To see the other types of publications on this topic, follow the link: Terminator Codon.

Journal articles on the topic 'Terminator Codon'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Terminator Codon.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Kozak, M. "Effects of intercistronic length on the efficiency of reinitiation by eucaryotic ribosomes." Molecular and Cellular Biology 7, no. 10 (October 1987): 3438–45. http://dx.doi.org/10.1128/mcb.7.10.3438.

Full text
Abstract:
Simian virus 40-based plasmids that direct the synthesis of preproinsulin during short-term transfection of COS cells have been used to probe the mechanism of reinitiation by eucaryotic ribosomes. Earlier studies from several laboratories had established that the ability of ribosomes to reinitiate translation at an internal AUG codon depends on having a terminator codon in frame with the preceding AUG triplet and upstream from the intended restart site. In the present studies, the position of the upstream terminator codon relative to the preproinsulin restart site has been systematically varied. The efficiency of reinitiation progressively improved as the intercistronic sequence was lengthened. When the upstream "minicistron" terminated 79 nucleotides before the preproinsulin start site, the synthesis of proinsulin was as efficient as if there were no upstream AUG codons. A mechanism is postulated that might account for this result, which is somewhat surprising inasmuch as bacterial ribosomes reinitiate less efficiently as the intercistronic gap is widened.
APA, Harvard, Vancouver, ISO, and other styles
2

Kozak, M. "Effects of intercistronic length on the efficiency of reinitiation by eucaryotic ribosomes." Molecular and Cellular Biology 7, no. 10 (October 1987): 3438–45. http://dx.doi.org/10.1128/mcb.7.10.3438-3445.1987.

Full text
Abstract:
Simian virus 40-based plasmids that direct the synthesis of preproinsulin during short-term transfection of COS cells have been used to probe the mechanism of reinitiation by eucaryotic ribosomes. Earlier studies from several laboratories had established that the ability of ribosomes to reinitiate translation at an internal AUG codon depends on having a terminator codon in frame with the preceding AUG triplet and upstream from the intended restart site. In the present studies, the position of the upstream terminator codon relative to the preproinsulin restart site has been systematically varied. The efficiency of reinitiation progressively improved as the intercistronic sequence was lengthened. When the upstream "minicistron" terminated 79 nucleotides before the preproinsulin start site, the synthesis of proinsulin was as efficient as if there were no upstream AUG codons. A mechanism is postulated that might account for this result, which is somewhat surprising inasmuch as bacterial ribosomes reinitiate less efficiently as the intercistronic gap is widened.
APA, Harvard, Vancouver, ISO, and other styles
3

Yarger, J. G., G. Armilei, and M. C. Gorman. "Transcription terminator-like element within a Saccharomyces cerevisiae promoter region." Molecular and Cellular Biology 6, no. 4 (April 1986): 1095–101. http://dx.doi.org/10.1128/mcb.6.4.1095.

Full text
Abstract:
We analyzed a cloned fragment of the yeast URA3 promoter region that contains a sequence of DNA capable of functioning as a highly efficient transcription terminator. BAL 31 deletions have shown the signal for the transcription termination activity is less than or equal to 110 base pairs and resides between bases 45 and 155 upstream of the URA3 primary ATG codon at base 227. In our in vivo assay system, the DNA fragment is able to terminate transcripts very efficiently in either orientation. The terminated transcripts bind to oligodeoxythymidylate cellulose columns and promote the synthesis of full-length cDNAs, suggesting that the transcripts are polyadenylated. The 110-base-pair region contains no sequence resembling terminator consensus sequences described by Zaret and Sherman (K.S. Zaret and F. Sherman, Cell, 28:563-573, 1982) or Henikoff and Cohen (S. Henikoff and E.H. Cohen, Mol. Cell. Biol., 4:1515-1520, 1984). We discuss the possible physiological relevance of this sequence to bona fide termination of transcription and to URA3 regulation in Saccharomyces cerevisiae.
APA, Harvard, Vancouver, ISO, and other styles
4

Yarger, J. G., G. Armilei, and M. C. Gorman. "Transcription terminator-like element within a Saccharomyces cerevisiae promoter region." Molecular and Cellular Biology 6, no. 4 (April 1986): 1095–101. http://dx.doi.org/10.1128/mcb.6.4.1095-1101.1986.

Full text
Abstract:
We analyzed a cloned fragment of the yeast URA3 promoter region that contains a sequence of DNA capable of functioning as a highly efficient transcription terminator. BAL 31 deletions have shown the signal for the transcription termination activity is less than or equal to 110 base pairs and resides between bases 45 and 155 upstream of the URA3 primary ATG codon at base 227. In our in vivo assay system, the DNA fragment is able to terminate transcripts very efficiently in either orientation. The terminated transcripts bind to oligodeoxythymidylate cellulose columns and promote the synthesis of full-length cDNAs, suggesting that the transcripts are polyadenylated. The 110-base-pair region contains no sequence resembling terminator consensus sequences described by Zaret and Sherman (K.S. Zaret and F. Sherman, Cell, 28:563-573, 1982) or Henikoff and Cohen (S. Henikoff and E.H. Cohen, Mol. Cell. Biol., 4:1515-1520, 1984). We discuss the possible physiological relevance of this sequence to bona fide termination of transcription and to URA3 regulation in Saccharomyces cerevisiae.
APA, Harvard, Vancouver, ISO, and other styles
5

Peabody, D. S., and P. Berg. "Termination-reinitiation occurs in the translation of mammalian cell mRNAs." Molecular and Cellular Biology 6, no. 7 (July 1986): 2695–703. http://dx.doi.org/10.1128/mcb.6.7.2695.

Full text
Abstract:
Many examples of internal translation initiation in eucaryotes have accumulated in recent years. In many cases terminators of upstream reading frames precede the internal initiation site, suggesting that translational reinitiation may be a mechanism for initiation at internal AUGs. To test this idea, a series of recombinants was constructed in the mammalian expression vector pSV2. Each contained a dicistronic transcription unit comprising the coding sequence for mouse dihydrofolate reductase (DHFR) followed by the gene for xanthine-guanine phosphoribosyl transferase (XGPRT) from Escherichia coli. Various versions of this pSV2dhfr-gpt recombinant plasmid altered the location at which the DHFR reading frame was terminated relative to the XGPRT initiation codon and demonstrated that this is a critical factor for the expression of XGPRT activity in transfected Cos-1 cells. Thus, when the DHFR frame terminated upstream or a very short distance downstream of the XGPRT initiator AUG, substantial levels of XGPRT activity were observed. When the DHFR frame terminated 50 nucleotides beyond the XGPRT initiator, activity was reduced about twofold. However, when the DHFR and XGPRT sequences were fused in-frame so that ribosomes which initiated at the DHFR AUG did not terminate until they encountered the XGPRT terminator, production of XGPRT activity was abolished. This dependence of internal translation initiation on the position of terminators of the upstream reading frame is consistent with the hypothesis that mammalian ribosomes are capable of translational reinitiation.
APA, Harvard, Vancouver, ISO, and other styles
6

Peabody, D. S., and P. Berg. "Termination-reinitiation occurs in the translation of mammalian cell mRNAs." Molecular and Cellular Biology 6, no. 7 (July 1986): 2695–703. http://dx.doi.org/10.1128/mcb.6.7.2695-2703.1986.

Full text
Abstract:
Many examples of internal translation initiation in eucaryotes have accumulated in recent years. In many cases terminators of upstream reading frames precede the internal initiation site, suggesting that translational reinitiation may be a mechanism for initiation at internal AUGs. To test this idea, a series of recombinants was constructed in the mammalian expression vector pSV2. Each contained a dicistronic transcription unit comprising the coding sequence for mouse dihydrofolate reductase (DHFR) followed by the gene for xanthine-guanine phosphoribosyl transferase (XGPRT) from Escherichia coli. Various versions of this pSV2dhfr-gpt recombinant plasmid altered the location at which the DHFR reading frame was terminated relative to the XGPRT initiation codon and demonstrated that this is a critical factor for the expression of XGPRT activity in transfected Cos-1 cells. Thus, when the DHFR frame terminated upstream or a very short distance downstream of the XGPRT initiator AUG, substantial levels of XGPRT activity were observed. When the DHFR frame terminated 50 nucleotides beyond the XGPRT initiator, activity was reduced about twofold. However, when the DHFR and XGPRT sequences were fused in-frame so that ribosomes which initiated at the DHFR AUG did not terminate until they encountered the XGPRT terminator, production of XGPRT activity was abolished. This dependence of internal translation initiation on the position of terminators of the upstream reading frame is consistent with the hypothesis that mammalian ribosomes are capable of translational reinitiation.
APA, Harvard, Vancouver, ISO, and other styles
7

Salmón, Marina, Guillem Paniagua, Carmen G. Lechuga, Fernando Fernández-García, Eduardo Zarzuela, Ruth Álvarez-Díaz, Monica Musteanu, et al. "KRAS4A induces metastatic lung adenocarcinomas in vivo in the absence of the KRAS4B isoform." Proceedings of the National Academy of Sciences 118, no. 30 (July 22, 2021): e2023112118. http://dx.doi.org/10.1073/pnas.2023112118.

Full text
Abstract:
In mammals, the KRAS locus encodes two protein isoforms, KRAS4A and KRAS4B, which differ only in their C terminus via alternative splicing of distinct fourth exons. Previous studies have shown that whereas KRAS expression is essential for mouse development, the KRAS4A isoform is expendable. Here, we have generated a mouse strain that carries a terminator codon in exon 4B that leads to the expression of an unstable KRAS4B154 truncated polypeptide, hence resulting in a bona fide Kras4B-null allele. In contrast, this terminator codon leaves expression of the KRAS4A isoform unaffected. Mice selectively lacking KRAS4B expression developed to term but died perinatally because of hypertrabeculation of the ventricular wall, a defect reminiscent of that observed in embryos lacking the Kras locus. Mouse embryonic fibroblasts (MEFs) obtained from Kras4B−/− embryos proliferated less than did wild-type MEFs, because of limited expression of KRAS4A, a defect that can be compensated for by ectopic expression of this isoform. Introduction of the same terminator codon into a KrasFSFG12V allele allowed expression of an endogenous KRAS4AG12V oncogenic isoform in the absence of KRAS4B. Exposure of Kras+/FSF4AG12V4B– mice to Adeno-FLPo particles induced lung tumors with complete penetrance, albeit with increased latencies as compared with control Kras+/FSFG12V animals. Moreover, a significant percentage of these mice developed proximal metastasis, a feature seldom observed in mice expressing both mutant isoforms. These results illustrate that expression of the KRAS4AG12V mutant isoform is sufficient to induce lung tumors, thus suggesting that selective targeting of the KRAS4BG12V oncoprotein may not have significant therapeutic consequences.
APA, Harvard, Vancouver, ISO, and other styles
8

Jenks, M. Harley, Thomas W. O'Rourke, and Daniel Reines. "Properties of an Intergenic Terminator and Start Site Switch That Regulate IMD2 Transcription in Yeast." Molecular and Cellular Biology 28, no. 12 (April 21, 2008): 3883–93. http://dx.doi.org/10.1128/mcb.00380-08.

Full text
Abstract:
ABSTRACT The IMD2 gene in Saccharomyces cerevisiae is regulated by intracellular guanine nucleotides. Regulation is exerted through the choice of alternative transcription start sites that results in synthesis of either an unstable short transcript terminating upstream of the start codon or a full-length productive IMD2 mRNA. Start site selection is dictated by the intracellular guanine nucleotide levels. Here we have mapped the polyadenylation sites of the upstream, unstable short transcripts that form a heterogeneous family of RNAs of ≈200 nucleotides. The switch from the upstream to downstream start sites required the Rpb9 subunit of RNA polymerase II. The enzyme's ability to locate the downstream initiation site decreased exponentially as the start was moved downstream from the TATA box. This suggests that RNA polymerase II's pincer grip is important as it slides on DNA in search of a start site. Exosome degradation of the upstream transcripts was highly dependent upon the distance between the terminator and promoter. Similarly, termination was dependent upon the Sen1 helicase when close to the promoter. These findings extend the emerging concept that distinct modes of termination by RNA polymerase II exist and that the distance of the terminator from the promoter, as well as its sequence, is important for the pathway chosen.
APA, Harvard, Vancouver, ISO, and other styles
9

Zhang, Haowei, Qin Li, Yongbin Li, and Sanfeng Chen. "The Serine Biosynthesis of Paenibacillus polymyxa WLY78 Is Regulated by the T-Box Riboswitch." International Journal of Molecular Sciences 22, no. 6 (March 16, 2021): 3033. http://dx.doi.org/10.3390/ijms22063033.

Full text
Abstract:
Serine is important for nearly all microorganisms in protein and downstream amino acids synthesis, however, the effect of serine on growth and nitrogen fixation was not completely clear in many bacteria, besides, the regulatory mode of serine remains to be fully established. In this study, we demonstrated that L-serine is essential for growth and nitrogen fixation of Paenibacillus polymyxa WLY78, but high concentrations of L-serine inhibit growth, nitrogenase activity, and nifH expression. Then, we revealed that expression of the serA whose gene product catalyzes the first reaction in the serine biosynthetic pathway is regulated by the T-box riboswitch regulatory system. The 508 bp mRNA leader region upstream of the serA coding region contains a 280 bp T-box riboswitch. The secondary structure of the T-box riboswitch with several conserved features: three stem-loop structures, a 14-bp T-box sequence, and an intrinsic transcriptional terminator, is predicted. Mutation and the transcriptional leader-lacZ fusions experiments revealed that the specifier codon of serine is AGC (complementary to the anticodon sequence of tRNAser). qRT-PCR showed that transcription of serA is induced by serine starvation, whereas deletion of the specifier codon resulted in nearly no expression of serA. Deletion of the terminator sequence or mutation of the continuous seven T following the terminator led to constitutive expression of serA. The data indicated that the T-box riboswitch, a noncoding RNA segment in the leader region, regulates expression of serA by a transcription antitermination mechanism.
APA, Harvard, Vancouver, ISO, and other styles
10

Levitin, Anastasia, and Charles Yanofsky. "Positions of Trp Codons in the Leader Peptide-Coding Region of the at Operon Influence Anti-Trap Synthesis and trp Operon Expression in Bacillus licheniformis." Journal of Bacteriology 192, no. 6 (January 8, 2010): 1518–26. http://dx.doi.org/10.1128/jb.01420-09.

Full text
Abstract:
ABSTRACT Tryptophan, phenylalanine, tyrosine, and several other metabolites are all synthesized from a common precursor, chorismic acid. Since tryptophan is a product of an energetically expensive biosynthetic pathway, bacteria have developed sensing mechanisms to downregulate synthesis of the enzymes of tryptophan formation when synthesis of the amino acid is not needed. In Bacillus subtilis and some other Gram-positive bacteria, trp operon expression is regulated by two proteins, TRAP (the tryptophan-activated RNA binding protein) and AT (the anti-TRAP protein). TRAP is activated by bound tryptophan, and AT synthesis is increased upon accumulation of uncharged tRNATrp. Tryptophan-activated TRAP binds to trp operon leader RNA, generating a terminator structure that promotes transcription termination. AT binds to tryptophan-activated TRAP, inhibiting its RNA binding ability. In B. subtilis, AT synthesis is upregulated both transcriptionally and translationally in response to the accumulation of uncharged tRNATrp. In this paper, we focus on explaining the differences in organization and regulatory functions of the at operon's leader peptide-coding region, rtpLP, of B. subtilis and Bacillus licheniformis. Our objective was to correlate the greater growth sensitivity of B. licheniformis to tryptophan starvation with the spacing of the three Trp codons in its at operon leader peptide-coding region. Our findings suggest that the Trp codon location in rtpLP of B. licheniformis is designed to allow a mild charged-tRNATrp deficiency to expose the Shine-Dalgarno sequence and start codon for the AT protein, leading to increased AT synthesis.
APA, Harvard, Vancouver, ISO, and other styles
11

Meng, Qi, and Robert L. Switzer. "Regulation of Transcription of the Bacillus subtilis pyrG Gene, Encoding Cytidine Triphosphate Synthetase." Journal of Bacteriology 183, no. 19 (October 1, 2001): 5513–22. http://dx.doi.org/10.1128/jb.183.19.5513-5522.2001.

Full text
Abstract:
ABSTRACT The B. subtilis pyrG gene, which encodes CTP synthetase, is located far from the pyrimidine biosynthetic operon on the chromosome and is independently regulated. The pyrGpromoter and 5′ leader were fused to lacZ and integrated into the chromosomes of several B. subtilis strains having mutations in genes of pyrimidine biosynthesis and salvage. These mutations allowed the intracellular pools of cytidine and uridine nucleotides to be manipulated by the composition of the growth medium. These experiments indicated that pyrGexpression is repressed by cytidine nucleotides but is largely independent of uridine nucleotides. The start ofpyrG transcription was mapped by primer extension to a position 178 nucleotides upstream of the translation initiation codon. A factor-independent termination hairpin lying between thepyrG promoter and its coding region is essential for regulation of pyrG expression. Primer-extended transcripts were equally abundant in repressed and derepressed cells when the primer bound upstream of the terminator, but they were much less abundant in repressed cells when the primer bound downstream of the terminator. Furthermore, deletion of the terminator from pyrG-lacZ fusions integrated into the chromosome yielded elevated levels of expression that was not repressible by cytidine. We suggest that cytidine repression of pyrGexpression is mediated by an antitermination mechanism in which antitermination by a putative trans-acting protein is reduced by elevated levels of cytidine nucleotides. Conservation of sequences and secondary structural elements in the pyrG5′ leaders of several other gram-positive bacteria indicates that theirpyrG genes are regulated by a similar mechanism.
APA, Harvard, Vancouver, ISO, and other styles
12

Johanesen, Priscilla A., Dena Lyras, Trudi L. Bannam, and Julian I. Rood. "Transcriptional Analysis of thetet(P) Operon from Clostridium perfringens." Journal of Bacteriology 183, no. 24 (December 15, 2001): 7110–19. http://dx.doi.org/10.1128/jb.183.24.7110-7119.2001.

Full text
Abstract:
ABSTRACT The Clostridium perfringens tetracycline resistance determinant from the 47-kb conjugative R-plasmid pCW3 is unique in that it consists of two overlapping genes, tetA(P) andtetB(P), which mediate resistance by different mechanisms. Detailed transcriptional analysis has shown that the inducible tetA(P) and tetB(P) genes comprise an operon that is transcribed from a single promoter, P3, located 529 bp upstream of the tetA(P) start codon. Deletion of P3 or alteration of the spacing between the −35 and −10 regions significantly reduced the level of transcription in a reporter construct. Induction was shown to be mediated at the level of transcription. Unexpectedly, a factor-independent terminator, T1, was detected downstream of P3 but before the start of thetetA(P) gene. Deletion or mutation of this terminator led to increased read-through transcription in the reporter construct. It is postulated that the T1 terminator is an intrinsic control element of the tet(P) operon and that it acts to prevent the overexpression of the TetA(P) transmembrane protein, even in the presence of tetracycline.
APA, Harvard, Vancouver, ISO, and other styles
13

LIU, HUAXUAN, LIYUN YAN, and GUOFANG JIANG. "The complete mitochondrial genome of the jumping grasshopper Sinopodisma pieli (Orthoptera: Acrididae) and the phylogenetic analysis of Melanoplinae." Zootaxa 4363, no. 4 (December 12, 2017): 506. http://dx.doi.org/10.11646/zootaxa.4363.4.3.

Full text
Abstract:
In this study, we reported the complete mitochondrial genome (mitogenome) of Sinopodisma pieli by polymerase chain reaction method for the first time, the type species of the genus Sinopodisma. Its mitogenome was a circular DNA molecule of 15,625 bp in length, with 76.0% A+T, and contained 13 protein-coding genes, 22 transfer RNA genes and two ribosomal RNA genes and one A+T control region. The overall base composition of the S. pieli mitogenome was 42.8% for A, 33.2% for T, 13.5% for C, and 10.5% for G, respectively. All 13 mitochondrial PCGs shared the start codon ATN. Twelve of the PCGs ended with termination codon TAA and TAG, while cytochrome coxidase subunit 1 (COI) utilized an incomplete T as terminator codon. All tRNA genes could be folded into the typical cloverleaf secondary structure, except trnS(AGN) lacking of dihydrouridine arm. The sizes of the large and small ribosomal RNA genes were 1379 bp and 794 bp, respectively. The A+T rich region was 798 bp in length and contained 88.5% AT content. A phylogenetic analysis based on 13 PCGs by using Bayesian inference (BI) and maximum likelihood (ML) revealed that Sinopodisma is not monophyletic group. We think that the name and taxonomic status of S. tsinlingensis are right, and it should not be moved into the genus Pedopodisma. These data will provide important information for a better understanding of the population genetics and species identification for Sinopodisma.
APA, Harvard, Vancouver, ISO, and other styles
14

Ohki, Reiko, Kozue Tateno, Teruaki Takizawa, Toshiko Aiso, and Makiko Murata. "Transcriptional Termination Control of a Novel ABC Transporter Gene Involved in Antibiotic Resistance in Bacillus subtilis." Journal of Bacteriology 187, no. 17 (September 1, 2005): 5946–54. http://dx.doi.org/10.1128/jb.187.17.5946-5954.2005.

Full text
Abstract:
ABSTRACT In members of one of the subfamilies of the bacterial ATP binding cassette (ABC) transporters, the two nucleotide binding domains are fused as a single peptide and the proteins have no membrane-spanning domain partners. Most of the ABC efflux transporters of this subfamily have been characterized in actinomycetes, producing macrolide, lincosamide, and streptogramin antibiotics. Among 40 ABC efflux transporters of Bacillus subtilis, five proteins belong to this subfamily. None of these proteins has been functionally characterized. We examined macrolide, lincosamide, and streptogramin antibiotic resistance in insertional disruptants of the genes that encode these proteins. It was found that only a disruptant of vmlR (formerly named expZ) showed hypersensitivity to virginiamycin M and lincomycin. Expression of the vmlR gene was induced by the addition of these antibiotics in growth medium. Primer extension analysis revealed that transcription of the vmlR gene initiates at an adenosine residue located 225 bp upstream of the initiation codon. From the analysis of the vmlR and lacZ fusion genes, a 52-bp deletion from +159 to +211 resulted in constitutive expression of the vmlR gene. In this region, a typical ρ-independent transcriptional terminator was found. It was suggested that the majority of transcription ends at this termination signal in the absence of antibiotics, whereas under induced conditions, RNA polymerase reads through the terminator, and transcription continues to the downstream vmlR coding region, resulting in an increase in vmlR expression. No stabilization of vmlR mRNA occurred under the induced conditions.
APA, Harvard, Vancouver, ISO, and other styles
15

Brill, Jeanette, Tamara Hoffmann, Harald Putzer, and Erhard Bremer. "T-box-mediated control of the anabolic proline biosynthetic genes of Bacillus subtilis." Microbiology 157, no. 4 (April 1, 2011): 977–87. http://dx.doi.org/10.1099/mic.0.047357-0.

Full text
Abstract:
Bacillus subtilis possesses interlinked routes for the synthesis of proline. The ProJ–ProA–ProH route is responsible for the production of proline as an osmoprotectant, and the ProB–ProA–ProI route provides proline for protein synthesis. We show here that the transcription of the anabolic proBA and proI genes is controlled in response to proline limitation via a T-box-mediated termination/antitermination regulatory mechanism, a tRNA-responsive riboswitch. Primer extension analysis revealed mRNA leader transcripts of 270 and 269 nt for the proBA and proI genes, respectively, both of which are synthesized from SigA-type promoters. These leader transcripts are predicted to fold into two mutually exclusive secondary mRNA structures, forming either a terminator or an antiterminator configuration. Northern blot analysis allowed the detection of both the leader and the full-length proBA and proI transcripts. Assessment of the level of the proBA transcripts revealed that the amount of the full-length mRNA species strongly increased in proline-starved cultures. Genetic studies with a proB–treA operon fusion reporter strain demonstrated that proBA transcription is sensitively tied to proline availability and is derepressed as soon as cellular starvation for proline sets in. Both the proBA and the proI leader sequences contain a CCU proline-specific specifier codon prone to interact with the corresponding uncharged proline-specific tRNA. By replacing the CCU proline specifier codon in the proBA T-box leader with UUC, a codon recognized by a Phe-specific tRNA, we were able to synthetically re-engineer the proline-specific control of proBA transcription to a control that was responsive to starvation for phenylalanine.
APA, Harvard, Vancouver, ISO, and other styles
16

Sharp, Josh S., and David H. Bechhofer. "Effect of Translational Signals on mRNA Decay in Bacillus subtilis." Journal of Bacteriology 185, no. 18 (September 15, 2003): 5372–79. http://dx.doi.org/10.1128/jb.185.18.5372-5379.2003.

Full text
Abstract:
ABSTRACT A 254-nucleotide model mRNA, designated ΔermC mRNA, was used to study the effects of translational signals and ribosome transit on mRNA decay in Bacillus subtilis. ΔermC mRNA features a strong ribosome-binding site (RBS) and a 62-amino-acid-encoding open reading frame, followed by a transcription terminator structure. Inactivation of the RBS or the start codon resulted in a fourfold decrease in the mRNA half-life, demonstrating the importance of ternary complex formation for mRNA stability. Data for the decay of ΔermC mRNAs with stop codons at positions increasingly proximal to the translational start site showed that actual translation—even the formation of the first peptide bond—was not important for stability. The half-life of an untranslated 3.2-kb ΔermC-lacZ fusion RNA was similar to that of a translated ΔermC-lacZ mRNA, indicating that the translation of even a longer RNA was not required for wild-type stability. The data are consistent with a model in which ribosome binding and the formation of the ternary complex interfere with a 5′-end-dependent activity, possibly a 5′-binding endonuclease, which is required for the initiation of mRNA decay. This model is supported by the finding that increasing the distance from the 5′ end to the start codon resulted in a 2.5-fold decrease in the mRNA half-life. These results underscore the importance of the 5′ end to mRNA stability in B. subtilis.
APA, Harvard, Vancouver, ISO, and other styles
17

Lu, Jian Zhang, Qin Guo, Mei Lin Cui, Lu Yang, Shan Shan Du, Hui Ruan, and Guo Qing He. "Construction of a Yeast Cell-Surface Display System and Expression of Trametes sp. laccase." Advanced Materials Research 347-353 (October 2011): 3635–40. http://dx.doi.org/10.4028/www.scientific.net/amr.347-353.3635.

Full text
Abstract:
Laccases (1.10.3.2, p-diphenol: dioxygen oxidoreductases) is a family of blue copper-containing oxidases that are commonly found in bacteria, fungi and plants. It is able to oxidize and degrade a variety of aromatic compounds and other organic compounds. Due to this ability, laccases can serve environmental bioremediation processes and industrial purposes. Cell-surface display of enzymes is one of the most attractive applications in yeast. It is a effective utilization to construct the whole cell biocatalyst. The cDNA sequence of Trametes sp. C30 LAC3 was optimized and synthesized according to the codon bias of Saccharomyces Italic textcerevisiae, because codon optimization has been proved to be effective to maximize production of heterologous proteins in yeast. The genes encoding galactokinase (GAL1) promoter, α-mating factor 1 (MFα1) pre-pro secretion signal, fully codon-optimized LAC3, the 320 amino acids of C terminal of α-agglutinin, alcohol dehydrogenase (ADH1) terminator and kanMX cassette were amplified and cloned into YEplac181 to construct a cell-surface display vector called pGMAAK-lac3 with α-agglutinin as an anchor. Then pGMAAK-lac3 was transformed into S. cerevisiae. The results show LAC3 was immobilized and actively expressed on S. cerevisiae. However, the substrate specifity and activity were obviously changed. The displayed LAC3 lost the activity to phenolic substrate (guaiacol) and its activity to non-phenolic substrate (ABTS) was greatly reduced. To our knowledge, this was the first attempt to construct and express laccase through cell-surface display technology.
APA, Harvard, Vancouver, ISO, and other styles
18

Lapteva, Y. S., O. E. Zolova, M. G. Shlyapnikov, I. M. Tsfasman, T. A. Muranova, O. A. Stepnaya, I. S. Kulaev, and I. E. Granovsky. "Cloning and Expression Analysis of Genes Encoding Lytic Endopeptidases L1 and L5 from Lysobacter sp. Strain XL1." Applied and Environmental Microbiology 78, no. 19 (August 3, 2012): 7082–89. http://dx.doi.org/10.1128/aem.01621-12.

Full text
Abstract:
ABSTRACTLytic enzymes are the group of hydrolases that break down structural polymers of the cell walls of various microorganisms. In this work, we determined the nucleotide sequences of theLysobactersp. strain XL1alpAandalpBgenes, which code for, respectively, secreted lytic endopeptidases L1 (AlpA) and L5 (AlpB).In silicoanalysis of their amino acid sequences showed these endopeptidases to be homologous proteins synthesized as precursors similar in structural organization: the mature enzyme sequence is preceded by an N-terminal signal peptide and a pro region. On the basis of phylogenetic analysis, endopeptidases AlpA and AlpB were assigned to the S1E family [clan PA(S)] of serine peptidases. Expression of thealpAandalpBopen reading frames (ORFs) inEscherichia coliconfirmed that they code for functionally active lytic enzymes. Each ORF was predicted to have the Shine-Dalgarno sequence located at a canonical distance from the start codon and a potential Rho-independent transcription terminator immediately after the stop codon. ThealpAandalpBmRNAs were experimentally found to be monocistronic; transcription start points were determined for both mRNAs. The synthesis of thealpAandalpBmRNAs was shown to occur predominantly in the late logarithmic growth phase. The amount ofalpAmRNA in cells ofLysobactersp. strain XL1 was much higher, which correlates with greater production of endopeptidase L1 than of L5.
APA, Harvard, Vancouver, ISO, and other styles
19

WANG, YING, JINJUN CAO, and WEIHAI LI. "The complete mitochondrial genome of the styloperlid stonefly species Styloperla spinicercia Wu (Insecta: Plecoptera) with family-level phylogenetic analyses of the Pteronarcyoidea." Zootaxa 4243, no. 1 (March 13, 2017): 125. http://dx.doi.org/10.11646/zootaxa.4243.1.5.

Full text
Abstract:
We present the complete mitochondrial (mt) genome sequence of the stonefly, Styloperla spinicercia Wu, 1935 (Plecoptera: Styloperlidae), the type species of the genus Styloperla and the first complete mt genome for the family Styloperlidae. The genome is circular, 16,129 base pairs long, has an A+T content of 70.7%, and contains 37 genes including the large and small ribosomal RNA (rRNA) subunits, 13 protein coding genes (PCGs), 22 tRNA genes and a large non-coding region (CR). All of the PCGs use the standard initiation codon ATN except ND1 and ND5, which start with TTG and GTG. Twelve of the PCGs stop with conventional terminal codons TAA and TAG, except ND5 which shows an incomplete terminator signal T. All tRNAs have the classic clover-leaf structures with the dihydrouridine (DHU) arm of tRNASer(AGN) forming a simple loop. Secondary structures of the two ribosomal RNAs are presented with reference to previous models. The structural elements and the variable numbers of tandem repeats are described within the control region. Phylogenetic analyses using both Bayesian (BI) and Maximum Likelihood (ML) methods support the previous hypotheses regarding family level relationships within the Pteronarcyoidea. The genetic distance calculated based on 13 PCGs and two rRNAs between Styloperla sp. and S. spinicercia is provided and interspecific divergence is discussed.
APA, Harvard, Vancouver, ISO, and other styles
20

Hiwasa-Tanase, Kyoko, Mpanja Nyarubona, Tadayoshi Hirai, Kazuhisa Kato, Takanari Ichikawa, and Hiroshi Ezura. "High-level accumulation of recombinant miraculin protein in transgenic tomatoes expressing a synthetic miraculin gene with optimized codon usage terminated by the native miraculin terminator." Plant Cell Reports 30, no. 1 (November 13, 2010): 113–24. http://dx.doi.org/10.1007/s00299-010-0949-y.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

De Gregorio, Eliana, Giustina Silvestro, Rossella Venditti, Maria Stella Carlomagno, and Pier Paolo Di Nocera. "Structural Organization and Functional Properties of Miniature DNA Insertion Sequences in Yersiniae." Journal of Bacteriology 188, no. 22 (September 8, 2006): 7876–84. http://dx.doi.org/10.1128/jb.00942-06.

Full text
Abstract:
ABSTRACT YPALs (Yersinia palindromic sequences) are miniature DNA insertions scattered along the chromosomes of yersiniae. The spread of these intergenic repeats likely occurred via transposition, as suggested by the presence of target site duplications at their termini and the identification of syntenic chromosomal regions which differ in the presence/absence of YPAL DNA among Yersinia strains. YPALs tend to be inserted closely downstream from the stop codon of flanking genes, and many YPAL targets overlap rho-independent transcriptional terminator-like sequences. This peculiar pattern of insertion supports the hypothesis that most of these repeats are cotranscribed with upstream sequences into mRNAs. YPAL RNAs fold into stable hairpins which may modulate mRNA decay. Accordingly, we found that YPAL-positive transcripts accumulate in Yersinia enterocolitica cells at significantly higher levels than homologous transcripts lacking YPAL sequences in their 3′ untranslated region.
APA, Harvard, Vancouver, ISO, and other styles
22

Cheng, J., P. Belgrader, X. Zhou, and L. E. Maquat. "Introns are cis effectors of the nonsense-codon-mediated reduction in nuclear mRNA abundance." Molecular and Cellular Biology 14, no. 9 (September 1994): 6317–25. http://dx.doi.org/10.1128/mcb.14.9.6317.

Full text
Abstract:
The translation of human triosephosphate isomerase (TPI) mRNA normally terminates at codon 249 within exon 7, the final exon. Frameshift and nonsense mutations of the type that cause translation to terminate prematurely at or upstream of codon 189 within exon 6 reduce the level of nuclear TPI mRNA to 20 to 30% of normal by a mechanism that is not a function of the distance of the nonsense codon from either the translation initiation or termination codon. In contrast, frameshift and nonsense mutations of another type that cause translation to terminate prematurely at or downstream of codon 208, also within exon 6, have no effect on the level of nuclear TPI mRNA. In this work, quantitations of RNA that derived from TPI alleles in which nonsense codons had been generated between codons 189 and 208 revealed that the boundary between the two types of nonsense codons resides between codons 192 and 195. The analysis of TPI gene insertions and deletions indicated that the positional feature differentiating the two types of nonsense codons is the distance of the nonsense codon upstream of intron 6. For example, the movement of intron 6 to a position downstream of its normal location resulted in a concomitant downstream movement of the boundary between the two types of nonsense codons. The analysis of intron 6 mutations indicated that the intron 6 effect is stipulated by the 88 nucleotides residing between the 5' and 3' splice sites. Since the deletion of intron 6 resulted in only partial abrogation of the nonsense codon-mediated reduction in the level of TPI mRNA, other sequences within TPI pre-mRNA must function in the effect. One of these sequences may be intron 2, since the deletion of intron 2 also resulted in partial abrogation of the effect. In experiments that switched introns 2 and 6, the replacement of intron 6 with intron 2 was of no consequence to the effect of a nonsense codon within either exon 1 or exon 6. In contrast, the replacement of intron 2 with intron 6 was inconsequential to the effect of a nonsense codon in exon 6 but resulted in partial abrogation of a nonsense codon in exon 1.
APA, Harvard, Vancouver, ISO, and other styles
23

Cheng, J., P. Belgrader, X. Zhou, and L. E. Maquat. "Introns are cis effectors of the nonsense-codon-mediated reduction in nuclear mRNA abundance." Molecular and Cellular Biology 14, no. 9 (September 1994): 6317–25. http://dx.doi.org/10.1128/mcb.14.9.6317-6325.1994.

Full text
Abstract:
The translation of human triosephosphate isomerase (TPI) mRNA normally terminates at codon 249 within exon 7, the final exon. Frameshift and nonsense mutations of the type that cause translation to terminate prematurely at or upstream of codon 189 within exon 6 reduce the level of nuclear TPI mRNA to 20 to 30% of normal by a mechanism that is not a function of the distance of the nonsense codon from either the translation initiation or termination codon. In contrast, frameshift and nonsense mutations of another type that cause translation to terminate prematurely at or downstream of codon 208, also within exon 6, have no effect on the level of nuclear TPI mRNA. In this work, quantitations of RNA that derived from TPI alleles in which nonsense codons had been generated between codons 189 and 208 revealed that the boundary between the two types of nonsense codons resides between codons 192 and 195. The analysis of TPI gene insertions and deletions indicated that the positional feature differentiating the two types of nonsense codons is the distance of the nonsense codon upstream of intron 6. For example, the movement of intron 6 to a position downstream of its normal location resulted in a concomitant downstream movement of the boundary between the two types of nonsense codons. The analysis of intron 6 mutations indicated that the intron 6 effect is stipulated by the 88 nucleotides residing between the 5' and 3' splice sites. Since the deletion of intron 6 resulted in only partial abrogation of the nonsense codon-mediated reduction in the level of TPI mRNA, other sequences within TPI pre-mRNA must function in the effect. One of these sequences may be intron 2, since the deletion of intron 2 also resulted in partial abrogation of the effect. In experiments that switched introns 2 and 6, the replacement of intron 6 with intron 2 was of no consequence to the effect of a nonsense codon within either exon 1 or exon 6. In contrast, the replacement of intron 2 with intron 6 was inconsequential to the effect of a nonsense codon in exon 6 but resulted in partial abrogation of a nonsense codon in exon 1.
APA, Harvard, Vancouver, ISO, and other styles
24

Xu-Guang, Zhai, Liu Yi, Wu Fang, Deng Guang-Bing, Pan Zhi-Fen, and Yu Mao-Qun. "Cloning and sequence analysis of the leucine-rich repeats region of cereal cyst nematode resistance gene." Chinese Journal of Agricultural Biotechnology 3, no. 3 (December 2006): 231–35. http://dx.doi.org/10.1079/cjb2006118.

Full text
Abstract:
AbstractAccording to the sequence of Rccn4, which is highly similar to the nucleotide-binding site (NBS) coding region of the cereal cyst nematode resistance gene, Cre3, three 3′ nested primers were designed to amplify its 3′ flanking region through single oligonucleotide nested polymerase chain reaction (SON-PCR). One 1264 bp band, Rccn-L, was amplified from E-10, a wheat–Aegilops variabilis translocation line containing the cereal cyst nematode resistance gene from Ae. variabilis. Sequence analysis showed that Rccn-L possesses the 3′ flanking sequence of Rccn4 and contains a 55 bp-sized consensus sequence with Rccn4. The coding region was 1026 bp, consisting of an incomplete open reading frame, a terminator codon and no initiation codon and intron; it encoded a peptide of 342 amino acid residues and shared 86% nucleotide sequence identity with Cre3. The peptide had a conserved leucine-rich repeat (LRR) domain, containing the imperfect repeats, XXLXXLXXL, comprising 17% leucine residues, and shares, respectively, 89% nucleotide sequence and 78% amino acid sequence identity with the LRR sequence of the Cre3 locus. In the present study, SON-PCR was used successfully, for the first time, in plant genome research and proved to be a valuable method in plant gene cloning. The acquirement of Rccn-L established the foundation for obtaining the complete Rccn gene and further structural and functional investigations.
APA, Harvard, Vancouver, ISO, and other styles
25

Salas-Marco, Joe, Hua Fan-Minogue, Adam K. Kallmeyer, Lawrence A. Klobutcher, Philip J. Farabaugh, and David M. Bedwell. "Distinct Paths To Stop Codon Reassignment by the Variant-Code Organisms Tetrahymena and Euplotes." Molecular and Cellular Biology 26, no. 2 (January 15, 2006): 438–47. http://dx.doi.org/10.1128/mcb.26.2.438-447.2006.

Full text
Abstract:
ABSTRACT The reassignment of stop codons is common among many ciliate species. For example, Tetrahymena species recognize only UGA as a stop codon, while Euplotes species recognize only UAA and UAG as stop codons. Recent studies have shown that domain 1 of the translation termination factor eRF1 mediates stop codon recognition. While it is commonly assumed that changes in domain 1 of ciliate eRF1s are responsible for altered stop codon recognition, this has never been demonstrated in vivo. To carry out such an analysis, we made hybrid proteins that contained eRF1 domain 1 from either Tetrahymena thermophila or Euplotes octocarinatus fused to eRF1 domains 2 and 3 from Saccharomyces cerevisiae. We found that the Tetrahymena hybrid eRF1 efficiently terminated at all three stop codons when expressed in yeast cells, indicating that domain 1 is not the sole determinant of stop codon recognition in Tetrahymena species. In contrast, the Euplotes hybrid facilitated efficient translation termination at UAA and UAG codons but not at the UGA codon. Together, these results indicate that while domain 1 facilitates stop codon recognition, other factors can influence this process. Our findings also indicate that these two ciliate species used distinct approaches to diverge from the universal genetic code.
APA, Harvard, Vancouver, ISO, and other styles
26

Barker, G. F., and K. Beemon. "Rous sarcoma virus RNA stability requires an open reading frame in the gag gene and sequences downstream of the gag-pol junction." Molecular and Cellular Biology 14, no. 3 (March 1994): 1986–96. http://dx.doi.org/10.1128/mcb.14.3.1986.

Full text
Abstract:
The intracellular accumulation of the unspliced RNA of Rous sarcoma virus was decreased when translation was prematurely terminated by the introduction of nonsense codons within its 5' proximal gene, the gag gene. Subcellular fractionation of transfected cells suggested that nonsense codon-mediated instability occurred in the cytoplasm. Analysis of constructs containing an in-frame deletion in the nucleocapsid domain of gag, which prevents interaction between the Gag protein and viral RNA, showed that an open reading frame extending to approximately 30 nucleotides from the natural gag termination codon was needed for RNA stability. Sequences at the gag-pol junction necessary for ribosomal frameshifting were not required for RNA stability; however, sequences located 100 to 200 nucleotides downstream of the natural gag termination codon were found to be necessary for stable RNA. The stability of RNAs lacking this downstream sequence was not markedly affected by premature termination codons. We propose that this downstream RNA sequence may interact with ribosomes translating gag to stabilize the RNA.
APA, Harvard, Vancouver, ISO, and other styles
27

Barker, G. F., and K. Beemon. "Rous sarcoma virus RNA stability requires an open reading frame in the gag gene and sequences downstream of the gag-pol junction." Molecular and Cellular Biology 14, no. 3 (March 1994): 1986–96. http://dx.doi.org/10.1128/mcb.14.3.1986-1996.1994.

Full text
Abstract:
The intracellular accumulation of the unspliced RNA of Rous sarcoma virus was decreased when translation was prematurely terminated by the introduction of nonsense codons within its 5' proximal gene, the gag gene. Subcellular fractionation of transfected cells suggested that nonsense codon-mediated instability occurred in the cytoplasm. Analysis of constructs containing an in-frame deletion in the nucleocapsid domain of gag, which prevents interaction between the Gag protein and viral RNA, showed that an open reading frame extending to approximately 30 nucleotides from the natural gag termination codon was needed for RNA stability. Sequences at the gag-pol junction necessary for ribosomal frameshifting were not required for RNA stability; however, sequences located 100 to 200 nucleotides downstream of the natural gag termination codon were found to be necessary for stable RNA. The stability of RNAs lacking this downstream sequence was not markedly affected by premature termination codons. We propose that this downstream RNA sequence may interact with ribosomes translating gag to stabilize the RNA.
APA, Harvard, Vancouver, ISO, and other styles
28

Lu, Jian Zhang, Mei Lin Cui, Shan Shan Du, Lu Yang, Qin Guo, Hui Ruan, and Guo Qing He. "Research on Construction of a Whole-Cell Biocatalyst for Xylan Degradation." Advanced Materials Research 347-353 (October 2011): 2599–603. http://dx.doi.org/10.4028/www.scientific.net/amr.347-353.2599.

Full text
Abstract:
Endo-1,4-β-xylanase (E.C.3.2.1.8) is a family of glycoside hydrolase. It is capable of hydrolyzing the backbone of substituted xylan polymers into fragments of random size. Due to this ability, xylanase can serve the degradation of lignocellulose, and facilitate the application of xylan. Cell-surface display of enzymes is one of the most attractive applications in yeast. It is a promising utilization in constructing the whole-cell biocatalyst of xylanase. For this purpose, a cDNA sequence of endo-1,4-β-xylanase B (XylB) from Aspergillus niger BCC14405 was optimized and synthesized according to the codon bias of Saccharomyces cerevisiae. The genes encoding galactokinase (GAL1) promoter, α-mating factor 1 (MFα1) pre-pro secretion signal, fully codon-optimized XylB, the 320 amino acids of C terminal of α-agglutinin, alcohol dehydrogenase (ADH1) terminator and kanMX cassette were amplified and cloned into YEplac181 to construct a cell-surface display vector called pGMAAK-XylB with α-agglutinin as an anchor. Then pGMAAK-XylB was transformed into S. cerevisiae. The results show XylB was immobilized and actively expressed on S. cerevisiae. Meanwhile, a secretion expression plasmid was also constructed using the above elements except α-agglutinin as a control strain in the study of characteristic of XylB. After an induction of 48 h by 2% galactose, the activity of displayed XylB reached 63 U/g dry-cell weight. The optimal pH of displayed XylB has changed from 5 to 6 and the optimal temperature has changed from 50 °C to 60 °C, comparing to the recombinant secretion XylB.
APA, Harvard, Vancouver, ISO, and other styles
29

Stolle, CA, MS Payne, and EJ Jr Benz. "Equal stabilities of normal beta globin and nontranslatable beta0 -39 thalassemic transcripts in cell-free extracts." Blood 70, no. 1 (July 1, 1987): 293–300. http://dx.doi.org/10.1182/blood.v70.1.293.bloodjournal701293.

Full text
Abstract:
Patients with beta zero thalassemia arising from premature terminator codon mutations in the gene for beta globin do not produce beta globin protein; these individuals also exhibit a decreased amount of beta globin mRNA in their erythroid cells. The absence of beta globin protein is readily explained by the inability of the beta zero-39 mRNA to be translated. The decrease in beta globin mRNA has been attributed to either decreased cytoplasmic stability of the nontranslatable decreased cytoplasmic stability of the nontranslatable mRNA or to an undefined nuclear lesion. To compare directly the relative stabilities of normal and beta zero-39 thalassemic globin transcripts, we prepared normal and thalassemic beta globin pre-mRNAs and mRNAs using cloned DNA templates and the SP6 promoter-polymerase system. The stability of the transcripts was assessed by incubation in various cell-free extracts. Our results indicate that although the stabilities of the beta globin transcripts varied considerably from one extract to another the stabilities of the beta zero-39 thalassemic pre-mRNAs and mRNAs were equal to those of normal beta globin mRNAs in every extract tested.
APA, Harvard, Vancouver, ISO, and other styles
30

Baldridge, Gerald D., Nicole Burkhardt, Michael J. Herron, Timothy J. Kurtti, and Ulrike G. Munderloh. "Analysis of Fluorescent Protein Expression in Transformants of Rickettsia monacensis, an Obligate Intracellular Tick Symbiont." Applied and Environmental Microbiology 71, no. 4 (April 2005): 2095–105. http://dx.doi.org/10.1128/aem.71.4.2095-2105.2005.

Full text
Abstract:
ABSTRACT We developed and applied transposon-based transformation vectors for molecular manipulation and analysis of spotted fever group rickettsiae, which are obligate intracellular bacteria that infect ticks and, in some cases, mammals. Using the Epicentre EZ::TN transposon system, we designed transposons for simultaneous expression of a reporter gene and a chloramphenicol acetyltransferase (CAT) resistance marker. Transposomes (transposon-transposase complexes) were electroporated into Rickettsia monacensis, a rickettsial symbiont isolated from the tick Ixodes ricinus. Each transposon contained an expression cassette consisting of the rickettsial ompA promoter and a green fluorescent protein (GFP) reporter gene (GFPuv) or the ompB promoter and a red fluorescent protein reporter gene (DsRed2), followed by the ompA transcription terminator and a second ompA promoter CAT gene cassette. Selection with chloramphenicol gave rise to rickettsial populations with chromosomally integrated single-copy transposons as determined by PCR, Southern blotting, and sequence analysis. Reverse transcription-PCR and Northern blots demonstrated transcription of all three genes. GFPuv transformant rickettsiae exhibited strong fluorescence in individual cells, but DsRed2 transformants did not. Western blots confirmed expression of GFPuv in R. monacensis and in Escherichia coli, but DsRed2 was expressed only in E. coli. The DsRed2 gene, but not the GFPuv gene, contains many GC-rich amino acid codons that are rare in the preferred codon suite of rickettsiae, possibly explaining the failure to express DsRed2 protein in R. monacensis. We demonstrated that our vectors provide a means to study rickettsia-host cell interactions by visualizing GFPuv-fluorescent R. monacensis associated with actin tails in tick host cells.
APA, Harvard, Vancouver, ISO, and other styles
31

Stolle, CA, MS Payne, and EJ Jr Benz. "Equal stabilities of normal beta globin and nontranslatable beta0 -39 thalassemic transcripts in cell-free extracts." Blood 70, no. 1 (July 1, 1987): 293–300. http://dx.doi.org/10.1182/blood.v70.1.293.293.

Full text
Abstract:
Abstract Patients with beta zero thalassemia arising from premature terminator codon mutations in the gene for beta globin do not produce beta globin protein; these individuals also exhibit a decreased amount of beta globin mRNA in their erythroid cells. The absence of beta globin protein is readily explained by the inability of the beta zero-39 mRNA to be translated. The decrease in beta globin mRNA has been attributed to either decreased cytoplasmic stability of the nontranslatable decreased cytoplasmic stability of the nontranslatable mRNA or to an undefined nuclear lesion. To compare directly the relative stabilities of normal and beta zero-39 thalassemic globin transcripts, we prepared normal and thalassemic beta globin pre-mRNAs and mRNAs using cloned DNA templates and the SP6 promoter-polymerase system. The stability of the transcripts was assessed by incubation in various cell-free extracts. Our results indicate that although the stabilities of the beta globin transcripts varied considerably from one extract to another the stabilities of the beta zero-39 thalassemic pre-mRNAs and mRNAs were equal to those of normal beta globin mRNAs in every extract tested.
APA, Harvard, Vancouver, ISO, and other styles
32

Fotheringham, I. G., S. A. Dacey, P. P. Taylor, T. J. Smith, M. G. Hunter, M. E. Finlay, S. B. Primrose, D. M. Parker, and R. M. Edwards. "The cloning and sequence analysis of the aspC and tyrB genes from Escherichia coli K12. Comparison of the primary structures of the aspartate aminotransferase and aromatic aminotransferase of E. coli with those of the pig aspartate aminotransferase isoenzymes." Biochemical Journal 234, no. 3 (March 15, 1986): 593–604. http://dx.doi.org/10.1042/bj2340593.

Full text
Abstract:
In this paper we describe the cloning and sequence analysis of the tyrB and aspC genes from Escherichia coli K12, which encode the aromatic aminotransferase and aspartate aminotransferase respectively. The tyrB gene was isolated from a cosmid carrying the nearby dnaB gene, identified by its ability to complement a dnaB lesion. Deletion and linker insertion analysis located the tyrB gene to a 1.7-kilobase NruI-HindIII-digest fragment. Sequence analysis revealed a gene encoding a 43 000 Da polypeptide. The gene starts with a GTG codon and is closely followed by a structure resembling a rho independent terminator. The aspC gene was cloned by screening gene banks, prepared from a prototrophic E. coli K12 strain, for plasmids able to complement the aspC tyrB lesions in the aminotransferase-deficient strain HW225. Sub-cloning and deletion analysis located the aspC gene on a 1.8-kilobase HincII-StuI-digest fragment. Sequence analysis revealed the presence of a gene encoding a 43 000 Da protein, the sequence of which is identical with that previously obtained for the aspartate aminotransferase from E. coli B. Considerable overproduction of the two enzymes was demonstrated. We compared the deduced protein sequences with those of the pig mitochondrial and cytoplasmic aspartate aminotransferases. From the extensive homology observed we are able to propose that the two E. coli enzymes possess subunit structures, subunit interactions and coenzyme-binding and substrate-binding sites that are very similar both to each other and to those of the mammalian enzymes and therefore must also have very similar catalytic mechanisms. Comparison of the aspC and tyrB gene sequences reveals that they appear to have diverged as much as is possible within the constraints of functionality and codon usage.
APA, Harvard, Vancouver, ISO, and other styles
33

Yang, Qian, Chien-Hung Yu, Fangzhou Zhao, Yunkun Dang, Cheng Wu, Pancheng Xie, Matthew S. Sachs, and Yi Liu. "eRF1 mediates codon usage effects on mRNA translation efficiency through premature termination at rare codons." Nucleic Acids Research 47, no. 17 (August 14, 2019): 9243–58. http://dx.doi.org/10.1093/nar/gkz710.

Full text
Abstract:
Abstract Codon usage bias is a universal feature of eukaryotic and prokaryotic genomes and plays an important role in regulating gene expression levels. A major role of codon usage is thought to regulate protein expression levels by affecting mRNA translation efficiency, but the underlying mechanism is unclear. By analyzing ribosome profiling results, here we showed that codon usage regulates translation elongation rate and that rare codons are decoded more slowly than common codons in all codon families in Neurospora. Rare codons resulted in ribosome stalling in manners both dependent and independent of protein sequence context and caused premature translation termination. This mechanism was shown to be conserved in Drosophila cells. In both Neurospora and Drosophila cells, codon usage plays an important role in regulating mRNA translation efficiency. We found that the rare codon-dependent premature termination is mediated by the translation termination factor eRF1, which recognizes ribosomes stalled on rare sense codons. Silencing of eRF1 expression resulted in codon usage-dependent changes in protein expression. Together, these results establish a mechanism for how codon usage regulates mRNA translation efficiency.
APA, Harvard, Vancouver, ISO, and other styles
34

Wiegert, Thomas, and Wolfgang Schumann. "SsrA-Mediated Tagging in Bacillus subtilis." Journal of Bacteriology 183, no. 13 (July 1, 2001): 3885–89. http://dx.doi.org/10.1128/jb.183.13.3885-3889.2001.

Full text
Abstract:
ABSTRACT A general mechanism in bacteria to rescue stalled ribosomes involves a stable RNA encoded by the ssrA gene. This RNA, termed tmRNA, encodes a proteolytic peptide tag which is cotranslationally added to truncated polypeptides, thereby targeting them for rapid proteolysis. To study this ssrA-mediated mechanism in Bacillus subtilis, a bipartite detection system was constructed that was composed of the HrcA transcriptional repressor and the bgaB reporter gene coding for a heat-stable β-galactosidase fused to an HrcA-controlled promoter. After the predicted proteolysis tag was fused to HrcA, the reporter β-galactosidase was expressed constitutively at a high level due to the instability of the tagged HrcA. Replacement of the two C-terminal alanine residues of the tag by aspartate rendered the repressor stable. Replacement of the hrcA stop codon by a transcriptional terminator sequence rendered the protein unstable; this was caused bytrans translational addition of the proteolytic tag. Inactivating the B. subtilis ssrA or smpB(yvaI) gene prevented the trans translational tagging reaction. Various protease-deficient strains of B. subtilis were tested for proteolysis of tagged HrcA. HrcA remained stable only in clpX or clpP knockouts, which suggests that this ATP-dependent protease is primarily responsible for the degradation of SsrA-tagged proteins in B. subtilis.
APA, Harvard, Vancouver, ISO, and other styles
35

Hirono, A., S. Miwa, H. Fujii, F. Ishida, K. Yamada, and K. Kubota. "Molecular study of eight Japanese cases of glucose-6-phosphate dehydrogenase deficiency by nonradioisotopic single-strand conformation polymorphism analysis [see comments]." Blood 83, no. 11 (June 1, 1994): 3363–68. http://dx.doi.org/10.1182/blood.v83.11.3363.3363.

Full text
Abstract:
Abstract Using a newly developed nonradioisotopic method of polymerase chain reaction single-strand conformation polymorphism (PCR-SSCP) analysis combined with the direct sequencing using the fluorescence-labeled terminator, we identified seven missense mutations, 527 A-->G, 1003 G-- >A, 1159 C--eT, 1160 G-->A, 1229 G-->A, 1246 G-->A, and 1361 G-->A, in eight Japanese patients with glucose-6-phosphate dehydrogenase (G6PD) deficiency. Except for the 527 A-->G, each mutation has been reported to cause variants G6PD Chatham, G6PD Guadalajara, G6PD Beverly Hills, G6PD “Japan”, G6PD Tokyo, and G6PD Andalus, respectively. In addition, a single base deletion in intron 5 was found in the patients with G6PD Guadalajara or G6PD Andalus. The variant with unique 527 A-->G was characterized and designated as G6PD Shinshu. We also characterized G6PD “Japan” and found that the variant had the striking resemblance with G6PD Riverside, bearing a missense mutation in the same codon, but causing a different amino acid substitution. Our modified PCR-SSCP analysis using minigel and ethidium bromide staining could detect six of the eight diverse mutations in the G6PD gene. Because it is easy and requires no special apparatus, this modified method will be useful for screening mutations in the G6PD gene.
APA, Harvard, Vancouver, ISO, and other styles
36

Hirono, A., S. Miwa, H. Fujii, F. Ishida, K. Yamada, and K. Kubota. "Molecular study of eight Japanese cases of glucose-6-phosphate dehydrogenase deficiency by nonradioisotopic single-strand conformation polymorphism analysis [see comments]." Blood 83, no. 11 (June 1, 1994): 3363–68. http://dx.doi.org/10.1182/blood.v83.11.3363.bloodjournal83113363.

Full text
Abstract:
Using a newly developed nonradioisotopic method of polymerase chain reaction single-strand conformation polymorphism (PCR-SSCP) analysis combined with the direct sequencing using the fluorescence-labeled terminator, we identified seven missense mutations, 527 A-->G, 1003 G-- >A, 1159 C--eT, 1160 G-->A, 1229 G-->A, 1246 G-->A, and 1361 G-->A, in eight Japanese patients with glucose-6-phosphate dehydrogenase (G6PD) deficiency. Except for the 527 A-->G, each mutation has been reported to cause variants G6PD Chatham, G6PD Guadalajara, G6PD Beverly Hills, G6PD “Japan”, G6PD Tokyo, and G6PD Andalus, respectively. In addition, a single base deletion in intron 5 was found in the patients with G6PD Guadalajara or G6PD Andalus. The variant with unique 527 A-->G was characterized and designated as G6PD Shinshu. We also characterized G6PD “Japan” and found that the variant had the striking resemblance with G6PD Riverside, bearing a missense mutation in the same codon, but causing a different amino acid substitution. Our modified PCR-SSCP analysis using minigel and ethidium bromide staining could detect six of the eight diverse mutations in the G6PD gene. Because it is easy and requires no special apparatus, this modified method will be useful for screening mutations in the G6PD gene.
APA, Harvard, Vancouver, ISO, and other styles
37

Buzina, Alla, and Marc J. Shulman. "Infrequent Translation of a Nonsense Codon Is Sufficient to Decrease mRNA Level." Molecular Biology of the Cell 10, no. 3 (March 1999): 515–24. http://dx.doi.org/10.1091/mbc.10.3.515.

Full text
Abstract:
In many organisms nonsense mutations decrease the level of mRNA. In the case of mammalian cells, it is still controversial whether translation is required for this nonsense-mediated RNA decrease (NMD). Although previous analyzes have shown that conditions that impede translation termination at nonsense codons also prevent NMD, the residual level of termination was unknown in these experiments. Moreover, the conditions used to impede termination might also have interfered with NMD in other ways. Because of these uncertainties, we have tested the effects of limiting translation of a nonsense codon in a different way, using two mutations in the immunoglobulin μ heavy chain gene. For this purpose we exploited an exceptional nonsense mutation at codon 3, which efficiently terminates translation but nonetheless maintains a high level of μ mRNA. We have shown 1) that translation of Ter462 in the double mutant occurs at only ∼4% the normal frequency, and 2) that Ter462 in cis with Ter3 can induce NMD. That is, translation of Ter462 at this low (4%) frequency is sufficient to induce NMD.
APA, Harvard, Vancouver, ISO, and other styles
38

Bergquist, P. L., V. S. J. Te'o, M. D. Gibbs, N. C. Curach, and K. M. H. Nevalainen. "Recombinant enzymes from thermophilic micro-organisms expressed in fungal hosts." Biochemical Society Transactions 32, no. 2 (April 1, 2004): 293–97. http://dx.doi.org/10.1042/bst0320293.

Full text
Abstract:
Cost-effective production of enzymes for industrial processes makes the appropriate selection of the host/vector expression system critical. We have tested two fungal systems for the bulk production of enzymes from thermophiles. The yeast Kluyveromyces lactis has been developed as a secretion host employing expression vectors based on the 2u-like plasmid pKD1 of Kluyveromyces drosophilarium. Our second system involves the filamentous fungus Trichoderma reesei. Signal and protein fusion vectors have been constructed using the strong cellobiohydrolase 1 (cbh1) promoter and recombinant plasmid DNAs introduced into various high-secreting T. reesei strains using biolistic particle delivery. In some cases (e.g. the xynB gene of Dictyoglomus thermophilum) we have reconstructed the genes according to Trichoderma codon preferences and demonstrated a dramatic increase in the production of the enzymes. The heterologous XynB enzyme is glycosylated differently in different Trichoderma strains. A proteomics approach has been taken to identify strongly expressed proteins produced by T. reesei under various cultivation conditions in order to identify condition-specific promoters driving the production of these proteins. Analyses indicated that HEX1, the major protein of the fungal Woronin body, is a dominant protein under both cellulase-inducing and -repressing conditions. The hex1 gene together with its promoter and terminator sequences has been isolated and the promoter function studied relative to cultivation time and medium.
APA, Harvard, Vancouver, ISO, and other styles
39

Li, Jun, Rui-Rui Lin, Yao-Yao Zhang, Kun-Jie Hu, Ya-Qi Zhao, Yan Li, Zhuo-Ran Huang, Xu Zhang, Xue-Xia Geng, and Jian-Hua Ding. "Characterization of the complete mitochondrial DNA of Theretra japonica and its phylogenetic position within the Sphingidae (Lepidoptera: Sphingidae)." ZooKeys 754 (May 3, 2018): 127–39. http://dx.doi.org/10.3897/zookeys.754.23404.

Full text
Abstract:
In the present study, the complete mitogenome of Theretrajaponica was sequenced and compared with other sequenced mitogenomes of Sphingidae species. The mitogenome of T.japonica, containing 37 genes (13 protein-coding genes, 22 tRNA genes, and two rRNA genes) and a region rich in adenine and thymine (AT-rich region), is a circular molecule with 15,399 base pairs (bp) in length. The order and orientation of the genes in the mitogenome are similar to those of other sequenced mitogenomes of Sphingidae species. All 13 protein-coding genes (PCGs) are initiated by ATN codons except for the cytochrome C oxidase subunit 1 gene (cox1) which is initiated by the codon CGA as observed in other lepidopteran insects. Cytochrome C oxidase subunit 2 gene (cox2) has the incomplete termination codon T and NADH dehydrogenase subunit 1 gene (nad1) terminates with TAG while the remainder terminates with TAA. Additionally, the codon distributions of the 13 PCGs revealed that Ile and Leu2 are the most frequently used codon families and codons CGG, CGC, CCG, CAG, and AGG are absent. The 431 bp AT-rich region includes the motif ATAGA followed by a 23 bp poly-T stretch, short tandem repeats (STRs) of TC and TA, two copies of a 28 bp repeat ‘ATTAAATTAATAAATTAA TATATTAATA’ and a poly-A element. Phylogenetic analyses within Sphingidae confirmed that T.japonica belongs to the Macroglossinae and showed that the phylogenetic relationship of T.japonica is closer to Ampelophagarubiginosa than Daphnisnerii. Phylogenetic analyses within Theretra demonstrate that T.japonica, T.jugurtha, T.suffusa, and T.capensis are clustered into one clade.
APA, Harvard, Vancouver, ISO, and other styles
40

Hung, Ming-Ni, Zhicheng Xia, Nien-Tai Hu, and Byong H. Lee. "Molecular and Biochemical Analysis of Two β-Galactosidases from Bifidobacterium infantisHL96." Applied and Environmental Microbiology 67, no. 9 (September 1, 2001): 4256–63. http://dx.doi.org/10.1128/aem.67.9.4256-4263.2001.

Full text
Abstract:
ABSTRACT Two genes encoding β-galactosidase isoenzymes,β-galI and β-galIII, fromBifidobacterium infantis HL96 were revealed on 3.6- and 2.4-kb DNA fragments, respectively, by nucleotide sequence analysis of the two fragments. β-galI (3,069 bp) encodes a 1,022-amino-acid (aa) polypeptide with a predicted molecular mass of 113 kDa. A putative ribosome binding site and a promoter sequence were recognized at the 5′ flanking region of β-galI. Further upstream a partial sequence of an open reading frame revealed a putative lactose permease gene transcribing divergently fromβ-galI. The β-galIII gene (2,076 bp) encodes a 691-aa polypeptide with a calculated molecular mass of 76 kDa. A rho-independent transcription terminator-like sequence was found 25 bp downstream of the termination codon. The amino acid sequences of β-GalI and β-GalIII are homologous to those found in the LacZ and the LacG families, respectively. The acid-base, nucleophilic, and substrate recognition sites conserved in the LacZ family were found in β-GalI, and a possible acid-base site proposed for the LacG family was located in β-GalIII, which featured a glutamate at residue 160. The coding regions of the β-galI andβ-galIII genes were each cloned downstream of a T7 promoter for overexpression in Escherichia coli. The molecular masses of the overexpressed proteins, as estimated by polyacrylamide gel electrophoresis on sodium dodecyl sulfate-polyacrylamide gels, agree with their predicted molecular weights. β-GalI and β-GalIII were specific for β-d-anomer-linked galactoside substrates. Both are more active in response to ONPG (o-nitrophenyl-β-d-galactopyranoside) than in response to lactose, particularly β-GalIII. The galacto-oligosaccharide yield in the reaction catalyzed by β-GalI at 37°C in 20% (wt/vol) lactose solution was 130 mg/ml, which is more than six times higher than the maximum yield obtained with β-GalIII. The structure of the major trisaccharide produced by β-GalI catalysis was characterized asO-β-d-galactopyranosyl-(1-3)-O-β-d-galactopyranosyl-(1-4)-d-glucopyranose (3′-galactosyl-lactose).
APA, Harvard, Vancouver, ISO, and other styles
41

Hoernes, Thomas Philipp, Nina Clementi, Michael Andreas Juen, Xinying Shi, Klaus Faserl, Jessica Willi, Catherina Gasser, et al. "Atomic mutagenesis of stop codon nucleotides reveals the chemical prerequisites for release factor-mediated peptide release." Proceedings of the National Academy of Sciences 115, no. 3 (January 3, 2018): E382—E389. http://dx.doi.org/10.1073/pnas.1714554115.

Full text
Abstract:
Termination of protein synthesis is triggered by the recognition of a stop codon at the ribosomal A site and is mediated by class I release factors (RFs). Whereas in bacteria, RF1 and RF2 promote termination at UAA/UAG and UAA/UGA stop codons, respectively, eukaryotes only depend on one RF (eRF1) to initiate peptide release at all three stop codons. Based on several structural as well as biochemical studies, interactions between mRNA, tRNA, and rRNA have been proposed to be required for stop codon recognition. In this study, the influence of these interactions was investigated by using chemically modified stop codons. Single functional groups within stop codon nucleotides were substituted to weaken or completely eliminate specific interactions between the respective mRNA and RFs. Our findings provide detailed insight into the recognition mode of bacterial and eukaryotic RFs, thereby revealing the chemical groups of nucleotides that define the identity of stop codons and provide the means to discriminate against noncognate stop codons or UGG sense codons.
APA, Harvard, Vancouver, ISO, and other styles
42

Zhang, Gaobo, Jian Yang, Fujun Qin, Congrui Xu, Jia Wang, Chengfeng Lei, Jia Hu, and Xiulian Sun. "A Reverse Genetics System for Cypovirus Based on a Bacmid Expressing T7 RNA Polymerase." Viruses 11, no. 4 (April 1, 2019): 314. http://dx.doi.org/10.3390/v11040314.

Full text
Abstract:
Dendrolimus punctatus cypovirus (DpCPV), belonging to the genus Cypovirus within the family Reoviridae, is considered the most destructive pest of pine forests worldwide. DpCPV has a genome consisting of 10 linear double-stranded RNA segments. To establish a reverse genetics system, we cloned cDNAs encoding the 10 genomic segments of DpCPV into three reverse genetics vectors in which each segment was transcribed under the control of a T7 RNA polymerase promoter and terminator tagged with a hepatitis delta virus ribozyme sequence. We also constructed a vp80-knockout Autographa californica multiple nucleopolyhedrovirus bacmid to express a T7 RNA polymerase codon-optimized for Sf9 cells. Following transfection of Sf9 cells with the three vectors and the bacmid, occlusion bodies (OBs) with the typical morphology of cypovirus polyhedra were observed by optical microscopy. The rescue system was verified by incorporation of a HindIII restriction enzyme site null mutant of the 9th genomic segment. Furthermore, when we co-transfected Sf9 cells with the reverse genetics vectors, the bacmid, and an additional vector bearing an egfp gene flanked with the 5′ and 3′ untranslated regions of the 10th genomic segment, aggregated green fluorescence co-localizing with the OBs was observed. The rescued OBs were able to infect Spodopetra exigua larvae, although their infectivity was significantly lower than that of wild-type DpCPV. This reverse genetics system for DpCPV could be used to explore viral replication and pathogenesis and to facilitate the development of novel bio-insecticides and expression systems for exogenous proteins.
APA, Harvard, Vancouver, ISO, and other styles
43

Nurpeisov, V., S. J. Hurwitz, and P. L. Sharma. "Fluorescent Dye Terminator Sequencing Methods for Quantitative Determination of Replication Fitness of Human Immunodeficiency Virus Type 1 Containing the Codon 74 and 184 Mutations in Reverse Transcriptase." Journal of Clinical Microbiology 41, no. 7 (July 1, 2003): 3306–11. http://dx.doi.org/10.1128/jcm.41.7.3306-3311.2003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Morozov, Igor Y., Susana Negrete-Urtasun, Joan Tilburn, Christine A. Jansen, Mark X. Caddick, and Herbert N. Arst. "Nonsense-Mediated mRNA Decay Mutation in Aspergillus nidulans." Eukaryotic Cell 5, no. 11 (September 8, 2006): 1838–46. http://dx.doi.org/10.1128/ec.00220-06.

Full text
Abstract:
ABSTRACT An Aspergillus nidulans mutation, designated nmdA1, has been selected as a partial suppressor of a frameshift mutation and shown to truncate the homologue of the Saccharomyces cerevisiae nonsense-mediated mRNA decay (NMD) surveillance component Nmd2p/Upf2p. nmdA1 elevates steady-state levels of premature termination codon-containing transcripts, as demonstrated using mutations in genes encoding xanthine dehydrogenase (hxA), urate oxidase (uaZ), the transcription factor mediating regulation of gene expression by ambient pH (pacC), and a protease involved in pH signal transduction (palB). nmdA1 can also stabilize pre-mRNA (unspliced) and wild-type transcripts of certain genes. Certain premature termination codon-containing transcripts which escape NMD are relatively stable, a feature more in common with certain nonsense codon-containing mammalian transcripts than with those in S. cerevisiae. As in S. cerevisiae, 5′ nonsense codons are more effective at triggering NMD than 3′ nonsense codons. Unlike the mammalian situation but in common with S. cerevisiae and other lower eukaryotes, A. nidulans is apparently impervious to the position of premature termination codons with respect to the 3′ exon-exon junction.
APA, Harvard, Vancouver, ISO, and other styles
45

Cohen, Sarit, Lior Kramarski, Shahar Levi, Noa Deshe, Oshrit Ben David, and Eyal Arbely. "Nonsense mutation-dependent reinitiation of translation in mammalian cells." Nucleic Acids Research 47, no. 12 (May 2, 2019): 6330–38. http://dx.doi.org/10.1093/nar/gkz319.

Full text
Abstract:
AbstractIn-frame stop codons mark the termination of translation. However, post-termination ribosomes can reinitiate translation at downstream AUG codons. In mammals, reinitiation is most efficient when the termination codon is positioned close to the 5′-proximal initiation site and around 78 bases upstream of the reinitiation site. The phenomenon was studied mainly in the context of open reading frames (ORFs) found within the 5′-untranslated region, or polycicstronic viral mRNA. We hypothesized that reinitiation of translation following nonsense mutations within the main ORF of p53 can promote the expression of N-truncated p53 isoforms such as Δ40, Δ133 and Δ160p53. Here, we report that expression of all known N-truncated p53 isoforms by reinitiation is mechanistically feasible, including expression of the previously unidentified variant Δ66p53. Moreover, we found that significant reinitiation of translation can be promoted by nonsense mutations located even 126 codons downstream of the 5′-proximal initiation site, and observed when the reinitiation site is positioned between 6 and 243 bases downstream of the nonsense mutation. We also demonstrate that reinitiation can stabilise p53 mRNA transcripts with a premature termination codon, by allowing such transcripts to evade the nonsense mediated decay pathway. Our data suggest that the expression of N-truncated proteins from alleles carrying a premature termination codon is more prevalent than previously thought.
APA, Harvard, Vancouver, ISO, and other styles
46

Väre, Ville Y. P., Ryan F. Schneider, Haein Kim, Erica Lasek-Nesselquist, Kathleen A. McDonough, and Paul F. Agris. "Small-Molecule Antibiotics Inhibiting tRNA-Regulated Gene Expression Is a Viable Strategy for Targeting Gram-Positive Bacteria." Antimicrobial Agents and Chemotherapy 65, no. 1 (October 19, 2020): e01247-20. http://dx.doi.org/10.1128/aac.01247-20.

Full text
Abstract:
ABSTRACTBacterial infections and the rise of antibiotic resistance, especially multidrug resistance, have generated a clear need for discovery of novel therapeutics. We demonstrated that a small-molecule drug, PKZ18, targets the T-box mechanism and inhibits bacterial growth. The T-box is a structurally conserved riboswitch-like gene regulator in the 5′ untranslated region (UTR) of numerous essential genes of Gram-positive bacteria. T-boxes are stabilized by cognate, unacylated tRNA ligands, allowing the formation of an antiterminator hairpin in the mRNA that enables transcription of the gene. In the absence of an unacylated cognate tRNA, transcription is halted due to the formation of a thermodynamically more stable terminator hairpin. PKZ18 targets the site of the codon-anticodon interaction of the conserved stem I and reduces T-box-controlled gene expression. Here, we show that novel analogs of PKZ18 have improved MICs, bactericidal effects against methicillin-resistant Staphylococcus aureus (MRSA), and increased efficacy in nutrient-limiting conditions. The analogs have reduced cytotoxicity against eukaryotic cells compared to PKZ18. The PKZ18 analogs acted synergistically with aminoglycosides to significantly enhance the efficacy of the analogs and aminoglycosides, further increasing their therapeutic windows. RNA sequencing showed that the analog PKZ18-22 affects expression of 8 of 12 T-box controlled genes in a statistically significant manner, but not other 5′-UTR regulated genes in MRSA. Very low levels of resistance further support the existence of multiple T-box targets for PKZ18 analogs in the cell. Together, the multiple targets, low resistance, and synergy make PKZ18 analogs promising drugs for development and future clinical applications.
APA, Harvard, Vancouver, ISO, and other styles
47

Castillo-Keller, Maria, Phu Vuong, and Rajeev Misra. "Novel Mechanism of Escherichia coli Porin Regulation." Journal of Bacteriology 188, no. 2 (January 15, 2006): 576–86. http://dx.doi.org/10.1128/jb.188.2.576-586.2006.

Full text
Abstract:
ABSTRACT A novel mechanism of Escherichia coli porin regulation was discovered from multicopy suppressors that permitted growth of cells expressing a mutant OmpC protein in the absence of DegP. Analyses of two suppressors showed that both substantially lowered OmpC expression. Suppression activities were confined to a short DNA sequence, which we designated ipeX for inhibition of porin expression, and to DNA containing a 3′-truncated ompR gene. The major effect of ipeX on ompC expression was exerted posttranscriptionally, whereas the truncated OmpR protein reduced ompC transcription. ipeX was localized within an untranslated region of 247 base pairs between the stop codon of nmpC—a remnant porin gene from the cryptic phage qsr′ (DLP12) genome—and its predicted Rho-independent transcriptional terminator. Interestingly, another prophage, PA-2, which encodes a porin similar to NmpC, known as Lc, has sequences downstream from lc identical to that of ipeX. PA-2 lysogenization leads to Lc expression and OmpC inhibition. Our data show that the synthesis of the lc transcript, whose 3′ end contains the corresponding ipeX sequence, inhibits OmpC expression. Overexpression of ipeX RNA inhibited both OmpC and OmpF expression but not that of OmpA. ompC-phoA chimeric gene constructs revealed a 248-bp untranslated region of ompC required for ipeX-mediated inhibition. However, no sequence complementarity was found between ipeX and this region of ompC, indicating that inhibition may not involve simple base pairing between the two RNA molecules. The effect of ipeX on ompC, but not on ompF, was independent of the RNA chaperone Hfq.
APA, Harvard, Vancouver, ISO, and other styles
48

Dai, Li-Shang, Sheng Li, Hui-min Yu, Guo-Qing Wei, Lei Wang, Cen Qian, Cong-Fen Zhang, et al. "Mitochondrial genome of the sweet potato hornworm, Agrius convolvuli (Lepidoptera: Sphingidae), and comparison with other Lepidoptera species." Genome 60, no. 2 (February 2017): 128–38. http://dx.doi.org/10.1139/gen-2016-0058.

Full text
Abstract:
In the present study, we sequenced the complete mitochondrial genome (mitogenome) of Agrius convolvuli (Lepidoptera: Sphingidae) and compared it with previously sequenced mitogenomes of lepidopteran species. The mitogenome was a circular molecule, 15 349 base pairs (bp) long, containing 37 genes. The order and orientation of genes in the A. convolvuli mitogenome were similar to those in sequenced mitogenomes of other lepidopterans. All 13 protein-coding genes (PCGs) were initiated by ATN codons, except for the cytochrome c oxidase subunit 1 (cox1) gene, which seemed to be initiated by the codon CGA, as observed in other lepidopterans. Three of the 13 PCGs had the incomplete termination codon T, while the remainder terminated with TAA. Additionally, the codon distributions of the 13 PCGs revealed that Asn, Ile, Leu2, Lys, Phe, and Tyr were the most frequently used codon families. All transfer RNAs were folded into the expected cloverleaf structure except for tRNASer(AGN), which lacked a stable dihydrouridine arm. The length of the adenine (A) + thymine (T)-rich region was 331 bp. This region included the motif ATAGA followed by a 19-bp poly-T stretch and a microsatellite-like (TA)8 element next to the motif ATTTA. Phylogenetic analyses (maximum likelihood and Bayesian methods) showed that A. convolvuli belongs to the family Sphingidae.
APA, Harvard, Vancouver, ISO, and other styles
49

Zhang, Jing, Xiaolei Sun, Yimei Qian, Jeffrey P. LaDuca, and Lynne E. Maquat. "At Least One Intron Is Required for the Nonsense-Mediated Decay of Triosephosphate Isomerase mRNA: a Possible Link between Nuclear Splicing and Cytoplasmic Translation." Molecular and Cellular Biology 18, no. 9 (September 1, 1998): 5272–83. http://dx.doi.org/10.1128/mcb.18.9.5272.

Full text
Abstract:
ABSTRACT Mammalian cells have established mechanisms to reduce the abundance of mRNAs that harbor a nonsense codon and prematurely terminate translation. In the case of the human triosephosphate isomerase (TPI gene), nonsense codons located less than 50 to 55 bp upstream of intron 6, the 3′-most intron, fail to mediate mRNA decay. With the aim of understanding the feature(s) of TPI intron 6 that confer function in positioning the boundary between nonsense codons that do and do not mediate decay, the effects of deleting or duplicating introns have been assessed. The results demonstrate that TPI intron 6 functions to position the boundary because it is the 3′-most intron. Since decay takes place after pre-mRNA splicing, it is conceivable that removal of the 3′-most intron from pre-mRNA “marks” the 3′-most exon-exon junction of product mRNA so that only nonsense codons located more than 50 to 55 nucleotides upstream of the “mark” mediate mRNA decay. Decay may be elicited by the failure of translating ribosomes to translate sufficiently close to the mark or, more likely, the scanning or looping out of some component(s) of the translation termination complex to the mark. In support of scanning, a nonsense codon does not elicit decay if some of the introns that normally reside downstream of the nonsense codon are deleted so the nonsense codon is located (i) too far away from a downstream intron, suggesting that all exon-exon junctions may be marked, and (ii) too far away from a downstream failsafe sequence that appears to function on behalf of intron 6, i.e., when intron 6 fails to leave a mark. Notably, the proposed scanning complex may have a greater unwinding capability than the complex that scans for a translation initiation codon since a hairpin structure strong enough to block translation initiation when inserted into the 5′ untranslated region does not block nonsense-mediated decay when inserted into exon 6 between a nonsense codon residing in exon 6 and intron 6.
APA, Harvard, Vancouver, ISO, and other styles
50

Jaafar, Fauziah Mohd, Houssam Attoui, Philippe de Micco, and Xavier de Lamballerie. "Termination and read-through proteins encoded by genome segment 9 of Colorado tick fever virus." Journal of General Virology 85, no. 8 (August 1, 2004): 2237–44. http://dx.doi.org/10.1099/vir.0.80019-0.

Full text
Abstract:
Genome segment 9 (Seg-9) of Colorado tick fever virus (CTFV) is 1884 bp long and contains a large open reading frame (ORF; 1845 nt in length overall), although a single in-frame stop codon (at nt 1052–1054) reduces the ORF coding capacity by approximately 40 %. However, analyses of highly conserved RNA sequences in the vicinity of the stop codon indicate that it belongs to a class of ‘leaky terminators’. The third nucleotide positions in codons situated both before and after the stop codon, shows the highest variability, suggesting that both regions are translated during virus replication. This also suggests that the stop signal is functionally leaky, allowing read-through translation to occur. Indeed, both the truncated ‘termination’ protein and the full-length ‘read-through’ protein (VP9 and VP9′, respectively) were detected in CTFV-infected cells, in cells transfected with a plasmid expressing only Seg-9 protein products, and in the in vitro translation products from undenatured Seg-9 ssRNA. The ratios of full-length and truncated proteins generated suggest that read-through may be down-regulated by other viral proteins. Western blot analysis of infected cells and purified CTFV showed that VP9 is a structural component of the virion, while VP9′ is a non-structural protein.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography