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.
Studies on the catalytic rate constant of ribosomal peptidyltransferase   总被引:3,自引:0,他引:3  
A detailed kinetic analysis of a model reaction for the ribosomal peptidyltransferase is described, using fMet-tRNA or Ac-Phe-tRNA as the peptidyl donor and puromycin as the acceptor. The initiation complex (fMet-tRNA X AUG X 70 S ribosome) or (Ac-Phe-tRNA X poly(U) X 70 S ribosome) (complex C) is isolated and then reacted with excess puromycin (S) to give fMet-puromycin or Ac-Phe-puromycin. This reaction (puromycin reaction) is first order at all concentrations of S tested. An important asset of this kinetic analysis is the fact that the relationship between the first order rate constant kobs and [S] shows hyperbolic saturation and that the value of kobs at saturating [S] is a measure of the catalytic rate constant (k cat) of peptidyltransferase in the puromycin reaction. With fMet-tRNA as the donor, this kcat of peptidyltransferase is 8.3 min-1 when the 0.5 M NH4Cl ribosomal wash is present, compared to 3.8 min-1 in its absence. The kcat of peptidyltransferase is 2.0 min-1 when Ac-Phe-tRNA replaces fMet-tRNA in the presence of the ribosomal wash and decreases to 0.8 min-1 in its absence. This kinetic procedure is the best method available for evaluating changes in the activity of peptidyltransferase in vitro. The results suggest that peptidyltransferase is subjected to activation by the binding of fMet-tRNA to the 70 S initiation complex.  相似文献   

2.
Yeast ribosomal protein L41 is dispensable in the yeast. Its absence had no effect on polyphenylalanine synthesis activity, and a limited effect on growth, translational accuracy, or the resistance toward the antibiotic paromomycin. Removal of L41 did not affect the 60:40 S ratio, but it reduced the amount of 80 S, suggesting that L41 is involved in ribosomal subunit association. However, the two most important effects of L41 were on peptidyltransferase activity and translocation. Peptidyltransferase activity was measured as a second-order rate constant (k(cat)/K(s)) corresponding to the rate of peptide bond formation; this k(cat)/K(s) was lowered 3-fold to 1.15 min(-1) mm(-1) in the L41 mutant compared with 3.46 min(-1) mm(-1) in the wild type. Translocation was also affected by L41. Elongation factor 2 (EF2)-dependent (enzymatic) translocation of Ac-Phe-tRNA from the A- to P-site was more efficient in the absence of L41, because 50% translocation was achieved at only 0.004 microm EF2 compared with 0.02 microm for the wild type. Furthermore, the EF2-dependent translocation was inhibited by 50% at 2.5 microm of the translocation inhibitor cycloheximide in the L41 mutant compared with 1.2 microm in the wild type. Finally, the rate of EF2-independent (spontaneous) translocation was increased in the absence of L41.  相似文献   

3.
Increased efficiencies of programmed -1 ribosomal frameshifting in yeast cells expressing mutant forms of ribosomal protein L3 are unable to maintain the dsRNA "Killer" virus. Here we demonstrate that changes in frameshifting and virus maintenance in these mutants correlates with decreased peptidyltransferase activities. The mutants did not affect Ty1-directed programmed +1 ribosomal frameshifting or nonsense-mediated mRNA decay. Independent experiments demonstrate similar programmed -1 ribosomal frameshifting specific defects in cells lacking ribosomal protein L41, which has previously been shown to result in peptidyltransferase defects in yeast. These findings are consistent with the hypothesis that decreased peptidyltransferase activity should result in longer ribosome pause times after the accommodation step of the elongation cycle, allowing more time for ribosomal slippage at programmed -1 ribosomal frameshift signals.  相似文献   

4.
N-(2-Nitro-4-azidobenzoyl)-[3H]puromycin (NAB-puromycin) was synthesized as a photoreactive derivative of puromycin in order to detect ribosomal proteins located near the peptidyltransferase centre of rat liver ribosomes. Irradiation of ribosome-NAB-puromycin complexes leads to covalent attachment of the affinity label to proteins of the large ribosomal subunit, in particular to proteins L28/29, and, to a somewhat lower extent, to proteins L4, L6, L10 and L24. The results are discussed in the light of earlier studies performed with other affinity labels that attacked the peptidyltransferase region of rat liver ribosomes.  相似文献   

5.
Virginiamycin M inhibits both peptide bond formation and binding of aminoacyl-tRNA to bacterial ribosomes, and induces a lasting inactivation of the 50 S subunit (50 S). In the present work, the effects of this antibiotic on the acceptor and donor sites of peptidyltransferase have been explored, in the presence of virginiamycin M as well as after its removal. Virginiamycin M inhibited the binding of puromycin to ribosomes and reduced both the enzymatic and nonenzymatic binding of Phe-tRNA to the A site by inducing its release from the ribosomes (similar effects were observed with 50 S), whereas the antibiotic had no effect on the binding of unacylated tRNAPhe to the same site. Moreover, virginiamycin M caused Ac-Phe-tRNA or Phe-tRNA to be released from the ribosomal P site, when complexes were incubated with unacylated tRNA, elongation factor G, and GTP (similar finding with 50 S). Instead, peptide bond formation between Ac-Phe-tRNA positioned at the P site and Phe-tRNA at the A site was found to take place, albeit at a very low rate, in the presence of the antibiotic. The overall conclusion is that both the acceptor and donor substrate binding sites of the peptidyltransferase, which interact with the aminoacyl moiety of tRNA, are permanently altered upon transient contact of ribosomes with virginiamycin M.  相似文献   

6.
The initiation complex (t-complex) formed in a cell-free system (E. coli) from Ac-Phe-tRNA, poly(U) and washed ribosomes in the presence of initiation factors (ribosomal wash) and GTP, contains the Ac-Phe-tRNA bound quantitatively in a puromycin-reactive state. The t-complex is irreversibly inactivated by spiramycin with respect to its reactivity toward puromycin. The inactivated t-complex retains all of the Ac-Phe-tRNA bound, but it does not react with puromycin (2 x10-minus-3M) within 32 min at 25 degrees. In the case of another inhibitor protein synthesis, sparsomycin, the permanently "modified" t-complex not only retains all the bound Ac-Phe-tRNA but it can still react with puromycin. In the continuous presence of sparsomycin (1 x 10-minus-7M) the bound Ac-Phe-tRNA reacts quantitatively at a rate which is one-tenth the rate at which the t-complex reacts with puromycin, at low (6.25 x 10-minus-5M) or high (2 x 10-minus-3M) concentrations. These results are not in agreement with current views according to which aparsomycin binds to the ribosome reversibly at a single site with a KI in the range of 10-minus6-10-minus-7 M and according to which this stie is at the A'-site (puromycin site) of peptidyl transferase.  相似文献   

7.
To ensure accurate and rapid protein synthesis, nearby and distantly located functional regions of the ribosome must dynamically communicate and coordinate with one another through a series of information exchange networks. The ribosome is ~2/3 rRNA and information should pass mostly through this medium. Here, two viable mutants located in the peptidyltransferase center (PTC) of yeast ribosomes were created using a yeast genetic system that enables stable production of ribosomes containing only mutant rRNAs. The specific mutants were C2820U (Escherichia coli C2452) and Ψ2922C (E. coli U2554). Biochemical and genetic analyses of these mutants suggest that they may trap the PTC in the ‘open’ or aa-tRNA bound conformation, decreasing peptidyl-tRNA binding. We suggest that these structural changes are manifested at the biological level by affecting large ribosomal subunit biogenesis, ribosomal subunit joining during initiation, susceptibility/resistance to peptidyltransferase inhibitors, and the ability of ribosomes to properly decode termination codons. These studies also add to our understanding of how information is transmitted both locally and over long distances through allosteric networks of rRNA–rRNA and rRNA–protein interactions.  相似文献   

8.
Using 125I-labeled phenol-alanine sparsomycin, an analogue of sparsomycin having higher biological activity than the unmodified antibiotic, we studied the requirements and the characteristics of its interaction with the ribosome. The drug does not bind to either isolated ribosomal subunits or reconstituted whole ribosomes. For sparsomycin binding to 70S and 80S ribosomes, the occupation of the peptidyltransferase P-site by an N-blocked aminoacyl-tRNA is a definitive requirement. The sparsomycin analogue binds to bacterial and yeast ribosomes with Ka values of around 10(6) M-1 and 0.6 x 10(6) M-1, respectively, but its affinity is probably affected by the character of the peptidyl-tRNA bound to the P-site. Chloramphenicol, lincomycin, and 16-atom ring macrolides compete with sparsomycin for binding to bacterial ribosomes, but streptogramins and 14-atom ring macrolides do not. Considering the reported low affinity of puromycin for bacterial ribosomes, this antibiotic is also a surprisingly good competitor of sparsomycin binding to these particles. In the case of yeast ribosomes, blasticidin is a relatively good competitor of sparsomycin interaction, but anisomycin, trichodermin, and narciclasin are not. As expected, puromycin is a poor competitor of the binding in this case. The results from competition studies carried out with different sparsomycin analogues reveal, in some cases, a discrepancy between the drug ribosomal affinity and its biological effects. This suggests that some intermediate step, perhaps a ribosomal conformational change, is required for the inhibition to take place.  相似文献   

9.
The ribosomal L12 protein gene of Sulfolobus solfataricus (SsoL12) has been subcloned and overexpressed in Escherichia coli. Five protein L12 mutants were designed: two NH2-terminal and two COOH-terminal truncated mutants and one mutant lacking the highly charged part of the COOH-terminal region. The mutant protein genes were overexpressed in E. coli and the products purified. The amino acid composition was verified and the NH2 terminally truncated mutants were subjected to Edman degradation. The SsoL12 protein was selectively removed from entire S. solfataricus ribosomes by an ethanol wash. The remaining ribosomal core particles showed a substantial decrease in the in vitro translational activity. S. solfataricus L12 protein overexpressed in E. coli (SsoL12e) was incorporated into these ribosomal cores and restored their translational activity. Mutants lacking any part of the COOH-terminal region could be incorporated into these cores, as proven by two-dimensional polyacrylamide gels of the reconstituted particles. Mutant SsoL12 MC2 (residue 1-70) was sufficient for dimerization and incorporation into ribosomes. In contrast to the COOH terminally truncated mutants, L12 proteins lacking the 12 highly conserved NH2-terminal residues or the entire NH2-terminal region (44 amino acids) are unable to bind to ribosomes, suggesting that the SsoL12 protein binds with its NH2-terminal portion to the ribosome. None of the mutants could significantly increase the translational activity of the core particles suggesting that every deleted part of the protein was needed directly or indirectly for translational activity. Our results suggest that the COOH terminally truncated mutants were bound to ribosomes but not functional for translation. Cores preincubated with these COOH terminally truncated mutants regained activity when a second incubation with the entire overexpressed SsoL12e protein followed. This finding suggests that archaebacterial L12 proteins are freely exchanged on the ribosome.  相似文献   

10.
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   

11.
The photoincorporation of p-azido[3H]puromycin [6-(dimethylamino)-9-[3'-deoxy-3'-[(p-azido-L-phenylalanyl)amino]-beta-D-ribofuranosyl]purine] into specific ribosomal proteins and ribosomal RNA [Nicholson, A. W., Hall, C. C., Strycharz, W. A., & Cooperman, B. S. (1982) Biochemistry (preceding paper in this issue)] is decreased in the presence of puromycin, thus demonstrating that labeling is site specific. The magnitudes of the decreases in incorporation into the major labeled 50S proteins found on addition of different potential ribosome ligands parallel the abilities of these same ligands to inhibit peptidyltransferase. This result provides evidence that p-azidopuromycin photoincorporation into these proteins occurs at the peptidyltransferase center of the 50S subunit, a conclusion supported by other studies of ribosome structure and function. A striking new finding of this work is that puromycin aminonucleoside is a competitive inhibitor of puromycin in peptidyltransferase. The photoincorporation of p-azidopuromycin is accompanied by loss of ribosomal function, but photoincorporated p-azidopuromycin is not a competent peptidyl acceptor. The significance of these results is discussed. Photolabeling of 30S proteins by p-azidopuromycin apparently takes place from sites of lower puromycin affinity than that of the 50S site. The possible relationship of the major proteins labeled, S18, S7, and S14, to tRNA binding is considered.  相似文献   

12.
It was previously shown that virginiamycin M produces in vivo an accumulation of pressure-sensitive (60 S) ribosomes, and in vitro an inactivation of the donor and acceptor sites of peptidyl transferase. The latter action, however, is expected to cause the accumulation in vivo of ribosome complexes carrying acylated tRNA species: such complexes are usually endowed with pressure resistance. However, present data indicate that poly(U).ribosome complexes carrying Phe-tRNA, Ac-Phe-tRNA or Ac-Phe-Phe-tRNA at either the A or the P site become pressure-sensitive after exposure to virginiamycin M in vitro. It is known also that uncoupled EF-G GTPase is stimulated by P-site-bound unacylated tRNA, not by the acylated species. Our data show, however, a stimulation of EF-G GTPase, when ribosomal complexes carrying Ac-Phe-tRNA or Ac-Phe-Phe-tRNA at the P site are incubated with virginiamycin M. The interpretation proposed to account for all these findings is that complexes carrying A- and P-site-bound aminoacyl-tRNA derivatives, which undergo a stable interaction with the peptidyl transferase, are endowed with ultracentrifugal stability, whereas complexes with unacylated tRNA (which does not interact with the enzyme) are pressure-sensitive. By inactivating the donor and acceptor sites of peptidyltransferase, virginiamycin M causes aminoacyl-tRNA.ribosome complexes to mimic tRNA.ribosome complexes in their pressure-lability and competence in EF-G GTPase stimulation. This interpretation is supported by the finding that the ribosome-promoted protection of aminoacyl-tRNA against spontaneous hydrolysis is suppressed by virginiamycin M.  相似文献   

13.
Z Gu  R Harrod  E J Rogers    P S Lovett 《Journal of bacteriology》1994,176(20):6238-6244
Inducible chloramphenicol resistance genes cat and cmlA are regulated by translation attenuation. For both genes, the leader codons that must be translated to deliver a ribosome to the induction site specify a peptide that inhibits peptidyltransferase in vitro. The antipeptidyltransferase activity of the peptides is thought to select the site of ribosome stalling that is essential for induction. Using variations of the cat-86 leader-encoded 5-mer peptide MVKTD, we demonstrate a correlation between the in vitro antipeptidyltransferase activity and the ability of the same peptide to support induction by chloramphenicol in vivo. MVKTD footprints to nucleotides 2058, 2059, and 2060 in 23S rRNA. In vivo methylation of nucleotide 2058 by the ermC methylase interferes neither with cat-86 induction nor with peptide inhibition of peptidyltransferase. The methylation eliminates the competition that normally occurs in vitro between erythromycin and MVKTD. MVKTD inhibits the peptidyltransferase of several eubacteria, a representative Archaea species, and the eukaryote Saccharomyces cerevisiae. Bacillus stearothermophilus supports the in vivo induction of cat-86, and the RNA that is phenol extracted from the 50S ribosomes of this gram-positive thermophile is catalytically active in the peptidyltransferase assay and sensitive to peptide inhibition. Our results indicate that peptidyltransferase inhibition by a cat leader peptide is essential to induction, and this activity can be altered by minor changes in the amino acid sequence of the peptide. The broad range of organisms shown to possess peptide-inhibitable peptidyltransferase suggests that the target is a highly conserved component of the ribosome and includes 23S rRNA.  相似文献   

14.
The puromycin reaction, catalyzed by the ribosomal peptidyltransferase, has been carried out so as to make the definition of two distinct parameters of this reaction possible. These are (a) the final degree of the reaction which gives the proportion of peptidyl (P)-site binding of the donor and (b) the reactivity of the donor substrate expressed as an apparent rate constant (kobs). This kobs varies with the concentration of puromycin; the maximal value (k3) of the kobs, at saturating concentrations of puromycin, gives the reactivity of the donor independently of the concentrations of both the donor and puromycin. k3 is also a measure of the activity of peptidyltransferase expressed as its catalytic rate constant (kcat). If we assume that the puromycin-reactive donor is bound at the ribosomal P site, we observe the following, depending on the conditions of the experiment: the proportion of P-site binding of the donor substrates AcPhe-tRNA or fMet-tRNA can be the same and close to 100%, while there is a tenfold increase in the reactivity of the donor (k3 = 0.8 min-1 versus 8.3 min-1). On the other hand there are conditions, under which the proportion of P-site binding increases from 30% to 100% while k3 remains low and equal to 0.8 min-1. Using the puromycin reaction it was also found that an increase of Mg2+ from 10 mM to 20 mM reduces the reactivity of the donor and, hence, the activity of peptidyltransferase, provided that this change in Mg2+ occurs during the binding of the donor but not when it occurs during peptide bond formation per se. The fact that the donor substrate may exist in various states of reactivity in this cell-free system raises the possibility that the rate of peptide bond formation may not be uniform during protein synthesis.  相似文献   

15.
Four mutant strains from Saccharomyces cerevisiae were used to study ribosome structure and function. They included a strain carrying deletions of the two genes encoding ribosomal protein L24, a strain carrying a mutation spb2 in the gene for ribosomal protein L39, a strain carrying a deletion of the gene for L39, and a mutant lacking both L24 and L39. The mutant lacking only L24 showed just 25% of the normal polyphenylalanine-synthesizing activity followed by a decrease in P-site binding, suggesting the possibility that protein L24 is involved in the kinetics of translation. Each of the two L39 mutants displayed a 4-fold increase of their error frequencies over the wild type. This was accompanied by a substantial increase in A-site binding, typical of error-prone mutants. The absence of L39 also increased sensitivity to paromomycin, decreased the ribosomal subunit ratio, and caused a cold-sensitive phenotype. Mutant cells lacking both ribosomal proteins remained viable. Their ribosomes showed reduced initial rates caused by the absence of L24 but a normal extent of polyphenylalanine synthesis and a substantial in vivo reduction in the amount of 80S ribosomes compared to wild type. Moreover, this mutant displayed decreased translational accuracy, hypersensitivity to the antibiotic paromomycin, and a cold-sensitive phenotype, all caused mainly by the deletion of L39. Protein L39 is the first protein of the 60S ribosomal subunit implicated in translational accuracy.  相似文献   

16.
Two mutants lacking protein L15 from the ribosome as determined by two dimensional gels were investigated using a number of different immunological methods. One strain was found to possess several protein L15 moieties which differed in net charge and in size. The other showed no evidence of L15 cross-reacting material (CRM) on the ribosome or in the supernatant. Ribosomes of this strain were used as a control in the process of the localisation of protein L15 on the surface of the large subunit of Escherichia coli ribosomes. Antigenic determinants mapped in the angle between the central protuberance and the L1 protuberance. Protein L15 has been assigned a central role in the large subunit in vitro assembly map, in peptidyltransferase activity and in the binding of erythromycin, so the significance of a mutant lacking this protein is discussed.  相似文献   

17.
Molecular dynamics simulation identified three highly conserved rRNA bases in the large subunit of the ribosome that form a three-dimensional (3D) "gate" that induces pausing of the aa-tRNA acceptor stem during accommodation into the A-site. A nearby fourth base contacting the "tryptophan finger" of yeast protein L3, which is involved in the coordinating elongation factor recruitment to the ribosome with peptidyltransfer, is also implicated in this process. To better understand the functional importance of these bases, single base substitutions as well as deletions at all four positions were constructed and expressed as the sole forms of ribosomes in yeast Saccharomyces cerevisiae. None of the mutants had strong effects on cell growth, translational fidelity, or on the interactions between ribosomes and tRNAs. However, the mutants did promote strong effects on cell growth in the presence of translational inhibitors, and differences in viability between yeast and Escherichia coli mutants at homologous positions suggest new targets for antibacterial therapeutics. Mutant ribosomes also promoted changes in 25S rRNA structure, all localized to the core of peptidyltransferase center (i.e., the proto-ribosome area). We suggest that a certain degree of structural plasticity is built into the ribosome, enabling it to ensure accurate translation of the genetic code while providing it with the flexibility to adapt and evolve.  相似文献   

18.
During protein synthesis, ribosomes become stalled on polyproline-containing sequences, unless they are rescued in archaea and eukaryotes by the initiation factor 5A (a/eIF-5A) and in bacteria by the homologous protein EF-P. While a structure of EF-P bound to the 70S ribosome exists, structural insight into eIF-5A on the 80S ribosome has been lacking. Here we present a cryo-electron microscopy reconstruction of eIF-5A bound to the yeast 80S ribosome at 3.9 Å resolution. The structure reveals that the unique and functionally essential post-translational hypusine modification reaches toward the peptidyltransferase center of the ribosome, where the hypusine moiety contacts A76 of the CCA-end of the P-site tRNA. These findings would support a model whereby eIF-5A stimulates peptide bond formation on polyproline-stalled ribosomes by stabilizing and orienting the CCA-end of the P-tRNA, rather than by directly contributing to the catalysis.  相似文献   

19.
We have measured the binding isotherms of C--A--C--C--A(3'NH)-[14C]Phe to the 70S ribosomes and 50S subunits of Escherichia coli and proposed a theoretical model for adsorption when cooperative interaction occurs between ligands that are adsorbed on ribosomes. Analysis of the experimental binding isotherms leads to the following conclusions. A ribosome (or subunit) binds two C--A--C--C--A(3'NH)-Phe molecules. The binding of C--A--C--C--A(3'NH)-Phe to a ribosome (or subunit) is a cooperative process, characterized by a cooperativity coefficient tau = 40 +/- 5 or more. The binding of C--A--C--C--A(3'NH)-AcPhe at the donor site of the peptidyltransferase center (association binding constant 1.5 X 10(6) M-1) and the binding of puromycin at the acceptor site also occur cooperatively with a coefficient of 10-25, the association binding constant of puromycin at the acceptor site being (1-2) X 10(4) M-1. The puromycin association binding constant at the donor site multiplied by the cooperativity coefficient of two interacting puromycin molecules absorbed on a ribosome equals 100-200 M-1.  相似文献   

20.
The activity of ribosomes from a clinical isolate of Escherichia coli, exposed to starvation for 7 days in sea salts medium, was investigated by measuring the kinetic parameters of ribosomal peptidyltransferase, by using the puromycin reaction as a model reaction. No alterations in the extent of peptide bond formation were observed during starvation. In contrast, a 50% reduction in the kmax/Ks ratio could be seen after 24 h of starvation; an additional 6 days of starvation resulted in a progressive but less abrupt decline in the kmax/Ks value. {kmax is the apparent catalytic rate constant of peptidyl transferase, and Ks is the dissociation constant of the encounter complex between acetyl (Ac)[3H]Phe-tRNA-poly(U)-ribosome and puromycin.} Although the distribution of ribosomal particles remained constant, a substantial decrease in the number of ribosomes per starved cell and a clear decline in the ability of ribosomes to bind AcPhe-tRNA were observed, particularly during the first day of starvation. Further analysis indicated that rRNA in general, but especially 23S rRNA, was rapidly degraded during the starvation period. In addition, the L12/L7 molar ratio decreased from 1.5 to 1 during the initial phase of starvation (up to 24 h) but remained constant during the subsequent starvation period. Ribosomes isolated from 24-h-starved cells, when artificially depleted of L7/L12 protein and reconstituted with L7/L12 protein from mid-logarithmic-phase cells, regenerated an L12/L7 molar ratio of 1.5 and restored the peptidyltransferase activity to a substantial level. An analogous effect of reconstitution on the efficiency of ribosomes in binding AcPhe-tRNA was evident not only during the initial phase but throughout the starvation period.  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