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 752 毫秒
1.
Expression of the plasmid gene cat-86 is induced in Bacillus subtilis by two antibiotics, chloramphenicol and the nucleoside antibiotic amicetin. We proposed that induction by either drug causes the destabilization of a stem-loop structure in cat-86 mRNA that sequesters the ribosome-binding site for the cat coding sequence. The destabilization event frees the ribosome-binding site, permitting the initiation of translation of cat-86 mRNA. cat-86 induction is due to the stalling of a ribosome in a leader region of cat-86 mRNA, which is located 5' to the RNA stem-loop structure. A stalled ribosome that is active in cat-86 induction has its aminoacyl site occupied by leader codon 6. To test the hypothesis that a leader site 5' to codon 6 permits a ribosome to stall in the presence of an inducing antibiotic, we inserted an extra codon between leader codons 5 and 6. This insertion blocked induction, which was then restored by the deletion of leader codon 6. Thus, induction seems to require the maintenance of a precise spatial relationship between an upstream leader site(s) and leader codon 6. Mutations in the ribosome-binding site for the cat-86 leader, RBS-2, which decreased its strength of binding to 16S rRNA, prevented induction. In contrast, mutations that significantly altered the sequence of RBS-2 but increased its strength of binding to 16S rRNA did not block induction by either chloramphenicol or amicetin. We therefore suspected that the proposed leader site that permitted drug-mediated stalling was located between RBS-2 and leader codon 6. This region of the cat-86 leader contains an eight-nucleotide sequence (conserved region I) that is largely conserved among all known cat leaders. The codon immediately 5' to conserved region I differs, however, between amicetin-inducible and amicetin-noninducible cat genes. In amicetin-inducible cat genes such as cat-86, the codon 5' to conserved region I is a valine codon, GTG. The same codon in amicetin-noninducible cat genes is a lysine codon, either AAA or AAG. When the GTG codon immediately 5' to conserved region I in cat-86 was changed to AAA, amicetin was no longer active in cat-86 induction, but chloramphenicol induction was unaffected by the mutation. The potential role of the GTG codon in amicetin induction is discussed.  相似文献   

2.
Placing a translation stop codon at the ribosomal pause site in the leader of the attenuation-regulated cat-86 gene activates cat expression in the absence of the inducer, chloramphenicol. Genetic experiments have shown that this phenomenon depends on the amino acid sequence of the leader-encoded peptide and could readily be explained if the peptide was an inhibitor of translation termination. Here we demonstrate that the cat-86 leader pentapeptide is an in vitro inhibitor of translation termination in addition to its previously described antipeptidyltransferase activity.  相似文献   

3.
Genes encoding chloramphenicol acetyltransferase in gram-positive bacteria are induced by chloramphenicol. Induction reflects an ability of the drug to stall a ribosome at a specific site in cat leader mRNA. Ribosome stalling at this site alters downstream RNA secondary structure, thereby unmasking the ribosome-binding site for the cat coding sequence. Here, we show that ribosome stalling in the cat-86 leader is a function of leader codons 2 through 5 and that stalling requires these codons to be presented in the correct reading frame. Codons 2 through 5 specify Val-Lys-Thr-Asp. Insertion of a second copy of the stall sequence 5' to the authentic stall sequence diminished cat-86 induction fivefold. Thus, the stall sequence can function in ribosome stalling when the stall sequence is displaced from the downstream RNA secondary structure. We suggest that the stall sequence may function in cat induction at two levels. First, the tetrapeptide specified by the stall sequence likely plays an active role in the induction strategy, on the basis of previously reported genetic suppression studies (W. W. Mulbry, N. P. Ambulos, Jr., and P.S. Lovett, J. Bacteriol. 171:5322-5324, 1989). Second, we show that embedded within the stall sequence of cat leaders is a region which is complementary to a sequence internal in 16S rRNA of Bacillus subtilis. This complementarity may guide a ribosome to the proper position on leader mRNA or potentiate the stalling event, or both. The region of complementarity is absent from Escherichia coli 16S rRNA, and cat genes induce poorly, or not at all, in E. coli.  相似文献   

4.
5.
6.
7.
The mutation sup-3 in Bacillus subtilis suppresses ochre (TAA) mutations at each of three codons in the 5' end of the cat-86 coding sequence. The suppressor is shown to insert lysine at ochre codons. The efficiency of suppression by sup-3 is about 15%, as determined by changing a cat-86 Lys codon (codon 12) to an ochre codon and measuring the level of CAT in the suppressor-containing strain. The results obtained are discussed in light of previous observations that ochre mutations at cat leader codons 2 and 3 can be phenotypically suppressed by sup-3, whereas ochre mutations at leader codons 4 and 5 cannot. Translation of the cat leader is essential to inducible expression of cat. Our data support the interpretation that the nature of amino acids 2 through 5 of the leader peptide contributes to determining whether chloramphenicol can stall a ribosome in the leader, which in turn leads to induction of cat expression.  相似文献   

8.
9.
10.
The chloramphenicol acetyltransferase gene cat-86 is induced through a mechanism that is a variation of classical attenuation. Induction results from the destabilization of an RNA stem-loop that normally sequesters the cat-86 ribosome-binding site. Destabilization of the stem-loop is due to the stalling of a ribosome in the leader region of cat-86 mRNA at a position that places the A site of the stalled ribosome at leader codon 6. Two events can stall ribosomes at the correct location to induce cat-86 translation: addition of chloramphenicol to cells and starvation of cells for the amino acid specified by leader codon 6. Induction by amino acid starvation is an anomaly because translation of the cat-86 coding sequence requires all 20 amino acids. To explain this apparent contradiction we postulated that amino acid starvation triggers intracellular proteolysis, thereby providing levels of the deprived amino acid sufficient for cat-86 translation. Here we show that a mutation in relA, the structural gene for stringent factor, blocks intracellular proteolysis that is normally triggered by amino acid starvation. The relA mutation also blocks induction of cat-86 by amino acid starvation, but the mutation does not interfere with chloramphenicol induction. Induction by amino acid starvation can be demonstrated in relA mutant cells if the depleted amino acid is restored at very low levels (e.g., 2 micrograms/ml). A mutation in relC, which may be the gene for ribosomal protein L11, blocks induction of cat-86 by either chloramphenicol or amino acid starvation. We believe this effect is due to a structural alteration of the ribosome resulting from the relC mutation and not to the relaxed phenotype of the cells.  相似文献   

11.
12.
Induction of the chloramphenicol acetyltransferase gene cat-86 in Bacillus subtilis results from the activation of translation of cat-86 mRNA. The inducers, chloramphenicol and amicetin, are thought to enable ribosomes to destabilize a stem-loop structure in cat-86 mRNA that sequesters the ribosome binding site for the cat-86 coding sequence, designated RBS-3. The region of cat-86 mRNA which is 5' to the stem-loop contained two additional ribosome binding sites, RBS-1 and RBS-2, located 84 and 56 nucleotides, respectively, upstream from RBS-3. RBS-1 and RBS-2 were each followed by a potential translation initiation codon and a short open reading frame. Bal 31-generated deletions into the 5' end of the regulatory region that removed RBS-1 but did not enter RBS-2 caused a fourfold decrease in the uninduced and chloramphenicol-induced level of cat-86 expression and a more than 10-fold reduction in the amicetin-induced level of expression. Deletions that removed both RBS-1 and RBS-2 but did not enter the stem-loop abolished both chloramphenicol- and amicetin-inducible expression. These data indicate that RBS-2 and sequences 3' to RBS-2 are minimally essential to chloramphenicol induction. However, the presence of RBS-1 in the mRNA elevated the maximum level of expression obtained during chloramphenicol induction. These studies also demonstrate that induction of cat-86 by amicetin is highly dependent on RBS-1. To determine whether a correlation existed between RBS-1 and amicetin inducibility, we examined the sequence of the regulatory regions for two natural variants of cat-86, cat-66 and cat-57, which are chloramphenicol inducible but are very poorly induced by amicetin. Both contained nucleotide sequence differences from cat-86 in the vicinity of RBS-1 that would prevent translation of the leader peptide associated with RBS-1 in cat-86. In contrast, the regulatory regions got the three genes were virtually identical in the vicinity of RBS-2. These data indicate that efficient induction by amicetin requires sequences that are not essential for induction by chloramphenicol.  相似文献   

13.
catA86 is the second gene in a constitutively transcribed, two-gene operon cloned from Bacillus pumilus . The region that intervenes between the upstream gene, termed the leader, and the catA86 coding sequence contains a pair of inverted repeat sequences which cause sequestration of the catA86 ribosome binding site in mRNA secondary structure. As a consequence, the catA86 coding sequence is untranslatable in the absence of inducer. Translation of the catA86 coding sequence is induced by chloramphenicol in Gram-positives and induction requires a function of the leader coding sequence. The leader-encoded peptide has been proposed to instruct its translating ribosome to pause at leader codon 6, enabling chloramphenicol to stall the ribosome at that site. Ribosome stalling causes destabilization of the RNA secondary structure, exposing the catA86 ribosome binding site, allowing activation of its translation. A comparable mechanism of induction by chloramphenicol has been proposed for the regulated cmlA gene from Gram-negative bacteria. The catA86 and cmlA leader-encoded peptides are in vitro inhibitors of peptidyl transferase, which is thought to be the basis for selection of the site of ribosome stalling. Both leader-encoded peptides have been shown to alter the secondary structure of Escherichia coli 23S rRNA in vitro. All peptide-induced changes in rRNA conformation are within domains IV and V, which contains the peptidyl transferase center. Here we demonstrate that the leader peptides alter the conformation of domains IV and V of large subunit rRNA from yeast and a representative of the Archaea. The rRNA target for binding the leader peptides is therefore conserved across kingdoms.  相似文献   

14.
The macrolide erythromycin binds to the large subunit of the prokaryotic ribosome near the peptidyltransferase center (PTC) and inhibits elongation of new peptide chains beyond a few amino acids. Nucleotides A2058 and A2059 (E. coli numbering) in 23S rRNA play a crucial role in the binding of erythromycin, and mutation of nucleotide A2058 confers erythromycin resistance in both gram-positive and gram-negative bacteria. There are high levels of sequence and structural similarity in the PTC of prokaryotic and eukaryotic ribosomes. However, eukaryotic ribosomes are resistant to erythromycin and the presence of a G at the position equivalent to E. coli nucleotide A2058 is believed to be the reason. To test this hypothesis, we introduced a G to A mutation at this position of the yeast Saccharomyces cerevisiae 25S rRNA and analyzed sensitivity toward erythromycin. Neither growth studies nor erythromycin binding assays on mutated yeast ribosomes indicated any erythromycin sensitivity in mutated yeast strains. These results suggest that the identity of nucleotide 2058 is not the only determinant responsible for the difference in erythromycin sensitivity between yeast and prokaryotes.  相似文献   

15.
The plasmid gene cat-86 and the cat gene resident on pC194 each encode chloramphenicol-inducible chloramphenicol acetyltransferase activity in Bacillus subtilis. Chloramphenicol induction has been proposed to result from chloramphenicol binding to ribosomes, which then permits the drug-modified ribosomes to perform events essential to induction. If this proposal were correct, B. subtilis mutants containing chloramphenicol-insensitive ribosomes should not permit chloramphenicol induction of either cat-86 or pC194 cat. However, we and others have been unable to isolate chloramphenicol-resistant ribosomal mutants of B. subtilis 168. We therefore developed a simple procedure for screening other antibiotics for the potential to induce cat-86 expression. One antibiotic, amicetin, was found to be an effective inducer of cat-86 but not of the cat gene on pC194. Amicetin and chloramphenicol each interact with the 50S ribosomal subunit, and the mechanism of cat-86 induction by both drugs may be similar. Amicetin-resistant mutants of B. subtilis were readily isolated, and in none of six mutants tested was cat-86 detectably inducible by amicetin, although the chloramphenicol-inducible phenotype was retained. The ami-1 mutation which is present in one of these amicetin-resistant mutants was mapped by PBS1 transduction to the "ribosomal gene cluster" adjacent to cysA. Additionally, ribosomes from cells harboring the ami-1 mutation contained an altered BL12a protein, as detected in two-dimensional polyacrylamide gel electrophoresis. Lastly, an in vitro protein-synthesizing system that uses ribosomes from an ami-1-containing cell line was more resistant to amicetin than a system that uses ribosomes from an amicetin-sensitive but otherwise isogenic strain. These results indicate that the host mutation, ami-1, which effectively abolished the inducibility of cat-86 by amicetin, altered a ribosomal component.  相似文献   

16.
17.
18.
19.
20.
A molecular genetic approach has been employed to investigate functional interactions within 23S rRNA. Each of the three base substitutions at guanine 2032 has been made. The 2032A mutation confers resistance to the antibiotics chloramphenicol and clindamycin, which interact with the 23S rRNA peptidyltransferase loop. All three base substitutions at position 2032 produce an erythromycin-hypersensitive phenotype. The 2032 substitutions were compared with and combined with a 12-bp deletion mutation in domain II and point mutations at positions 2057 and 2058 in the peptidyltransferase region of domain V that also confer antibiotic resistance. Both the domain II deletion and the 2057A mutation relieve the hypersensitive effect of the 2032A mutation, producing an erythromycin-resistant phenotype; in addition, the combination of the 2032A and 2057A mutations confers a higher level of chloramphenicol resistance than either mutation alone. 23S rRNAs containing mutations at position 2058 that confer clindamycin and erythromycin resistance become deleterious to cell growth when combined with the 2032A mutation and, additionally, confer hypersensitivity to erythromycin and sensitivity to clindamycin and chloramphenicol. Introduction of the domain II deletion into these double-mutation constructs gives rise to erythromycin resistance. The results are interpreted as indicating that position 2032 interacts with the peptidyltransferase loop and that there is a functional connection between domains II and V.  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