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 46 毫秒
1.
Chloramphenicol is thought to interfere competitively with the binding of the aminoacyl-tRNA 3′-terminus to ribosomal A-site. However, noncompetitive or mixed-noncompetitive inhibition, often observed to be dependent on chloramphenicol concentration and ionic conditions, leaves some doubt about the precise mode of action. Here, we examine further the inhibition effect of chloramphenicol, using a model system derived from Escherichia coli in which a peptide bond is formed between puromycin and AcPhe-tRNA bound at the P-site of poly(U)-programmed ribosomes, under ionic conditions (6 mM Mg2+, 100 mM NH4+, 100 µM spermine) more closely resembling the physiological status. Kinetics reveal that chloramphenicol (I) reacts rapidly with AcPhe-tRNA·poly(U)·70S ribosomal complex (C) to form the encounter complex CI which is then isomerized slowly to a more tight complex, C*I. A similar inhibition pattern is observed, if complex C modified by a photoreactive analogue of spermine, reacts in buffer free of spermine. Spermine, either reversibly interacting with or covalently attached to ribosomes, enhances the peptidyltransferase activity and increases the chloramphenicol potency, without affecting the isomerization step. As indicated by photoaffinity labeling, the peptidyltransferase center at which chloramphenicol binds, is one of the preferred cross-linking sites for polyamines. This fact may explain the effect of spermine on chloramphenicol binding to ribosomes.  相似文献   

2.
A photoreactive analogue of spermine, N1-azidobenzamidino (ABA)-spermine, was covalently attached after irradiation to Escherichia coli 30S ribosomal subunits or naked 16S rRNA. By means of RNase H digestion and primer extension, the cross-linking sites of ABA-spermine in naked 16S rRNA were characterised and compared with those identified in 30S subunits. The 5′ domain, the internal and terminal loops of helix H24, as well as the upper part of helix H44 in naked 16S rRNA, were found to be preferable binding sites for polyamines. Association of 16S rRNA with ribosomal proteins facilitated its interaction with photoprobe, except for 530 stem–loop nt, whose modification by ABA-spermine was abolished. Association of 30S with 50S subunits, poly(U) and AcPhe-tRNA (complex C) further altered the susceptibility of ABA-spermine cross-linking to 16S rRNA. Complex C, modified in its 30S subunit by ABA-spermine, reacted with puromycin similarly to non-photolabelled complex. On the contrary, poly(U)-programmed 70S ribosomes reconstituted from photolabelled 30S subunits and untreated 50S subunits bound AcPhe-tRNA more efficiently than untreated ribosomes, but were less able to recognise and reject near cognate aminoacyl-tRNA. The above can be interpreted in terms of conformational changes in 16S rRNA, induced by the incorporation of ABA-spermine.  相似文献   

3.
Polyamine binding to 23S rRNA was investigated, using a photoaffinity labeling approach. This was based on the covalent binding of a photoreactive analog of spermine, N1-azidobenzamidino (ABA)-spermine, to Escherichia coli ribosomes or naked 23S rRNA under mild irradiation conditions. The cross-linking sites of ABA-spermine in 23S rRNA were determined by RNase H digestion and primer-extension analysis. Domains I, II, IV and V in naked 23S rRNA were identified as discrete regions of preferred cross-linking. When 50S ribosomal subunits were targeted, the interaction of the photoprobe with the above 23S rRNA domains was elevated, except for helix H38 in domain II whose susceptibility to cross-linking was greatly reduced. In addition, cross-linking sites were identified in domains III and VI. Association of 30S with 50S subunits, poly(U), tRNAPhe and AcPhe-tRNA to form a post-translocation complex further altered the cross-linking, in particular to helices H11–H13, H21, H63, H80, H84, H90 and H97. Poly(U)-programmed 70S ribosomes, reconstituted from photolabeled 50S subunits and untreated 30S subunits, bound AcPhe-tRNA in a similar fashion to native ribosomes. However, they exhibited higher reactivity toward puromycin and enhanced tRNA-translocation efficiency. These results suggest an essential role for polyamines in the structural and functional integrity of the large ribosomal subunit.  相似文献   

4.
5S rRNA is an integral component of the large ribosomal subunit in virtually all living organisms. Polyamine binding to 5S rRNA was investigated by cross-linking of N1-azidobenzamidino (ABA)-spermine to naked 5S rRNA or 50S ribosomal subunits and whole ribosomes from Escherichia coli cells. ABA-spermine cross-linking sites were kinetically measured and their positions in 5S rRNA were localized by primer extension analysis. Helices III and V, and loops A, C, D and E in naked 5S rRNA were found to be preferred polyamine binding sites. When 50S ribosomal subunits or poly(U)-programmed 70S ribosomes bearing tRNAPhe at the E-site and AcPhe-tRNA at the P-site were targeted, the susceptibility of 5S rRNA to ABA-spermine was greatly reduced. Regardless of 5S rRNA assembly status, binding of spermine induced significant changes in the 5S rRNA conformation; loop A adopted an apparent ‘loosening’ of its structure, while loops C, D, E and helices III and V achieved a more compact folding. Poly(U)-programmed 70S ribosomes possessing 5S rRNA cross-linked with spermine were more efficient than control ribosomes in tRNA binding, peptidyl transferase activity and translocation. Our results support the notion that 5S rRNA serves as a signal transducer between regions of 23S rRNA responsible for principal ribosomal functions.  相似文献   

5.
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   

6.
The effect of two photoreactive analogues of spermine, N(1)-azidobenzamidino- (ABA-) spermine and N(1)-azidonitrobenzoyl- (ANB-) spermine, on ribosomal functions was studied in a cell-free system derived from Escherichia coli. In the dark, both analogues stimulated the binding of AcPhe-tRNA to poly(U)-programmed ribosomes, enhanced the stability of the ternary complex AcPhe-tRNA.poly(U).ribosome (complex C), and caused stimulatory and inhibitory effects on peptidyltransferase activity. ABA-spermine exhibited more pronounced effects than ANB-spermine. Each photoprobe was covalently attached after irradiation to both ribosomal subunits and also to free rRNA isolated from 70S ribosomes. Photolabeled complex C showed a reactivity toward puromycin, similar to that exhibited by complex C reacting reversibly with photoprobes free in solution. The distribution of the incorporated radioactivity among the ribosomal components was determined under two experimental conditions, one stimulating and the other inhibiting peptidyltransferase activity. Under both conditions, ABA-spermine was the strongest cross-linker. Upon stimulatory conditions, 14% of ABA-[(14)C]spermine cross-linked to complex C was bound to the protein fraction. The proteins primarily labeled were identified as S3, S4, L2, L3, L6, L15, L17, and L18. Upon inhibitory conditions, a higher percent of the incorporated radioactivity was found in ribosomal proteins, while the pattern of protein labeling was characterized by a remarkable decrease of cross-linked proteins L2, L3, L6, L15, L17. and L18 and by an increase of cross-linked proteins S9, S18, L1, L16, L22, L23, and L27. On the basis of these results and literature data, the involvement of spermine in the conformation and important functions of ribosomes is discussed.  相似文献   

7.
The effect of spermine on the binding of AcPhe-tRNA to poly(U)-programmed ribosomes (step 1) and on the puromycin reaction (step 2) has been studied in a cell-free system, derived from E. coli.In the absence of ribosomal wash (FWR fraction) and at suboptimal concentration of Mg++ (6 mM), spermine stimulated the binding of AcPhe-tRNA at least five fold, while at 10 mM Mg++ there was a three fold stimulation. The above stimulatory effect was decreased at 6 mM Mg++, or was abolished at 10 mM Mg++ by the presence of FWR during the binding. Beside the stimulatory effect, spermine enhanced the stability of initiation complex AcPhe-tRNA-poly(U)-ribosome.In step 2, spermine affected the final degree of puromycin reaction and the activity status of peptidyltransferase. Both stimulatory and inhibitory effects have been observed, depending on the experimental conditions followed during the binding of the donor and during the peptide bond formation.  相似文献   

8.
Poly(U)-programmed 70S ribosomes can be shown to be 80% to 100% active in binding the peptidyl-tRNA analogue AcPhe-tRNA to their A or P sites, respectively. Despite this fact, only a fraction of such ribosomes primed with AcPhe-tRNA participate in poly(U)-directed poly(Phe) synthesis (up to 65%) at 14 mM Mg2+ and 160 mM NH4+. Here it is demonstrated that the apparently 'inactive' ribosomes (greater than or equal to 35%) are able to participate in peptide-bond formation, but lose their nascent peptidyl-tRNA at the stage of Ac(Phe)n-tRNA, with n greater than or equal to 2. The relative loss of early peptidyl-tRNAs is largely independent of the degree of initial saturation with AcPhe-tRNA and is observed in a poly(A) system as well. This observation resolves a current controversy concerning the active fraction of ribosomes. The loss of Ac(Phe)n-tRNA is reduced but still significant if more physiological conditions for Ac(Phe)n synthesis are applied (3 mM Mg2+, 150 mM NH4+, 2 mM spermidine, 0.05 mM spermine). Chloramphenicol (0.1 mM) blocks the puromycin reaction with AcPhe-tRNA as expected but, surprisingly, does not affect the puromycin reaction with Ac(Phe)2-tRNA nor peptide bond formation between AcPhe-tRNA and Phe-tRNA. The drug facilitates the release of Ac(Phe)2-4-tRNA from ribosomes at 14 mM Mg2+ while it hardly affects the overall synthesis of poly(Phe) or poly(Lys).  相似文献   

9.
The effects of spermine on peptidyltransferase inhibition by an aminohexosylcytosine nucleoside, blasticidin S, and by a macrolide, spiramycin, were investigated in a model system derived from Escherichia coli, in which a peptide bond is formed between puromycin and AcPhe-tRNA bound at the P-site of poly(U)-programmed ribosomes. Kinetics revealed that blasticidin S, after a transient phase of interference with the A-site, is slowly accommodated near to the P-site so that peptide bond is still formed but with a lower catalytic rate constant. At high concentrations of blasticidin S (>10 x K(i)), a second drug molecule binds to a weaker binding site on ribosomes, and this may account for the onset of a subsequent mixed-noncompetitive inhibition phase. Spermine enhances the blasticidin S inhibitory effect by facilitating the drug accommodation to both sites. On the other hand, spiramycin (A) was found competing with puromycin for the A-site of AcPhe-tRNA.poly(U).70 S ribosomal complex (C) via a two-step mechanism, according to which the fast formation of the encounter complex CA is followed by a slow isomerization to a tighter complex, termed C(*)A. In contrast to that observed with blasticidin S, spermine reduced spiramycin potency by decreasing the formation and stability of complex C(*)A. Polyamine effects on drug binding were more pronounced when a mixture of spermine and spermidine was used, instead of spermine alone. Our kinetic results correlate well with cross-linking and crystallographic data and suggest that polyamines bound at the vicinity of the antibiotic binding pockets modulate diversely the interaction of these drugs with ribosomes.  相似文献   

10.
The aim of the present study was to investigate homoharringtonine alkaloid effect on: (i) the nonenzymatic and eEF-1-dependent Phe-tRNAPhe binding to poly(U)-programmed human placenta 80 S ribosomes; (ii) diphenylalanine synthesis accompanying nonenzymatic Phe-tRNAPhe binding; and (iii) acetylphenylalanyl-puromycin formation. Neither nonenzymatic nor eEF-1-dependent Phe-tRNAPhe binding were noticeably affected by the alkaloid, whereas diphenylalanine synthesis and puromycin reaction were strongly inhibited by homoharringtonine. It has been proposed that the site of homoharringtonine binding on 80 S ribosomes should overlap or coincide with the acceptor site of the ribosome.  相似文献   

11.
Association constants for tRNA binding to poly(U) programmed ribosomes were assessed under standardized conditions with a single preparation of ribosomes, tRNAs, and elongation factors, respectively, at 15 and 10 mM Mg2+. Association constants were determined by Scatchard plot analysis (the constants are given in units of [10(7)/M] measured at 15 mM Mg2+): the ternary complex Phe-tRNA.elongation factor EF-Tu.GTP (12 +/- 3), Phe-tRNA (1 +/- 0.4), AcPhe-tRNA (0.7 +/- 0.3), and deacylated tRNA(Phe) (0.4 +/- 0.15) bind with decreasing affinity to the A site of poly(U)-programmed ribosomes. tRNA(Phe) (7.2 +/- 0.8) binds to the P site with higher affinity than AcPhe-tRNA (3.7 +/- 1.3). The affinity of the E site for deacylated tRNA(Phe) (1 +/- 0.2) is about the same as that of the A site for AcPhe-tRNA (0.7 +/- 0.3). At lower Mg2+ concentrations the affinity of the E site ligand becomes stronger relative to the affinities of the A site ligands. Phe-tRNA and ternary complexes can occupy the A site at 0 degrees C in the presence of poly(U) even if the P site is free, whereas, as already known, deacylated tRNA or AcPhe-tRNA bind first to the P site of programmed ribosomes. Hill plot analyses of the binding data confirm an allosteric linkage between A and E sites in the sense of a negative cooperativity.  相似文献   

12.
Binding of yeast tRNAPhe anticodon arm to Escherichia coli 30 S ribosomes   总被引:7,自引:0,他引:7  
A 15-nucleotide fragment of RNA having the sequence of the anticodon arm of yeast tRNAPhe was constructed using T4 RNA ligase. The stoichiometry and binding constant of this oligomer to poly(U)-programmed 30 S ribosomes was found to be identical to that of deacylated tRNAPhe. The anticodon arm and tRNAPhe also compete for the same binding site on the ribosome. These data indicate that the interaction of tRNAPhe with poly(U)-programmed 30 S ribosomes is primarily a result of contacts in the anticodon arm region and not with other parts of the transfer RNA. Since similar oligomers which cannot form a stable helical stem do not bind ribosomes, a clear requirement for the entire anticodon arm structure is demonstrated.  相似文献   

13.
The standard technique for determination of the ribosomal site location of bound tRNA, viz. the puromycin reaction, has been analyzed with regard to its applicability under tRNA saturation conditions. The criteria derived have been used to re-examine the exclusion principle for peptidyl-tRNA binding, which states that only one peptidyl-tRNA (AcPhe-tRNA) can be bound per ribosome although in principle two sites (A and P site) are available. The following results were obtained. The puromycin reaction is only appropriate for a site determination if the reaction conditions prevent one ribosome from performing more than one puromycin reaction. With an excess of AcPhe-tRNA over ribosomes, and in the absence of EF-G, this criterion is fulfilled at 0 degree C, where the P-site-bound material reacts with puromycin (quantitative reaction after 50 h), while the A-site-bound material does not. In contrast, at 37 degrees C the extent of the puromycin reaction can exceed the binding values by 2-4-fold ('repetitive reaction'). In the presence of EF-G a repetitive puromycin reaction is seen even at 0 degree C, i.e. EF-G can already promote a translocation reaction at 0 degree C. However, the extent of translocation becomes negligibly low for short incubation times (up to 60 min) at 0 degree C, if only catalytic amounts of EF-G are used. Using the criteria outlined above, the validity of the exclusion principle for Escherichia coli ribosomes was confirmed pursuing two different experimental strategies. Ribosomes were saturated with AcPhe-tRNA at one molecule per 70S ribosome, and a quantitative puromycin reaction demonstrated the exclusive P-site location of the AcPhe-tRNA. The same result was also found in the presence of viomycin, which blocks the translocation reaction. These findings also indicate that here nearly 100% of the ribosomes participate in AcPhe-tRNA binding to the P site. Precharging the P sites of 70S ribosomes with one Ac[14C]Phe-tRNA molecule per ribosome prevented additional Ac[3H]Phe-tRNA binding. In contrast, 70S particles carrying one molecule of [14C]tRNAPhe per ribosome were able to bind up to a further 0.64 molecule Ac[3H]Phe-tRNA per ribosome.  相似文献   

14.
Arginine inhibits the formation of acetylleucyl-puromycin from C(U)-A-C-C-A-LeuAc and puromycin ('fragment reaction'), catalized by Escherichia coli and yeast ribosomes. From 18 different L-amino acids assayed, arginine was the most effective in producing inhibition (50% inhibition at 20 mM, with 1 mM puromycin). L-Argininamide and D-arginine gave about the same inhibition as L-arginine. The inhibition by L-arginine is competitive with respect to puromycin. The plot of the slopes obtained in a Lineweaver and Burk representation versus [Arg]2, and the plot of 1/v versus [Arg]2 at a fixed concentration of puromycin, are linear, which seems to indicate that two arginine molecules must interact at the puromycin binding site to produce inhibition. In addition to the 'fragment reaction', arginine inhibits the non-enzymatic binding of AcPhe-tRNA, C(U)-A-C-C-A-Leu and C(U)-A-C-C-A-LeuAc to ribosomes. However, it does not inhibit poly(U)-directed polyphenylalanine synthesis or the reaction of puromycin with AcPhe-tRNA previously bound to the peptidyl site. The results agree with arginine binding to the acceptor site, and with a sequential mechanism for the 'fragment reaction', puromycin binding first.  相似文献   

15.
Kinetic analyses of tRNA binding to the ribosome and of the translocation reaction showed the following results. 1) The activation energy for the P site binding of AcPhe-tRNA to poly(U)-programmed ribosomes is relatively high (Ea = 72 kJ mol-1; 15 mM Mg2+). If only the P site is occupied with deacylated tRNA(Phe), then the E site can be filled more easily with tRNA(Phe) (no activation energy measurable) than the A site with AcPhe-tRNA (Ea = 47 kJ mol-1; 15 mM Mg2+). 2) A ribosome with blocked P and E sites represents a standard state of the elongation cycle, in contrast to a ribosome with only a filled P site. The two states differ in that AcPhe-tRNA binding to the A site of a ribosome with prefilled P and E sites requires much higher activation energy (87 versus 47 kJ mol-1). The latter reaction simulates the allosteric transition from the post- to the pretranslocational state, whereby the tRNA(Phe) is released from the E site upon occupation of the A site (Rheinberger, H.-J., and Nierhaus, K. H. (1986) J. Biol. Chem. 261, 9133-9139). The reversed transition from the pre- to the posttranslocational state (translocation reaction) requires about the same activation energy (90 kJ mol-1). 3) Both elongation factors EF-Tu and EF-G drastically reduce the respective activation energies. 4) The rate of the A site occupation is slower than the rate of translocation in the presence of the respective elongation factors. The data suggest that the A site occupation rather than, as generally assumed, the translocation reaction is the rate-limiting step of the elongation cycle.  相似文献   

16.
Puromycin inhibits the interaction of peptidyl-tRNA analogs AcPhe-tRNA Phe ox-red, AcPhe-tRNA Phe and FMet-tRNA f Met with the donor (P) site of Escherichia coli ribosomes. It affects both template-free and poly(U)-dependent systems. The inhibition is apparently due to direct competition for the P-site. On isolated 30S ribosomal subunits it was shown that the puromycin binding site is situated far from the peptidyl transferase center. Quantitative measurements of the inhibition revealed that the affinity constant of puromycin for the P-site is not less than its affinity for the A-moiety of the peptidyl transferase center [1.1 divided by 3.8) X 10(3) M-1).  相似文献   

17.
Plautia stali intestine virus (PSIV) has an internal ribosome entry site (IRES) at the intergenic region of the genome. The PSIV IRES initiates translation with glutamine rather than the universal methionine. To analyze the mechanism of IRES-mediated initiation, binding of IRES RNA to salt-washed ribosomes in the absence of translation factors was studied. Among the three pseudoknots (PKs I, II and III) within the IRES, PK III was the most important for ribosome binding. Chemical footprint analyses showed that the loop parts of the two stem–loop structures in Domain 2, which are highly conserved in related viruses, are protected by 40S but not by 60S ribosomes. Because PK III is close to the two loops, these structural elements were considered to be important for binding of the 40S subunit. Competitive binding analyses showed that the IRES RNA does not bind poly(U)-programmed ribosomes preincubated with tRNAPhe or its anticodon stem– loop (ASL) fragment. However, Domain 3-deleted IRES bound to programmed ribosomes preincubated with the ASL, suggesting that Domains 1 and 2 have roles in IRES binding to 40S subunits and that Domain 3 is located at the ribosome decoding site.  相似文献   

18.
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   

19.
Hygromycin A, a novel inhibitor of ribosomal peptidyltransferase   总被引:1,自引:0,他引:1  
In cell-free systems from Escherichia coli, hygromycin A inhibits polypeptide synthesis directed by either poly(U) or phage R 17 RNA, and the reaction of puromycin with either natural peptidyl-tRNA, or AcPhe-tRNA, or the 3'-terminal fragment of AcLeu-tRNA (C-A-C-C-A-LeuAc). In contrast, the antibiotic does no inhibit the enzymatic binding of Phe-tRNA to ribosomes or the translocation of AcPhe-tRNA. It is concluded that hygromycin A is a specific inhibitor of the peptide bond formation step of protein synthesis. The action of hygromycin A on peptidyl transfer is similar to that of chloramphenicol, an antibiotic that shares some common structural features with hygromycin A. Both antibiotics inhibit the binding of C-A-C-C-A-Leu to the acceptor site of peptidyl transferase and stimulate that of C-A-C-C-A-LeuAc to the donor site of the enzyme. Moreover, hygromycin A blocks the binding of chloramphenicol to ribosomes, indicating that the binding sites of the antibiotics may be closely related. Hygromycin A is a more potent agent than chloramphenicol and binds quite strongly to ribosomes.  相似文献   

20.
In a medium of high ionic strength, rat liver rough microsomes can be nondestructively disassembled into ribosomes and stripped membranes if nascent polypeptides are discharged from the bound ribosomes by reaction with puromycin. At 750 mM KCl, 5 mM MgCl2, 50 mM Tris·HCl, pH 7 5, up to 85% of all bound ribosomes are released from the membranes after incubation at room temperature with 1 mM puromycin. The ribosomes are released as subunits which are active in peptide synthesis if programmed with polyuridylic acid. The ribosome-denuded, or stripped, rough microsomes (RM) can be recovered as intact, essentially unaltered membranous vesicles Judging from the incorporation of [3H]puromycin into hot acid-insoluble material and from the release of [3H]leucine-labeled nascent polypeptide chains from bound ribosomes, puromycin coupling occurs almost as well at low (25–100 mM) as at high (500–1000 mM) KCl concentrations. Since puromycin-dependent ribosome release only occurs at high ionic strength, it appears that ribosomes are bound to membranes via two types of interactions: a direct one between the membrane and the large ribosomal subunit (labile at high KCl concentration) and an indirect one in which the nascent chain anchors the ribosome to the membrane (puromycin labile). The nascent chains of ribosomes specifically released by puromycin remain tightly associated with the stripped membranes. Some membrane-bound ribosomes (up to 40%) can be nondestructively released in high ionic strength media without puromycin; these appear to consist of a mixture of inactive ribosomes and ribosomes containing relatively short nascent chains. A fraction (~15%) of the bound ribosomes can only be released from membranes by exposure of RM to ionic conditions which cause extensive unfolding of ribosomal subunits, the nature and significance of these ribosomes is not clear.  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