scispace - formally typeset
Search or ask a question

Showing papers on "RNA-dependent RNA polymerase published in 1985"


Journal ArticleDOI
TL;DR: A coupled system that permits the exclusive expression of genes under the control of a T7 RNA polymerase promoter and its use to express high levels of phage T7 gene 5 protein, a subunit of T7 DNA polymerase is described.
Abstract: The RNA polymerase gene of bacteriophage T7 has been cloned into the plasmid pBR322 under the inducible control of the lambda PL promoter. After induction, T7 RNA polymerase constitutes 20% of the soluble protein of Escherichia coli, a 200-fold increase over levels found in T7-infected cells. The overproduced enzyme has been purified to homogeneity. During extraction the enzyme is sensitive to a specific proteolysis, a reaction that can be prevented by a modification of lysis conditions. The specificity of T7 RNA polymerase for its own promoters, combined with the ability to inhibit selectively the host RNA polymerase with rifampicin, permits the exclusive expression of genes under the control of a T7 RNA polymerase promoter. We describe such a coupled system and its use to express high levels of phage T7 gene 5 protein, a subunit of T7 DNA polymerase.

3,214 citations


Journal ArticleDOI
01 Sep 1985-Cell
TL;DR: The nucleotide sequence of two yeast RNA polymerase genes, RPO21 and RPO31, which encode the largest subunits of RNA polymerases II and III, respectively are determined.

612 citations


Journal ArticleDOI
TL;DR: The in vitro synthesis of extraneous RNA sequences by SP6 and T7 RNA polymerases from specific DNA templates is described and sequences copied from the noncoding template strand were among the extraneous transcripts.
Abstract: The in vitro synthesis of extraneous RNA sequences by SP6 and T7 RNA polymerases from specific DNA templates is described. Transcription of templates prepared by digestion with restriction enzymes that leave 3' protruding ends resulted in the production of significant amounts of long, template-sized RNA transcripts which hybridized to vector DNA. Sequences copied from the noncoding template strand were among the extraneous transcripts. The presence of these sequences in probe preparations were detected in Southern and RNase protection hybridization assays. In contrast, transcription of DNA templates with blunt or 5' protruding ends yielded few RNA products as extraneous sequences.

303 citations


Journal ArticleDOI
03 Jan 1985-Nature
TL;DR: Evidence is provided that subgenomic RNA arises by internal initiation on the (−)-strand of genomic RNA, and it is believed that this also represents the first in vitro demonstration of a replicase from a eukaryotic (+)-stranded RNA virus capable of initiating synthesis of (+)-sense RNA.
Abstract: The genomes of many (+)-stranded RNA viruses, including plant viruses and alphaviruses, consist of polycistronic RNAs whose internal genes are expressed via subgenomic messenger RNAs. The mechanism(s) by which these subgenomic mRNAs arise are poorly understood. Based on indirect evidence, three models have been proposed: (1) internal initiation by the replicase on the (-)-strand of genomic RNA, (2) premature termination during (-)-strand synthesis, followed by independent replication of the subgenomic RNA and (3) processing by nuclease cleavage of genome-length RNA. Using an RNA-dependent RNA polymerase (replicase) preparation from barley leaves infected with brome mosaic virus (BMV) to synthesize the viral subgenomic RNA in vitro, we now provide evidence that subgenomic RNA arises by internal initiation on the (-)-strand of genomic RNA. We believe that this also represents the first in vitro demonstration of a replicase from a eukaryotic (+)-stranded RNA virus capable of initiating synthesis of (+)-sense RNA.

280 citations


Journal ArticleDOI
TL;DR: The data suggest that antisense RNA will be useful in identifying the function of specific mRNA sequences during early development of the frog.
Abstract: The bacteriophage SP6 promoter and RNA polymerase were used to synthesize sense and antisense RNAs coding for the enzymes thymidine kinase (TK) and chloramphenicol acetyl transferase (CAT). Injection of antisense CAT RNA into frog oocytes inhibited expression of sense CAT mRNA. Similarly, antisense TK RNA inhibited expression of sense TK mRNA. Antisense RNAs were stable in oocytes and had no detectable effect on either the expression of endogenous proteins or on the expression of nonhomologous RNA transcripts. CAT activity expressed from a plasmid transcribed in the oocyte nucleus was also inhibited by antisense RNA injected into the oocyte cytoplasm. The data suggest that antisense RNA will be useful in identifying the function of specific mRNA sequences during early development of the frog.

237 citations


Journal ArticleDOI
TL;DR: The study of the intracellular RNA species of B1 virus revealed that probably all of the virus-specific subgenomic mRNA species contained the body sequences of strain A59 but the leader sequences of JHM, indicating that the JHM leader RNA could determine the site of initiation for coronavirus mRNAs.
Abstract: We have isolated a recombinant virus between the A59 and JHM strains of mouse hepatitis virus, which contain a single species of nonsegmented RNA genome. This recombinant was derived by mixed infection of DBT cells with temperature-sensitive mutants of A59 and JHM at nonpermissive temperature. Viruses recovered at this temperature were screened by oligonucleotide fingerprinting of their genomic RNAs. One recombinant virus, B1, was found to contain mostly A59-derived sequences, but the 3 kilobases at the 5' end of the genomic RNA was derived from JHM. Thus, the crossover point in the B1 genome is located within gene A, which codes for the viral RNA polymerases. The study of the intracellular RNA species of B1 virus revealed that probably all of the virus-specific subgenomic mRNA species contained the body sequences of strain A59 but the leader sequences of JHM. This result indicates that the JHM leader RNA, which differs from the A59 leader RNA, could be fused to the mRNAs of a different virus strain during RNA transcription. Furthermore, B1 virus-infected cells contain an additional subgenomic mRNA species which is transcribed from a new initiation site within gene C, suggesting that the leader RNA could determine the site of initiation for coronavirus mRNAs. These data represent a first report of RNA recombination between viruses, other than picornaviruses, which contain nonsegmented RNA genomes.

215 citations


Journal ArticleDOI
TL;DR: The level of RNA polymerase II associated with several heat shock genes increased dramatically in response to heat shock, whereas the level associated with the copia genes decreased, indicating that both induction of heat shock gene expression and repression of theCopia gene expression by heat shock occur at the transcriptional level.
Abstract: We describe a method for examining the in vivo distribution of a protein on specific eucaryotic DNA sequences. In this method, proteins are cross-linked to DNA in intact cells, and the protein-DNA adducts are isolated by immunoprecipitation with antiserum against the protein. Characterization of the DNA cross-linked to the precipitated protein identifies the sequences with which the protein is associated in vivo. Here, we applied these methods to detect RNA polymerase II-DNA interactions in heat-shocked and untreated Drosophila melanogaster Schneider line 2 cells. The level of RNA polymerase II associated with several heat shock genes increased dramatically in response to heat shock, whereas the level associated with the copia genes decreased, indicating that both induction of heat shock gene expression and repression of the copia gene expression by heat shock occur at the transcriptional level. Low levels of RNA polymerase II were present on DNA outside of the transcription units, and for at least two genes, hsp83 and hsp26, RNA polymerase II initiated binding near the transcription start site. Moreover, for hsp70, the density of RNA polymerase II on sequences downstream of the polyadenylate addition site was much lower than that observed on the gene internal sequences. Examination of the amount of specific restriction fragments cross-linked to RNA polymerase II provides a means of detecting RNA polymerase II on individual members of multigene families. This analysis shows that RNA polymerase II is associated with only one of the two cytoplasmic actin genes.

215 citations


Journal ArticleDOI
01 May 1985-Virology
TL;DR: The genome of infectious bursal disease virus (IBDV) strain 002-73 was found to consist of two segments of double-stranded RNA which could encode proteins of MW 125K while the MW of the translated products was 169K suggesting that a precursor-product relationship exists between some of the translation products.

204 citations


Journal ArticleDOI
TL;DR: It is suggested that lacUV5 mRNA synthesis is primed by a promoter-bound enzyme complex that synthesizes the initial nine or ten bases in the mRNA chain.

196 citations


Journal ArticleDOI
15 Jan 1985-Virology
TL;DR: A model in which flavivirus RF functions as a recycling template for semiconservative and asymmetric replication, on which only one nascent strand is synthesized per cycle is formed.

188 citations


Journal ArticleDOI
TL;DR: It is suggested that phosphate groups in RNA polymerase II-binding proteins may regulate transcription by modulating the interaction of RNA polymerases II with other regulatory proteins that possess sequence recognition specificity.

Journal ArticleDOI
TL;DR: A method for localizing and quantitating m6A within a large RNA molecule, the genomic RNA of Rous sarcoma virus, and observing a high frequency of methylation at PuGm6ACU sequences is described.
Abstract: N6-methyladenosine (m6A) residues are present as internal base modifications in most higher eucaryotic mRNAs; however, the biological function of this modification is not known. We describe a method for localizing and quantitating m6A within a large RNA molecule, the genomic RNA of Rous sarcoma virus. Specific fragments of 32P-labeled Rous sarcoma virus RNA were isolated by hybridization with complementary DNA restriction fragments spanning nucleotides 6185 to 8050. RNA was digested with RNase and finger-printed, and individual oligonucleotides were analyzed for the presence of m6A by paper electrophoresis and thin-layer chromatography. With this technique, seven sites of methylation in this region of the Rous sarcoma virus genome were localized at nucleotides 6394, 6447, 6507, 6718, 7414, 7424, and 8014. Further, m6A was observed at two additional sites whose nucleotide assignments remain ambiguous. A clustering of two or more m6A residues was seen at three positions within the RNA analyzed. Modification at certain sites was found to be heterogeneous, in that different molecules of RNA appeared to be methylated differently. Previous studies have determined that methylation occurs only in the sequences Gm6AC and Am6AC. We observed a high frequency of methylation at PuGm6ACU sequences. The possible involvement of m6A in RNA splicing events is discussed.

Journal ArticleDOI
01 Jul 1985-Cell
TL;DR: Chromatographic fractionation and complementation analysis of extracts from mock- and virus-infected cells indicated that the factor(s) responsible for the enhanced activity was localized entirely in the fraction containing transcription factor IIIC.

Journal ArticleDOI
TL;DR: Three lines of evidence suggest that VAI RNA facilitates translation by preventing inactivation of the function of eukaryotic initiation factor 2 (eIF-2) by relieving the mutant's translational defect.
Abstract: The virus-associated VAI RNA of adenovirus is a small, RNA polymerase III-transcribed species required for efficient translation of mRNAs late after infection. Deletion mutant dl331 fails to produce this RNA and, as a result, grows poorly. Three lines of evidence suggest that VAI RNA facilitates translation by preventing inactivation of the function of eukaryotic initiation factor 2 (eIF-2). First, the mutant's translational defect can be relieved by addition of eIF-2 or eIF-2B (GTP recycling factor). Second, extracts of mutant-infected cells exhibit enhanced protein P1/eIF-2 alpha subunit kinase activity. Third, dl331 can grow with nearly normal kinetics in cells that do not express the kinase.

Journal ArticleDOI
01 Mar 1985-Cell
TL;DR: It is shown that binding of RNA I to the homologous RNA II is inhibited by an RNA I specified by a plasmid of different compatibility, and the copy numbers of both ColE1 and RSF1030 are increased when both plasmids are present in the same cell.

Journal ArticleDOI
01 Aug 1985-Cell
TL;DR: This work shows that this nucleotide-sequence-specific interaction of Q protein and RNA polymerase can occur while the enzyme is pausing after 16 nucleotides of the late gene transcript have been made, and suggests that the ability of Q-modifiedRNA polymerase to escape this pause reflects the change that allows it to go through terminators.

Journal ArticleDOI
TL;DR: Structural analysis of this RNA showed that it is RNA I, the RNA that inhibits plasmid DNA synthesis, and it is remarkable that nine nucleotides around the cleavage sites are identical in both substrates.

Journal ArticleDOI
TL;DR: Primer extension analysis of tk RNA isolated from COS cells transfected with prHuTK reveals that transcription from the RNA polymerase I promoter is highly efficient and initiates at the same position used for the synthesis of endogenous rRNA in HeLa cells, suggesting that prHuK transcription is under the control of the human rRNA promoter and RNA polymerases I.
Abstract: We studied RNA polymerase I transcription in cells transfected with a plasmid, prHuTK, containing the herpes simplex virus tk gene fused to a human rRNA promoter. Primer extension analysis of tk RNA isolated from COS cells transfected with prHuTK reveals that transcription from the RNA polymerase I promoter is highly efficient and initiates at the same position used for the synthesis of endogenous rRNA in HeLa cells. The RNA products derived from prHuTK are distinguishable from normal RNA polymerase II transcripts of tk in that they are not polyadenylated, are extremely unstable, and are found predominantly in the nucleus. Moreover, the transcription observed is resistant to 300 micrograms of alpha-amanitin per ml. These results strongly suggest that prHuTK transcription is under the control of the human rRNA promoter and RNA polymerase I. To further characterize the activity of the human rDNA promoter in vivo, a series of 5' and 3' deletion mutants was tested in this transfection assay. The deletion analysis indicates that a core region of ca. 40 base pairs overlapping the initiation site is critical for transcription. In addition, a region between nucleotides -234 and -131 upstream from the core sequence serves to modulate the efficiency of transcription. Insertion into prHuTK of additional ribosomal nontranscribed spacer DNA or the simian virus 40 enhancer element has no apparent effect on the promoter activity. Surprisingly, RNA polymerase II transcripts synthesized at low levels from two start sites within the core control element of the wild-type RNA polymerase I promoter are activated upon deletion of upstream RNA polymerase I promoter sequences. However, these RNA polymerase II transcripts are not expressed from the endogenous rRNA promoter.

Journal ArticleDOI
TL;DR: The complete nucleotide sequence of the carnation mottle genomic RNA (4003 nucleotides) was determined for cloned cDNA copies of viral RNA containing over 99% of the sequence and was completed by direct sequence analysis of RNA and cDNA transcripts.
Abstract: The complete nucleotide sequence of carnation mottle genomic RNA (4003 nucleotides) is presented. The sequence was determined for cloned cDNA copies of viral RNA containing over 99% of the sequence and was completed by direct sequence analysis of RNA and cDNA transcripts. The sequence contains two long open reading frames which together can account for observed translation products. One translation product would arise by suppression of an amber termination codon and the sequence raises the possibility that a second suppression event could also occur. Sequence homology exists between a portion of the carnation mottle virus sequence and that of putative RNA polymerases from other RNA viruses.

Journal ArticleDOI
TL;DR: Results suggest that the NS protein is involved in the RNA chain elongation step of RNA synthesis, and possibly interacts with both the L protein and the template N-RNA and unwinds the latter to facilitate the movement of L protein on the template RNA.

Journal ArticleDOI
13 Jun 1985-Nature
TL;DR: It is proposed that the micRNA immune system provides an effective means of preventing viral infection as well as the expression of harmful genes in both prokaryotes and eukaryotes.
Abstract: The "operon' theory of gene regulation, in which protein repressor molecules bind to the operator site of a gene to prevent its transcription, is now well established. Recently, however, cases have been discovered in which gene expression is regulated by complementary RNA molecules that are able to bind to the transcripts of particular genes and consequently prevent their translation. For example, the synthesis of OmpF protein (a major outer membrane protein) in Escherichia coli is regulated by a short RNA complementary to a region of ompF RNA encompassing the "Shine-Dalgarno' sequence and the translation initiation codon. This RNA has been termed micRNA (messenger-RNA-interfering complementary RNA), and its discovery has prompted us to construct an artificial micRNA system designed to regulate gene expression in E. coli. A given target gene can be repressed by artificially producing an RNA (micRNA) complementary to the mRNA encoded by that gene. A micRNA system has also been used successfully in tissue-cultured mammalian cells. The use of artificial micRNAs to specifically regulate individual genes has great potential as a novel cellular immune system for blocking bacteriophage or virus infection. Here, we report that on induction of micRNAs directed against the coat protein and/or the replicase of the E. coli bacteriophage SP, phage proliferation was effectively prevented. We propose that the micRNA immune system provides an effective means of preventing viral infection as well as the expression of harmful genes in both prokaryotes and eukaryotes.

Journal ArticleDOI
TL;DR: Sequencing of parts of this cloned DNA show that these three intergenic regions contain a common 11-nucleotide sequence, which may contain a binding site for a leader RNA that hybridizes to negative-strand viral RNA at the beginning of each gene to prime mRNA synthesis.
Abstract: cDNA clones that represent various portions of the coronavirus mouse hepatitis virus strain A59 genome RNA have been constructed. cDNAs were synthesized by transcription of genome RNA by using either oligo(dT) or random oligomers of calf thymus DNA as primers. These cDNAs were converted into double-stranded DNA and cloned into pBR322 by standard techniques. The resulting cloned viral DNA fragments were mapped to viral genes by hybridization with Northern blots of intracellular RNA from mouse hepatitis virus strain A59-infected cells. These cDNA clones map in six of the seven viral genes. Clone g344, 1.8 kilobases, is the largest and encompasses gene 5 (which encodes a nonstructural protein) and gene 6 (which encodes the E1 viral glycoprotein) as well as the intergenic regions preceding genes 5, 6, and 7. Sequencing of parts of this cloned DNA show that these three intergenic regions contain a common 11-nucleotide sequence. This sequence shares homology with the 39 end of the viral mRNA leader sequence. Thus, this common intergenic sequence may contain a binding site for a leader RNA that hybridizes to negative-strand viral RNA at the beginning of each gene to prime mRNA synthesis. The different degrees of homology between the leader and its putative binding site may influence the differential rates of transcription of the various viral mRNAs. Images

Journal ArticleDOI
TL;DR: There is no significant base sequence homology among the countertranscripts of pT181, ColE1, and R1/NR1/R6-5, suggesting that the structural parallelism has risen by convergent molecular evolution.
Abstract: A transcription map of the replication control region of the Staphylococcus aureus plasmid pT181 has been constructed. Two major leftward transcripts, RNA III and RNA IV, start at positions 339 and 413, respectively. These two RNAs can serve as mRNAs for a plasmid-specific replication protein RepC. Two short rightward transcripts, RNA I and RNA II, approximately 85 and 150 nucleotides long, respectively, start at position 246. These rightward transcripts (referred to as countertranscripts) do not appear to be translated but act directly as negative regulators of plasmid replication, probably by interfering with translation of the RepC mRNAs. There is no significant base sequence homology among the countertranscripts of pT181, ColE1, and R1/NR1/R6-5, suggesting that the structural parallelism has risen by convergent molecular evolution.

Journal ArticleDOI
TL;DR: The entire translation products of RNA 1 of CMV and BMV could be aligned with significant homology and less prominent homologies were found with alfalfa mosaic virus RNA 1 translation product and with tobacco mosaic virus Mr-126000 protein.
Abstract: The nucleotide sequence of the 3389 residues of RNA 1 (Mr1.15 × 106) of the Q strain of cucumber mosaic virus (CMV) was determined, completing the primary structure of the CMV genome (8617 nucleotides). CMV RNA 1 was sequenced by the dideoxy-chain-termination method using M13 clones carrying RNA 1 sequences as well as synthetic oligonucleotide primers on RNA 1 as a template. At the 5′ end of the RNA there are 97 noncoding residues between the cap structure and the first AUG (98–100), which is the start of a single long openreading frame. This reading frame encodes a translation product of 991 amino acid residues (Mr 110791) and stops 319 nucleotide residues from the 3′ end of RNA 1. In addition to the conserved 3′ region present in all CMV RNAs (307 residues in RNA 1), RNAs 1 and 2 have highly homologous 5′ leader sequences, a 12-nucleotide segment of which is also conserved in the corresponding RNAs of brome mosaic virus (BMV). CMV satellite RNA can form stable base pairs with a region of CMV RNAs 1 and 2 including this 12-nucleotide sequence, implying a regulatory function. This conserved sequence is part of a hairpin structure in RNAs 1 and 2 of CMV and BMV and in CMV satellite RNA. The entire translation products of RNA 1 of CMV and BMV could be aligned with significant homology. Less prominent homologies were found with alfalfa mosaic virus RNA 1 translation product and with tobacco mosaic virus Mr-126000 protein.

Journal ArticleDOI
TL;DR: Data support the leader-primed mechanism for coronavirus transcription and suggest that one or more free leader RNAs are used as primers of mRNA synthesis, indicating that leader RNA synthesis is distinct from the transcription of mRNAs.

Journal ArticleDOI
TL;DR: The process of Darwinian selection in the self-replication of single-stranded RNA by Q beta replicase was investigated by analytical and computer-simulation methods, and the relative population change of the competing species was found to be a useful definition of selection value.
Abstract: The process of Darwinian selection in the self-replication of single-stranded RNA by Q beta replicase was investigated by analytical and computer-simulation methods. For this system, the relative population change of the competing species was found to be a useful definition of selection value, calculable from measurable kinetic parameters and concentrations of each species. Critical differences in the criteria for selection were shown to pertain for replicase/RNA ratios greater than or less than 1, for the case that formation of double-stranded RNA occurs and when comparisons are made of closed with open systems. At a large excess of enzyme, RNA species grow exponentially without interfering with each other, and selection depends only on the fecundity of the species, i.e., their overall replication rates. For RNA concentrations greater than the replicase concentration, the selection of species is governed by their abilities to compete for enzyme. Under conditions where formation of double strands occurs, competition leads to a coexistence of the species; the selection values vanish, and the concentration ratios depend only on the template binding and double-strand formation rates. The approach to coexistence is rapid, because when its competitors are in a steady state, a species present in trace amount is amplified exponentially. When formation of hybrid double strands occurs at a substantial rate, coexistence of hybridizing species is essentially limited to cases where the formation rate of heterologous double strands is smaller than the geometric mean of the formation rates of the homologous double strands. At limiting cases, e.g. in the steady states, simple analytical expressions for the main aspects of the selection process were found. Experimental data support the analytical expressions and the simulations.

Journal ArticleDOI
TL;DR: An in vitro transcription system which utilized a semisynthetic DNA template was developed and used to compare RNA chain elongation by wild type and mutant RNA polymerases II of Drosophila and indicated that a mutation to amanitin resistance reduces the rate at which the enzyme elongates transcripts.

Journal ArticleDOI
TL;DR: Intact primer RNA for discontinuous DNA replication of Escherichia coli has been detected by specific labeling in vitro of its 5'-terminal tri- (or di-) phosphate group with vaccinia guanylyltransferase and [alpha-32P]GTP.

Journal ArticleDOI
TL;DR: Analysis of the data from both enzymes suggests features that may be important for polymerase function, including a region that differs between the T3 and T7 enzymes that exhibits significant homology to the bi-helical domain that is common to many sequence-specific DNA binding proteins.
Abstract: The RNA polymerases encoded by bacteriophages T3 and T7 have similar structures, but exhibit nearly exclusive template specificities. We have determined the nucleotide sequence of the region of T3 DNA that encodes the T3 RNA polymerase (the gene 1.0 region), and have compared this sequence with the corresponding region of T7 DNA. The predicted amino acid sequence of the T3 RNA polymerase exhibits very few changes when compared to the T7 enzyme (82% of the residues are identical). Significant differences appear to cluster in three distinct regions in the amino-terminal half of the protein. Analysis of the data from both enzymes suggests features that may be important for polymerase function. In particular, a region that differs between the T3 and T7 enzymes exhibits significant homology to the bi-helical domain that is common to many sequence-specific DNA binding proteins. The region that flanks the structural gene contains a number of regulatory elements including: a promoter for the E. coli RNA polymerase, a potential processing site for RNase III and a promoter for the T3 polymerase. The promoter for the T3 RNA polymerase is located only 12 base pairs distal to the stop codon for the structural gene.

Journal ArticleDOI
01 Jun 1985-Cell
TL;DR: Analysis of td-encoded transcripts provides evidence in favor of maturation at the RNA level, consistent with an RNA processing model that would account for the direct gene transcript serving as precursor for both free intron RNA and a spliced mRNA that is colinear with the thymidylate synthase product.