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.
Tales of poly(A): a review   总被引:31,自引:0,他引:31  
D Munroe  A Jacobson 《Gene》1990,91(2):151-158
  相似文献   

2.
Possible involvement of poly(A) in protein synthesis.   总被引:22,自引:6,他引:16       下载免费PDF全文
The experiments of this paper have re-evaluated the possibility that poly(A) is involved in protein synthesis by testing whether purified poly(A) might competitively inhibit in vitro protein synthesis in rabbit reticulocyte extracts. We have found that poly(A) inhibits the rate of translation of many different poly(A)+ mRNAs and that comparable inhibition is not observed with other ribopolymers. Inhibition by poly(A) preferentially affects the translation of adenylated mRNAs and can be overcome by increased mRNA concentrations or by translating mRNPs instead of mRNA. The extent of inhibition is dependent on the size of the competitor poly(A) as well as on the translation activity which a lysate has for poly(A)+ RNA. In light of our results and numerous experiments in the literature, we propose that poly(A) has a function in protein synthesis and that any role in the determination of mRNA stability is indirect.  相似文献   

3.
In Saccharomyces cerevisiae the function of the RN A1 gene is believed to be required for the transport of newly synthesized mRNA from the nucleus to the cytoplasm. Nuclear poly(A)+ RNAs accumulate and cytoplasmic mRNAs decay after the temperature-sensitive (ts) rna1.1 mutant is shifted from 25 degrees C to 37 degrees C. In this study the 3' poly(A) upon poly(A)+ RNA synthesized after expression of rna1.1 was shown to be appreciably longer than the poly(A) normally present on yeast cytoplasmic mRNA. This increased poly(A) length is due to rna1.1, since it was found only in this mutant after a 25 degrees C to 37 degrees C heat shock, not an intragenic non-ts revertant of rna1.1, wild-type (RN A1+) cells or a RN A1+, rna2.1 mutant subjected to equivalent heat shocks. It may be an indication that the normal shortening of the poly(A) on mRNAs does not occur in the nucleus, but happens only with transport to the cytoplasm. Alterations in the mean size of poly(A) may be a relatively simple marker for mRNA transport defects.  相似文献   

4.
The mRNA species which exist in the HeLa cell polyribisomes in a form devoid of A sequences longer than 8 nucleotides constitute the poly(A)-free class of mRNA. The rapidly labelled component of this mRNA class shares no measurable sequence homology with poly(A)-containing RNA. If poly(A)-free mRNA larger than 12 S labelled for 2 h in vivo is hybridized with total cellular DNA, it hybridizes primarily with single-copy DNA. When a large excess of steady poly(A)-containing RNA is added before hybridization of labelled poly(A)-free RNA, no inhibition of hybridization occurs. This indicates the existence of a class of poly(A)-free mRNA with no poly(A)-containing counterpart. Some mRNA species can exist solely as poly(A)-containing mRNAs. These mRNAs in HeLa cells are found almost exclusively in the mRNA species present only a few times per cell (scarce sequences). Some mRNA species can exist in two forms, poly(A)containing and lacking, as evidenced by the translation data in vitro of Kaufmann et al. [Proc. Natl Acad. Sci. U.S.A. 74, 4801--4805 (1977)]. In addition, if cDNA to total poly(A)-containing mRNA is fractionated into abundant and scarce classes, 47% of the scarce class cDNA can be readily hybridized with poly(A)-free mRNA. 10% of the abundant cDNA to poly(A)-containing mRNA will hybridize with poly(A)-free sequences very rapidly while the other 90% hybridize 160 times more slowly, indicating two very different frequency distributions. The cytoplasmic metabolism of these three distinct mRNA classes is discussed.  相似文献   

5.
Isolation of poly(A)+ RNA by paper affinity chromatography   总被引:16,自引:0,他引:16  
Poly(A)+ RNA was isolated from in vitro short-term-labeled total cytoplasmic RNA of Ehrlich ascites tumor cells by oligo(dT) cellulose chromatography. This poly(A)+ RNA fraction was compared with a poly(A)+ RNA fraction isolated by a new procedure which involves specific binding of poly(A)+ RNA to messenger affinity paper (mAP) and its release in hot (70 degrees C) water. In typical experiments 10-11 micrograms (2.3%) of poly(A)+ RNA can be retained from 500 micrograms of total cytoplasmic RNA per cm2 of mAP in a quick one-step procedure. The poly(A)+ RNA preparations isolated by the two methods proved to be almost identical with respect to their fraction in total cytoplasmic RNA, specific radioactivities, sucrose gradient profiles, and translation assays. Since the isolation of poly(A)+ RNA by mAP is much less time consuming than that by oligo(dT) column chromatography and since the poly(A)+ RNA can be recovered from mAP in small volumes, which avoids further loss during precipitations, it can be advantageously used for preparative isolation of poly(A)+ RNA.  相似文献   

6.
The content, biosynthesis and template activity of poly(A)+ RNA in the early stages of sea urchin development have been studied. The amount of poly(A)+ RNA reaches a maximum at the middle blastula stage in polyribosomes and at the 8-blastomere stage in the cytoplasm. Poly(A)+ RNA synthesis becomes noticeable at the 64-blastomere stage and the spectrum of newly synthesized molecules is different from that at the middle blastula stage. The products of translation in vitro of poly(A)+ RNA at all the stages studied show insignificant differences and contain a major group of polypeptides of molecular mass 10-20 kDa.  相似文献   

7.
Mouse follicles were labeled with [3H]uridine and then cultured in vitro for 3 days. When oocytes were disrupted, about 40% of the total radiolabeled RNA could be sedimented at 9,000g. Fractionation of this RNA on poly(U)-Sepharose revealed that about 30% and 60% of the total amount of radiolabeled poly(A)- and poly(A)+ RNA, respectively, were in the pellet fraction. Treatments that disrupt protein structure reduced the amount of 9,000g sedimentable RNA and affected to the same extent the distribution of Poly(A)- and poly(A)+ RNA in the pellet and supernatant fractions. CsCl centrifugation of formaldehyde-fixed pellets revealed that virtually all of the radiolabeled RNA had a density significantly lower than that of ribosomes. The sedimentable RNA appeared not to be polysomal, membrane bound or associated wih a cytoskeleton. Agarose gel electrophoresis after poly(U)-Sepharose fractionation of either the pellet or supernatant revealed the presence of 28S, 18S, 5S + 4S, and heterodisperse poly(A)+ RNA. The size of distribution of poly(A)+ RNA in the pellet and supernatant fractions was fairly similar. Pulse-chase experiments revealed that the stability of poly(A)- RNA in the pellet and supernatant fractions was the same within the experimental error and a similar situation was found for poly(A)+ RNA. RNA in pellet translated in vitro coded for discrete size classes of protein. Since the relative band intensities were similar for both total and pellet RNA translated in vitro there seemed to be no major partitioning of specific size classes of mRNA into the pellet fraction. These results are discussed in terms of a possible composition of the lattice structures that accumulate during mouse oocyte growth and have been postulated to be a storage form for ribosome (Burkholder et al., '71).  相似文献   

8.
L Gedamu  K Iatrou  G H Dixon 《Cell》1977,10(3):443-451
Poly(A)+ protamine mRNA was isolated from trout testis cells in a very pure form, and artificial poly(A)- protamine mRNA molecules were derived from it by enzymatic deadenylation with RNAase H from calf thymus after hybridization with oligo(dT). The deadenylated protamine mRNA was found to be active in a wheat germ cell-free system and yielded a labeled product which co-migrated with authentic protamine. These deadenylated mRNA molecules were subsequently used as markers on denaturing polyacrylamide gels to identify and allow the purification of the poly(A)- protamine components known to exist in vivo in the total cellular poly(A)- RNA. RNA species of molecular weights similar to the enzymatically deadenylated subcomponents of protamine mRNA were observed in the natural poly(A)-RNA population of the testis cells. These naturally occurring poly(A)- protamine mRNAs were isolated by preparative gel electrophoresis and further characterized by 3H-poly(U) hybridization assay, by hybridization to complementary DNA made against highly purified poly(A)+ protamine mRNA, and by their ability to direct protamine synthesis in a cell-free system.  相似文献   

9.
As an approach to understanding the structures and mechanisms which determine mRNA decay rates, we have cloned and begun to characterize cDNAs which encode mRNAs representative of the stability extremes in the poly(A)+ RNA population of Dictyostelium discoideum amoebae. The cDNA clones were identified in a screening procedure which was based on the occurrence of poly(A) shortening during mRNA aging. mRNA half-lives were determined by hybridization of poly(A)+ RNA, isolated from cells labeled in a 32PO4 pulse-chase, to dots of excess cloned DNA. Individual mRNAs decayed with unique first-order decay rates ranging from 0.9 to 9.6 h, indicating that the complex decay kinetics of total poly(A)+ RNA in D. discoideum amoebae reflect the sum of the decay rates of individual mRNAs. Using specific probes derived from these cDNA clones, we have compared the sizes, extents of ribosome loading, and poly(A) tail lengths of stable, moderately stable, and unstable mRNAs. We found (i) no correlation between mRNA size and decay rate; (ii) no significant difference in the number of ribosomes per unit length of stable versus unstable mRNAs, and (iii) a general inverse relationship between mRNA decay rates and poly(A) tail lengths. Collectively, these observations indicate that mRNA decay in D. discoideum amoebae cannot be explained in terms of random nucleolytic events. The possibility that specific 3'-structural determinants can confer mRNA instability is suggested by a comparison of the labeling and turnover kinetics of different actin mRNAs. A correlation was observed between the steady-state percentage of a given mRNA found in polysomes and its degree of instability; i.e., unstable mRNAs were more efficiently recruited into polysomes than stable mRNAs. Since stable mRNAs are, on average, "older" than unstable mRNAs, this correlation may reflect a translational role for mRNA modifications that change in a time-dependent manner. Our previous studies have demonstrated both a time-dependent shortening and a possible translational role for the 3' poly(A) tracts of mRNA. We suggest, therefore, that the observed differences in the translational efficiency of stable and unstable mRNAs may, in part, be attributable to differences in steady-state poly(A) tail lengths.  相似文献   

10.
Total cellular RNA extracted from Rhodospirillum rubrum cultured in butyrate-containing medium under strict photosynthetic conditions to the stationary phase of growth has been fractionated on an oligodeoxythymidylic acid-cellulose column into polyadenylated [poly(A)+] RNA and poly(A)- RNA fractions. The poly(A)+ fraction was 9 to 10% of the total bulk RNA isolated. Analysis of the poly(A)+ RNA on a denaturing urea-polyacrylamide gel revealed four sharp bands of RNA distributed in heterodisperse fashion between 16S and 9S. Similar fractionation of the poly(A)- RNA resulted in the separation of 23, 16, and 5S rRNAs and 4S tRNA. Poly(A)+ fragments isolated after combined digestion with pancreatic A and T1 RNases and analysis by denaturing gel electrophoresis demonstrated two major components of 80 and 100 residues. Alkaline hydrolysis of the nuclease-resistant, purified residues showed AMP-rich nucleotides. Through the use of snake venom phosphodiesterase, poly(A) tracts were placed at the 3' end of poly(A)+ RNA. Stimulation of [3H]leucine incorporation into hot trichloroacetic acid-precipitable polypeptides in a cell-free system from wheat germ primed by the poly(A)+ RNA mixture was found to be 220-fold higher than that for poly(A)- RNAs (on a unit mass basis), a finding which demonstrated that poly(A)+ RNAs in R. rubrum are mRNAs. Gel electrophoretic analysis of the translation mixture revealed numerous 3H-labeled products including a major band (Mr, 52,000). The parent protein was precipitated by antibodies to ribulose bisphosphate carboxylase-oxygenase and comprised 6.5% of the total translation products.  相似文献   

11.
12.
The two-dimensional gel electrophoresis of polypeptides synthesized in vitro from poly(A)+ RNA showed that mRNA populations change during sporulation of Physarum polycephalum. The differential hybridization of a cDNA library prepared from poly(A)+ RNA isolated from sporulating cells revealed that of 846 clones, 64 corresponded to sporulation-specific mRNAs. Further analysis demonstrated that these clones contained seven different sequences: three abundant sequences composing 3.2, 1.8, and 1.2% of the library and four other less abundant sequences. It is probable that all the major mRNAs specifically expressed in early stages of sporulation were identified. The most abundant mRNA from this group coded for a hydrophobic protein that contained a signal peptide. This protein is 47% similar to another Physarum protein, which was encoded by the most abundant plasmodium-specific mRNA. The plasmodial mRNA was degraded during sporulation and was replaced by the sporulation mRNA. These two proteins are thus encoded by members of a gene family whose expression is developmentally regulated.  相似文献   

13.
14.
Detachment of flagella in Chlamydomonas reinhardii stimulates a rapid accumulation of tubulin mRNAs. The induced tubulin mRNAs are normally rapidly degraded following flagellar regeneration, but inhibition of protein synthesis with cycloheximide prevents their degradation. alpha-Tubulin poly(A) tail lengths were measured during normal accumulation and degradation, and in cycloheximide-treated cells. To measure alpha-tubulin mRNA poly(A) chain lengths with high resolution, specific 3' fragments of alpha 1- and alpha 2-tubulin mRNAs, generated by RNase H digestion of mRNA-oligonucleotide hybrids, were sized by Northern analysis. Both alpha-tubulin mRNAs have a newly synthesized poly(A) chain of about 110 adenylate residues. The poly(A) tails shorten with time, and show an average length of 40 to 60 adenylate residues by 90 minutes after deflagellation, at which time induced alpha-tubulin mRNA is being rapidly degraded. Poly(A) loss is significantly accelerated in cycloheximide-treated cells, and this loss is not attributible simply to the longer time the stabilized molecules spend in the cytoplasm. A large fraction of alpha-tubulin mRNA accumulates as mRNA with very short poly(A) tails (less than 10 residues) in the presence of cycloheximide, indicating that deadenylated alpha-tubulin mRNAs can be stable in vivo, at least in the absence of protein synthesis. The rate and extent of poly(A) loss in cycloheximide are greater for alpha 2-tubulin mRNA than for alpha 1-tubulin mRNA. This difference cannot be attributed to differential ribosome loading. This finding is interesting in that the two mRNAs are very similar in sequence with the exception of their 3' untranslated regions.  相似文献   

15.
A variety of rapidly growing mammalian cells contain a substantial portion of their actin mRNA in a poly(A)- form. We have used DNA-driven hybridization of a cloned actin cDNA-containing plasmid with pulse-labeled RNA from mouse S-180 ascites cells to examine newly synthesized actin mRNA. Our results indicate that the same proportion of newly synthesized and steady-state actin mRNA (approx. 40%) exists in a poly(A)- deficient form. This suggests that the poly(A)- form arises by some process other than slow cytoplasmic de-adenylation of a poly(A)+ precursor. We have also examined cell cycle-enriched populations of S-180 ascites cells for the presence of poly(A)- actin mRNA. Results from these experiments indicate that cells in G1 phase of the cell cycle contain predominantly poly(A)+ actin mRNA, while the poly(A)- form is restricted to late-S and post-S phase cells.  相似文献   

16.
Pollination induces many physiological responses in the flower, including deterioration and death in specific pistil cell types. It is shown here that within the style of tobacco, pollination-induced cell deterioration was restricted to the transmitting tissue while the surrounding cortical tissue was not affected. It was distinct from general senescence since exogenously applying the senescence-inducing hormone ethylene, or its precursor aminocyclopropane-1-carboxylic acid (ACC), to the flower or the pistil induced overall deterioration in the entire flower. Furthermore, both pollen tube growth and ethylene action were needed for the entire spectrum of cellular changes associated with this pollination-induced transmitting tissue deterioration process. It is also shown that pollination-induced mRNA poly(A) tail-shortening for at least three major classes of transmitting tissue-specific mRNAs. As is commonly observed for poly(A) tail-shortened mRNAs, the levels of two of these three mRNA classes decline after pollination. On the other hand, the third class of mRNAs, transmitting tissue-specific (TTs) mRNAs, was maintained at a very high level subsequent to pollination, even after substantial poly(A)-tail shortening. TTS mRNAs encode a pollen tube growth-promoting and -attracting protein needed for optimal in vivo pollen tube growth. The specific preservation of TTS mRNAs in the deteriorating transmitting tissue cells suggests that these cells can distinguish molecules needed in the pollinated styles from those that are dispensable, and protect them from degradation. It is suggested that the pollination-induced mRNA poly(A) tail-shortening and cell death are programmed processes suited to the post-pollination transmitting tissue environment. Results showing that ACC is a candidate signal molecule for the pollination-induced mRNA-shortening which is accentuated by ethylene and mediated via a protein phosphorylation-dependent signal transduction pathway are also presented.  相似文献   

17.
Total cellular poly(A+)-RNA was isolated from a lactating cow mammary gland. The poly(A+)-RNA molecules exhibit a heterogeneous distribution from 500 to 5000 nucleotides (average size--1600 nucleotides) and are made up of three main fractions (1550, 950 and 600 nucleotides) possessing a high template activity during translation in vitro. Optimal conditions for poly(A+)-RNA translation in a cell-free protein-synthesizing system from wheat embryos were elaborated. Immunochemical analysis of translation products revealed that 30% of the synthesized polypeptides are precipitated by immunoglobulins against cow milk proteins. Using hybridization with homologous cDNA, the kinetic complexity and heterogeneity of total cellular poly(A+)-RNA were investigated. This population was shown to consist of four classes differing in the diversity of their nucleotide sequences and the number of copies per cell. The total amount of the poly(A+)-RNA species in the cells of a lactating cow mammary gland is 9200, i.e., 0.46% of the genome complexity.  相似文献   

18.
The complexity of rat liver poly (A)+ messenger RNA (mRNA) has been measured by analysis of the kinetics of hydridization with both complementary DNA (cDNA) and single copy DNA. The complementary DNA-poly(A)+ mRNA hybridization reaction demonstrates the existence of three abundance classes representing 18, 37, and 45% of the cDNA and 4, 290, and 24 000 different 1800-nucleotide sequences respectively. The poly(A)+ mRNA driven single copy DNA hybridization reaction reveals a single major transition accounting for 1.9% of the haploid rat genome. The kinetics of the poly(A)+ mRNA driven single copy DNA reaction suggest that approximately 45% of the mass of the mRNA population contains over 95% of the complexity. Although higher than previous estimates, the base sequence complexities of rat liver poly(A)+ mRNA measured in these two ways are in good agreement, suggesting that the technique of poly(A)+ mRNA-cDNA hybridization may be used in approximating the complexity as well as abundance of a messenger RNA population. DNA-driven cDNA reactions reveal that about 10% of rat liver poly(A)+ mRNA is transcribed from repetitive sequences in the rat genome.  相似文献   

19.
The steady state changes in total rat hepatic cytoplasmic RNA, poly(A)+ RNA and poly(A)-RNA were assessed in response to turpentine induced inflammation. From 18 to 24 h after injury, cytoplasmic RNA doubled, while poly(A)+ RNA peaked at 24 h, 3.5 times over control animals. Cell-free translation showed significant increases in messenger RNA levels beginning at 18 h. Gel electrophoresis of translation products revealed significant increases in several polypeptides and a decrease in others. Poly(A)-RNA from control and injured rats translated to an insignificant level and the electrophoretic gel patterns of their proteins were similar. Furthermore, no change had occurred in the 3' poly(A)-sequences during the course of inflammation.  相似文献   

20.
The poly(A) content of early mouse embryos fluctuates widely: after a transient increase in the one-cell embryo, there is a 70% drop in the two-cell and an approximately fivefold increase between the two-cell and early blastocyst stages (L. Pikó and K. B. Clegg, 1982, Dev. Biol.89, 362–378). To shed light on the significance of these changes, we analyzed the size distribution of total poly(A) from embryos at different stages of development by gel electrophoresis and hybridization with [3H]poly(U). The number-average size of poly(A) tracts varies only slightly, from 61 to 77 nucleotides, indicating that the changes in poly(A) content are due primarily to changes in the number of poly(A) sequences, i.e., the number of poly(A)+ mRNA. From these data, the number of poly(A)+ mRNA can be estimated as follows: ovulated egg, 1.7 × 107; one-cell embryo, 2.4 × 107; late two-cell, 0.7 × 107; late eight-cell, 1.3 × 107; and early blastocyst, 3.4 × 107. These results suggest the elimination of the bulk of maternal poly(A)+ mRNA at the two-cell stage, to be replaced by newly synthesized mRNA derived from the embryonic genome. To study the synthesis of poly(A)+ mRNA, we cultured mouse embryos in vitro with [3H]adenosine and analyzed the labeled poly(A)+ RNA as to molecular size, length of the poly(A) tail, and relative distribution of label in poly(A) vs internal locations. We observed an active incorporation of label into large-molecular-weight (average size about 2 kb) poly(A)+ RNA at all stages from the one-cell to the blastocyst. However, in the one-cell embryo, about 70% of the label was localized in the poly(A) tail, suggesting cytoplasmic polyadenylation, and only about 30% was localized in the remainder of the molecule, suggesting the complete new synthesis of a small amount of poly(A)+ RNA. Differences in the size distribution of the labeled poly(A) as compared with the total poly(A) in the one-cell embryo indicate that the labeling is not due to a general turnover of poly(A) tails, but rather to the polyadenylation of previously nonpolyadenylated, stored RNA. Significant new synthesis of poly(A)+ RNA is evident from the two-cell stage onward and most likely accounts for the sharp rise in the number of poly(A)+ RNA molecules by the early blastocyst stage.  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