Stop codon

Last updated
Stop codon (red dot) of the human mitochondrial DNA MT-ATP8 gene, and start codon (blue circle) of the MT-ATP6 gene. For each nucleotide triplet (square brackets), the corresponding amino acid is given (one-letter code), either in the +1 reading frame for MT-ATP8 (in red) or in the +3 frame for MT-ATP6 (in blue). In this genomic region, the two genes overlap. Homo sapiens-mtDNA~NC 012920-ATP8+ATP6 Overlap.svg
Stop codon (red dot) of the human mitochondrial DNA MT-ATP8 gene, and start codon (blue circle) of the MT-ATP6 gene. For each nucleotide triplet (square brackets), the corresponding amino acid is given (one-letter code), either in the +1 reading frame for MT-ATP8 (in red) or in the +3 frame for MT-ATP6 (in blue). In this genomic region, the two genes overlap.

In molecular biology, a stop codon (or termination codon) is a codon (nucleotide triplet within messenger RNA) that signals the termination of the translation process of the current protein. [1] Most codons in messenger RNA correspond to the addition of an amino acid to a growing polypeptide chain, which may ultimately become a protein; stop codons signal the termination of this process by binding release factors, which cause the ribosomal subunits to disassociate, releasing the amino acid chain.

Contents

While start codons need nearby sequences or initiation factors to start translation, a stop codon alone is sufficient to initiate termination.

Properties

Standard codons

In the standard genetic code, there are three different termination codons:

Codon Standard code
(Translation table 1)
Name
DNARNA
TAGUAGSTOP = Ter(*)"amber"
TAAUAASTOP = Ter(*)"ochre"
TGAUGASTOP = Ter(*)"opal" (or "umber")

Alternative stop codons

There are variations on the standard genetic code, and alternative stop codons have been found in the mitochondrial genomes of vertebrates, [2] Scenedesmus obliquus , [3] and Thraustochytrium . [4]

Table of alternative stop codons and comparison with the standard genetic code
Genetic codeTranslation
table
CodonTranslation
with this code
Standard translation
DNARNA
Vertebrate mitochondrial 2AGAAGASTOP = Ter(*)Arg(R)
AGGAGGSTOP = Ter(*)Arg(R)
Scenedesmus obliquus mitochondrial 22TCAUCASTOP = Ter(*)Ser(S)
Thraustochytrium mitochondrial 23TTAUUASTOP = Ter(*)Leu(L)
Amino-acid biochemical propertiesNonpolarPolarBasicAcidicTermination: stop codon

Reassigned stop codons

The nuclear genetic code is flexible as illustrated by variant genetic codes that reassign standard stop codons to amino acids. [5]

Table of conditional stop codons and comparison with the standard genetic code
Genetic codeTranslation
table
CodonConditional
translation
Standard translation
DNARNA
Karyorelict nuclear 27TGAUGATer(*)orTrp(W)Ter(*)
Condylostoma nuclear 28TAAUAATer(*)orGln(Q)Ter(*)
TAGUAGTer(*)orGln(Q)Ter(*)
TGAUGATer(*)orTrp(W)Ter(*)
Blastocrithidia nuclear 31TAAUAATer(*)orGlu(E)Ter(*)
TAGUAGTer(*)orGlu(E)Ter(*)

Translation

In 1986, convincing evidence was provided that selenocysteine (Sec) was incorporated co-translationally. Moreover, the codon partially directing its incorporation in the polypeptide chain was identified as UGA also known as the opal termination codon. [6] Different mechanisms for overriding the termination function of this codon have been identified in prokaryotes and in eukaryotes. [7] A particular difference between these kingdoms is that cis elements seem restricted to the neighborhood of the UAG codon in prokaryotes while in eukaryotes this restriction is not present. Instead such locations seem disfavored albeit not prohibited. [8]

In 2003, a landmark paper described the identification of all known selenoproteins in humans: 25 in total. [9] Similar analyses have been run for other organisms.

The UAG codon can translate into pyrrolysine (Pyl) in a similar manner.

Genomic distribution

Distribution of stop codons within the genome of an organism is non-random and can correlate with GC-content. [10] [11] For example, the E. coli K-12 genome contains 2705 TAA (63%), 1257 TGA (29%), and 326 TAG (8%) stop codons (GC content 50.8%). [12] Also the substrates for the stop codons release factor 1 or release factor 2 are strongly correlated to the abundance of stop codons. [11] Large scale study of bacteria with a broad range of GC-contents shows that while the frequency of occurrence of TAA is negatively correlated to the GC-content and the frequency of occurrence of TGA is positively correlated to the GC-content, the frequency of occurrence of the TAG stop codon, which is often the minimally used stop codon in a genome, is not influenced by the GC-content. [13]

Recognition

Recognition of stop codons in bacteria have been associated with the so-called 'tripeptide anticodon', [14] a highly conserved amino acid motif in RF1 (PxT) and RF2 (SPF). Even though this is supported by structural studies, it was shown that the tripeptide anticodon hypothesis is an oversimplification. [15]

Nomenclature

Stop codons were historically given many different names, as they each corresponded to a distinct class of mutants that all behaved in a similar manner. These mutants were first isolated within bacteriophages (T4 and lambda), viruses that infect the bacteria Escherichia coli . Mutations in viral genes weakened their infectious ability, sometimes creating viruses that were able to infect and grow within only certain varieties of E. coli.

amber mutations (UAG)

They were the first set of nonsense mutations to be discovered, isolated by Richard H. Epstein and Charles Steinberg and named after their friend and graduate Caltech student Harris Bernstein, whose last name means "amber" in German (cf. Bernstein). [16] [17]

Viruses with amber mutations are characterized by their ability to infect only certain strains of bacteria, known as amber suppressors. These bacteria carry their own mutation that allows a recovery of function in the mutant viruses. For example, a mutation in the tRNA that recognizes the amber stop codon allows translation to "read through" the codon and produce a full-length protein, thereby recovering the normal form of the protein and "suppressing" the amber mutation. [18] Thus, amber mutants are an entire class of virus mutants that can grow in bacteria that contain amber suppressor mutations. Similar suppressors are known for ochre and opal stop codons as well.

tRNA molecules carrying unnatural aminoacids have been designed to recognize the amber stop codon in bacterial RNA. This technology allows for incorporation of orthogonal aminoacids (such as p-azidophenylalanine) at specific locations of the target protein.

ochre mutations (UAA)

It was the second stop codon mutation to be discovered. Reminiscent of the usual yellow-orange-brown color associated with amber, this second stop codon was given the name of "ochre", an orange-reddish-brown mineral pigment. [17]

Ochre mutant viruses had a property similar to amber mutants in that they recovered infectious ability within certain suppressor strains of bacteria. The set of ochre suppressors was distinct from amber suppressors, so ochre mutants were inferred to correspond to a different nucleotide triplet. Through a series of mutation experiments comparing these mutants with each other and other known amino acid codons, Sydney Brenner concluded that the amber and ochre mutations corresponded to the nucleotide triplets "UAG" and "UAA". [19]

opal or umber mutations (UGA)

The third and last stop codon in the standard genetic code was discovered soon after, and corresponds to the nucleotide triplet "UGA". [20]

To continue matching with the theme of colored minerals, the third nonsense codon came to be known as "opal", which is a type of silica showing a variety of colors. [17] Nonsense mutations that created this premature stop codon were later called opal mutations or umber mutations.

Mutations and disease

Nonsense

Nonsense mutations are changes in DNA sequence that introduce a premature stop codon, causing any resulting protein to be abnormally shortened. This often causes a loss of function in the protein, as critical parts of the amino acid chain are no longer assembled. Because of this terminology, stop codons have also been referred to as nonsense codons.

Nonstop

A nonstop mutation, also called a stop-loss variant, is a point mutation that occurs within a stop codon. Nonstop mutations cause the continued translation of an mRNA strand into what should be an untranslated region. Most polypeptides resulting from a gene with a nonstop mutation lose their function due to their extreme length and the impact on normal folding. Nonstop mutations differ from nonsense mutations in that they do not create a stop codon but, instead, delete one. Nonstop mutations also differ from missense mutations, which are point mutations where a single nucleotide is changed to cause replacement by a different amino acid. Nonstop mutations have been linked with many inherited diseases including endocrine disorders, [21] eye disease, [22] and neurodevelopmental disorders. [23] [24]

Hidden stops

An example of a single base deletion forming a stop codon. Frameshift deletion (13062713935).jpg
An example of a single base deletion forming a stop codon.

Hidden stops are non-stop codons that would be read as stop codons if they were frameshifted +1 or −1. These prematurely terminate translation if the corresponding frame-shift (such as due to a ribosomal RNA slip) occurs before the hidden stop. It is hypothesised that this decreases resource wastage on nonfunctional proteins and the production of potential cytotoxins. Researchers at Louisiana State University propose the ambush hypothesis , that hidden stops are selected for. Codons that can form hidden stops are used in genomes more frequently compared to synonymous codons that would otherwise code for the same amino acid. Unstable rRNA in an organism correlates with a higher frequency of hidden stops. [25] However, this hypothesis could not be validated with a larger data set. [26]

Stop-codons and hidden stops together are collectively referred as stop-signals. Researchers at University of Memphis found that the ratios of the stop-signals on the three reading frames of a genome (referred to as translation stop-signals ratio or TSSR) of genetically related bacteria, despite their great differences in gene contents, are much alike. This nearly identical genomic-TSSR value of genetically related bacteria may suggest that bacterial genome expansion is limited by their unique stop-signals bias of that bacterial species. [27]

Translational readthrough

Stop codon suppression or translational readthrough occurs when in translation a stop codon is interpreted as a sense codon, that is, when a (standard) amino acid is 'encoded' by the stop codon. Mutated tRNAs can be the cause of readthrough, but also certain nucleotide motifs close to the stop codon. Translational readthrough is very common in viruses and bacteria, and has also been found as a gene regulatory principle in humans, yeasts, bacteria and drosophila. [28] [29] This kind of endogenous translational readthrough constitutes a variation of the genetic code, because a stop codon codes for an amino acid. In the case of human malate dehydrogenase, the stop codon is read through with a frequency of about 4%. [30] The amino acid inserted at the stop codon depends on the identity of the stop codon itself: Gln, Tyr, and Lys have been found for the UAA and UAG codons, while Cys, Trp, and Arg for the UGA codon have been identified by mass spectrometry. [31] Extent of readthrough in mammals have widely variable extents, and can broadly diversify the proteome and affect cancer progression. [32]

Use as a watermark

In 2010, when Craig Venter unveiled the first fully functioning, reproducing cell controlled by synthetic DNA he described how his team used frequent stop codons to create watermarks in RNA and DNA to help confirm the results were indeed synthetic (and not contaminated or otherwise), using it to encode authors' names and website addresses. [33]

See also

Related Research Articles

<span class="mw-page-title-main">Genetic code</span> Rules by which information encoded within genetic material is translated into proteins

The genetic code is the set of rules used by living cells to translate information encoded within genetic material into proteins. Translation is accomplished by the ribosome, which links proteinogenic amino acids in an order specified by messenger RNA (mRNA), using transfer RNA (tRNA) molecules to carry amino acids and to read the mRNA three nucleotides at a time. The genetic code is highly similar among all organisms and can be expressed in a simple table with 64 entries.

<span class="mw-page-title-main">Mutation</span> Alteration in the nucleotide sequence of a genome

In biology, a mutation is an alteration in the nucleic acid sequence of the genome of an organism, virus, or extrachromosomal DNA. Viral genomes contain either DNA or RNA. Mutations result from errors during DNA or viral replication, mitosis, or meiosis or other types of damage to DNA, which then may undergo error-prone repair, cause an error during other forms of repair, or cause an error during replication. Mutations may also result from insertion or deletion of segments of DNA due to mobile genetic elements.

The coding region of a gene, also known as the coding sequence (CDS), is the portion of a gene's DNA or RNA that codes for a protein. Studying the length, composition, regulation, splicing, structures, and functions of coding regions compared to non-coding regions over different species and time periods can provide a significant amount of important information regarding gene organization and evolution of prokaryotes and eukaryotes. This can further assist in mapping the human genome and developing gene therapy.

<span class="mw-page-title-main">Codon usage bias</span> Genetic bias in coding DNA

Codon usage bias refers to differences in the frequency of occurrence of synonymous codons in coding DNA. A codon is a series of three nucleotides that encodes a specific amino acid residue in a polypeptide chain or for the termination of translation.

<span class="mw-page-title-main">Translation (biology)</span> Cellular process of protein synthesis

In biology, translation is the process in living cells in which proteins are produced using RNA molecules as templates. The generated protein is a sequence of amino acids. This sequence is determined by the sequence of nucleotides in the RNA. The nucleotides are considered three at a time. Each such triple results in addition of one specific amino acid to the protein being generated. The matching from nucleotide triple to amino acid is called the genetic code. The translation is performed by a large complex of functional RNA and proteins called ribosomes. The entire process is called gene expression.

<span class="mw-page-title-main">Frameshift mutation</span> Mutation that shifts codon alignment

A frameshift mutation is a genetic mutation caused by indels of a number of nucleotides in a DNA sequence that is not divisible by three. Due to the triplet nature of gene expression by codons, the insertion or deletion can change the reading frame, resulting in a completely different translation from the original. The earlier in the sequence the deletion or insertion occurs, the more altered the protein. A frameshift mutation is not the same as a single-nucleotide polymorphism in which a nucleotide is replaced, rather than inserted or deleted. A frameshift mutation will in general cause the reading of the codons after the mutation to code for different amino acids. The frameshift mutation will also alter the first stop codon encountered in the sequence. The polypeptide being created could be abnormally short or abnormally long, and will most likely not be functional.

<span class="mw-page-title-main">Point mutation</span> Replacement, insertion, or deletion of a single DNA or RNA nucleotide

A point mutation is a genetic mutation where a single nucleotide base is changed, inserted or deleted from a DNA or RNA sequence of an organism's genome. Point mutations have a variety of effects on the downstream protein product—consequences that are moderately predictable based upon the specifics of the mutation. These consequences can range from no effect to deleterious effects, with regard to protein production, composition, and function.

In genetics, a nonsense mutation is a point mutation in a sequence of DNA that results in a nonsense codon, or a premature stop codon in the transcribed mRNA, and leads to a truncated, incomplete, and possibly nonfunctional protein product. Nonsense mutations are not always harmful; the functional effect of a nonsense mutation depends on many aspects, such as the location of the stop codon within the coding DNA. For example, the effect of a nonsense mutation depends on the proximity of the nonsense mutation to the original stop codon, and the degree to which functional subdomains of the protein are affected. As nonsense mutations leads to premature termination of polypeptide chains; they are also called chain termination mutations.

<span class="mw-page-title-main">Silent mutation</span> DNA mutation with no observable effect on an organisms phenotype

Silent mutations are mutations in DNA that do not have an observable effect on the organism's phenotype. They are a specific type of neutral mutation. The phrase silent mutation is often used interchangeably with the phrase synonymous mutation; however, synonymous mutations are not always silent, nor vice versa. Synonymous mutations can affect transcription, splicing, mRNA transport, and translation, any of which could alter phenotype, rendering the synonymous mutation non-silent. The substrate specificity of the tRNA to the rare codon can affect the timing of translation, and in turn the co-translational folding of the protein. This is reflected in the codon usage bias that is observed in many species. Mutations that cause the altered codon to produce an amino acid with similar functionality are often classified as silent; if the properties of the amino acid are conserved, this mutation does not usually significantly affect protein function.

In genetics, a missense mutation is a point mutation in which a single nucleotide change results in a codon that codes for a different amino acid. It is a type of nonsynonymous substitution.

Genetics, a discipline of biology, is the science of heredity and variation in living organisms.

<span class="mw-page-title-main">Synonymous substitution</span>

A synonymous substitution is the evolutionary substitution of one base for another in an exon of a gene coding for a protein, such that the produced amino acid sequence is not modified. This is possible because the genetic code is "degenerate", meaning that some amino acids are coded for by more than one three-base-pair codon; since some of the codons for a given amino acid differ by just one base pair from others coding for the same amino acid, a mutation that replaces the "normal" base by one of the alternatives will result in incorporation of the same amino acid into the growing polypeptide chain when the gene is translated. Synonymous substitutions and mutations affecting noncoding DNA are often considered silent mutations; however, it is not always the case that the mutation is silent.

<span class="mw-page-title-main">Start codon</span> First codon of a messenger RNA translated by a ribosome

The start codon is the first codon of a messenger RNA (mRNA) transcript translated by a ribosome. The start codon always codes for methionine in eukaryotes and archaea and a N-formylmethionine (fMet) in bacteria, mitochondria and plastids.

A suppressor mutation is a second mutation that alleviates or reverts the phenotypic effects of an already existing mutation in a process defined synthetic rescue. Genetic suppression therefore restores the phenotype seen prior to the original background mutation. Suppressor mutations are useful for identifying new genetic sites which affect a biological process of interest. They also provide evidence between functionally interacting molecules and intersecting biological pathways.

A nonsense suppressor is a factor which can inhibit the effect of the nonsense mutation. Nonsense suppressors can be generally divided into two classes: a) a mutated tRNA which can bind with a termination codon on mRNA; b) a mutation on ribosomes decreasing the effect of a termination codon. It is believed that nonsense suppressors keep a low concentration in the cell and do not disrupt normal translation most of the time. In addition, many genes do not have only one termination codon, and cells commonly use ochre codons as the termination signal, whose nonsense suppressors are usually inefficient.

Missense mRNA is a messenger RNA bearing one or more mutated codons that yield polypeptides with an amino acid sequence different from the wild-type or naturally occurring polypeptide. Missense mRNA molecules are created when template DNA strands or the mRNA strands themselves undergo a missense mutation in which a protein coding sequence is mutated and an altered amino acid sequence is coded for.

<span class="mw-page-title-main">Expanded genetic code</span> Modified genetic code

An expanded genetic code is an artificially modified genetic code in which one or more specific codons have been re-allocated to encode an amino acid that is not among the 22 common naturally-encoded proteinogenic amino acids.

A nonsynonymous substitution is a nucleotide mutation that alters the amino acid sequence of a protein. Nonsynonymous substitutions differ from synonymous substitutions, which do not alter amino acid sequences and are (sometimes) silent mutations. As nonsynonymous substitutions result in a biological change in the organism, they are subject to natural selection.

<span class="mw-page-title-main">DNA and RNA codon tables</span> List of standard rules to translate DNA encoded information into proteins

A codon table can be used to translate a genetic code into a sequence of amino acids. The standard genetic code is traditionally represented as an RNA codon table, because when proteins are made in a cell by ribosomes, it is messenger RNA (mRNA) that directs protein synthesis. The mRNA sequence is determined by the sequence of genomic DNA. In this context, the standard genetic code is referred to as translation table 1. It can also be represented in a DNA codon table. The DNA codons in such tables occur on the sense DNA strand and are arranged in a 5-to-3 direction. Different tables with alternate codons are used depending on the source of the genetic code, such as from a cell nucleus, mitochondrion, plastid, or hydrogenosome.

This glossary of cellular and molecular biology is a list of definitions of terms and concepts commonly used in the study of cell biology, molecular biology, and related disciplines, including genetics, biochemistry, and microbiology. It is split across two articles:

References

  1. Griffiths AJF, Miller JH, Suzuki DT, Lewontin RC, Gelbart WM (2000). "Chapter 10 (Molecular Biology of Gene Function): Genetic code: Stop codons". An Introduction to Genetic Analysis. W.H. Freeman and Company.
  2. Barrell, B. G.; Bankier, A. T.; Drouin, J. (1979-11-08). "A different genetic code in human mitochondria". Nature. 282 (5735): 189–194. Bibcode:1979Natur.282..189B. doi:10.1038/282189a0. ISSN   0028-0836. PMID   226894. S2CID   4335828.
  3. A. M. Nedelcu, R. W. Lee, G. Lemieux, M. W. Gray, G. Burger (June 2000). "The complete mitochondrial DNA sequence of Scenedesmus obliquus reflects an intermediate stage in the evolution of the green algal mitochondrial genome". Genome Research. 10 (6): 819–831. doi:10.1101/gr.10.6.819. PMC   310893 . PMID   10854413.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  4. Wideman, Jeremy G.; Monier, Adam; Rodríguez-Martínez, Raquel; Leonard, Guy; Cook, Emily; Poirier, Camille; Maguire, Finlay; Milner, David S.; Irwin, Nicholas A. T.; Moore, Karen; Santoro, Alyson E. (2019-11-25). "Unexpected mitochondrial genome diversity revealed by targeted single-cell genomics of heterotrophic flagellated protists". Nature Microbiology. 5 (1): 154–165. doi:10.1038/s41564-019-0605-4. hdl: 10871/39819 . ISSN   2058-5276. PMID   31768028. S2CID   208279678.
  5. Swart, Estienne Carl; Serra, Valentina; Petroni, Giulio; Nowacki, Mariusz (2016). "Genetic Codes with No Dedicated Stop Codon: Context-Dependent Translation Termination". Cell. 166 (3): 691–702. doi:10.1016/j.cell.2016.06.020. PMC   4967479 . PMID   27426948.
  6. Zinoni, F; Birkmann, A; Stadtman, T; Böck, A (1986). "Nucleotide sequence and expression of the selenocysteine-containing polypeptide of formate dehydrogenase (formate-hydrogen-lyase-linked) from Escherichia coli". Proceedings of the National Academy of Sciences. 83 (13): 4650–4654. Bibcode:1986PNAS...83.4650Z. doi: 10.1073/pnas.83.13.4650 . PMC   323799 . PMID   2941757.
  7. Böck, A (2013). "Selenoprotein Synthesis". Encyclopedia of Biological Chemistry. pp. 210–213. doi:10.1016/B978-0-12-378630-2.00025-6. ISBN   9780123786319 . Retrieved 23 August 2021.
  8. Mix, H; Lobanov, A; Gladyshev, V (2007). "SECIS elements in the coding regions of selenoprotein transcripts are functional in higher eukaryotes". Nucleic Acids Research. 35 (2): 414–423. doi:10.1093/nar/gkl1060. PMC   1802603 . PMID   17169995.
  9. Kryukov, G; Gladyshev, V (2003). "Characterization of mammalian selenoproteomes". Science. 300 (5624): 1439–1443. Bibcode:2003Sci...300.1439K. doi:10.1126/science.1083516. PMID   12775843. S2CID   10363908.
  10. Povolotskaya IS, Kondrashov FA, Ledda A, Vlasov PK (2012). "Stop codons in bacteria are not selectively equivalent". Biology Direct. 7: 30. doi: 10.1186/1745-6150-7-30 . PMC   3549826 . PMID   22974057.
  11. 1 2 Korkmaz, Gürkan; Holm, Mikael; Wiens, Tobias; Sanyal, Suparna (2014). "Comprehensive Analysis of Stop Codon Usage in Bacteria and Its Correlation with Release Factor Abundance". The Journal of Biological Chemistry. 289 (44): 775–806. doi: 10.1074/jbc.M114.606632 . PMC   4215218 . PMID   25217634.
  12. "Escherichia coli str. K-12 substr. MG1655, complete genome [Genbank Accession Number: U00096]". GenBank. NCBI. Retrieved 2013-01-27.
  13. Wong, Tit-Yee; Fernandes, Sanjit; Sankhon, Naby; Leong, Patrick P; Kuo, Jimmy; Liu, Jong-Kang (2008). "Role of Premature Stop Codons in Bacterial Evolution". Journal of Bacteriology. 190 (20): 6718–6725. doi:10.1128/JB.00682-08. PMC   2566208 . PMID   18708500.
  14. Ito, Koichi; Uno, Makiko; Nakamura, Yoshikazu (1999). "A tripeptide 'anticodon' deciphers stop codons in messenger RNA". Nature. 403 (6770): 680–684. doi:10.1038/35001115. PMID   10688208. S2CID   4331695.
  15. Korkmaz, Gürkan; Sanyal, Suparna (2017). "R213I mutation in release factor 2 (RF2) is one step forward for engineering an omnipotent release factor in bacteria Escherichia coli". Journal of Biological Chemistry. 292 (36): 15134–15142. doi: 10.1074/jbc.M117.785238 . PMC   5592688 . PMID   28743745.
  16. Stahl FW (1995). "The amber mutants of phage T4". Genetics. 141 (2): 439–442. doi:10.1093/genetics/141.2.439. PMC   1206745 . PMID   8647382.
  17. 1 2 3 Lewin, Benjamin; Krebs, Jocelyn E.; Goldstein, Elliott S.; Kilpatrick, Stephen T. (2011-04-18). Lewin's Essential GENES. Jones & Bartlett Publishers. ISBN   978-1-4496-4380-5.
  18. Robin Cook. "Amber, Ocher, and Opal Mutations Summary". World of Genetics. Gale.
  19. Brenner, S.; Stretton, A. O. W.; Kaplan, S. (1965). "Genetic Code: The 'Nonsense' Triplets for Chain Termination and their Suppression". Nature. 206 (4988): 994–8. Bibcode:1965Natur.206..994B. doi:10.1038/206994a0. PMID   5320272. S2CID   28502898.
  20. Brenner, S.; Barnett, L.; Katz, E. R.; Crick, F. H. C. (1967). "UGA: A Third Nonsense Triplet in the Genetic Code". Nature. 213 (5075): 449–50. Bibcode:1967Natur.213..449B. doi:10.1038/213449a0. PMID   6032223. S2CID   4211867.
  21. Pang S.; Wang W.; et al. (2002). "A novel nonstop mutation in the stop codon and a novel missense mutation in the type II 3beta-hydroxysteroid dehydrogenase (3beta-HSD) gene causing, respectively, nonclassic and classic 3beta-HSD deficiency congenital adrenal hyperplasia". J Clin Endocrinol Metab. 87 (6): 2556–63. doi: 10.1210/jcem.87.6.8559 . PMID   12050213.
  22. Doucette, L.; et al. (2011). "A novel, non-stop mutation in FOXE3 causes an autosomal dominant form of variable anterior segment dysgenesis including Peters anomaly". European Journal of Human Genetics. 19 (3): 293–299. doi:10.1038/ejhg.2010.210. PMC   3062009 . PMID   21150893.
  23. Torres-Torronteras, J.; Rodriguez-Palmero, A.; et al. (2011). "A novel nonstop mutation in TYMP does not induce nonstop mRNA decay in a MNGIE patient with severe neuropathy" (PDF). Hum. Mutat. 32 (4): E2061–E2068. doi:10.1002/humu.21447. PMID   21412940. S2CID   24446773.
  24. Spaull, R; Steel, D; Barwick, K; Prabhakar, P; Wakeling, E; Kurian, MA (2022-07-23). "STXBP1 Stop‐Loss Mutation Associated with Complex Early Onset Movement Disorder without Epilepsy". Movement Disorders Clinical Practice. 9 (6): 837–840. doi:10.1002/mdc3.13509. ISSN   2330-1619. PMC   9346254 . PMID   35937496.
  25. Seligmann, Hervé; Pollock, David D. (2004). "The Ambush Hypothesis: Hidden Stop Codons Prevent Off-Frame Gene Reading". DNA and Cell Biology. 23 (10): 701–5. doi:10.1089/1044549042476910. PMID   15585128.
  26. Cavalcanti, Andre; Chang, Charlotte H.; Morgens, David W. (2013). "Ambushing the ambush hypothesis: predicting and evaluating off-frame codon frequencies in Prokaryotic Genomes". BMC Genomics. 14 (418): 1–8. doi: 10.1186/1471-2164-14-418 . PMC   3700767 . PMID   23799949.
  27. Wong, Tit-Yee; Schwartzbach, Steve (2015). "Protein mis-termination initiates genetic diseases, cancers, and restricts bacterial genome expansion". Journal of Environmental Science and Health, Part C. 33 (3): 255–85. doi:10.1080/10590501.2015.1053461. PMID   26087060. S2CID   20380447.
  28. Namy O, Rousset JP, Napthine S, Brierley I (2004). "Reprogrammed genetic decoding in cellular gene expression". Molecular Cell. 13 (2): 157–68. doi: 10.1016/S1097-2765(04)00031-0 . PMID   14759362.
  29. Schueren F, Lingner T, George R, Hofhuis J, Gartner J, Thoms S (2014). "Peroxisomal lactate dehydrogenase is generated by translational readthrough in mammals". eLife. 3: e03640. doi: 10.7554/eLife.03640 . PMC   4359377 . PMID   25247702.
  30. Hofhuis J, Schueren F, Nötzel C, Lingner T, Gärtner J, Jahn O, Thoms S (2016). "The functional readthrough extension of malate dehydrogenase reveals a modification of the genetic code". Open Biol. 6 (11): 160246. doi:10.1098/rsob.160246. PMC   5133446 . PMID   27881739.
  31. Blanchet S, Cornu D, Argentini M, Namy O (2014). "New insights into the incorporation of natural suppressor tRNAs at stop codons in Saccharomyces cerevisiae". Nucleic Acids Res. 42 (15): 10061–72. doi:10.1093/nar/gku663. PMC   4150775 . PMID   25056309.
  32. Ghosh, Souvik; Guimaraes, Joao C; Lanzafame, Manuela; Schmidt, Alexander; Syed, Afzal Pasha; Dimitriades, Beatrice; Börsch, Anastasiya; Ghosh, Shreemoyee; Mittal, Nitish; Montavon, Thomas; Correia, Ana Luisa; Danner, Johannes; Meister, Gunter; Terracciano, Luigi M; Pfeffer, Sébastien; Piscuoglio, Salvatore; Zavolan, Mihaela (15 September 2020). "Prevention of dsRNA‐induced interferon signaling by AGO1x is linked to breast cancer cell proliferation". The EMBO Journal. 39 (18): e103922. doi:10.15252/embj.2019103922. PMC   7507497 . PMID   32812257.
  33. "Watch me unveil "synthetic life"". 21 May 2010.