Fact-checked by Grok 2 weeks ago

Stop codon

A stop codon, also known as a termination codon, is a trinucleotide sequence in (mRNA) that signals the end of protein synthesis during in the . In the standard , there are three stop codons: UAA (), UAG (), and UGA (), which do not encode any but instead instruct the to halt polypeptide chain . These codons are essential for defining the precise length of proteins, ensuring that translation terminates correctly at the intended positions specified by the mRNA sequence. During the translation process, when a stop codon occupies the A site of the , it is recognized by specialized protein s rather than (tRNA) molecules. In prokaryotes (), release factor 1 (RF1) binds to UAA or UAG, while release factor 2 (RF2) binds to UAA or UGA; release factor 3 (RF3) assists in the process. In eukaryotes, eukaryotic release factor 1 (eRF1) recognizes all three stop codons, aided by eRF3. These factors facilitate the of the bond linking the completed polypeptide to the tRNA in the and releasing the newly synthesized protein from the . This mechanism prevents erroneous continuation of beyond the coding sequence, maintaining the fidelity of . Stop codons play a critical role in the , where 61 of the 64 possible trinucleotides specify and the remaining three serve as termination signals, reflecting the evolutionary optimization of the code for efficient . that introduce premature stop codons, known as mutations, can lead to truncated, non-functional proteins and are associated with various genetic disorders, underscoring their importance in and . While the standard code is nearly universal, some organisms and organelles exhibit variations where stop codons are reassigned to code for , highlighting the plasticity of termination.

Definition and Fundamentals

Role in Protein Synthesis

Stop codons are nucleotide triplets in messenger RNA (mRNA) that signal the end of the coding sequence during protein synthesis, specifically UAA, UAG, and UGA, which do not specify any amino acid. These codons function as termination signals in the process of translation, where the ribosome reads the mRNA sequence to assemble amino acids into a polypeptide chain. Upon encountering a stop codon in the ribosomal A site, translation halts, and the completed polypeptide is released from the ribosome, marking the conclusion of protein synthesis. In the structure of the , which consists of 64 possible codons formed by the four bases (A, U, G, C) in groups of three, 61 codons encode while the three stop codons occupy specific positions without corresponding transfer RNAs (tRNAs) to deliver . The absence of tRNAs complementary to stop codons ensures that these sequences cannot recruit , instead directing the translational machinery to terminate efficiently. This design allows precise control over protein length, as the stop codon defines the boundary of the in the mRNA. The role of stop codons as termination signals was elucidated in the early 1960s through genetic studies in bacteriophage T4, led by , Richard Garen, and colleagues, who identified UAA, UAG, and UGA as nonsense mutations causing premature chain termination. Concurrently, Marshall Nirenberg and his team at the used cell-free systems and synthetic polynucleotides to demonstrate that these result in termination of polypeptide . Their 1964 filter-binding with Philip further confirmed that no tRNAs bind to these codons, supporting their assignment as stop signals and contributing to the full deciphering of the by 1966. This work on the earned Nirenberg the 1968 Nobel Prize in Physiology or Medicine.

Standard Stop Codon Sequences

In the standard genetic code, the three stop codons are UAA (also known as ), UAG (), and UGA (opal or umber). These triplets occur in (mRNA) and direct the to terminate protein synthesis. Their DNA counterparts in the genome are TAA, TAG, and TGA, respectively. All three stop codons share a common : they begin with the dinucleotide UA or UG and end with a purine base, either A or G. This configuration facilitates specific recognition by class I release factors (RF1 or RF2 in prokaryotes, eRF1 in eukaryotes) that bind directly to the ribosomal A site, rather than by aminoacyl-tRNAs. Consequently, the wobble base pairing mechanism—typically operative at the third position of sense codons to allow flexible anticodon-codon interactions—does not apply to stop codons, as no tRNA decoding occurs. Unlike the 61 sense codons that specify , stop codons act as punctuation signals within the mRNA , precisely delineating the conclusion of the (ORF) and halting ribosomal translocation to avoid inappropriate translation of the downstream 3' (UTR). In the , stop codon usage among protein-coding genes shows a toward UGA at approximately 50%, followed by UAA at 28% and UAG at 22%, patterns shaped by , translational efficiency, and selective forces in eukaryotes.

Variations Across Genetic Codes

Alternative Stop Codons

In various non-universal genetic codes, certain codons that encode in the standard code function as stop signals, diverging from the UAA, UAG, and UGA terminators. These alternative stop codons typically arise in organelles or specialized lineages where codon reassignments optimize efficiency or adapt to genomic constraints, such as reduced tRNA sets. For instance, in the mitochondrial genetic code of vertebrates, the codons and AGG, which specify in the nuclear standard code, serve as stop codons alongside UAA and UAG, while UGA codes for . This reassignment expands the set of terminators to four, facilitating precise protein synthesis in compact mitochondrial genomes. Similar variations occur in other mitochondrial systems. In the green alga Scenedesmus obliquus, the codon UCA acts as a stop signal in addition to the standard UAA, UAG, and UGA, representing a rare case where a serine-encoding codon in the universal code is repurposed for termination. In eukaryotic cells broadly, UGA retains its primary role as a stop codon but can be contextually recoded to incorporate () at specific sites via a specialized (SELB or eEFSec) and a SECIS in the mRNA, though it functions as a terminator elsewhere in the . Bacterial lineages exhibit exceptions where standard stop codons are reassigned, necessitating reliance on alternatives for termination. In species, such as Mycoplasma capricolum, UGA encodes instead of serving as a stop, leaving UAA and UAG as the sole terminators; this deviation is supported by a dedicated tRNA^Trp with a UCA anticodon that decodes UGA. Such reassignments reduce the number of stop signals to two, potentially increasing susceptibility to but aligning with the bacteria's AT-rich, minimal genomes. These alternative stop codons often emerge evolutionarily through sense-to-stop reassignments, typically following the loss of tRNAs for low-usage codons, which allows unassigned to be captured as terminators without disrupting essential proteins. This process is facilitated by ambiguous decoding phases where codons transition from sense to stop functions. Stop codon variations, including such alternatives, are documented in approximately 60% of known genetic codes, highlighting their role in code diversification while preserving overall translational fidelity.

Reassigned and Non-Standard Stop Codons

In certain organisms, standard stop codons are naturally reassigned to encode non-standard amino acids through specialized translational machinery. For instance, the opal codon UGA, which typically signals termination, is recoded to incorporate selenocysteine (Sec), the 21st amino acid, in eukaryotes and some bacteria. This process relies on a dedicated selenocysteine tRNA (tRNASec) that recognizes UGA, paired with a selenocysteine insertion sequence (SECIS) element in the mRNA's 3' untranslated region, which recruits a specialized elongation factor to promote Sec insertion over termination. Similarly, the amber codon UAG is reassigned to encode pyrrolysine (Pyl), the 22nd amino acid, in methanogenic archaea such as species of the genus Methanosarcina. This reassignment is facilitated by the pylT gene, which encodes a unique tRNAPyl that decodes UAG, along with a dedicated pyrrolysyl-tRNA synthetase that charges the tRNA with Pyl, enabling its role in methylamine metabolism. A more extreme variation occurs in certain , such as Condylostoma magnum and Parduczia sp., where all three standard stop codons (UAA, UAG, UGA) are reassigned to —typically UAA and UAG to , and UGA to —resulting in no dedicated stop codons across all 64 triplets. In these codes, translation termination is context-dependent, relying on mRNA features like 3' end structures or poly(A) tails to signal release, with efficiency around 90-98% and minimal readthrough (mean <1.8%). Such reassignments are rare in nature but can have pathogenic implications in humans. In some cancers, aberrant suppression of the opal codon UGA occurs, allowing translational readthrough that produces extended protein isoforms and alters protein function, potentially contributing to disease progression. In synthetic biology, stop codons are deliberately reassigned to expand the genetic code for incorporating unnatural amino acids (UAAs) into proteins. The amber codon UAG is commonly suppressed using orthogonal tRNA-aminoacyl-tRNA synthetase pairs, which are engineered to be independent of the host's machinery; these pairs charge the suppressor tRNA with a desired UAA, enabling site-specific insertion during translation. This approach has been widely adopted to introduce over 40 UAAs, such as photocrosslinkers or fluorescent probes, for applications in protein engineering and therapeutics. Comparative genomics methods, including computational screening of codon usage and phylogenetic analysis across thousands of genomes, have been instrumental in detecting these reassignments. Such studies reveal that natural stop codon reassignments occur in a small fraction—estimated at less than 1% for specific variants like pyrrolysine—of sequenced microbial genomes, highlighting their evolutionary rarity and context-specific adaptation.

Molecular Recognition and Termination

Recognition by Release Factors

In bacteria, translation termination is initiated when a stop codon enters the ribosomal A-site, recruiting class I release factors RF1 or RF2 for specific recognition. RF1 decodes UAA and UAG codons, while RF2 decodes UAA and UGA codons, ensuring precise identification of termination signals. These factors bind as ternary complexes with RF3, a GTPase that enhances dissociation but does not directly participate in codon recognition. The specificity of stop codon recognition by RF1 and RF2 arises from a conserved tripeptide motif in their N-terminal domain, which mimics the anticodon loop of tRNA and inserts into the ribosomal decoding center. In RF1, the PAT (Pro-Ala-Thr) motif forms hydrogen bonds with the first two nucleotides of UAA or UAG, while in RF2, the SPF (Ser-Pro-Phe) motif interacts similarly with UAA or UGA, discriminating against sense codons by over six orders of magnitude. Upon binding, the release factor's domain 2 and 3 position the conserved GGQ motif near the peptidyl transferase center (PTC), but initial recognition triggers a conformational shift in the 30S subunit, stabilizing the A-site interaction without immediate hydrolysis. In eukaryotes, a single class I release factor, eRF1, recognizes all three stop codons (UAA, UAG, and UGA) in the ribosomal A-site, forming a ternary complex with eRF3·GTP to facilitate binding. The N-domain of eRF1 contains a flexible mini-domain that adopts distinct conformations to accommodate any stop codon, with key residues like Tyr125 and Gln184 forming universal interactions via a GTS motif analogous to bacterial tripeptides. eRF3, like bacterial RF3, acts as a GTPase to promote eRF1 recruitment and conformational activation, but eRF1's broader specificity enables omnipotent decoding across eukaryotic lineages. Release factors exhibit structural conservation across bacteria, archaea, and eukaryotes, with the N-domain serving as the core for stop codon recognition despite sequence divergence. Crystal structures, such as those of the Thermus thermophilus 70S ribosome bound to RF1/RF2 (PDB: 3D5A, 2WH1), reveal conserved interactions between the RF N-domain and the A-site helix 44 of 16S rRNA, highlighting key residues like Arg192 in RF1 for base-specific contacts. Similarly, the human eRF1 structure (PDB: 1DT9) shows homologous domain architecture, underscoring evolutionary preservation of the decoding mechanism.

Translation Termination Mechanism

Following stop codon recognition by release factors, the translation termination mechanism proceeds through a series of biochemical steps that ensure efficient polypeptide release and ribosomal recycling, preventing stalling and enabling reuse for new translation cycles. In prokaryotes, this involves class I release factors RF1 or RF2, which recognize specific stop codons, and class II factor RF3, a GTPase. In eukaryotes, the analogous factors are eRF1 and eRF3, respectively, forming a ternary complex with GTP upon initial binding. The process is highly conserved, driven primarily by GTP hydrolysis, and achieves near-complete fidelity without direct ATP involvement in the core termination events. The first step entails RF binding-induced ribosome stalling and subsequent GTP hydrolysis by the class II factor. In bacteria, RF1 or RF2 accommodates into the ribosomal A site, triggering a conformational shift that stalls elongation; after peptidyl-tRNA hydrolysis, RF3 then binds GTP and associates with the complex, hydrolyzing GTP to promote the dissociation of RF1 or RF2 from the peptidyl transferase center (PTC) and their eventual recycling. Similarly, in eukaryotes, eRF3-GTP hydrolysis, stimulated by eRF1's interaction with the ribosome, occurs on a millisecond timescale and drives eRF1 accommodation for catalysis. This GTP-dependent step provides the thermodynamic energy for conformational rearrangements, ensuring precise timing without ATP consumption. Next, water-mediated hydrolysis cleaves the ester bond linking the completed polypeptide to the peptidyl-tRNA in the P site. The GGQ motif of the class I release factor () inserts into the PTC, where it positions a catalytic water molecule to perform a nucleophilic attack on the ester carbonyl, liberating the nascent chain while leaving deacylated tRNA bound. This reaction proceeds rapidly, with kinetic rates on the order of milliseconds, and exhibits near 100% efficiency in standard cellular conditions, minimizing incomplete terminations. Finally, ribosome recycling dissociates the post-termination ribosomal complex into subunits for reuse. In bacteria, the ribosome recycling factor (RRF) binds the stalled 70S ribosome in a tRNA-mimetic manner, and together with initiation factor 3 (IF3) and EF-G-GTP, induces subunit splitting via GTP hydrolysis, releasing mRNA and deacylated tRNA from the 30S subunit. In eukaryotes, ABCE1, an ATP-binding protein, hydrolyzes ATP to separate the 60S and 40S subunits from the 80S complex, with subsequent mRNA and tRNA release facilitated by initiation factors like eIF1 and eIF3. These recycling steps maintain translational throughput by rapidly clearing the ribosome.

Nomenclature and Historical Naming

Amber Codon (UAG)

The amber codon, UAG, was first identified in the early 1960s through studies of conditional lethal mutants in bacteriophage T4 conducted by Richard H. Epstein and Seymour Benzer. These mutants exhibited rapid chain termination during protein synthesis in non-permissive hosts like Escherichia coli strain B, but could propagate in permissive strains such as E. coli K12, leading to their characterization as nonsense mutations. The name "amber" was selected by Epstein and colleagues to honor their graduate student colleague Harris Bernstein, whose surname translates to "amber" in German, after Bernstein isolated one of the initial mutants during a late-night screening session. This nomenclature highlighted the mutants' distinctive phenotype and facilitated their use in fine-structure genetic mapping of the T4 genome. In standard genetic codes, the UAG codon is the least frequently used stop signal, accounting for approximately 16-20% of termination sites across diverse prokaryotic and eukaryotic genomes, with its prevalence showing minimal dependence on genomic GC content compared to UAA and UGA. This lower usage may contribute to its heightened susceptibility to suppression in genetic screens, where suppressor mutations can restore function more readily than with other stops. Amber mutations introduce premature UAG stops in coding sequences, truncating polypeptides and rendering them nonfunctional, a property exploited in foundational studies of translational suppression. In E. coli, the supE mutation in the glnV gene (also known as glnX) encodes a glutamine-inserting tRNA that specifically recognizes UAG, enabling phenotypic rescue of amber mutants and allowing dissection of gene function and tRNA anticodon interactions. In contemporary biotechnology, the amber codon is preferentially employed for site-directed incorporation of non-natural amino acids into proteins via orthogonal tRNA/synthetase pairs that suppress UAG without interfering with endogenous translation. This approach, pioneered in E. coli systems, has enabled precise protein engineering for applications in structural biology and therapeutics, with efficiencies optimized through release factor 1 (prfA) attenuation.

Ochre Codon (UAA)

The ochre codon, designated as UAA, received its name in 1965 through studies on suppressible nonsense mutations in bacteriophage T4 and Escherichia coli, where Sydney Brenner and Jonathan Beckwith identified a new class of chain-terminating mutants distinct from previously known amber (UAG) mutants. To maintain a thematic nomenclature based on colors, they termed these UAA mutants "ochre," drawing from the earthy pigment associated with yellow-orange hues, which paralleled the amber naming convention established earlier for UAG in 1963. This discovery built on prior work showing that such nonsense mutations led to truncated proteins and could be suppressed by specific tRNA alterations, revealing UAA's role as a universal termination signal. In eukaryotic genomes, the UAA codon is a prevalent stop signal, accounting for approximately 30% of natural termination sites in human genes, though this frequency varies across taxa with UGA often dominating in higher eukaryotes. It contributes to efficient recognition by release factors, making it a strong terminator in translation. Unlike UAG or UGA, UAA's bias toward high usage in certain contexts, such as highly expressed genes in some organisms, underscores its evolutionary optimization for rapid chain release. The UAA codon frequently emerges via C-to-T transitions in DNA, particularly from glutamine (CAA) or glutamic acid (GAA) codons, which represent common mutational pathways due to spontaneous cytosine deamination or 5-methylcytosine alterations at CpG hotspots. This transition bias explains UAA's prevalence among nonsense mutations in genetic diseases, comprising about 18% of reported cases in humans. Additionally, UAA arises commonly in UV-induced mutagenesis because pyrimidine dimer formation at hotspots preferentially generates C-to-T changes that convert sense codons to ochre terminators, accounting for a significant portion of UV-specific nonsense mutations in model systems like E. coli and phage T4. Early experiments leveraging ochre mutants played a pivotal role in codon assignment during the 1960s, as suppression patterns in E. coli confirmed UAA as a non-coding terminator rather than an ambiguous sense codon, aligning with biochemical assays using synthetic polynucleotides that identified it as a chain-end signal. These studies, including crosses between ochre and amber mutants, demonstrated that UAA could not revert via single base changes to amber without altering the reading frame, solidifying its distinct identity in the genetic code.

Opal or Umber Codon (UGA)

The UGA codon serves as a stop signal in the standard genetic code and is referred to by the dual nomenclature "opal" or "umber," a convention rooted in the colorful naming scheme for nonsense mutations established during the elucidation of the genetic code in the 1960s and 1970s. The "opal" designation, adopted in the 1970s, evokes the iridescent gemstone, extending the thematic analogy from the "amber" (UAG) name—coined after Caltech graduate student Harris Bernstein, whose surname translates to "amber" in German—and the "ochre" (UAA) label, inspired by earthy pigments. The alternative "umber" term, denoting a dark brown pigment, emerged from early studies on Escherichia coli mutants and is used interchangeably, though less commonly today. UGA was identified as the third stop codon in 1967 through genetic analysis of suppressor mutants in bacteriophage T4, which distinguished it from UAA and UAG by failing to suppress UGA-induced chain termination, confirming its role as a nonsense triplet that does not encode an amino acid. This discovery completed the set of three termination signals, with UGA recognized by release factor 2 (RF2) in prokaryotes and both RF1 and RF2 in eukaryotes. In terms of properties, UGA is one of the most frequent stop codons across diverse genomes, accounting for approximately 50% of terminations in humans and varying similarly in many other organisms, often preferred due to its AT-rich composition suiting GC-biased mutational patterns. Its versatility stands out, as UGA is reassigned to encode in numerous prokaryotes, eukaryotes, and archaea via a specialized elongation factor ( or ) and a stem-loop structure () in the mRNA, enabling incorporation of this rare 21st amino acid without altering the standard termination function elsewhere. In some methanogenic archaea, UGA can also be contextually decoded as , highlighting its evolutionary flexibility, though reassignment typically involves rather than UGA. Usage of UGA exhibits bias toward higher prevalence in prokaryotes with elevated GC content, where it terminates up to 40-50% of genes in high-GC species like , compared to lower usage in AT-rich genomes favoring UAA. Termination efficiency at UGA is highly context-dependent, modulated by the 3' flanking nucleotides immediately following the codon; for instance, a purine (A or G) at the +4 position enhances release factor binding and reduces readthrough by up to 10-fold in E. coli, while pyrimidine contexts weaken it, influencing overall translational fidelity.

Genomic and Evolutionary Patterns

Distribution in Genomes

Stop codon usage varies significantly across genomes, reflecting differences in mutational biases, selective pressures, and evolutionary histories. In vertebrates and other higher eukaryotes, UGA is the most prevalent stop codon, followed by UAA, with UAG being the least frequent; for example, in human genes, UGA accounts for approximately 50% of terminations, UAA for 28%, and UAG for 22%. In contrast, bacterial genomes typically show UAA as the dominant stop codon (around 50-60% in low-GC species like ), followed by UGA and then UAG, though this order reverses in high-GC bacteria such as those in the phylum, where UGA exceeds UAA due to compositional constraints. Tools like facilitate the analysis of these patterns by computing codon usage indices, including relative synonymous codon usage (RSCU) for stop codons, across large genomic datasets. The distribution of stop codons is strongly influenced by genomic GC content, with AT-rich UAA favored in low-GC genomes (e.g., <40% GC in many Firmicutes) to minimize energy costs in replication and transcription, while GC-richer UGA predominates in high-GC environments (e.g., >60% GC in ). Additionally, acts to optimize termination efficiency, favoring UAA—the most efficient stop codon recognized by both release factors RF1 and RF2 in —as the primary terminator in highly expressed genes, whereas less efficient UGA and UAG are under stronger purifying selection to avoid errors. Evolutionarily, stop codon usage is highly conserved in core housekeeping genes across prokaryotes and eukaryotes, ensuring reliable termination in essential pathways, but shows greater variation in organellar genomes; for instance, mitochondrial DNA often reassigns UGA to , relying on incomplete stop codons like or for termination. genomic studies reveal shifts in stop codon preferences in multiple eukaryotic lineages, with independent reassignments (e.g., UAA/UAG to ) occurring in at least 10-15 distinct clades, including and dinoflagellates, highlighting the plasticity of the genetic code under niche-specific pressures. Recent metagenomic surveys of uncultured microbes, analyzing over 250,000 bacterial and archaeal genomes from diverse environments, confirm these trends while uncovering novel variations; for example, standard stop codon usage predominates (99.8% of cases), but rare reassignments like UGA to appear in phyla such as SR1 from microbiomes, expanding our understanding of code diversity in uncultured lineages.

Hidden Stop Codons

Hidden stop codons, also referred to as out-of-frame stop codons, are instances of the standard termination signals (UAA, UAG, or UGA) that appear in the +1 or -1 s of a protein-coding sequence, relative to the primary (ORF). These codons do not disrupt the of the intended protein but instead function to rapidly terminate any that may occur due to ribosomal frameshifting or erroneous in alternative frames. This masking effect ensures that only the correct is productively translated under normal conditions. The primary functional role of stop codons is to mitigate the risks associated with translational errors, such as frameshifts caused by ribosomal slippage, which could otherwise lead to the synthesis of aberrant, nonfunctional, or cytotoxic polypeptides. By providing an "" mechanism for early termination, they minimize cellular resource expenditure on wasteful protein production and serve as an evolutionary safeguard, buffering against deleterious mutations that might otherwise extend erroneous ORFs. This selective pressure is evident in the overrepresentation of codons that contribute to stops among synonymous alternatives, particularly in sequences prone to frameshift errors. In genomes, hidden stop codons play a crucial role in managing overlapping s, which are common for maximizing coding capacity in compact DNA or . For example, in bacteriophage φX174, multiple overlapping genes utilize out-of-frame stop codons to delimit boundaries and prevent unintended protein extensions during programmed frameshifts. Eukaryotic examples illustrate how hidden stop codons contribute to regulated and genomic efficiency. In the human gene, which encodes the catalytic subunit of gamma, an overlapping initiated by a CUG codon produces the accessory protein POLGARF, with hidden stops in non-primary frames ensuring precise termination and avoiding interference with the main function. Such configurations are also observed in mitochondrial genes across vertebrates, where hidden stops correlate with ribosomal stability and protect against frameshift-induced errors in high-mutation environments. Detection of hidden stop codons relies on bioinformatics approaches that scan coding sequences for ORFs across all six reading . Tools like NCBI's ORFfinder identify potential start and stop sites in alternative frames, revealing stops that would terminate off-frame translation prematurely. Analyses using such methods, often combined with statistical models like , demonstrate that hidden stops occur frequently in coding regions—typically every 50–100 codons in alternative frames—and are significantly enriched compared to intergenic sequences, underscoring their adaptive significance.

Mutations and Associated Diseases

Nonsense Mutations

A nonsense mutation is a point mutation in the DNA sequence that changes a sense codon into one of the three stop codons (UAG, UAA, or UGA), resulting in a premature termination codon (PTC) in the mRNA transcript. This alteration causes the ribosome to halt translation prematurely, producing a truncated polypeptide that is often non-functional or unstable. For instance, a single nucleotide substitution in the codon CAG (encoding glutamine) to TAG introduces an amber stop codon (UAG in mRNA), exemplifying how such changes disrupt protein synthesis. Nonsense mutations account for approximately 11% of all known human genetic disease-causing variants. These mutations are implicated in various genetic disorders, particularly those involving truncated proteins essential for cellular function. In , the G542X in the CFTR gene introduces a PTC, leading to absent or minimal functional CFTR protein and severe disease manifestations in affected individuals. Similarly, in (DMD), in the DMD gene, which disrupt production, comprise about 13% of all cases, contributing to progressive muscle degeneration. The primary consequence of nonsense mutations is the activation of nonsense-mediated mRNA decay (NMD), a surveillance pathway that recognizes and degrades mRNAs containing PTCs located more than 50-55 upstream of an exon-exon junction. This degradation typically reduces steady-state mRNA levels by 50-90%, severely limiting the production of full-length protein and exacerbating the loss-of-function . In cases where NMD is inefficient, the resulting truncated proteins may also be subject to ubiquitin-mediated proteasomal degradation, further diminishing functional protein levels. Therapeutic strategies targeting focus on promoting ribosomal of PTCs to restore some full-length protein production. , a small-molecule drug, received conditional marketing authorization from the in 2014 for treating DMD in patients aged 5 years and older, enabling partial suppression of the PTC in responsive patients. However, the authorization was not renewed as of March 2025 due to insufficient of clinical from confirmatory studies, resulting in its in the . Such interventions hold potential for the ~10% of rare genetic diseases attributable to , though varies by context and disease.

Nonstop Mutations

Nonstop mutations, also known as stop-loss mutations, occur through deletions, insertions, or single base-pair substitutions that eliminate one of the three stop codons (UAA, UAG, or UGA) in a gene's coding sequence. This alteration prevents normal termination, allowing the to continue translating into the 3' (UTR) of the mRNA. Without an in-frame stop codon, the may stall upon reaching the polyadenylated tail, incorporating or other polybasic sequences from the poly(A) tract, which can lead to the production of aberrant, extended polypeptides. The primary cellular consequence of nonstop mutations is the activation of the nonstop decay (NSD) pathway, a mechanism that rapidly degrades the affected mRNA to prevent synthesis of potentially harmful proteins. In this process, the stalled at the mRNA's 3' end recruits factors that target the transcript for exonucleolytic degradation, resulting in mRNA half-lives as short as 2 minutes in model systems. Additionally, any translated nonstop proteins are often unstable and subject to proteasomal degradation, but if produced in sufficient quantities, the polybasic extensions can cause toxicity by disrupting cellular membranes or aggregating. Unlike premature stop codons, which trigger (NMD), nonstop transcripts evade NMD because they lack an upstream of the missing termination site. Nonstop mutations are rare contributors to human genetic diseases, accounting for a small of loss-of-function variants, with a identifying 119 such mutations across 87 . They are typically associated with severe phenotypes due to complete loss of protein function or dominant-negative effects from aberrant products. For instance, a nonstop in the TYMP gene underlies mitochondrial neurogastrointestinal encephalomyopathy (MNGIE), a fatal disorder characterized by gastrointestinal dysmotility and neuropathy, where the extended protein fails to undergo efficient in affected cells. Similarly, nonstop variants in RPS19 cause Diamond-Blackfan anemia, a congenital bone marrow failure syndrome leading to severe and increased cancer risk. In eukaryotes, NSD surveillance relies on specific factors to detect and eliminate nonstop mRNAs, distinct from other decay pathways. In yeast, the GTPase Ski7 binds the stalled 40S ribosomal subunit and recruits the Ski complex (Ski2, Ski3, Ski8) along with the RNA exosome for 3'-to-5' degradation. Mammalian cells employ homologous components, including the Hbs1l-Pelota complex to release the stalled ribosome, followed by recruitment of the exosome via factors like NEXT (nuclear exosome targeting complex) or SKIV2L2. This machinery ensures that nonstop transcripts are efficiently cleared, minimizing proteotoxic stress, though defects in these factors can exacerbate disease severity in nonstop mutation carriers.

Advanced Biological Phenomena

Translational Readthrough

Translational readthrough refers to the natural process by which ribosomes bypass a stop codon during , allowing to continue into downstream sequences and produce extended or full-length proteins. This phenomenon occurs at low basal efficiencies, typically around 0.1% or lower for natural termination codons in eukaryotes, but can be modulated by specific mRNA contexts to reach 1-5% in regulated cases. In essential genes, such low basal rates help maintain translational fidelity while permitting adaptive responses. The primary mechanisms of translational involve competition between release factors (RFs) and near-cognate tRNAs at the ribosomal A-site. Near-cognate tRNAs, such as tRNA^Gln with a single mismatch in its anticodon, can mispair with stop codons like UAG or UAA, inserting an and continuing . Alternatively, RF slippage—where eRF1/eRF3 in eukaryotes or RF1/RF2 in prokaryotes dissociate prematurely—can occur, particularly under suboptimal conditions. In synthesis, UGA stop codons are recoded as () through a specialized readthrough mechanism involving the SECIS (selenium incorporation sequence) element in the 3' UTR, which recruits Sec-tRNA^Sec and inhibits RF activity, achieving efficiencies up to 10-20% in specific transcripts. This regulated bypassing is context-dependent, often promoted by 3' mRNA sequences that form or stem-loops, as seen in viruses like Moloney (MuLV), where a downstream stimulates ~5-10% readthrough of the gag UAG stop codon to produce the Gag-Pol polyprotein essential for . Several factors influence readthrough efficiency across organisms. In eukaryotes, the proximity of the stop codon to the poly(A) tail and binding by poly(A)-binding protein (PABP) can modulate RF recruitment; while PABP generally enhances termination, certain configurations reduce it and favor in stress contexts. In , tmRNA (transfer-messenger ) rescues ribosomes stalled on non-stop mRNAs or damaged transcripts by trans-translation, which adds a degradation tag to the nascent and terminates , preventing prolonged stalling. Recent research from the 2020s highlights readthrough's role in stress responses: metabolic or acid stress increases stop codon readthrough by 2- to 10-fold through altered RF activity or tRNA competition, promoting phenotypic heterogeneity and survival in and mammalian cells, with basal rates in essential genes maintained at 0.1-1% to balance fidelity and adaptability.

Stop Codon Suppression

Stop codon suppression refers to engineered genetic or pharmacological interventions that enable the to bypass premature termination signals, typically introduced by mutations, allowing continued of the mRNA into a full-length or near-full-length protein. These methods have been pivotal in studying genetic mechanisms and developing therapies for diseases caused by mutations, such as and . Suppressor tRNAs represent one of the earliest genetic tools for stop codon suppression, discovered in the through bacterial auxotrophic screens where mutants resistant to certain conditions incorporated at stop codons. For instance, suppressors like supF, derived from tRNA^Tyr, insert at UAG codons by altering the anticodon to CUA, enabling in systems. Similarly, opal suppressors target UGA, though with varying efficiencies depending on the tRNA source. These suppressors have been instrumental in mapping genes and understanding translational fidelity since their initial isolation in auxotrophs. Pharmacological suppression often employs antibiotics, such as gentamicin, which bind to the and stabilize near-cognate tRNA pairing at stop codons, promoting at efficiencies of approximately 5-20% in mammalian cells for models. This approach has shown promise in preclinical studies for rescuing protein function in nonsense-mediated diseases, with gentamicin's effects modulated by the stop codon identity and flanking mRNA sequence—UAA being most responsive. Clinical trials have explored its use, though dosing is limited by . More recent small molecules, such as (an anti-inflammatory drug), have shown promise in promoting of premature stop codons in preclinical models as of 2025. Additionally, tRNA-based therapeutics, like those developed by companies targeting universal suppression of , are advancing toward clinical applications for diseases including and . Advanced has expanded suppression capabilities through orthogonal tRNA-aminoacyl-tRNA synthetase (aaRS) pairs, which selectively charge suppressor tRNAs with non-natural at stop codons, enabling the incorporation of a 21st or beyond into proteins. These systems, evolved via techniques recognized in the 2018 , allow precise site-specific modifications without interfering with canonical translation. Recent 2020s developments integrate technologies, such as CRISPR-dCas13 targeted to mRNA regions downstream of stop codons, to induce transcript-specific by modulating ribosomal pausing or activity. Despite these advances, stop codon suppression faces significant limitations, including off-target at endogenous stop codons, which can produce aberrant proteins and disrupt cellular . Efficiency also varies markedly by codon and context, with UGA proving the most recalcitrant due to dual recognition by release factors RF1 and RF2, often yielding lower suppression rates than UAA or UAG. Pharmacological agents like aminoglycosides exacerbate this through , while orthogonal systems require cell-type-specific optimization to minimize with native machinery.

Applications in Biotechnology

Use as Genetic Watermarks

Stop codons can be integrated into synthetic DNA constructs as elements of genetic watermarks to authenticate , track dissemination, and deter unauthorized use, typically by embedding them in non-coding regions or as structural delimiters where they do not interfere with . This approach leverages the termination signal of stop codons (UAA, UAG, UGA) to mark boundaries or signatures without producing functional proteins, ensuring in host organisms. A prominent example is the DNA-Crypt algorithm, which embeds cryptographic watermarks in DNA by alternating blocks of plain text and cipher text strands, each terminated by a stop codon to delineate message segments for secure encoding and decoding via DNA polymerase. This method allows the watermark to propagate faithfully during replication while maintaining error correction through integrated codes like Hamming or WDH, facilitating detection via amplification and sequencing with knowledge of specific primers. In large-scale de novo genome synthesis, such as the Synthetic Yeast Genome Project (Sc2.0) initiated in the 2010s, all TAG (UAG) stop codons are systematically replaced with TAA across synthetic chromosomes, creating a uniform codon usage pattern that distinguishes synthetic DNA from natural genomes during verification by whole-genome sequencing. This recoding not only frees the UAG codon for potential genetic code expansion but also acts as an inherent identifier for intellectual property protection and misuse prevention in eukaryotic synthetic biology. In applications, UAG stop codons are incorporated into constructs as conditional safety signals; for instance, in systems for non-canonical incorporation, the UAG acts as a default unless suppressed by orthogonal tRNAs, halting expression of engineered proteins if components fail and thereby containing potential biohazards. Detection relies on targeted sequencing to confirm the presence of these embedded stops, which remain inert to native translational machinery unless intentionally overridden. Advantages include seamless without phenotypic disruption, in replication, and robust traceability, making stop codon-based watermarks valuable for in advanced synthetic constructs.

Engineering in Synthetic Biology

In synthetic biology, engineers have manipulated stop codons to expand the beyond the standard 20 , enabling the incorporation of unnatural amino acids (UAAs) and the creation of orthogonal systems. A key approach involves recoding bacterial genomes to eliminate stop codons entirely, freeing them for reassignment. For instance, the Synthetic Yeast Genome Project (Sc2.0) and parallel bacterial efforts, such as the 2016 recoding of to remove the UAG stop codon (creating Syn61 with 61 codons), demonstrated viability by replacing all 321 UAG instances with UAA and deleting release factor 1 (RF1). Building on this, the 2025 development of Syn57 recoded to use only 57 codons by eliminating the stop codon (UAG) and six sense codons (four for serine and two for ) through over 100,000 genome edits, achieving robust growth and virus resistance while maintaining essential functions. These recoded strains, part of the broader 2020s Sync project, allow stop codons to be repurposed without terminating , supporting multiplexed UAA incorporation and providing incompatibility with natural phages for enhanced . Amber suppression systems exemplify stop codon engineering for protein modification, where the UAG (amber) codon serves a dual role: as a stop in natural contexts or a UAA cue via orthogonal tRNA/aminoacyl-tRNA synthetase (aaRS) pairs. Pioneered in the early , this method uses engineered Methanocaldococcus jannaschii tyrosyl-tRNA synthetase variants to charge amber-suppressing tRNAs with UAAs, enabling site-specific insertion at UAG sites in response to amber codons introduced by . Since then, over 100 distinct UAAs—including photocaged, fluorescent, and bioorthogonal groups—have been incorporated into proteins in bacteria, yeast, and mammalian cells, facilitating applications like protein dynamics studies and . In optimized strains, such as RF1-deleted E. coli, suppression efficiencies reach up to 90% for full-length proteins, minimizing truncation and enhancing yield through reduced competition from release factors. These engineered systems have broad applications in . In biosensors, amber suppression incorporates UAAs like p-azidophenylalanine for click chemistry-based detection, enabling real-time monitoring of protein interactions with sensitivities improved by 10-fold over natural reporters. For therapeutics, UAA-modified proteins serve as next-generation biologics, such as antibodies with site-specific payloads for , reducing off-target effects in cancer treatments. Stop codon recoding also enhances production; recoded E. coli strains like Syn61 safely propagate attenuated viruses by introducing amber stops that halt replication in non-engineered hosts, yielding non-infectious particles for with up to 80% while preserving immunogenicity.30214-9) This approach ensures , preventing viral escape during manufacturing. Ongoing research in 2025 targets 57-codon frameworks for multiplexed expression, where all stop codons are reassigned to enable simultaneous incorporation of multiple UAAs in a single , potentially tripling the diversity of synthetic proteins for complex circuits. These efforts, exemplified by Syn57's , promise to accelerate the design of novel enzymes and materials, with prototypes achieving 70% viability in multi-UAA regimes.

References

  1. [1]
    Stop Codon - National Human Genome Research Institute
    A stop codon is a sequence of three nucleotides (a trinucleotide) in DNA or messenger RNA (mRNA) that signals a halt to protein synthesis in the cell.
  2. [2]
    From RNA to Protein - Molecular Biology of the Cell - NCBI Bookshelf
    Stop Codons Mark the End of Translation​​ Proteins known as release factors bind to any ribosome with a stop codon positioned in the A site, and this binding ...
  3. [3]
    Unusual base pairing during the decoding of a stop codon ... - PubMed
    Aug 1, 2013 · During normal translation, the binding of a release factor to one of the three stop codons (UGA, UAA or UAG) results in the termination of
  4. [4]
    Nonsense Mutation - National Human Genome Research Institute
    A nonsense mutation is a DNA change that creates a stop codon, causing a protein to end its translation early, resulting in a non-functional protein.
  5. [5]
    Multiple independent genetic code reassignments of the UAG stop ...
    The standard genetic code uses three stop codons (UAA, UAG, and UGA) to terminate translation and 61 sense codons to encode an amino acid [1].
  6. [6]
    START and STOP Codons - News-Medical
    There are 3 STOP codons in the genetic code - UAG, UAA, and UGA. These codons signal the end of the polypeptide chain during translation. These codons are also ...<|separator|>
  7. [7]
    Stages of translation (article) | Khan Academy
    Termination happens when a stop codon in the mRNA (UAA, UAG, or UGA) enters the A site. Stop codons are recognized by proteins called release factors, which fit ...
  8. [8]
    Protein Synthesis (Translation) - Geosciences LibreTexts
    Sep 24, 2019 · Termination of translation occurs when the ribosome encounters a stop codon, which does not code for a tRNA. Release factors cause the ...
  9. [9]
    Stops making sense – For the people? - PMC - NIH
    May 18, 2023 · The standard genetic code contains 64 codons, 61 of which (sense) code for specific amino acids and three define stop (nonsense) codons.
  10. [10]
    The Information in DNA Determines Cellular Function via Translation
    The codons UAA, UAG, and UGA are the stop codons that signal the termination of translation. Figure 2 shows the 64 codon combinations and the amino acids or ...
  11. [11]
    The End of Translation: stop codons looking for something they ...
    No tRNAs in the cell have anticodons that complement any of the three possible stop codons. Therefore, stop codons are able to end the translation process when ...
  12. [12]
    Expanding the Genetic Code for Biological Studies - PMC - NIH
    There are three stop codons that do not code for amino acids but function as a signal for termination of translation. By reassigning a “blank” stop codon to an ...
  13. [13]
    Deciphering the Genetic Code - National Historic Chemical Landmark
    In 1964 Nirenberg and Philip Leder, a postdoctoral fellow at NIH, discovered a way to determine the sequence of the letters in each triplet word for amino acids ...Missing: stop | Show results with:stop
  14. [14]
    [PDF] Marshall Nirenberg - Nobel Lecture
    These dramatic results showed that "non- sense" codons correspond to the termination of protein synthesis.
  15. [15]
    A 'mad race to the finish'
    Feb 1, 2012 · With this assay, Nirenberg's group deciphered most of the codons by 1966. In 1968, Nirenberg shared the Nobel Prize in physiology or medicine ...
  16. [16]
    Stop Codon Usage as a Window into Genome Evolution - NIH
    Most organisms also have three alternative options for the stop codon (UAA, UGA, and UAG in mRNA or TAA, TGA, and TAG in genomic sequence). Like synonymous ...
  17. [17]
    Dual stop codon suppression in mammalian cells with genomically ...
    Nov 20, 2023 · The amber codon is the least abundant and most efficiently suppressed stop codon in mammalian cells. The other two stop codons, TAA (ochre) and ...Missing: names umbersource:
  18. [18]
    Two-step model of stop codon recognition by eukaryotic release ...
    Feb 22, 2013 · Recognition of the first and second nucleotides of stop codon (UA or UG). Second A and G are decoded at different sites of eRF1. ( Step 2) ...
  19. [19]
    Untranslated regions of mRNAs | Genome Biology | Full Text
    Feb 28, 2002 · Indeed, normal stop codons and the 3' UTR are usually located in the last exon of the sequence and thus are not followed by a splicing junction.Abstract · Control Of Translation... · Regulation Of Mrna StabilityMissing: punctuation | Show results with:punctuation
  20. [20]
    Selective forces and mutational biases drive stop codon usage in ...
    May 17, 2016 · The three stop codons UAA, UAG, and UGA signal the termination of mRNA translation. As a result of a mechanism that is not adequately understood ...<|control11|><|separator|>
  21. [21]
    Genetic Codes - NCBI
    In other organisms, AGA/AGG code for either arginine or serine and in vertebrate mitochondria they code a STOP. Evidence for glycine translation of AGA/AGG has ...The Vertebrate Mitochondrial... · The Ascidian Mitochondrial... · Rhabdopleuridae...
  22. [22]
    Evolution of the mitochondrial genetic code. I. Origin of AGR serine ...
    AGA and AGG (AGR) are arginine codons in the universal genetic code. These codons are read as serine or are used as stop codons in metazoan mitochondria.
  23. [23]
    Genetic Codes with No Dedicated Stop Codon: Context-Dependent ...
    Jul 14, 2016 · We provide evidence, based on ribosomal profiling and “stop” codon depletion shortly before coding sequence ends, that mRNA 3′ ends may contribute to ...
  24. [24]
    UGA is read as tryptophan in Mycoplasma capricolum - PubMed
    We show that UGA also codes for tryptophan in Mycoplasma capricolum, a wall-less bacterium having a genome only 20-25% the size of the Escherichia coli genome.
  25. [25]
    [PDF] The Evolution of Alternative Genetic Codes - Harvard DASH
    These clades all have very low genomic GC content, an evolutionary force which likely drove GC-rich codons to rarity and allowed for their reassignment. The CGA ...
  26. [26]
    Driving change: the evolution of alternative genetic codes - Cell Press
    The RF would lose the ability to recognize the codon being captured by the tRNA, enabling its full reassignment to a sense codon (Figure 3). Codon reassignment ...
  27. [27]
    The Molecular Biology of Selenocysteine - PMC - PubMed Central
    The incorporation of this amino acid turns out to be a fascinating problem in molecular biology because Sec is encoded by a stop codon, UGA. Layered on top of ...
  28. [28]
    Decoding apparatus for eukaryotic selenocysteine insertion
    Decoding UGA as selenocysteine requires a unique tRNA, a specialized elongation factor, and specific secondary structures in the mRNA, termed SECIS elements ...
  29. [29]
    Methanogenic archaea encoding Pyrrolysine maintain ambiguous ...
    Pyrrolysine (Pyl) is encoded by the amber codon (TAG/UAG) and is widespread in archaea, where it is required for methylamine-mediated methanogenesis, an ...
  30. [30]
    Activation of the Pyrrolysine Suppressor tRNA Requires Formation ...
    Monomethylamine methyltransferase of the archaeon Methanosarcina barkeri contains a rare amino acid, pyrrolysine, encoded by the termination codon UAG.Missing: reassignment | Show results with:reassignment
  31. [31]
    Genome-scale quantification and prediction of pathogenic stop ...
    Aug 22, 2024 · Premature termination codons (PTCs) cause ~10–20% of inherited diseases and are a major mechanism of tumor suppressor gene inactivation in ...
  32. [32]
    Suppressive cancer nonstop extension mutations increase C ... - NIH
    The STOP codon was mutated to each of the 20 proteinogenic amino acids showing a positive correlation of hydrophobicity with the loss of protein expression. n = ...
  33. [33]
    Fine-tuning Interaction between Aminoacyl-tRNA Synthetase and ...
    Orthogonal tRNA-aminoacyl-tRNA synthetase pairs are widely used for the site-specific incorporation of nearly 80 unnatural amino acids (unAAs) in Escherichia ...
  34. [34]
    Using genetically incorporated unnatural amino acids to control ...
    The amber stop codon (UAG) is often used as the blank codon due to its minimal occurrence in most organisms. Mechanism of genetic code expansion for site- ...
  35. [35]
    Reducing the genetic code induces massive rearrangement of the ...
    Less than 1% of all sequenced genomes encode an operon that reassigns the stop codon UAG to pyrrolysine (Pyl), a genetic code variant that results from the ...
  36. [36]
    A computational screen for alternative genetic codes in over ... - eLife
    Nov 9, 2021 · These are candidates for new codon reassignments, but could also include inference errors. For stop codons, 99.80% out of a total of 145,855 ...
  37. [37]
    Recognition of the amber UAG stop codon by release factor RF1
    Jun 29, 2010 · Upon recognition of a stop codon, release factors RF1 and RF2 promote hydrolysis of peptidyl‐tRNA in the peptidyl‐transferase center (PTC), ...
  38. [38]
    4V89: Crystal Structure of Release Factor RF3 Trapped in the GTP ...
    Jul 9, 2014 · The 3.3 Å crystal structure of the RF3·GDPNP·ribosome complex provides a high-resolution description of interactions and structural ...
  39. [39]
    A tripeptide discriminator for stop codon recognition - ScienceDirect
    Mar 6, 2002 · Only recently has it been established that a tripeptide in the bacterial release factors (RFs), RF1 and RF2, is responsible for the stop codon recognition.
  40. [40]
    The accuracy of codon recognition by polypeptide release factors
    RF1 and RF2 discriminate against sense codons related to stop codons by between 3 and more than 6 orders of magnitude. This high level of accuracy is obtained ...Sign Up For Pnas Alerts · Results · Rfs Terminate With High...
  41. [41]
    Stop Codon Recognition by Release Factors Induces Structural ...
    Nov 30, 2007 · In bacteria, the three stop codons are decoded by two class I RFs with overlapping specificity: RF1 recognizes UAG and UAA while RF2 recognizes ...
  42. [42]
    Structural insights into eRF3 and stop codon recognition by eRF1
    Eukaryotic translation termination is mediated by two interacting release factors, eRF1 and eRF3, which act cooperatively to ensure efficient stop codon ...
  43. [43]
    Origin of the omnipotence of eukaryotic release factor 1 - Nature
    Nov 10, 2017 · When a stop codon is presented at the decoding site, eRF1 binds to the ribosome in ternary complex with the class-2 release factor eRF3 and GTP.
  44. [44]
    Distinct eRF3 Requirements Suggest Alternate eRF1 Conformations ...
    The eukaryotic class II release factor eRF3 facilitates eRF1 stop codon recognition and carries out GTP hydrolysis prior to polypeptide chain release ...
  45. [45]
    The Origin and Evolution of Release Factors: Implications for ... - MDPI
    Bacteria possess two paralogous RF-PHs, RF1 and RF2, which are nearly absolutely conserved across bacteria and are responsible for recognizing distinct sets of ...
  46. [46]
  47. [47]
  48. [48]
  49. [49]
  50. [50]
  51. [51]
  52. [52]
  53. [53]
    Amber Mutants of Bacteriophage T4D: Their Isolation and Genetic ...
    Amber mutants arise from a restricted class of mutational event. They are rarely, if ever, induced by proflavine.
  54. [54]
    Amber Mutant - an overview | ScienceDirect Topics
    Harris had the nickname Immer Wieder Bernstein ('Forever Amber' in German). That night we isolated several of the desired mutants and named them amber mutants.
  55. [55]
    The Genome of Bacteriophage T4: An Archeological Dig - PMC - NIH
    Dick Epstein had by then mapped 26 amber mutants, and I had mapped 28 ts mutations. Thirteen of the identified genes contained both amber and ts mutations. We ...
  56. [56]
    Purifying and positive selection in the evolution of stop codons
    Jun 18, 2018 · The frequency of UAA and UGA stop codons strongly depends on the genomic GC-content, whereas the frequency of UAG appears to be independent of ...Missing: percentages | Show results with:percentages
  57. [57]
    Evidence that the supE44 Mutation of Escherichia coli Is an Amber ...
    Earlier studies have shown that supE44 is a weak amber suppressor and that its efficiency varies up to 35-fold depending on the reading context of the stop ...
  58. [58]
    Ochre mutants, a new class of suppressible nonsense mutants
    Ochre mutants, a new class of suppressible nonsense mutants. Author links ... Brenner and Stretton, 1965. S. Brenner, A.O.W. Stretton. J. Mol. Biol., 13 ...
  59. [59]
    Mutations to nonsense codons in human genetic disease
    terminate protein synthesis at the ends of human genes. In human cells, natural termination codon usage divides UAG (23%), UAA. (30%) and UGA (47%) (10-12).
  60. [60]
    UV-induced mutation hotspots occur at DNA damage hotspots - Nature
    Jul 8, 1982 · The results reported here reveal the presence of UV-induced base damage hotspots which include the nonsense mutation hotspots.
  61. [61]
    A breakthrough from 60 years ago: “General nature of the genetic ...
    Jun 7, 2021 · In 1967, the last of the 64 codons was deciphered, in a paper co-authored by Brenner and Crick [47]. This was the so-called opal codon, UGA.BACKGROUND · THE ROLE OF SUPPRESSORS · WRITING THE ARTICLE
  62. [62]
    UGA suppression by a mutant RNA of the large ribosomal subunit.
    Several avenues of action of the suppressor mutation are suggested, including altered interactions with release factors, ribosomal protein L11, or 16S rRNA.Missing: stop | Show results with:stop
  63. [63]
    UGA: A Third Nonsense Triplet in the Genetic Code - Nature
    The UGA triplet of the bases uracil, guanine and adenine does not code for an amino-acid and is therefore also a “nonsense triplet”.Missing: discovery | Show results with:discovery
  64. [64]
    Selective forces and mutational biases drive stop codon usage ... - NIH
    May 17, 2016 · There are differences among the usage of the three stop codons in eukaryotes: UAA is the most frequent in lower organisms, UGA is prevalent in ...
  65. [65]
    Distinct genetic code expansion strategies for selenocysteine and ...
    Selenocysteine is synthesized via a tRNA-dependent pathway and decodes UGA (opal) codons. The incorporation of selenocysteine requires the concerted action of ...
  66. [66]
    Comprehensive Analysis of Stop Codon Usage in Bacteria and Its ...
    In the standard bacterial codon table, there are three stop codons, TAG, TGA, and TAA (UAG, UGA, and UAA on mRNA), which are recognized by two class I release ...
  67. [67]
    A direct estimation of the context effect on the efficiency of termination
    A multitude of in vivo experiments has shown that the relative efficiency of decoding of stop codons depends on their immediate nucleotide context (reviewed by ...
  68. [68]
    Correspondence Analysis of Codon Usage
    Apr 15, 2005 · CodonW is a programme designed to simplify the Multivariate analysis (correspondence analysis) of codon and amino acid usage.Readme file · Correspondence Analysis · CodonW downloadMissing: stop | Show results with:stop
  69. [69]
    Evolution and codon usage bias of mitochondrial and nuclear ...
    We found that patterns of codon usage bias at gene level are more similar between mitogenomes of different species than the mitogenome and nuclear genome of the ...Evolution And Codon Usage... · Results · Patterns Of Codon Usage Bias...<|control11|><|separator|>
  70. [70]
    hidden stop codons prevent off-frame gene reading - PubMed - NIH
    Our "ambush" hypothesis suggests that hidden stops are sometimes selected for. Codons of many amino acids can contribute to hidden stops.Missing: immunoglobulin eukaryotes
  71. [71]
    SHIFT: Server for hidden stops analysis in frame-shifted translation
    Feb 23, 2013 · Hidden stop codons occur naturally in coding sequences among all organisms. These codons are associated with the early termination of ...
  72. [72]
    Natural selection retains overrepresented out-of-frame stop codons ...
    Sep 9, 2010 · Out-of-frame stop codons (OSCs) occur naturally in coding sequences of all organisms, providing a mechanism of early termination of translation ...
  73. [73]
    Overlapping genes in natural and engineered genomes
    Oct 5, 2021 · Most overlapping genes in eukaryotes are classified as such because their 5′ or 3′ UTRs overlap. Of those with overlaps between the start and ...
  74. [74]
  75. [75]
  76. [76]
    ORFfinder Home - NCBI - NIH
    ORF finder searches for open reading frames (ORFs) in the DNA sequence you enter. The program returns the range of each ORF, along with its protein translation.Missing: hidden | Show results with:hidden
  77. [77]
    Engineered tRNAs suppress nonsense mutations in cells and in vivo
    May 31, 2023 · One strategy to suppress nonsense mutations is to use natural tRNAs with altered anticodons to base-pair to the newly emerged PTC and promote ...
  78. [78]
    Nonsense Mutation - an overview | ScienceDirect Topics
    Nonsense mutations create a premature stop codon thus producing a truncated, usually non-functioning, protein. It is estimated that nonsense mutations account ...
  79. [79]
    Recoding of Nonsense Mutation as a Pharmacological Strategy - PMC
    Approximately 11% of genetic human diseases are caused by nonsense mutations that introduce a premature termination codon (PTC) into the coding sequence.
  80. [80]
    Pharmacological Responses of the G542X-CFTR to CFTR Modulators
    Jun 23, 2022 · G542X-CFTR, a premature termination codon (PTC) mutation, is the most common disease-associated mutation found in the remaining 10% of patients ...Abstract · Introduction · Materials and Methods · Results
  81. [81]
    Exon skipping for nonsense mutations in Duchenne muscular ...
    Jun 1, 2012 · Areas covered: Approximately 15% of DMD cases are caused by a nonsense mutation. Although patient databases have previously been surveyed for ...Missing: prevalence | Show results with:prevalence
  82. [82]
    Suppression of Nonsense Mutations by New Emerging Technologies
    Jun 20, 2020 · A point mutation (indicated) occurs in a gene, creating a premature termination codon (PTC) at the site >50–55 nts upstream of the exon–intron ...
  83. [83]
    An approach for suppressing nonsense mutations in disease genes
    Feb 9, 2023 · Nonsense mutations create premature termination codons (PTCs), activating the nonsense-mediated mRNA decay (NMD) pathway to degrade most ...<|separator|>
  84. [84]
    Nonsense-mediated mRNA decay uses complementary ... - NIH
    Our data indicate that NMD limits the accumulation of proteins encoded by NMD substrates by mechanisms beyond mRNA degradation, such that even when NMD- ...
  85. [85]
    Ataluren—Promising Therapeutic Premature Termination Codon ...
    Among them, only ataluren was approved in several countries to treat nonsense mutation Duchenne muscular dystrophy (DMD) patients. This review summarizes ...
  86. [86]
    Ataluren treatment of patients with nonsense mutation ...
    To treat genetic disorders due to a nonsense mutation, ataluren (PTC124) has been developed as a first-in-class, investigational new drug designed to enable ...
  87. [87]
    Degradation of mRNAs that lack a stop codon: A decade of nonstop ...
    Nonstop decay is the mechanism of identifying and disposing aberrant transcripts that lack in-frame stop codons.
  88. [88]
    A meta-analysis of single base-pair substitutions in translational ...
    May 1, 2011 · 'Nonstop' mutations are single base-pair substitutions that occur within translational termination (stop) codons and which can lead to the ...
  89. [89]
    Nonsense mRNA suppression via nonstop decay - eLife
    Jan 8, 2018 · Nonstop mRNA decay degrades mRNAs with a premature stop codon after such mRNAs are targeted by the nonsense-mediated decay machinery.
  90. [90]
    A meta-analysis of single base-pair substitutions in translational ...
    We have performed a meta-analysis of the 119 nonstop mutations (in 87 different genes) known to cause human inherited disease.
  91. [91]
  92. [92]
  93. [93]
    The Hbs1-Dom34 Protein Complex Functions in Non-stop mRNA ...
    Non-stop mRNA decay is characterized by its dependence on translation in yeast (23, 43). To confirm that the decay observed for non-stop mRNA in mammalian cells ...
  94. [94]
    Stop-Codon Readthrough in Therapeutic Protein Candidates ...
    Among the 85 polypeptide chains in the 48 proteins, 2 chains used UAA stop codon, 58 chains used UAG stop codon, and 25 used UGA stop codon. The UAG stop codon ...
  95. [95]
    Transcriptome-wide investigation of stop codon readthrough in ...
    The trend in eRF1 mutant cells is consistent with the hypothesis that proximity of a stop codon to PABP enhances termination efficiency [38]. The fact that ...<|control11|><|separator|>
  96. [96]
    canonical translation mechanisms reinterpreted - Oxford Academic
    Sep 12, 2019 · This review summarizes the recent advances in understanding the mechanisms of three types of recoding events: stop-codon readthrough, –1 ribosome frameshifting ...
  97. [97]
    Translational readthrough potential of natural termination codons in ...
    A number of natural mechanisms that suppress translation termination exist. One of them is STOP codon readthrough, the process that enables the ribosome to pass ...
  98. [98]
    Evidence of efficient stop codon readthrough in four mammalian genes
    Jul 10, 2014 · Stop codon readthrough is used extensively by viruses to expand their gene expression. Until recent discoveries in Drosophila, only a very ...
  99. [99]
    PABP enhances release factor recruitment and stop codon ... - NIH
    Jul 14, 2016 · Several studies suggest that PABP interferes with NMD and that PABP deletion increases read-through of stop codons, thereby indirectly providing ...
  100. [100]
    Metabolic stress promotes stop-codon readthrough and phenotypic ...
    Aug 24, 2020 · Termination of translation is signaled by stop codons and catalyzed by release factors. Occasionally, stop codons can be suppressed by near ...
  101. [101]
    Genome-wide screening reveals metabolic regulation of stop-codon ...
    Sep 6, 2023 · It has also been shown that acid stress increases stop-codon readthrough on the ribosome by compromising the activity of release factors (16).
  102. [102]
    Aminoglycoside-induced mutation suppression (stop codon ... - NIH
    Here we review nonsense mutation suppression by aminoglycosides as a therapeutic strategy to treat DMD with special emphasis on gentamicin-induced readthrough.
  103. [103]
    The Origins and Recent Promise of Nonsense Suppressor tRNAs
    Sep 7, 2023 · Nonsense suppressor tRNAs, or sup-tRNAs, read through premature stop codons in the mRNA caused by nonsense mutations, allowing protein synthesis to continue.
  104. [104]
  105. [105]
    Amber, ochre and opal suppressor tRNA genes derived from a ...
    Amber, ochre and opal suppressor tRNA genes have been generated by using oligonucleotide directed site‐specific mutagenesis to change one or two nucleotides ...Missing: Su7 | Show results with:Su7
  106. [106]
    An <i>Amber</i> Suppressor of <i>Escherichia coli</i> Strain KO1
    Like su2, su7 also inserts glutamine, but they differ from each other in mapping position and suppression effi ciency. This result, however, does not deny the ...<|separator|>
  107. [107]
    Aminoglycoside-mediated rescue of a disease-causing nonsense ...
    Aminoglycoside antibiotics, mostly gentamicin, have been shown to suppress clinically relevant premature stop codons in in vitro cell systems and in two in vivo ...
  108. [108]
    Statistical Analysis of Readthrough Levels for Nonsense Mutations ...
    We demonstrate that the presence of a uracil residue immediately upstream the stop codon is associated with a stronger response to gentamicin treatment.
  109. [109]
    Suppression of Nonsense Mutations As A Therapeutic Approach To ...
    First, the efficiency of suppressing PTCs is greatly influenced by the identity of the stop codon and the surrounding mRNA sequence. Various aminoglycosides ...
  110. [110]
    Genetic Code Expansion: Recent Developments and Emerging ...
    Dec 31, 2024 · Selenocysteine (Sec) is naturally co-translationally incorporated into proteins by recoding the UGA opal codon with a specialized elongation ...
  111. [111]
    Expanding the genetic code: a non-natural amino acid story
    Mar 1, 2023 · b) The orthogonal tRNA can only work with the orthogonal aminoacyl-tRNA (aaRS) synthetase and the engineered tRNA with the engineered aaRS.
  112. [112]
    Transcript-specific induction of stop codon readthrough using a ...
    Mar 18, 2024 · The CRISPR-dCas13 system can be targeted to an mRNA downstream of its stop codon to induce stop codon readthrough during translation.
  113. [113]
    Context effects of genetic code expansion by stop codon suppression
    Absence of off target stop codon suppression reveals that native terminal stop-codons have high termination efficiency. •. Practical considerations for ...
  114. [114]
    Repurposing tRNAs for nonsense suppression - Nature
    Jun 22, 2021 · Three stop codons (UAA, UAG and UGA) terminate protein synthesis and are almost exclusively recognized by release factors.
  115. [115]
    Dual stop codon suppression in mammalian cells with genomically ...
    Nov 6, 2023 · The relatively lower efficiency of ochre and opal stop codon suppression limits the overall yield of dual suppressed protein.
  116. [116]
    DNA-based watermarks using the DNA-Crypt algorithm
    May 29, 2007 · After such a block of plain and cipher text, there is a stop codon (Figure 1). The DNA polymerase completes the plain and cipher text ...
  117. [117]
    Design of a synthetic yeast genome - Science
    Mar 10, 2017 · Stop codon recoding/stop swaps​​ Similar to “REcoli,” the designed Sc2. 0 genome replaces all UAG stop codons with UAA; de novo synthesis rather ...
  118. [118]
    Construction of a synthetic Saccharomyces cerevisiae pan-genome ...
    Jun 24, 2022 · 0 project. This included the substitution of TAA for TAG stop-codons, the introduction of oligonucleotide watermarks within 36 ORFs ( ...
  119. [119]
    Advanced and Safe Synthetic Microbial Chassis with Orthogonal ...
    Aug 16, 2024 · Through the use of CRISPR-assisted transposition, we have engineered a safe Escherichia coli chassis that integrates an orthogonal ...
  120. [120]
    RF1 attenuation enables efficient non-natural amino acid ...
    Jun 8, 2017 · Amber codon suppression for the insertion of non-natural amino acids (nnAAs) is limited by competition with release factor 1 (RF1).