosome.  相似文献   

15.
Summary Continuation of early evolutionary bonding between tRNAs would provide a solution to residence time problems between peptidyl-tRNA and mRNA. It could also improve the speed of peptide bond formation by holding the amino acid close to the growing peptide.The tRNA clover leaf structure would allow each tRNA to from a TC(GA)-loop bond to one side and a D-loop bond to the other, hence fixing itself within a group of tRNAs, all attached to the mRNA. This can be developed into a system for peptide elongation in which bonds are made and broken in an ordered sequence, with each step triggering the next. This leads to a model system that fits with some recent propsals for a three-site ribosome.  相似文献   

16.
Following peptide bond formation, transfer RNAs (tRNAs) and messenger RNA (mRNA) are translocated through the ribosome, a process catalyzed by elongation factor EF-G. Here, we have used a combination of chemical footprinting, peptidyl transferase activity assays, and mRNA toeprinting to monitor the effects of EF-G on the positions of tRNA and mRNA relative to the A, P, and E sites of the ribosome in the presence of GTP, GDP, GDPNP, and fusidic acid. Chemical footprinting experiments show that binding of EF-G in the presence of the non-hydrolyzable GTP analog GDPNP or GDP.fusidic acid induces movement of a deacylated tRNA from the classical P/P state to the hybrid P/E state. Furthermore, stabilization of the hybrid P/E state by EF-G compromises P-site codon-anticodon interaction, causing frame-shifting. A deacylated tRNA bound to the P site and a peptidyl-tRNA in the A site are completely translocated to the E and P sites, respectively, in the presence of EF-G with GTP or GDPNP but not with EF-G.GDP. Unexpectedly, translocation with EF-G.GTP leads to dissociation of deacylated tRNA from the E site, while tRNA remains bound in the presence of EF-G.GDPNP, suggesting that dissociation of tRNA from the E site is promoted by GTP hydrolysis and/or EF-G release. Our results show that binding of EF-G in the presence of GDPNP or GDP.fusidic acid stabilizes the ribosomal intermediate hybrid state, but that complete translocation is supported only by EF-G.GTP or EF-G.GDPNP.  相似文献   

17.
The bacterial ssrA gene codes for a dual function RNA, tmRNA, which possesses tRNA-like and mRNA-like regions. The tmRNA appends an oligopeptide tag to the polypeptide on the P-site tRNA by a trans-translation process that rescues ribosomes stalled on the mRNAs and targets the aberrant protein for degradation. In cells, processing of the stalled ribosomes is also pioneered by drop-off of peptidyl-tRNAs. The ester bond linking the peptide to tRNA is hydrolyzed by peptidyl-tRNA hydrolase (Pth), an essential enzyme, which releases the tRNA and the aberrant peptide. As the trans-translation mechanism utilizes the peptidyl-transferase activity of the stalled ribosomes to free the tRNA (as opposed to peptidyl-tRNA drop-off), the need for Pth to recycle such tRNAs is bypassed. Thus, we hypothesized that tmRNA may rescue a defect in Pth. Here, we show that overexpression of tmRNA rescues the temperature-sensitive phenotype of Escherichia coli (pthts). Conversely, a null mutation in ssrA enhances the temperature-sensitive phenotype of the pthts strain. Consistent with our hypothesis, overexpression of tmRNA results in decreased accumulation of peptidyl-tRNA in E.coli. Furthermore, overproduction of tmRNA in E.coli strains deficient in ribosome recycling factor and/or lacking the release factor 3 enhances the rescue of pthts strains. We discuss the physiological relevance of these observations to highlight a major role of tmRNA in decreasing cellular peptidyl-tRNA load.  相似文献   

18.
The upstream open reading frame (uORF) in the mRNA encoding S-adenosylmethionine decarboxylase is a cis-acting element that confers feedback control by cellular polyamines on translation of this message. Recent studies demonstrated that elevated polyamines inhibit synthesis of the peptide encoded by the uORF by stabilizing a ribosome paused in the vicinity of the termination codon. These studies suggested that polyamines act at the termination step of uORF translation. In this paper, we demonstrate that elevated polyamines stabilize an intermediate in the termination process, the complete nascent peptide linked to the tRNA that decodes the final codon. The peptidyl-tRNA molecule is found associated with the ribosome fraction, and decay of this molecule correlated with release of the paused ribosome from the message. Furthermore, the stability of this complex is influenced by the same parameters that influence regulation by the uORF in vivo, namely the concentration of polyamines and the sequence of the uORF-encoded peptide. These results suggest that the regulated step in uORF translation is after formation of the peptidyl-tRNA molecule but before hydrolysis of the peptidyl-tRNA bond. This regulation may involve an interaction between the peptide, polyamines, and a target in the translational apparatus.  相似文献   

19.
Macrolides, lincosamides and type B synergimycins are powerful inhibitors of protein synthesis in vivo, but many of them were found to be inactive in vitro. In the present work, we confirm that virginiamycin S (a type B synergimycin) and erythromycin (a 14-membered macrolide) have no effect on poly(U)-directed poly(Phe) synthesis. However, the amino-acid polymerization reactions directed by poly(U,G), poly(U,C), poly(A,G) and poly(A,C) were increasingly inhibited (20-50%) by both antibiotics. The action of these inhibitors proved to be template-dependent and favored by the incorporation of proline and of basic amino acids into peptides. Under these conditions, virginiamycin S and erythromycin markedly stimulated a release of peptidyl-tRNA from the ribosomes. In the poly(A,C) model system, these antibiotics produced a 50% inhibition of amino-acid incorporation into total peptides, a 70% release of ribosome-bound peptidyl-tRNA, and a 95% repression of the synthesis of long peptide chains. The production of equivalent effects at saturating concentrations of these antibiotics in the four model systems examined is suggestive of a similarity in their mode of action. Our results indicate that 14-membered macrolides and type B synergimycins can act on ribosomes during the whole elongation process. The functional block produced by both antibiotics is usually reversible, but may result in a premature release of peptidyl-tRNA when the stability of ribosomal complexes is lowered by the incorporation of basic amino acids.  相似文献   

20.
Dinos GP  Kalpaxis DL 《Biochemistry》2000,39(38):11621-11628
The inhibition of peptide bond formation by tylosin, a 16-membered ring macrolide, was studied in a model system derived from Escherichia coli. In this cell-free system, a peptide bond is formed between puromycin (acceptor substrate) and AcPhe-tRNA (donor substrate) bound at the P-site of poly(U)-programmed ribosomes. It is shown that tylosin inhibits puromycin reaction as a slow-binding, slowly reversible inhibitor. Detailed kinetic analysis reveals that tylosin (I) reacts rapidly with complex C, i.e., the AcPhe-tRNA. poly(U).70S ribosome complex, to form the encounter complex CI, which then undergoes a slow isomerization and is converted to a tight complex, CI, inactive toward puromycin. These events are described by the scheme C + I <==> (K(i)) CI <==> (k(4), k(5)) CI. The K(i), k(4), and k(5) values are equal to 3 microM, 1.5 min(-1), and 2.5 x 10(-3) min(-1), respectively. The extremely low value of k(5) implies that the inactivation of complex C by tylosin is almost irreversible. The irreversibility of the tylosin effect on peptide bond formation is significant for the interpretation of this antibiotic's therapeutic properties; it also renders the tylosin reaction a useful tool in the study of other macrolides failing to inhibit the puromycin reaction but competing with tylosin for common binding sites on the ribosome. Thus, the tylosin reaction, in conjunction with the puromycin reaction, was applied to investigate the erythromycin mode of action. It is shown that erythromycin (Er), like tylosin, interacts with complex C according to the kinetic scheme C + Er <==> (K(er)) CEr <==> (k(6), k(7)) C*Er and forms a tight complex, CEr, which remains active toward puromycin. The determination of K(er), k(6), and k(7) enables us to classify erythromycin as a slow-binding ligand of ribosomes.  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