Fact-checked by Grok 2 weeks ago

Messenger RNA

Messenger RNA (mRNA) is a single-stranded ribonucleic acid (RNA) molecule that carries genetic information from deoxyribonucleic acid (DNA) to ribosomes, where it serves as a template for protein synthesis during translation. In eukaryotes, mRNA is transcribed from a gene's DNA template via RNA polymerase II and transported from the nucleus to the cytoplasm. mRNA encodes the amino acid sequence of proteins using a series of three-nucleotide codons that specify particular amino acids or translation signals. As a key intermediary in the central dogma of molecular biology, mRNA enables the expression of genetic information stored in DNA to produce functional proteins essential for cellular processes. The discovery of mRNA occurred in 1961, when Sydney Brenner, François Jacob, and Matthew Meselson demonstrated that it acts as an unstable, short-lived carrier of genetic information from DNA to protein-synthesizing ribosomes in bacteria. This finding built on earlier hypotheses by Jacob and Monod and resolved how genetic instructions are transferred without direct DNA involvement in translation. In eukaryotes, mRNA production involves transcription in the nucleus followed by extensive processing of the initial transcript, known as pre-mRNA, to generate mature mRNA ready for export and translation. Eukaryotic pre-mRNA processing includes three major steps: addition of a 5' cap (a 7-methylguanosine ) to protect the mRNA and facilitate binding; splicing to remove non-coding introns and join coding exons; and cleavage and at the 3' end, adding a poly-A tail for and . The mature mRNA typically consists of a (UTR), the coding sequence, a 3' UTR, the 5' cap, and the poly-A tail, with the overall length varying from hundreds to thousands of nucleotides depending on the gene. These modifications ensure mRNA , efficient nuclear through nuclear pore complexes, and accurate translation, where decode the mRNA sequence in coordination with (tRNA) molecules. Beyond its fundamental role in , mRNA has gained prominence in , particularly in mRNA vaccines that instruct cells to produce viral proteins for immune response training, as seen in vaccines. Dysregulation of mRNA processing or stability is implicated in various diseases, including cancer and neurodegenerative disorders, highlighting its critical regulatory functions. Over 150,000 unique mRNAs have been identified in cells, enabling the diversity of the from a limited through mechanisms like .

Introduction

Definition and Discovery

Messenger RNA (mRNA) is a single-stranded ribonucleic acid () molecule transcribed from a DNA template that serves as an intermediary carrying the to ribosomes for directing protein synthesis. In eukaryotes, transcription occurs in the , with mRNA exported to ribosomes in the ; in prokaryotes, both transcription and take place in the . This process aligns with the , which posits that genetic information flows from DNA to RNA to proteins. The concept of mRNA emerged in 1961 when François Jacob and proposed it as an unstable intermediary in bacterial , particularly in their studies of the , where it was envisioned as a short-lived that transmits regulatory signals from to ribosomes for rapid protein production. Their model explained the observed quick turnover of in , contrasting with the stability of other RNA types, and laid the groundwork for understanding inducible gene systems. This proposal was experimentally confirmed through pulse-labeling studies in the early 1960s, notably by , , and , who demonstrated that a small fraction of rapidly labeled, unstable RNA becomes associated with ribosomes during protein in bacteriophage-infected , directly linking it to the induction. These findings established mRNA's role as the specific carrier of genetic information, distinguishing it from (tRNA) and (rRNA), which primarily function in the translational apparatus rather than encoding protein sequences.

Role in Gene Expression

Messenger RNA (mRNA) occupies a central position in the central dogma of molecular biology, which posits a unidirectional flow of genetic information from DNA to RNA to proteins. In this framework, mRNA is transcribed from DNA templates in the nucleus (in eukaryotes) or cytoplasm (in prokaryotes), capturing the genetic sequence as a single-stranded RNA molecule complementary to the template strand of DNA (and thus matching the coding strand, with uracil replacing thymine). This process, known as transcription, ensures that the information encoded in genes is transferred to a portable form that can direct protein synthesis. Once produced, mRNA serves as the template during translation, where ribosomes bind to it and decode its nucleotide triplets—codons—into a specific sequence of amino acids, thereby producing functional proteins essential for cellular processes. A key distinction in mRNA function arises from its organization in different organisms. In eukaryotes, mRNAs are predominantly monocistronic, meaning each encodes a single polypeptide chain, which facilitates independent regulation of individual proteins and aligns with the compartmentalized nature of eukaryotic cells. Conversely, prokaryotic mRNAs are often polycistronic, derived from operons that group multiple genes under a single promoter, allowing coordinated of several proteins from one mRNA transcript to support rapid responses to environmental changes, such as nutrient availability. This polycistronic strategy is exemplified in bacterial operons like the , where enzymes are expressed together. In prokaryotes, mRNA is translated directly without the extensive processing seen in eukaryotes. mRNA integrates deeply with broader cellular regulatory networks, serving as a focal point for control at both transcriptional and post-transcriptional levels. Transcriptional regulation modulates mRNA production rates through factors like promoters and enhancers, while post-transcriptional mechanisms— including mRNA degradation, splicing, and interactions with RNA-binding proteins—adjust its availability, stability, and translation efficiency to achieve precise spatiotemporal gene expression. These layers of regulation allow cells to adapt dynamically, for instance, by rapidly degrading unnecessary mRNAs during stress responses. The functional role of mRNA in exhibits remarkable evolutionary conservation, underpinned by the near-universal that interprets its codons consistently across , , and eukaryotes. This code, deciphered through pioneering experiments using synthetic mRNAs, assigns specific or stop signals to each of the 64 possible triplets, enabling seamless machinery compatibility across diverse life forms and highlighting mRNA's ancient origins in the . Minor variations in the code occur in certain organelles and organisms, but the core triplet-based decoding remains invariant, underscoring mRNA's fundamental conservation.

Structure

Core Components

Mature messenger RNA (mRNA) in eukaryotes exhibits a fundamental linear consisting of a 5' cap, a coding sequence, and a 3' end, flanked by untranslated regions. This core structure ensures the mRNA's stability, export from the , and efficient into proteins. The 5' cap is added post-transcriptionally and consists of a 7-methylguanosine (m⁷G) moiety linked via a 5'-5' triphosphate bridge to the first of the transcript, protecting the mRNA from exonucleolytic degradation and facilitating binding. The coding sequence (CDS), also known as the open reading frame (ORF), spans from the start codon AUG to a stop codon (UAA, UAG, or UGA), directly encoding the sequence of the polypeptide. Unlike DNA, mRNA incorporates uracil (U) in place of thymine (T) within its nucleotide composition of (A), (G), (C), and uracil, with eukaryotic mRNAs typically ranging from 500 to 10,000 in length to accommodate the CDS and flanking elements. At the 3' end, mRNA maturation involves endonucleolytic cleavage at a site downstream of the signal, most commonly the hexanucleotide AAUAAA in eukaryotes, followed by the addition of a poly(A) tail. This cleavage and tailing process defines the mature 3' terminus, contributing to mRNA stability and translational efficiency. Additionally, mRNA achieves a closed-loop configuration through interactions between the 5' cap and the poly(A)-binding protein (PABP) at the 3' end, mediated by the scaffold protein eIF4G, which enhances mRNA circularization and promotes ribosome recycling for sustained .

Untranslated Regions

Messenger RNA (mRNA) untranslated regions (UTRs) are non-coding segments flanking the coding sequence (CDS) that play crucial roles in regulating translation initiation, mRNA stability, and localization. The 5' UTR is located upstream of the start codon, while the 3' UTR is downstream of the stop codon; both contain sequence elements and structures that modulate gene expression without being translated into protein. The 5' UTR serves as the primary site for ribosome recruitment and initiation codon recognition. In prokaryotes, it typically harbors the Shine-Dalgarno sequence, a purine-rich motif approximately 5-10 nucleotides upstream of the AUG start codon, which base-pairs with the anti-Shine-Dalgarno sequence in the 16S rRNA to position the ribosome accurately. Prokaryotic 5' UTRs are generally short, averaging 20-30 nucleotides, reflecting the streamlined nature of bacterial translation. In eukaryotes, the 5' UTR contains the Kozak consensus sequence surrounding the start codon (e.g., GCCA/GCC AUG G), which enhances recognition by the scanning 40S ribosomal subunit. Eukaryotic 5' UTRs average 100-200 nucleotides in length and facilitate the scanning mechanism, where the 43S pre-initiation complex binds near the 5' cap and moves downstream to identify the first suitable AUG codon. The 3' UTR exerts control over mRNA stability and translational efficiency through embedded regulatory motifs. It often includes AU-rich elements (AREs), sequences rich in and uracil (e.g., AUUUA repeats), that bind proteins to promote or inhibit decay, thereby fine-tuning transcript . Additionally, 3' UTRs serve as binding platforms for microRNAs (miRNAs), where sequences complementary to miRNA guide the (RISC) to repress or induce degradation. Eukaryotic 3' UTRs vary widely in length, typically ranging from 100 to 2000 , with averages around 1000 in humans, allowing for layered regulatory inputs. The poly-A tail, appended to the 3' end, interacts with elements in the adjacent 3' UTR to stabilize the mRNA and facilitate circularization during . Secondary structures, such as stem-loops formed by base-pairing within UTRs, significantly influence mRNA functionality. In the 5' UTR, stable hairpins can impede ribosomal scanning, reducing efficiency, while moderate structures may enhance by positioning the . In the 3' UTR, stem-loops can shield or expose regulatory sites, affecting miRNA access or protein binding that modulates stability and decay rates. Prokaryotic UTRs are characteristically shorter and simpler, with fewer regulatory elements suited to rapid, constitutive expression in unicellular organisms. In contrast, eukaryotic UTRs are longer and more complex, incorporating diverse motifs for sophisticated post-transcriptional control that supports multicellular and environmental responses.

Modifications and Variants

Messenger RNA undergoes various post-transcriptional modifications that influence its stability, localization, and function in gene expression. One of the most prevalent internal modifications is N6-methyladenosine (m⁶A), which marks adenosine residues within the mRNA sequence and is the most abundant modification in eukaryotic mRNAs. This modification is dynamically regulated by writer proteins such as METTL3 and erasers like FTO, affecting multiple aspects of mRNA metabolism, including alternative splicing through interactions with splicing factors and nuclear export via recognition by YTHDC1 protein. In eukaryotic processing, m⁶A sites are enriched in 3' untranslated regions and near stop codons, contributing to fine-tuning of mRNA fate without altering the primary sequence. Another key modification is the addition of a poly(A) tail at the 3' end, consisting of 50-250 residues in mammalian mRNAs, which is enzymatically synthesized by poly(A) polymerase during nuclear processing. This homopolymeric tail binds multiple copies of poly(A)-binding protein (PABP), enhancing mRNA stability by protecting against 3' exonucleolytic degradation and promoting efficiency through circularization of the mRNA via PABP-eIF4G interactions. The length of the poly(A) tail is tightly controlled, with longer tails correlating with increased translational output and cytoplasmic persistence. mRNA exists in distinct structural variants, primarily linear and circular forms. Linear mRNA, the canonical form, features 5' cap, coding sequence, and 3' poly(A) tail, rendering it susceptible to exonucleases but optimized for ribosomal translation. In contrast, circular mRNA (circRNA) forms through back-splicing, where a splice donor joins an upstream splice acceptor, creating a covalently closed loop that resists degradation by exonucleases due to the absence of free ends. Most circRNAs are derived from exonic sequences, though some retain intronic elements (EIcircRNAs) or arise purely from introns (ciRNAs), and they primarily function in post-transcriptional regulation, such as acting as miRNA sponges or modulating protein activity, rather than serving as templates for protein synthesis. Beyond endogenous forms, synthetic mRNA variants have emerged, particularly in therapeutic applications. In vitro transcribed (IVT) mRNA mimics endogenous linear mRNA but is engineered with optimized untranslated regions and capping analogs for enhanced stability and immunogenicity control, as seen in vaccines like those encoding . Circular synthetic mRNAs, produced via or ribozyme-mediated strategies, offer advantages over linear IVT counterparts, including greater resistance to degradation and prolonged expression, positioning them as next-generation platforms for vaccines and gene therapies. These variants highlight the versatility of mRNA engineering while preserving core functional principles of natural transcripts.

Biosynthesis

Transcription Initiation and Elongation

In prokaryotic transcription, the sigma (σ) factor associates with the core RNA polymerase to form the holoenzyme, which specifically recognizes and binds to promoter regions on the DNA. The promoter typically features conserved sequences known as the -10 box (TATAAT consensus) and the -35 box (TTGACA consensus), located upstream of the transcription start site at +1. Upon binding, the holoenzyme unwinds the DNA to form an open complex, initiating RNA synthesis at the +1 site by incorporating the first nucleotide, usually a purine. The σ factor is then released, allowing the core polymerase to enter the elongation phase, where it synthesizes the RNA transcript at an average rate of approximately 50 nucleotides per second. Eukaryotic transcription of messenger RNA (mRNA) precursors is carried out by (Pol II), which requires the assembly of a preinitiation (PIC) at the core promoter. The (TBP), a subunit of the transcription factor IID (TFIID) , binds to the , a core promoter element typically located 25-35 base pairs upstream of the transcription start site. Additional general transcription factors (TFIIA, TFIIB, TFIIE, TFIIF, and TFIIH) join to recruit Pol II, while the Mediator bridges the PIC with gene-specific activators to stabilize assembly and facilitate promoter opening. proceeds following of the C-terminal (CTD) of Pol II's largest subunit by TFIIH's subunit, which releases Pol II from the promoter and promotes processive RNA chain extension. Promoter elements dictate the specificity and efficiency of transcription . The core promoter, encompassing sequences like the , initiator (Inr), and downstream promoter element (DPE) in eukaryotes—or the -10 and -35 boxes in prokaryotes—directly interacts with the transcription machinery to define the start site and directionality. Enhancers, in contrast, are distal regulatory elements that loop to contact the core promoter via and other coactivators, enhancing transcription rates but not altering the primary site. Transcription directionality is established by the orientation of these elements relative to the antisense () strand, which is read 3' to 5' to synthesize the RNA strand 5' to 3'. A key feature of initiation in both prokaryotes and eukaryotes is , where repeatedly synthesizes and releases short RNA transcripts (typically 2-15 ) without clearing the promoter. This non-productive cycling allows the enzyme to probe promoter conformation until stable promoter clearance occurs, transitioning to productive ; during synthesis, uracil is incorporated opposite on the template strand.

Termination and Primary Transcript

In prokaryotes, transcription termination occurs through two primary mechanisms: Rho-independent and Rho-dependent. Rho-independent termination, also known as , involves the formation of a stable RNA structure in the nascent transcript, followed by a run of residues (U-run) that weakens the RNA-DNA hybrid, causing to dissociate from the DNA template. This process does not require additional protein factors and is driven solely by the sequence-specific folding of the RNA and its interaction with the polymerase. In contrast, Rho-dependent termination relies on the Rho protein, a hexameric RNA that binds to specific rut (Rho utilization) sites on the emerging RNA, translocates along the transcript in a 5' to 3' direction using , and unwinds the transcription elongation complex, leading to release. This mechanism is particularly important for terminating transcription at sites lacking strong intrinsic signals and helps prevent unwanted into downstream genes. In eukaryotes, transcription termination by RNA polymerase II (Pol II) is more complex and tightly linked to the processing of the primary transcript. Termination is triggered by the polyadenylation signal (typically AAUAAA) located in the 3' untranslated region of the pre-mRNA, which causes Pol II to pause approximately 1-2 kilobases downstream. The cleavage and polyadenylation specificity factor (CPSF) complex then recognizes this signal and recruits endonucleases, such as CPSF-73, to cleave the RNA at the poly(A) site, separating the upstream pre-mRNA from the downstream fragment. The 5'-3' exoribonuclease Xrn2 (also known as Rat1 in yeast) subsequently degrades the downstream cleaved RNA, acting as a "torpedo" that catches up to the paused Pol II, destabilizes the elongation complex, and promotes polymerase release through allosteric changes and dephosphorylation of the C-terminal domain. This process ensures efficient termination and prevents the production of aberrant extended transcripts. The primary transcript, often referred to as pre-mRNA or heterogeneous nuclear RNA (hnRNA) in eukaryotes, is the initial, unprocessed product of transcription that includes both exons and s, along with extended 5' and 3' untranslated regions beyond the mature mRNA boundaries. In prokaryotes, the primary transcript is typically mature mRNA without introns, but in eukaryotes, it encompasses the full gene sequence transcribed by Pol II, with exons representing the coding and regulatory segments (averaging 50-250 base pairs each) interspersed by introns that can span hundreds to thousands of base pairs. Eukaryotic primary transcripts can reach lengths of up to 100 kilobases or more, reflecting the expansive intron content that constitutes about 95% of the total in many protein-coding genes. These transcripts also feature temporary 5' extensions from promoter-proximal regions and 3' extensions downstream of the poly(A) site, which are later trimmed during processing. Transcription termination is functionally coupled to pre-mRNA in eukaryotes to enhance and , with termination factors like CPSF recruiting processing machinery such as capping enzymes and splicing components during . This co-transcriptional ensures that 3' end cleavage facilitates Xrn2-mediated termination while simultaneously enabling and export signals, reducing the risk of defective transcripts. In prokaryotes, termination more directly coordinates with , but the eukaryotic underscores the compartmentalized nature of .

Processing

5' Capping and Export Signals

The 5' capping of (mRNA) occurs co-transcriptionally shortly after transcription initiation, typically when the nascent transcript reaches a length of 20-30 , allowing the 5' end to emerge from the (Pol II) exit channel. This process begins with the RNA triphosphatase removing the γ-phosphate from the 5' triphosphate end of the pre-mRNA, followed by the guanylyltransferase component of the (CE, also known as RNGTT in humans) transferring a (GMP) moiety from GTP to form an unusual 5'-5' triphosphate linkage, resulting in GpppN at the 5' end. Subsequent steps involve the RNA guanine-7-methyltransferase (RNMT) adding a to the N7 position of the to produce m7GpppN, while cap methyltransferases 1 and 2 (CMTR1 and CMTR2) catalyze 2'-O-ribose on the first and second , respectively, yielding the mature cap 0 (m7GpppN) or cap 1 (m7GpppNm) structures essential for mRNA stability and function. These enzymes associate directly with the phosphorylated C-terminal domain of Pol II and the paused elongation complex, ensuring efficient coupling of capping to transcription. The primary functions of the 5' include protecting the mRNA from degradation by 5' to 3' exonucleases, such as Xrn1, thereby enhancing transcript stability during processing and export. Additionally, the cap promotes initiation by serving as a for the 4E (), which is part of the eIF4F complex that recruits the ribosomal subunit to the mRNA 5' end, facilitating scanning to the . This cap-eIF4E is critical for efficient ribosome recruitment and is modulated by of 4E-BP proteins, underscoring the cap's role in translational control. In terms of export signals, the mature 5' cap is immediately recognized by the nuclear cap-binding complex (), composed of CBP80 (NCBP1) and CBP20 (NCBP2), which binds the m7G structure with high affinity and shields it from exonucleases while recruiting the (transcription-export) complex. The CBC-TREX interaction, mediated by components like ALYREF, couples the capped mRNA to the nuclear pore complex for passage into the , ensuring that only properly capped transcripts are exported efficiently. This cap-dependent signaling also briefly coordinates with splicing factors to promote intron removal, though the primary export linkage occurs via . Unlike eukaryotic mRNA, prokaryotic transcripts lack a 5' due to the absence of Pol II-like capping machinery, relying instead on direct binding of the 30S ribosomal subunit to the Shine-Dalgarno sequence upstream of the for without cap-mediated protection or recruitment.

Splicing and Intron Removal

Splicing is a critical post-transcriptional process in eukaryotic cells that removes non-coding from pre-mRNA and ligates the coding exons to produce mature mRNA. This process is carried out by the , a large ribonucleoprotein complex composed of five small nuclear ribonucleoproteins (snRNPs: U1, , U4/U6, and U5) and numerous associated proteins. The spliceosome assembles stepwise on the pre-mRNA, recognizing specific sequence motifs at intron boundaries and internal sites to ensure precise excision and joining.00146-9) Spliceosome assembly begins with the recognition of the 5' splice site by the U1 snRNP, which base-pairs with the conserved dinucleotide at the intron-exon junction, adhering to the GU rule established from early sequence analyses of splice junctions. Subsequently, the U2 snRNP binds the branch point sequence, typically located 20–50 upstream of the 3' splice site and featuring a conserved (A) residue within a YNCURAC consensus (where Y is , N any , R ), forming base pairs with U2 snRNA to stabilize the commitment complex. The 3' splice site is marked by an AG dinucleotide, also recognized through interactions involving U2 and later U5 snRNPs, completing the early recognition phase. The splicing mechanism proceeds via two sequential transesterification reactions. In the first step, the 2'-OH group of the branch point attacks the at the 5' splice site, cleaving the 5' and forming a intermediate where the is looped via a 2'-5' . The second transesterification involves the 3'-OH of the freed 5' attacking the 3' splice site, ligating the exons and releasing the . These reactions are catalyzed within the spliceosome's , with Prp8 serving as a central that positions substrates and coordinates catalysis across both steps. Prp16, an associated with the U5 , drives conformational rearrangements and proofreading after the first step to ensure fidelity before the second transesterification. Alternative splicing allows a single pre-mRNA to generate multiple mRNA isoforms by varying inclusion, such as through , mutually exclusive exons, or retention, thereby expanding diversity. In humans, approximately 95% of multi-exon genes undergo , producing numerous isoforms that can differ in function, localization, or stability. This regulation often involves sequence elements like exonic or intronic splicing enhancers/silencers and is influenced by splicing factors that modulate splice site choice during assembly. In contrast to spliceosomal splicing, certain introns in organellar genomes, such as those in mitochondria and chloroplasts, can undergo self-splicing without proteins, relying on the RNA's intrinsic activity. Group I introns, common in fungal and plant organelles, initiate splicing with an exogenous cofactor attacking the 5' splice site, followed by exon , as first demonstrated in rRNA.90176-3.pdf) Group II introns, prevalent in bacterial and organellar genomes, mirror the spliceosomal pathway more closely by forming a lariat intermediate via attack, with self-splicing observed in mitochondrial introns. These self-splicing mechanisms highlight evolutionary links between ancient and the modern .

Polyadenylation and 3' End Formation

In eukaryotic mRNA processing, the polyadenylation signal, typically the hexanucleotide sequence AAUAAA located 10-30 nucleotides upstream of the site, is recognized by the and specificity () . Downstream of this signal, GU- or U-rich , situated approximately 20-30 nucleotides after the AAUAAA , are bound by the stimulation (), which helps position the machinery. The pre-mRNA is then cleaved endonucleolytically by the CPSF-associated endonuclease between these signals, generating the 3' end for subsequent . Following cleavage, poly(A) polymerase (PAP) catalyzes the addition of a poly(A) tail, consisting of approximately 200-250 adenine residues, to the newly exposed 3' hydroxyl group. The nuclear poly(A)-binding protein 1 (PABPN1) binds to the growing tail, stimulating PAP activity and ensuring controlled elongation until the optimal length is reached, after which it inhibits further addition to prevent over-adenylation. This length regulation is critical, as tails shorter or longer than this range can impair mRNA function. The poly(A) tail serves multiple essential functions in mRNA maturation and utilization. It promotes nuclear export by facilitating the recruitment of export adaptors such as ALYREF, which links the mRNA to the NXF1/NXT1 export receptor at the complex. In the , the tail bound by cytoplasmic poly(A)-binding protein (PABP) enhances mRNA stability by shielding the 3' end from exonucleolytic degradation, thereby extending the mRNA's half-life.01137-6.pdf) Additionally, PABP interacts with the translation initiation factor eIF4G, forming a closed-loop structure with the 5' cap that stimulates recruitment and enhances efficiency. A notable variant occurs in replication-dependent histone mRNAs, which lack a poly(A) tail and instead terminate in a conserved stem-loop structure formed through a distinct processing pathway. In this case, the U7 small nuclear ribonucleoprotein (snRNP) recognizes a specific binding site downstream of the stem-loop, directing cleavage and ligation to generate the mature 3' end, which regulates histone mRNA stability in a cell cycle-dependent manner.

RNA Editing

RNA editing refers to post-transcriptional enzymatic modifications that alter the nucleotide sequence of messenger RNA (mRNA), thereby expanding the proteome and influencing gene expression without changing the genomic DNA. These changes are catalyzed by deaminase enzymes and occur primarily in specific contexts to fine-tune protein function, stability, and regulatory interactions. The most prevalent form of RNA editing in eukaryotes is adenosine-to-inosine (A-to-I) editing, mediated by adenosine deaminases acting on RNA (ADAR) enzymes, which deaminate adenosine residues to inosine; during translation, inosine is recognized as guanosine (G) by the ribosome. ADAR1, ADAR2, and ADAR3 are the primary enzymes involved, with ADAR1 and ADAR2 being catalytically active; ADAR2 is particularly abundant in the brain, where it contributes to transcriptome diversity by editing neuronal mRNAs, such as those encoding glutamate receptors, to modulate synaptic plasticity and neurotransmitter signaling. A-to-I editing often targets double-stranded RNA structures formed by inverted Alu repeats in primates, leading to synonymous or nonsynonymous codon changes that can affect protein isoforms. In contrast, -to- (C-to-U) editing is less common and primarily mediated by the APOBEC1 enzyme, which deaminates to in specific mRNA targets. A example occurs in the (apoB) mRNA in the mammalian , where APOBEC1, in complex with cofactors like ACF, edits a CAA codon to UAA at position 6666, introducing a premature that truncates the protein to produce the shorter ApoB48 isoform essential for lipid transport, rather than the full-length ApoB100. This editing is tissue-specific and requires mooring sequence elements downstream of the target for enzyme recruitment. Genome-wide studies have identified thousands of RNA editing sites in the transcriptome, with over 14,000 A-to-I sites in more than 1,400 mRNAs reported early on, predominantly in Alu elements, though recent analyses reveal up to 189,000 cell-type-specific sites, particularly in the . These edits influence various processes, including by altering splice site recognition, coding sequence changes that modify protein function, and modulation of microRNA binding sites to affect mRNA stability and translation. For instance, A-to-I editing can recode subunits, enhancing calcium permeability in neurons. Regulation of RNA editing occurs at multiple levels, with ADAR enzymes localized differently: nuclear isoforms like edit pre-mRNAs, potentially integrating with splicing machinery to influence inclusion, while cytoplasmic forms such as target mature mRNAs or RNAs. Dysregulation is linked to diseases; for example, mutations or downregulation of ADAR2 lead to inefficient editing of the GluA2 receptor Q/R site in (ALS), causing in motor neurons and contributing to neurodegeneration.

Translation

Initiation Complex Formation

In prokaryotes, translation initiation commences with the binding of the ribosomal subunit to the messenger RNA (mRNA) at the Shine-Dalgarno (SD) sequence located in the (UTR), which base-pairs with the complementary anti-Shine-Dalgarno () sequence (CCUCC) near the 3' end of the (rRNA). This interaction aligns the , typically , in proximity to the ribosomal , ensuring accurate positioning for initiator tRNA binding. The process is facilitated by three initiation factors: IF1, which occupies the A site to block non-initiator tRNAs and stabilize the subunit; IF3, which promotes mRNA binding and prevents premature association with the 50S subunit to maintain fidelity; and IF2, a that delivers the initiator formylmethionyl-tRNA^fMet (fMet-tRNA^fMet) to the codon in the . Upon by IF2, the 50S subunit joins to form the complete 70S complex, releasing the initiation factors. In eukaryotes, initiation begins with the assembly of the 43S preinitiation complex (PIC), comprising the ribosomal subunit associated with eukaryotic initiation factors (eIFs) eIF1, eIF1A, and eIF3, along with the ternary complex of -GTP-bound initiator methionyl-tRNA^i (Met-tRNA^i). specifically recognizes and stabilizes Met-tRNA^i in the ternary complex, delivering it to the subunit's in a partially accommodated . The 43S PIC is then recruited to the mRNA's 5' cap structure (m^7GpppN) via the eIF4F complex, which includes the cap-binding protein , the multifunctional scaffold eIF4G, and the ATP-dependent RNA eIF4A; eIF4G bridges eIF4E and eIF3 to tether the ribosome to the mRNA. From this cap-bound position, the 43S PIC scans the 5' UTR in a 5'-to-3' direction, unwinding secondary structures with eIF4A's activity, until it identifies the start AUG codon. Optimal recognition of the eukaryotic depends on its surrounding sequence context, known as the Kozak consensus: GCCRCCAUGG, where R denotes a (A or G) at the -3 position relative to the , and the +4 position is preferably G; this motif enhances initiation efficiency by stabilizing codon-anticodon pairing and PIC accommodation. Mutations deviating from this consensus reduce accuracy and efficiency, as the -3 and +4 G positions interact directly with ribosomal elements and eIFs to promote GTP by and release of eIFs. The 5' UTR influences this scanning process by providing binding sites for regulatory factors that modulate movement. While most eukaryotic mRNAs rely on this cap-dependent scanning mechanism, certain viral mRNAs and cellular transcripts under conditions utilize internal ribosome entry sites (IRES) for cap-independent initiation. IRES elements, often complex RNA structures in the 5' UTR, directly recruit the subunit and associated eIFs to an internal without prior cap binding or scanning, enabling translation when cap-dependent pathways are inhibited, as first demonstrated in RNA. This alternative mode supports viral replication and cellular adaptation to stressors like .

Elongation and Codon Decoding

During elongation, the ribosome moves along the mRNA in the 5' to 3' direction, adding to the growing polypeptide chain one at a time. This process begins after the formation of the initiation complex, where the initiator tRNA occupies the and the A site is empty. Each cycle of elongation involves decoding of the mRNA codon in the A site, formation of a , and translocation of the mRNAs and tRNAs relative to the . Codon-anticodon pairing occurs when the anticodon of an incoming (aa-tRNA) base-pairs with the mRNA codon in the ribosomal A site. According to the wobble hypothesis, the third position of the codon allows for non-standard base pairing, enabling a single tRNA to recognize multiple synonymous codons and reducing the number of required tRNAs. This flexibility arises from modifications in the anticodon's first position (corresponding to the codon's third), such as pairing with A, C, or U. Selection of the aa-tRNA for the A-site codon is facilitated by . In prokaryotes, elongation factor (EF-Tu) forms a ternary complex with GTP and aa-tRNA, delivering it to the A site where codon recognition induces , releasing EF-Tu-GDP and allowing of the aa-tRNA. In eukaryotes, the homologous 1A (eEF1A) performs an analogous role, binding GTP and aa-tRNA to promote accurate decoding via induced-fit conformational changes in the upon pairing. by eEF1A ensures through kinetic , rejecting near- tRNAs. Peptide bond formation is catalyzed by the center (PTC) in the large ribosomal subunit, which is composed entirely of (rRNA) acting as a . The 23S rRNA in prokaryotes (or 28S rRNA in eukaryotes) positions the peptidyl-tRNA in the and the aa-tRNA in the A site, facilitating nucleophilic attack by the A-site amino group on the P-site ester bond without requiring protein . This rRNA-mediated reaction transfers the nascent chain to the A-site aa-tRNA. Following formation, translocation shifts the deacylated tRNA to the E site, the peptidyl-tRNA to the , and advances the mRNA by one codon to expose the next codon in the A site. In prokaryotes, elongation factor G (), bound to GTP, binds the and promotes this movement; GTP by EF-G accelerates the conformational changes in the ribosome and tRNAs, resolving hybrid states and ensuring efficient translocation. The eukaryotic counterpart, elongation factor 2 (), operates similarly, using GTP to drive tRNA and mRNA movement within the 80S . The speed of elongation varies between organisms and is influenced by codon usage. In prokaryotes, ribosomes typically incorporate 10-20 amino acids per second under optimal conditions. Eukaryotic translation is generally slower, at approximately 5-6 amino acids per second, with additional pauses at rare codons due to limited availability of corresponding tRNAs, which can regulate co-translational protein folding and quality control.

Termination and Ribosome Release

Translation termination occurs when the ribosome encounters one of three stop codons—UAA, UAG, or UGA—in the mRNA, signaling the end of protein synthesis and triggering the release of the completed polypeptide chain. In prokaryotes, RF1 recognizes UAA and UAG, while RF2 recognizes UAA and UGA; both possess peptidyl-tRNA activity that cleaves the bond linking the nascent to the tRNA in the . In eukaryotes, eRF1 decodes all three stop codons and catalyzes the , functioning in a ternary complex with GTP-bound eRF3, a that enhances termination efficiency. The GTP by RF3 or eRF3 promotes the of the class I release factors (RF1/RF2 or eRF1) from the , ensuring rapid progression to the next phase. Following peptide release, the post-termination ribosomal must be disassembled to components for new rounds of . In both prokaryotes and eukaryotes, the ATP-binding cassette protein ABCE1 plays a central role in splitting the into its / and 60S/50S subunits, facilitating the release of the deacylated tRNA and mRNA. In eukaryotes, this process is assisted by initiation factors such as eIF1 and eIF1A, which help in subunit separation and prevent premature reinitiation on the same mRNA. The freed mRNA can then be d for additional cycles or marked for , depending on cellular conditions. Although stop codons generally halt translation, certain mechanisms allow read-through in specific contexts. Suppressor tRNAs with anticodons complementary to stop codons can occasionally insert an amino acid, enabling translation to continue, though this is rare and often mutagenic. A notable exception is the recoding of UGA as selenocysteine in selenoproteins, where a selenocysteine insertion sequence (SECIS) element in the 3' untranslated region recruits selenocysteyl-tRNA^Sec and elongation factor SelB/eEFSec to decode UGA without terminating translation. Such programmed read-through is essential for incorporating this rare amino acid and exemplifies how mRNA context can override standard termination signals.

Localization and Stability

Nuclear Export Mechanisms

In eukaryotic cells, the nuclear export of messenger RNA (mRNA) is a tightly regulated process that ensures only mature, properly processed transcripts are transported from the to the for . This translocation occurs through nuclear pore complexes (NPCs), large protein assemblies embedded in the , and involves the formation of export-competent messenger ribonucleoprotein (mRNP) particles. The process is essential for , as it separates transcription in the from in the , preventing premature translation of immature mRNAs. A key player in this pathway is the (transcription-export) complex, a conserved multisubunit assembly that couples mRNA transcription and processing to export. In and mammals, the TREX complex, which includes the THO subcomplex and the RNA Sub2 (or UAP56 in humans), is recruited to nascent mRNA during transcription elongation and splicing. This recruitment facilitates the loading of the primary mRNA export receptor, NXF1 (Mex67 in ) bound to NXT1 (Mtr2 in ), onto the mRNP, directing it to the nuclear basket of the NPC for translocation. The THO component of TREX prevents formation during transcription, ensuring smooth handover to export factors, while Sub2 unwinds secondary structures to promote NXF1 binding. Seminal studies have shown that TREX disrupts mRNA export, leading to accumulation and cellular defects. Directionality of mRNA export is achieved independently of the classical Ran-GTP gradient that drives most nuclear transport, relying instead on asymmetric localization and ATP-dependent remodeling at the NPC. Although the Ran-GTP/GDP gradient maintains overall nuclear-cytoplasmic asymmetry, the NXF1-NXT1 mediated export of mRNPs does not directly require Ran-GTP for translocation. Instead, the DEAD-box ATPase Dbp5 (DDX19 in humans), anchored to the cytoplasmic of the NPC via Nup159 (Nup214), uses to unwind mRNP complexes upon arrival at the cytoplasmic side, releasing mature mRNA into the and export factors back to the . This Dbp5 cycle, stimulated by Gle1 and hexakisphosphate (IP6), ensures unidirectional transport and prevents back-diffusion of mRNPs. Quality control during export is mediated by the (EJC), a multiprotein assembly deposited 20-24 upstream of exon-exon junctions by the splicing machinery. The EJC, consisting of core components eIF4A3, MAGOH, Y14, and MLN51, marks spliced mRNAs as export-competent and distinguishes them from unspliced or aberrantly processed transcripts, which are retained in the . EJCs recruit and NXF1, enhancing export efficiency, and also flag mRNAs for post-export surveillance, such as (NMD) if premature stop codons are detected. This mechanism ensures that only high-quality mRNAs proceed to . The 5' cap and poly(A) tail, added during , briefly facilitate export by serving as sites for adaptor proteins like CBP80 and PABPN1, which indirectly to NXF1 and promote mRNP remodeling. In prokaryotes, nuclear export is irrelevant due to the absence of a ; instead, transcription and are directly coupled in the , with ribosomes nascent mRNA as it emerges from .

Cytoplasmic Trafficking and Localization

Once in the cytoplasm, messenger RNAs (mRNAs) are assembled into messenger ribonucleoprotein (mRNP) complexes that facilitate their trafficking and localization to specific subcellular sites, enabling spatially restricted protein synthesis. This process is crucial for cellular asymmetry, such as in polarized cells like neurons and oocytes. Localization signals, often termed "zipcodes," are primarily located in the 3' untranslated region (3' UTR) of mRNAs and serve as recognition motifs for RNA-binding proteins (RBPs) that direct mRNPs to target destinations.00126-3) For instance, the β-actin mRNA contains a 54-nucleotide zipcode in its 3' UTR that binds the RBP ZBP1 (zipcode-binding protein 1), which mediates transport to neuronal dendrites, supporting actin cytoskeleton dynamics at synaptic sites. These interactions ensure that mRNAs are packaged into transport-competent mRNPs shortly after nuclear export.00651-7) Directed transport of mRNPs relies on motor proteins that move along the , particularly . motors, such as kinesin-1, drive plus-end-directed transport toward the periphery, while powers minus-end-directed movement toward the microtubule-organizing center. In asymmetric distribution, these motors coordinate to position mRNAs; for example, in oocytes, transports gurken mRNA to the anterior-dorsal region for eggshell patterning, while kinesin-1 relocates oskar mRNA to the posterior pole for specification.01302-7) This motor-driven mechanism is essential for long-distance trafficking in large s, where mRNPs form granules visible by and associate with via adaptor proteins.00602-X) mRNP granules play key roles in cytoplasmic regulation and storage during trafficking. Processing bodies (P-bodies) sequester mRNAs for translational repression or decay, acting as hubs for mRNA surveillance and without directly driving localization.00643-X) granules, induced by cellular , temporarily store mRNAs by halting , allowing rapid resumption upon stress relief; they often dock with P-bodies, facilitating mRNA exchange and contributing to spatiotemporal control in the cytoplasm.01027-9) In contrast to directed , shorter mRNAs typically rely on passive for local positioning, whereas longer or structurally complex mRNAs favor active, motor-mediated delivery to overcome cytoplasmic barriers.01213-8) This dichotomy ensures efficient , with diffusion sufficing for and directed mechanisms enabling precise asymmetry.

Degradation

Prokaryotic mRNA Decay Pathways

In prokaryotes, particularly bacteria like , mRNA decay is a rapid process that ensures quick adaptation to environmental changes, with an average mRNA of approximately 3-7 minutes under conditions. This turnover is primarily mediated by a combination of endonucleolytic and exonucleolytic activities, often coupled to , and contrasts with the longer-lived eukaryotic mRNAs. The core machinery includes ribonucleases such as RNase E, polynucleotide phosphorylase (PNPase), and RNase II, which collectively degrade mRNA from internal sites or the ends. A major initiation pathway involves endonucleolytic cleavage by RNase E, a key enzyme in the RNA degradosome complex, which targets unstructured regions, stem-loops, or monosome-bound mRNAs. RNase E preferentially cleaves at A/U-rich sites downstream of the 5' end, often in a translation-independent manner, generating fragments that are subsequently susceptible to exonucleolytic attack. In polycistronic mRNAs, common in , such cleavages can differentially destabilize individual cistrons, allowing coordinated yet modular gene expression. Following endonucleolytic cuts or direct 3' end processing, degradation proceeds via 3'-5' exonucleases like PNPase and RNase II, which require prior shortening of the 3' end. Unlike eukaryotes, prokaryotic mRNAs lack extensive poly(A) tails; instead, limited polyadenylation by poly(A) polymerase I (PAP I) adds short A-tails to facilitate processive degradation by these exonucleases. PNPase, a phosphorolytic enzyme, degrades from the 3' end using phosphate as a cofactor, while RNase II hydrolyzes phosphodiester bonds but stalls at stem-loops. An alternative 5'-3' decay pathway begins with the RNA pyrophosphohydrolase RppH, which converts the 5'-triphosphate end of primary transcripts to a monophosphate, priming the mRNA for exonucleolytic degradation. This RppH-mediated is often translation-coupled, as ribosome protection hinders access, and is followed by enhanced endonucleolytic cleavage primarily by RNase E, with subsequent 3'-5' exonucleolytic degradation by enzymes such as PNPase. The efficiency of this pathway depends on 5' end accessibility and can be modulated by upstream open reading frames or secondary structures. mRNA stability in bacteria is further regulated by small regulatory RNAs (sRNAs) that base-pair with target mRNAs, often facilitated by the chaperone protein Hfq, leading to enhanced recruitment of RNases like RNase E for accelerated decay. For instance, Hfq-sRNA complexes can expose cleavage sites or block translation, thereby promoting endonucleolytic initiation and shortening mRNA lifespan in response to stress. This post-transcriptional control layer allows fine-tuned regulation without altering transcription rates.

Eukaryotic mRNA Turnover Processes

In eukaryotic cells, mRNA turnover is a tightly regulated process that determines transcript stability and levels, primarily occurring in the through a deadenylation-dependent pathway that contrasts with the more rapid, translation-coupled decay seen in prokaryotes. This basal pathway ensures the removal of mRNAs after their functional lifespan, recycling and preventing accumulation of potentially harmful transcripts. The initial and rate-limiting step in most eukaryotic mRNA decay is deadenylation, where the poly(A) tail is progressively shortened by deadenylase complexes. The CCR4-NOT complex, a major multi-subunit deadenylase, plays a central role by recruiting to the mRNA via interactions with poly(A)-binding proteins (PABPs) and catalyzing the removal of adenylate residues through its catalytic subunits and Caf1 (also known as Pop2). This process typically reduces the poly(A) tail length from over 200 to a stub of 10-20 adenines, which destabilizes the mRNA and triggers subsequent decay steps. Shortening of the poly(A) tail promotes , the of the 5' cap structure (m7GpppN) by the Dcp1/Dcp2 heterodimeric . Dcp2 provides the catalytic activity, while Dcp1 acts as a cofactor that enhances decapping efficiency and recruits other factors. Once the cap is removed, the mRNA body becomes accessible to the 5'-3' exoribonuclease Xrn1, which rapidly degrades the transcript from the 5' end in a processive manner. This decapping-dependent 5'-3' pathway accounts for the majority of bulk mRNA turnover in eukaryotes. In parallel or as an alternative route, particularly for aberrant or unadenylated mRNAs, the RNA exosome complex mediates 3'-5' exonucleolytic . The cytoplasmic exosome, assisted by the cofactor Ski7, targets non-polyadenylated or prematurely deadenylated transcripts for and , ensuring of defective mRNAs. In the nucleus, the exosome subunit Rrp6 (also known as R) contributes to the processing and of aberrant transcripts before export, further supporting mRNA . Eukaryotic mRNA half-lives vary widely, typically ranging from several hours to days, allowing for fine-tuned control of protein synthesis. Factors such as —where optimal codons correlate with increased stability—and mRNA secondary structure in the coding sequence influence decay rates by modulating translation efficiency and accessibility to decay factors. For instance, mRNAs enriched in optimal codons exhibit longer half-lives, while structured regions can protect against rapid degradation.

Regulatory Decay Mechanisms

Messenger RNA (mRNA) degradation serves as a critical regulatory mechanism to fine-tune by targeting specific transcripts for rapid decay under physiological conditions. These pathways, distinct from constitutive turnover, respond to sequence features or cellular signals to selectively eliminate mRNAs, thereby controlling protein levels in processes like , response, and immune . Key examples include surveillance systems that detect aberrant transcripts and pathways that silence endogenous or viral genes. Nonsense-mediated decay (NMD) is a pathway that targets mRNAs containing premature termination codons (PTCs) for degradation, preventing the production of truncated proteins. In eukaryotes, NMD recognizes PTCs located more than 50 upstream of an exon-exon junction, where the (EJC) is deposited during splicing. The UPF1 RNA , along with UPF2 and UPF3, forms a that interacts with the EJC; UPF2 and UPF3 bridge UPF1 to the EJC, stimulating UPF1's activity to unwind the mRNA and recruit decay factors. AU-rich elements (), often found in the 3' untranslated regions (UTRs) of mRNAs encoding cytokines and proto-oncogenes, mediate rapid decay to limit inflammatory responses. The zinc finger protein tristetraprolin (TTP) binds directly to these , such as in tumor factor-alpha (TNF-α) mRNA, and recruits the deadenylation machinery to shorten the poly(A) tail, thereby promoting and exonucleolytic degradation. This TTP-ARE interaction is regulated by , which modulates TTP's binding affinity and decay-promoting activity. MicroRNAs (miRNAs) regulate post-transcriptionally by guiding the (RISC), which includes proteins, to complementary sites in the 3' UTR of target mRNAs. Binding of -loaded miRISC to the 3' UTR recruits GW182 (also known as TNRC6), which interacts with deadenylation complexes like CCR4-NOT to trigger poly(A) tail removal, followed by decapping and 5'-to-3' exonucleolytic decay, often without significant translational repression in animals. This mechanism silences hundreds of genes involved in and . Small interfering RNAs (siRNAs) mediate precise mRNA silencing through RISC in both and , primarily for antiviral defense and endogenous regulation. In , siRNAs derived from viral double-stranded RNA direct proteins in RISC to cleave complementary viral or endogenous transcripts via perfect base-pairing. In , siRNAs contribute to antiviral responses by targeting viral genomes and also silence endogenous transposons or repetitive elements, enhancing genome stability.

Regulation and Functions

Post-Transcriptional Regulation

Post-transcriptional regulation of messenger RNA (mRNA) encompasses mechanisms that fine-tune after transcription, including spatial localization that controls where and when occurs. mRNA localization to specific subcellular compartments enables precise spatial regulation of protein synthesis, ensuring proteins are produced at the right time and place to support cellular functions such as and . For instance, in animal embryos, maternally deposited mRNAs are localized and translationally repressed until fertilization, allowing coordinated activation to drive early developmental processes like axis formation in or cell fate specification in vertebrates. This spatial control restricts translation to targeted sites, preventing ectopic protein production and enhancing efficiency in resource-limited environments. RNA-binding proteins (RBPs) play a central role in by modulating mRNA stability and through interactions with untranslated regions (UTRs). The RBP HuR binds to AU-rich elements () in the 3' UTRs of target mRNAs, promoting their stabilization and increasing protein output, as seen in the regulation of inflammatory cytokines like TNF-α where HuR competes with destabilizing factors to extend mRNA . In contrast, tristetraprolin (TTP) recognizes similar to recruit machinery, accelerating mRNA and suppressing excessive immune responses; for example, TTP targets mRNAs encoding inhibitors of , maintaining by preventing overproduction. These antagonistic actions of HuR and TTP exemplify how RBPs achieve dynamic control over mRNA fate via UTR sequences. Biomolecular phase separation further contributes to localized mRNA regulation by forming membraneless condensates that compartmentalize mRNA-ribonucleoprotein (mRNP) complexes. RNAs and RBPs drive liquid-liquid to create dynamic droplets enriched in specific mRNAs, which sequester transcripts for localized control of and , as observed in cytoplasmic granules that buffer mRNA stoichiometries and restrict access to ribosomes. In human embryonic stem cells, for instance, FXR1-containing condensates spatially organize mRNPs to influence by concentrating regulatory RNAs and proteins. These condensates provide a for efficient, insulated reactions, enhancing spatiotemporal precision in . Feedback loops involving mRNAs that encode regulators of their own processing represent another layer of autoregulation, ensuring balanced expression of splicing factors and other RBPs. Splicing factors often bind their own pre-mRNAs to promote events that include premature stop codons, triggering (NMD) to autoregulate levels and prevent toxic accumulation, as demonstrated in networks involving and hnRNPs. For example, RBPMS, a master regulator of smooth muscle splicing, engages in such loops to maintain during . These auto-regulatory mechanisms, including positive feedbacks with transcription factors, robustly coordinate post-transcriptional events during development.

Non-Coding and Emerging Roles

Circular RNAs (circRNAs) derived from mRNA loci represent a major class of non-coding transcripts with regulatory functions distinct from protein synthesis. These circRNAs are produced via back-splicing of pre-mRNA exons, resulting in stable, closed-loop structures resistant to exonuclease degradation. A key example is ciRS-7 (also known as CDR1as), generated from the CDR1 locus, which functions primarily as a microRNA (miRNA) sponge. ciRS-7 harbors more than 70 conserved binding sites for miR-7, sequestering the miRNA and derepressing its targets, such as those involved in neuronal function; this was first identified through high-throughput sequencing and functional assays in human and mouse brain tissues. Such sponging activity exemplifies how circRNAs from protein-coding genes modulate post-transcriptional gene regulation without translating into proteins. While predominantly non-coding, certain circRNAs exhibit protein-coding potential, challenging traditional classifications. Translation occurs via cap-independent mechanisms, including internal ribosome entry sites (IRES) or N6-methyladenosine (m6A) modifications that recruit to the circular . For instance, circ-ZNF609, derived from the ZNF609 mRNA locus, encodes a short protein that promotes myoblast and during muscle development, as demonstrated by and knockout studies in models. This capability has been observed in a subset of circRNAs, where the encoded peptides regulate cellular processes independently of their linear counterparts. Brief reference to their structural origins highlights how back-splicing events from linear mRNA precursors enable these diverse roles. Linear mRNAs also contribute non-coding functions by serving as scaffolds for protein complexes in signaling pathways. Such scaffolding roles extend mRNA utility beyond coding, organizing ribonucleoprotein complexes for efficient cellular responses. Emerging evidence positions mRNA export as a cellular stress sensor. Under conditions like heat shock, global mRNA export is inhibited through the inactivation and dissociation of export adaptors and guard proteins from the export receptor NXF1 (also known as TAP), such as via phosphorylation of Nab2 by the MAPK kinase Slt2 in yeast, retaining bulk transcripts in the nucleus while permitting selective export of stress-inducible mRNAs (e.g., heat shock proteins). This adaptive response prioritizes survival gene expression and was elucidated through studies on nuclear retention dynamics in yeast and mammalian cells. For viral infections, export can be modulated differently, often through viral interference with export factors. mRNAs further participate in phase-separated organelles, membraneless compartments formed by liquid-liquid phase separation (LLPS). In stress granules and , mRNAs act as scaffolds or modulators, recruiting RNA-binding proteins like G3BP1 to drive condensate assembly and sequester stalled complexes. Specific mRNA secondary structures influence LLPS specificity, as shown in polyglutamine-driven systems where mRNA length and sequence dictate partitioning into droplets. This role enhances mRNA stability and regulates under stress, with implications for neurodegeneration. Distinguishing mRNAs from long non-coding RNAs (lncRNAs) relies on coding potential: mRNAs contain a typically encoding proteins of at least 100 , enabling ribosomal , whereas lncRNAs (>200 ) lack substantial ORFs and primarily exert regulatory effects. Despite this, functional overlap exists, as some mRNAs perform lncRNA-like roles (e.g., ) without relying on their CDS, underscoring convergent evolutionary adaptations in RNA functionality.

History and Applications

Historical Milestones

The concept of messenger RNA (mRNA) as an intermediary carrier of genetic information from DNA to protein synthesis was first proposed in 1961 by and , who described it within their model of in , suggesting that mRNA serves as a transient template dictated by structural genes and regulated by operator regions. This theoretical framework was experimentally validated later that year through pulse-labeling experiments by , , and , which demonstrated the existence of a short-lived, rapidly turning-over species in that correlates with β-galactosidase synthesis. In the , the discovery of heterogeneous nuclear RNA (hnRNA) revealed large, rapidly labeled nuclear transcripts in eukaryotic cells, identified by Sheldon Penman and colleagues as polydisperse RNAs with sizes up to 100 kb, serving as precursors to mature mRNA. Concurrently, Phillip A. Sharp and independently uncovered split genes and in 1977 while studying adenovirus transcripts, showing that non-coding introns are removed from pre-mRNA to form continuous coding exons, a finding that earned them the 1993 in or . Their work demonstrated that eukaryotic genes are discontinuous, with splicing enabling diverse protein isoforms from single genes. The 1970s also saw the elucidation of mRNA modifications essential for stability and processing. Aaron J. Shatkin and Yasuhiro Furuichi identified the 5' cap structure in 1975, a 7-methylguanosine linked via a 5'-5' triphosphate bridge to the mRNA's first , initially observed in reovirus mRNA and later confirmed in eukaryotic cellular mRNAs to protect against exonucleases and facilitate . James E. Darnell and coworkers discovered the poly(A) tail in 1971, a 3' addition of 100–250 residues to most eukaryotic mRNAs, and by the 1980s established its roles in nuclear export, enhancement, and mRNA stability through studies on cell transcripts. During the 1990s and 2000s, advances in sequencing technologies enabled the cataloging of patterns, with B. R. Graveley and others compiling comprehensive databases from expressed sequence tags (ESTs), revealing that over 60% of human genes undergo to generate proteomic diversity, as detailed in early genome-wide analyses like those from the era. The discovery of microRNAs (miRNAs) as key post-transcriptional regulators stemmed from Victor Ambros's identification of lin-4 in 1993, but gained mechanistic insight through Andrew Z. Fire and Craig C. Mello's 1998 experiments in C. elegans, showing that double-stranded RNA triggers sequence-specific mRNA degradation via (RNAi), for which they received the 2006 in or . In the , the epitranscriptome emerged as a dynamic layer of mRNA regulation, with N6-methyladenosine (m6A) modifications mapped genome-wide; key studies by Dan Dominissini, , and Samie R. Jaffrey in 2012–2013 identified m6A as the most abundant internal mRNA modification, influencing splicing, , and through (e.g., METTL3), reader (e.g., YTHDF2), and (e.g., FTO) proteins. Concurrently, CRISPR-based tools expanded to , with Omar O. Abudayyeh and Feng Zhang's 2017 discovery of Cas13 enabling programmable cleavage and base editing of mRNAs without altering the , advancing targeted transcript modulation in eukaryotic systems.

Biotechnological and Therapeutic Uses

Messenger RNA (mRNA) has emerged as a versatile platform in and therapeutics, enabling rapid production of s and proteins for and treatments. The most prominent application is in , where synthetic mRNA instructs cells to produce viral proteins, triggering immune responses without using live pathogens. This approach accelerated during the , with mRNA from Pfizer-BioNTech and receiving emergency authorization in late 2020. By 2025, over 13 billion doses of , including a significant number of mRNA-based doses, have been administered globally, significantly reducing severe illness and hospitalizations. Self-amplifying mRNA , which encode replicase enzymes to amplify production within cells, offer potential for lower dosing and longer-lasting immunity; for instance, ARCT-154 demonstrated superior persistence compared to conventional mRNA in phase 3 trials completed by 2025 and received authorization in in 2023. By 2025, mRNA platforms have expanded to other respiratory viruses, with Moderna's mRNA-1345 receiving FDA approval for expanded use in preventing RSV lower disease in adults aged 18–59 at increased risk. Moderna's mRNA-1010 quadrivalent seasonal also reported positive phase 3 efficacy data in mid-2025, showing relative vaccine efficacy against A and B strains, paving the way for potential regulatory approval. In therapeutic applications, mRNA enables targeted protein expression for disease treatment. Personalized cancer immunotherapies represent a key advance, with developing individualized mRNA vaccines that encode neoantigens derived from patient tumor mutations to stimulate T-cell responses. For example, autogene cevumeran (BNT122) has advanced to phase 2 trials for pancreatic and other cancers, showing durable T-cell activation in three-year follow-up data from phase 1 studies. mRNA also facilitates protein replacement therapies, particularly for ischemic conditions. AZD8601, an mRNA encoding VEGF-A delivered via intramyocardial injection, has been evaluated in clinical trials for patients with undergoing coronary artery bypass grafting (CABG), promoting to reduce myocardial ischemia in preclinical models and early human studies, with potential applications for refractory . These therapies leverage mRNA's transient expression to avoid long-term risks associated with vectors. mRNA serves as a critical tool in research, produced via transcription (IVT) for various applications. IVT mRNA encoding reporter proteins like or GFP is widely used in assays to monitor efficiency, mRNA stability, and cellular responses in high-throughput screens. In genome editing, IVT mRNA delivers CRISPR-Cas9 guide RNAs as ribonucleoproteins, enabling precise, temporary modifications without genomic integration, though chemical modifications are often required to mitigate innate immune activation via RIG-I pathways. employs mRNA circuits for cellular behaviors, such as inducible protein expression in response to transcription factors, facilitating the of logic gates and metabolic pathways in mammalian cells. Delivery and stability challenges have driven innovations essential to mRNA's success. Lipid nanoparticles (LNPs) encapsulate mRNA, shielding it from and promoting endosomal escape for cytosolic delivery, as validated in the lipid formulations of approved vaccines. Incorporation of modified nucleosides, such as , further enhances performance by reducing recognition by Toll-like receptors and RIG-I, thereby minimizing inflammatory responses while boosting translation yields up to tenfold in human cells. These modifications, combined with optimized 5' caps and poly-A tails in IVT processes, have enabled mRNA's transition from research reagent to scalable therapeutic modality.

References

  1. [1]
    Messenger RNA (mRNA)
    Messenger RNA (mRNA) is a single-stranded RNA involved in protein synthesis, carrying protein information from DNA to the cytoplasm. It acts as a translator ...
  2. [2]
    Ribonucleic Acid (RNA) Fact Sheet
    May 24, 2024 · The functions of RNA are broad and include carrying biological information, providing structure, facilitating chemical reactions and ...
  3. [3]
    1961: mRNA Ferries Information
    Apr 26, 2013 · Sydney Brenner, Francois Jacob, and Matthew Meselson discovered that mRNA is the molecule that takes information from DNA in the nucleus to the protein-making ...
  4. [4]
    Who discovered messenger RNA? - PubMed
    Jun 29, 2015 · The announcement of the discovery of messenger RNA (mRNA) and the cracking of the genetic code took place within weeks of each other in a climax of scientific ...
  5. [5]
    From DNA to RNA - Molecular Biology of the Cell - NCBI Bookshelf
    Transcription and translation are the means by which cells read out, or express, the genetic instructions in their genes. Because many identical RNA copies ...
  6. [6]
    RNA Processing and Turnover - The Cell - NCBI Bookshelf - NIH
    The processing of mRNA involves modification of the 5′ terminus by capping with 7-methylguanosine (m7G), modification of the 3′ terminus by polyadenylation, and ...Processing of Ribosomal and... · Processing of mRNA in... · Splicing Mechanisms
  7. [7]
    What are mRNA vaccines and how do they work? - MedlinePlus
    Nov 21, 2022 · Messenger RNA is a type of RNA that is necessary for protein production. Once cells finish making a protein, they quickly break down the mRNA.
  8. [8]
    3′ end mRNA processing: molecular mechanisms and implications ...
    Messenger RNA 3′ end processing is a well-orchestrated process that involves components of the transcription, the splicing and the translation machinery.The Eukaryotic Mrna 3′... · Mutations Of Sequence... · Role Of Trans-Acting Factors
  9. [9]
    An Unstable Intermediate Carrying Information from Genes ... - Nature
    An unstable intermediate carrying information from genes to ribosomes for protein synthesis. Nature volume 190, pages 576–581 (1961)
  10. [10]
    Central Dogma of Molecular Biology - Nature
    Aug 8, 1970 · Central Dogma of Molecular Biology. FRANCIS CRICK. Nature volume 227, pages 561–563 (1970)Cite this article.
  11. [11]
  12. [12]
  13. [13]
    mRNA cap regulation in mammalian cell function and fate - PMC - NIH
    7-Methylguanosine is linked to the first transcribed nucleotide via a 5′ to 5′ triphosphate bridge. The first transcribed nucleotide is methylated on the O-2 ...
  14. [14]
    Translation: DNA to mRNA to Protein | Learn Science at Scitable
    Translation involves two steps: first, DNA is transcribed into mRNA. Then, mRNA is read to assemble proteins using the genetic code.
  15. [15]
    Biochemistry, RNA Structure - StatPearls - NCBI Bookshelf - NIH
    Jul 29, 2023 · Three main types of RNA are involved in protein synthesis. They are messenger RNA (mRNA), transfer RNA (tRNA), and ribosomal RNA (rRNA).
  16. [16]
    Protein Synthesis - An On-Line Biology Book
    mRNA molecules are long (500- 10,000 nucleotides). Ribosomes are the organelle (in all cells) where proteins are synthesized. They consist of two-thirds rRNA ...
  17. [17]
    Single-molecule visualization of mRNA circularization during ... - NIH
    Jan 31, 2023 · Most mRNAs are circularized via the eIF4E–eIF4G–PABP interaction, which stabilizes mRNAs and enhances translation by recycling ribosomes.
  18. [18]
    The role of m6A modification in the biological functions and diseases
    Feb 21, 2021 · N6-methyladenosine (m6A) is the most prevalent, abundant and conserved internal cotranscriptional modification in eukaryotic RNAs ...
  19. [19]
    YTHDC1 mediates nuclear export of N6-methyladenosine ... - eLife
    Oct 6, 2017 · YTHDC1 incorporates target mRNAs into the nuclear export pathway by interaction with SRSF3, delivering RNA to the splicing and adaptor protein.
  20. [20]
    Control of poly(A) tail length - Eckmann - 2011 - WIREs RNA
    Nov 17, 2010 · Poly(A) tails of mRNAs have an initial length of 70–80 nucleotides in yeast and ∼250 nucleotides in mammalian cells. These long tails have a ...Missing: adenine | Show results with:adenine
  21. [21]
    Roles of mRNA poly(A) tails in regulation of eukaryotic gene ...
    Mar 13, 2023 · The poly(A) tail contributes to both the translational status and stability of mRNAs, and it functions as a master regulator of gene expression ...
  22. [22]
    Circular RNA: metabolism, functions and interactions with proteins
    Dec 14, 2020 · Circular RNAs (CircRNAs) are single-stranded, covalently closed RNA molecules that are ubiquitous across species ranging from viruses to mammals.
  23. [23]
    Review The Biogenesis, Functions, and Challenges of Circular RNAs
    Aug 2, 2018 · Circular intronic RNAs (ciRNAs) represent another class of circular RNA molecules. They are derived from lariat introns of Pol II transcripts ...
  24. [24]
    In Vitro Transcribed RNA-Based Platform Vaccines - NIH
    Oct 16, 2023 · This review presents an overview of the mRNA platform and a critical discussion of the more modern self-amplifying mRNA, trans-amplifying mRNA, and circular ...
  25. [25]
    The interaction between bacterial transcription factors and RNA ...
    In the initiation complex the σ2.4 and σ4.2 bind to the −10 and −35 promoter element respectively and in this conformation σ2.2 binds to the CH domain of the β' ...
  26. [26]
    15.2: Prokaryotic Transcription - Biology LibreTexts
    Apr 9, 2022 · Although promoters vary among prokaryotic genomes, a few elements are conserved. At the -10 and -35 regions upstream of the initiation site ...
  27. [27]
    RNA Polymerase Elongation Factors - PMC - PubMed Central - NIH
    The average transcription rate of E. coli RNAP is in the range of 50–100 nucleotides per second, equaling a translocation step time of 10–20 ms. But the ...
  28. [28]
    Structure and mechanism of the RNA polymerase II transcription ...
    TBP (TATA binding protein) has ancient origins as a transcription factor; in eukaryotes, TBP regulates Pol II transcription as a component of the TFIID complex ...
  29. [29]
    Eukaryotic core promoters and the functional basis of transcription ...
    RNA polymerase II (Pol II) core promoters are specialized DNA sequences at transcription start sites of protein-coding and non-coding genes.
  30. [30]
    Assembly of RNA polymerase II transcription initiation complexes - NIH
    TFIID together with Mediator helps stabilize TFIIH to facilitate promoter opening and CTD phosphorylation. Thus, PIC assembly is finely choreographed and is ...
  31. [31]
    Article TFIIH Phosphorylation of the Pol II CTD Stimulates Mediator ...
    May 22, 2014 · TFIIH phosphorylation of the CTD causes Mediator dissociation, thereby permitting rapid promoter escape of Pol II from the preinitiation complex.Missing: box | Show results with:box
  32. [32]
    The Determinants of Directionality in Transcriptional Initiation - PMC
    Core promoter elements work synergistically to establish transcriptional directionality. Core promoter elements are vital components in determining whether ...
  33. [33]
    Promoters - Addgene
    The Pribnow box (TATAAT) is located at the -10 position and is essential for transcription initiation. The -35 position, simply titled the -35 element ...
  34. [34]
    Promoter clearance by RNA polymerase II - PMC - PubMed Central
    Apr 17, 2025 · For bacterial RNA polymerase, transcription passes through many cycles of abortive initiation in which short RNAs, typically up to about 10 nt, ...
  35. [35]
    Abortive Initiation - an overview | ScienceDirect Topics
    Abortive initiation is a non-productive step in transcription where RNA polymerase synthesizes and releases short RNA transcripts, which are released without ...
  36. [36]
    RNA - Transcription - Chemistry LibreTexts
    Jul 4, 2022 · One major difference is that the heterocyclic amine, adenine, on DNA codes for the incorporation of uracil in RNA rather than thymine as in DNA.
  37. [37]
    Rho-dependent transcription termination: mechanisms and roles in ...
    Intrinsic (Rho-independent) termination relies on specific transcript motifs—a stable hairpin structure followed by a run of uracil residues—sufficient to ...
  38. [38]
    Bacterial Transcription Terminators: The RNA 3′-End Chronicles
    Rho-dependent terminators are sites of dissociation mediated by an RNA helicase called Rho. Despite decades of study, the molecular mechanisms of both intrinsic ...
  39. [39]
    The bacterial transcription termination factor Rho coordinates Mg2+ ...
    The bacterial protein Rho triggers transcription termination at the ends of many operons and when transcription and translation become uncoupled.Missing: prokaryotic | Show results with:prokaryotic
  40. [40]
    RNA polymerase II termination: Means to an end
    c | In poly(A)-dependent termination in humans, pausing of human Pol II is induced when CPSF bound to the body of Pol II recognizes the AAUAAA signal sequence ...
  41. [41]
    Xrn2 Accelerates RNA Polymerase II Termination by CPSF73 Activity
    They generated human cell lines from which the 5′-to-3′ exoribonuclease Xrn2 or the poly(A) signal endoribonuclease CPSF73 can be rapidly controlled and show ...
  42. [42]
    An end in sight? Xrn2 & RNA polymerase II termination
    This torpedo model posits that the degrading exonuclease chases down Pol II and somehow signals termination. The role of a molecular torpedo in the termination ...
  43. [43]
    Knowing when to stop: Transcription termination on protein-coding ...
    A functional mRNA polyadenylation signal is required for transcription termination by RNA polymerase II. ... Xrn2 accelerates termination by RNA polymerase II ...
  44. [44]
    Transcriptomes and Proteomes - Genomes - NCBI Bookshelf - NIH
    Unspliced pre-mRNA forms the nuclear RNA fraction called heterogenous nuclear RNA (hnRNA). Some eukaryotic pre-rRNAs and pre-tRNAs also contain introns, as ...
  45. [45]
    Gene expression and regulation - Autoimmunity - NCBI Bookshelf
    Prokaryote cells have a specific sequence that is 3 to 9 nucleotides in length (UAAGGAGG) in 5'UTR mRNA called Shine–Dalgarno, which is complementary to the 3' ...Missing: simpler | Show results with:simpler
  46. [46]
    Reflections on the history of pre-mRNA processing and highlights of ...
    This review discusses more than thirty years of experimental data relating to the processing of pre-messenger RNA.
  47. [47]
    Non-coding RNAs: the architects of eukaryotic complexity - PMC
    In humans, introns account for ∼95% of the pre-mRNA transcripts of protein coding genes, and are generally of high sequence complexity. As far as can be judged ...
  48. [48]
    Coupling pre-mRNA processing to transcription on the RNA ... - NIH
    The vast majority of precursor mRNAs (pre-mRNAs) undergo 5′ capping, splicing and 3′ polyadenylation before export to the cytoplasm for subsequent translation ...
  49. [49]
    Coupling mRNA processing with transcription in time and space - PMC
    The co-transcriptional nature of mRNA processing has permitted the evolution of coupling mechanisms that coordinate transcription with mRNA capping, splicing, ...
  50. [50]
    Structures of co-transcriptional RNA capping enzymes on paused ...
    May 30, 2024 · Our structures indicate that RNGTT and CMTR1 directly bind the paused elongation complex, illuminating the mechanism by which 5'-end capping of pre-mRNA during ...
  51. [51]
    Regulation and function of CMTR1‐dependent mRNA cap methylation
    mRNA is modified co‐transcriptionally at the 5′ end by the addition of an inverted guanosine cap structure which can be methylated at several positions.
  52. [52]
    Structural insights into human co-transcriptional capping
    Jul 20, 2023 · Here, we provide mechanistic insights into the three major steps of human co-transcriptional pre-mRNA capping based on six different cryoelectron microscopy ( ...
  53. [53]
    mRNA capping: biological functions and applications
    In addition to its essential role of cap-dependent initiation of protein synthesis, the mRNA cap also functions as a protective group from 5′ to 3′ exonuclease ...
  54. [54]
    Cap-dependent eukaryotic initiation factor-mRNA interactions ...
    Cap-dependent ribosome recruitment to eukaryotic mRNAs during translation initiation is stimulated by the eukaryotic initiation factor (eIF) 4F complex and eIF ...
  55. [55]
    Photocaged 5′ cap analogues for optical control of mRNA ... - Nature
    Jun 20, 2022 · The primary hallmark of eukaryotic mRNAs is their 5′ cap, whose molecular contacts to the eukaryotic translation initiation factor eIF4E govern ...
  56. [56]
    Distinct Functions of the Cap-Binding Complex in ... - PubMed - NIH
    Apr 2, 2019 · Cap-binding complex (CBC) associates cotranscriptionally with the cap structure at the 5' end of nascent mRNA to protect it from exonucleolytic degradation.
  57. [57]
    Human mRNA Export Machinery Recruited to the 5′ End of mRNA
    We show that the human TREX complex is recruited to a region near the 5′ end of mRNA, with the TREX component Aly bound closest to the 5′ cap.
  58. [58]
    The nuclear cap-binding complex as choreographer of gene ...
    The largely nuclear cap-binding complex (CBC) binds to the 5′ caps of RNA polymerase II (RNAPII)-synthesized transcripts and serves as a dynamic interaction ...
  59. [59]
    A catalogue of splice junction sequences - PMC - NIH
    Splice junction sequences from a large number of nuclear and viral genes encoding protein have been collected. The sequence CAAG/GTAGAGT was found to be a ...
  60. [60]
    Excision of an intact intron as a novel lariat structure ... - PubMed
    Abstract. To study the mechanisms of RNA splicing we have analyzed the products generated by in vitro processing of a truncated 32P-labeled human beta-globin ...
  61. [61]
    Prp8 protein: At the heart of the spliceosome - PMC - PubMed Central
    Pre-mRNA splicing involves two trans-esterification reactions within the highly dynamic spliceosome complex. A vast amount of mainly biochemical data led to a ...Missing: seminal papers
  62. [62]
    Deep surveying of alternative splicing complexity in the human ...
    By combining mRNA-Seq and EST-cDNA sequence data, we estimate that transcripts from approximately 95% of multiexon genes undergo alternative splicing and that ...Missing: prevalence | Show results with:prevalence
  63. [63]
    Mechanism and regulation of mRNA polyadenylation
    Nearly every known mRNA contains a polyadenylation signal sequence, the hexanucleotide AAUAAA, 10–30 bases upstream of the cleavage/polyadenylation site. AAUAAA ...
  64. [64]
    Determinants and Implications of mRNA Poly(A) Tail Size - NIH
    The nuclear poly(A)-binding protein (PABPN1) binds to the newly formed poly(A) tail and to allow rapid addition of A-residues until the tail is about 200-250 nt ...
  65. [65]
    Role of poly (A) tail as an identity element for mRNA nuclear export
    RNA-binding adaptor proteins, including Aly/REF and shuttling SR proteins, are first recruited to mRNA, and these adaptor proteins in turn recruit mRNA export ...Assay Of Rna Export Coupled... · Results · Promoter Sequence Per Se...
  66. [66]
    impact of mRNA poly(A) tail length on eukaryotic translation stages
    Jun 14, 2024 · In the immediate after polyadenylation, the average length of mammalian poly(A) tails ranges from 150 to 250 nucleotides (nt) (2,3). However ...
  67. [67]
    U7 deciphered: the mechanism that forms the unusual 3′ end of ...
    Aug 5, 2021 · In animal cells, replication-dependent histone mRNAs end with a highly conserved stem–loop structure followed by a 4- to 5-nucleotide ...
  68. [68]
    The Genomic Landscape and Clinical Relevance of A-to-I RNA ...
    Oct 1, 2015 · ADAR-mediated A-to-I RNA editing represents a widespread, phylogenetically conserved, post-transcriptional mechanism to engender genomic ...
  69. [69]
    A-to-I RNA editing – thinking beyond the single nucleotide - PMC
    Adenosine-to-inosine RNA editing is a conserved process, which is performed by ADAR enzymes. By changing nucleotides in coding regions of genes and altering ...
  70. [70]
    Adenosine-to-inosine RNA editing and human disease
    Nov 29, 2013 · A-to-I RNA editing is a post-transcriptional modification that converts adenosines to inosines in both coding and noncoding RNA transcripts.
  71. [71]
    Functions and Regulation of RNA Editing by ADAR Deaminases
    RNA editing has several advantages over gene mutations. As with alternative splicing, the extent of RNA editing can be differentially regulated (ranging from no ...
  72. [72]
    Widespread A-to-I RNA Editing of Alu-Containing mRNAs in the ...
    We find that 1,445 human mRNAs (1.4%) are subject to RNA editing at more than 14,500 sites, and our data further suggest that the vast majority of pre-mRNAs ( ...
  73. [73]
    APOBEC1 mediated C-to-U RNA editing: target sequence and trans ...
    Mammalian C-to-U RNA editing was described more than 30 yr ago as a single nucleotide modification in small intestinal Apob RNA, later shown to be mediated ...
  74. [74]
    Genome-wide identification and functional analysis of Apobec-1 ...
    Jun 19, 2014 · The prototype for mammalian C-to-U RNA editing is apolipoprotein B (apoB) RNA, where Apobec-1-mediated deamination of a CAA codon introduces a ...
  75. [75]
    C-to-U RNA Editing: Mechanisms Leading to Genetic Diversity
    ... C-to-U RNA editing express apobec-1 mRNA, the transcript encoding the catalytic deaminase of the apoB RNA editing holoenzyme. In addition, a consensus apobec-1 ...
  76. [76]
    Genome-Wide Identification of Human RNA Editing Sites by Parallel ...
    May 29, 2009 · Thirteen edited genes are known in the nonrepetitive portion of the human genome, but the overall prevalence of RNA editing is unclear. Li et al ...
  77. [77]
    Cellular and genetic drivers of RNA editing variation in the human ...
    May 30, 2022 · RNA editing patterns are highly cell type-specific, with 189,229 cell type-associated sites. The cellular specificity for thousands of sites is ...
  78. [78]
    Complex regulation of ADAR-mediated RNA-editing across tissues
    Jan 15, 2016 · It has been suggested that the nuclear Adar proteins edit pre-mRNAs prior to splicing, while the cytoplasmic p150 isoform may edit viral RNAs ...
  79. [79]
    RNA Editing: A New Therapeutic Target in Amyotrophic Lateral ...
    RNA editing deficiency at the Q/R site in GluA2 due to downregulation of ADAR2 has been demonstrated in the motor neurons of patients with sporadic ALS and ...
  80. [80]
    Structural basis of sequestration of the anti-Shine-Dalgarno ...
    The SD base-pairs with the 30S subunit's anti-Shine-Dalgarno (ASD) sequence, a pyrimidine-rich element (CCUCC) near the 3′ end of 16S rRNA.
  81. [81]
    Control of translation initiation involves a factor-induced ...
    The SD pairs with the 3′ end of 16S ribosomal RNA (rRNA), termed anti-Shine–Dalgarno sequence (ASD), and this mRNA–rRNA interaction facilitates positioning of ...
  82. [82]
    Initiation of mRNA translation in bacteria: structural and dynamic ...
    Three proteins, the initiation factors (IFs) IF1, IF2, and IF3, determine the kinetics and fidelity of the overall initiation process. The three IFs are bound, ...
  83. [83]
    Protein Synthesis Initiation in Eukaryotic Cells - PMC
    Scanning describes the movement of the 43S PIC along the mRNA from the m7G cap, in a 3′ direction, searching for an AUG initiation codon in a suitable context ...An Overview Of The... · 43s Pic Formation · Eifs-1, -1a, And -3 Promote...
  84. [84]
    Mechanism of Translation Initiation in Eukaryotes - NCBI - NIH
    The initiation of protein synthesis consists in the recruitment of a ribosome·initiator tRNA complex to the initiation codon of a messenger RNA.
  85. [85]
    Rapid 40S scanning and its regulation by mRNA structure ... - NIH
    How the eukaryotic 43S preinitiation complex scans along the 5′ untranslated region (5′UTR) of a capped mRNA to locate the correct start codon remains elusive.
  86. [86]
    IRES-mediated cap-independent translation, a path leading to ...
    Sep 3, 2019 · Most eukaryotic mRNAs are translated in a cap-dependent fashion; however, under stress conditions, the cap-independent translation driven by ...
  87. [87]
    Codon--anticodon pairing: the wobble hypothesis - PubMed
    Codon--anticodon pairing: the wobble hypothesis. J Mol Biol. 1966 Aug;19(2):548-55. doi: 10.1016/s0022-2836(66)80022-0. Author. F H Crick. PMID: 5969078; DOI ...Missing: Francis original
  88. [88]
    Mechanism of activation of elongation factor Tu by ribosome
    In this work we have studied the mechanism of GTP hydrolysis catalyzed by the ribosome:EF-Tu:aa-tRNA complex with both wild-type and mutant EF-Tu proteins. This ...
  89. [89]
    The Elongation, Termination, and Recycling Phases of Translation ...
    Codon recognition by the tRNA triggers GTP hydrolysis by eEF1A, releasing the factor and enabling the aminoacyl-tRNA to be accommodated into the A site.
  90. [90]
    Unusual resistance of peptidyl transferase to protein extraction ...
    Jun 5, 1992 · Peptidyl transferase, the ribosomal activity responsible for catalysis of peptide bond formation, is resistant to vigorous procedures that are conventionally ...
  91. [91]
    Hydrolysis of GTP by elongation factor G drives tRNA movement on ...
    Jan 2, 1997 · EF-G-dependent GTP hydrolysis is shown to precede, and greatly accelerate, the rearrangement of the ribosome that leads to translocation.
  92. [92]
    Translational Control by Ribosome Pausing in Bacteria: How a Non ...
    ... rate of translation in bacteria is about 10–20 amino acids/s. Thus, rapid folding events that are much faster than amino acid addition to the nascent chain ...
  93. [93]
    Distinct stages of the translation elongation cycle revealed by ... - eLife
    May 9, 2014 · Comparing our relative occupancy values to an estimated bulk elongation rate of 5.6 amino acids per second (Ingolia et al., 2011), our model ...
  94. [94]
    Synonymous but not Silent: The Codon Usage Code for Gene ...
    Optimal codons speed up the translation elongation rate, while rare codons slow it down, due to differential expression levels of corresponding tRNAs. Codon ...
  95. [95]
    Atomic mutagenesis of stop codon nucleotides reveals the chemical ...
    Jan 3, 2018 · Class I release factors (RFs) are in charge of recognizing stop codons and consequently hydrolyzing the peptidyl-tRNA at the ribosomal P site.
  96. [96]
    Single amino acid substitution in prokaryote polypeptide release ...
    Translation termination generally requires two codon-specific polypeptide release factors (RFs), RF1 (for UAG/UAA) and RF2 (for UGA/UAA), in prokaryotes (1, 2) ...
  97. [97]
    Selective inhibition of human translation termination by a drug-like ...
    Oct 2, 2020 · Translation termination occurs when an mRNA stop codon enters the ribosome A site. In eukaryotes, termination is mediated by eRF1 and eRF3 ( ...Results · Erf1 Conformation In The... · Dna Constructs And In Vitro...
  98. [98]
    Kinetic analysis reveals the ordered coupling of translation ... - PNAS
    Dec 5, 2011 · The departure of the termination factors ensures that ribosome recycling does not commence until after completion of peptide release.
  99. [99]
    The accuracy of codon recognition by polypeptide release factors
    Termination of protein synthesis in prokaryotes depends on RF1 and RF2. RF1 normally terminates at UAA and UAG and RF2 at UAA and UGA (10).
  100. [100]
    Crystal structures of the human elongation factor eEFSec suggest a ...
    Oct 6, 2016 · Selenocysteine is the only proteinogenic amino acid encoded by a recoded in-frame UGA codon that does not operate as the canonical opal stop ...
  101. [101]
    A highly efficient form of the selenocysteine insertion sequence ...
    ... UGA is read as the Sec codon, and translation continues until the true stop signal. The ratio of full-length and truncated forms of fusion proteins that ...
  102. [102]
    Mechanisms of nuclear mRNA export: a structural perspective - PMC
    Sep 12, 2019 · One mechanism that targets export-competent mRNPs to the NPC nuclear basket is transcription-coupled mRNA export, mediated by the yeast TREX-2 ...Figure 2 · 3. Mrnp Targeting To The Npc · 4.1. The Mrna Export...
  103. [103]
    The role of TREX in gene expression and disease - Portland Press
    TRanscription and EXport (TREX) is a conserved multisubunit complex essential for embryogenesis, organogenesis and cellular differentiation throughout life.
  104. [104]
    A Conserved mRNA Export Machinery Coupled to pre-mRNA Splicing
    Considering that Ran does not appear to play a role in general mRNA export, the Ran-GTP/GDP gradient that is thought to impart directionality on the ...Main Text · A Conserved Mrna Export... · Other Conserved Mrna Export...<|control11|><|separator|>
  105. [105]
    Recruitment of the human TREX complex to mRNA during splicing
    Together, our data indicate that recruitment of the human TREX complex to spliced mRNA occurs by a splicing-coupled mechanism rather than by the direct ...
  106. [106]
    mRNA nuclear export at a glance | Journal of Cell Science
    Jun 15, 2009 · The poly(A) tail is added by a poly(A) polymerase and bound by a ... Mex67p, a novel factor for nuclear mRNA export, binds to both poly(A)+ RNA ...Key Steps In Mrna Processing · Recruitment Of Mrna Export... · Links To Disease And...
  107. [107]
    Dbp5p, a cytosolic RNA helicase, is required for poly(A)+ RNA export
    It has long been proposed that the energy requirement for mRNA export is derived from GTP hydrolysis by Ran (reviewed in Koepp and Silver, 1996). We find that ...
  108. [108]
    The Dbp5 cycle at the nuclear pore complex during mRNA export II
    The Dbps are RNA-dependent ATPases that act via helicase activity for unwinding RNA duplexes and/or remodeling activity for altering the protein composition of ...Missing: gradient | Show results with:gradient
  109. [109]
    Exon Junction Complexes: Supervising the Gene Expression ...
    The EJC regulates many post-transcriptional steps of gene expression, including mRNA export and translation and nonsense-mediated mRNA decay.
  110. [110]
    A Day in the Life of the Exon Junction Complex - PubMed Central
    Jun 5, 2020 · The exon junction complex (EJC) is an abundant messenger ribonucleoprotein (mRNP) component that is assembled during splicing and binds to mRNAs upstream of ...
  111. [111]
    Coupled Transcription-Translation in Prokaryotes - PubMed Central
    Coupled transcription-translation (CTT) is a hallmark of prokaryotic gene expression. CTT occurs when ribosomes associate with and initiate translation of mRNAs ...Missing: export | Show results with:export
  112. [112]
    Huntingtin mediates dendritic transport of β-actin mRNA in rat neurons
    Nov 3, 2011 · The zipcode sequence in the 3′UTR of β-actin mRNA is sufficient for dendritic targeting and co-localization with Htt in rat neurons. The ...β-Actin Mrna Transport Is... · Kinesin-1 And Dynein Are... · Zipcode-Zbp1 Pathway Of...
  113. [113]
    Microtubule-based transport is essential to distribute RNA and ...
    Oct 27, 2021 · RNAs exhibit restricted diffusion and directed transport in live myotubes. Our observations suggest that mRNAs in myofibers cannot escape ...
  114. [114]
    Global analysis of mRNA decay and abundance in Escherichia coli ...
    A wide range of stabilities was observed for individual mRNAs of E. coli, although ≈80% of all mRNAs had half-lives between 3 and 8 min. Genes having ...
  115. [115]
    Messenger RNA Degradation in Bacterial Cells - PubMed Central
    This review will focus on mRNA turnover in bacterial cells, including the ribonucleases and RNA elements that govern mRNA decay.
  116. [116]
    Direct entry by RNase E is a major pathway for the degradation and ...
    Sep 18, 2014 · We found that a sample enriched for mRNA was cleaved extensively by both NTH-RNase E and T170V; moreover, there was no obvious difference in the ...
  117. [117]
    Initiation of mRNA decay in bacteria | Cellular and Molecular Life ...
    Sep 25, 2013 · mRNA decay in bacteria initiates with an endonucleolytic cleavage, and the 5' end and translation initiation can precede or favor this process.
  118. [118]
    RNA polyadenylation and its consequences in prokaryotes - Journals
    Nov 5, 2018 · Numerous studies carried out over the last three decades have shown that polyadenylation greatly contributes to the control of prokaryotic gene expression.
  119. [119]
    Nucleotide specificity in bacterial mRNA recycling - PMC - NIH
    In 2008, Belasco and coworkers reported the existence of an RNA pyrophosphohydrolase enzyme (RppH) that triggers mRNA degradation in E. coli (8). This activity ...
  120. [120]
    Specificity of RppH-dependent RNA degradation in Bacillus subtilis
    Here we report that purified B. subtilis RppH requires at least two unpaired nucleotides at the 5′ end of its RNA substrates and prefers three or more.
  121. [121]
    Bacterial Small RNA-based Negative Regulation: Hfq and Its ...
    A large group of bacterial small regulatory RNAs (sRNAs) use the Hfq chaperone to mediate pairing with and regulation of mRNAs.
  122. [122]
    Regulation of mRNA Stability During Bacterial Stress Responses
    Polynucleotide phosphorylase promotes the stability and function of Hfq-binding sRNAs by degrading target mRNA-derived fragments. Nucleic Acids Res. 47 ...RNases and Other... · mRNA Stabilization as a... · The Importance of RNA Decay...
  123. [123]
    Regulation of eukaryotic mRNA deadenylation and degradation by ...
    Apr 19, 2023 · Recruitment of the Ccr4-Not complex to a target mRNA results in deadenylation mediated by the Caf1 and Ccr4 catalytic subunits of the complex.Structure of the Ccr4-Not... · Regulation of mRNA... · Recruitment of the Ccr4-Not...
  124. [124]
    A direct interaction between DCP1 and XRN1 couples mRNA ...
    Nov 11, 2012 · The removal of the mRNA 5' cap structure by the decapping enzyme DCP2 leads to rapid 5'→3' mRNA degradation by XRN1, suggesting that the two ...
  125. [125]
    RNA surveillance and the exosome - PMC - PubMed Central - NIH
    This complex acts together with Ski7 as a cofactor for the cytoplasmic exosome in mRNA turnover and surveillance. Yeast Ski7 shows homology to GTPases that ...
  126. [126]
  127. [127]
    Codon bias confers stability to human mRNAs | EMBO reports
    Recently, Presnyak and colleagues showed that mRNA half‐lives are correlated with optimal codon content based on a metric, the codon stabilization coefficient ( ...
  128. [128]
    mRNA structure regulates protein expression through changes in ...
    Nov 11, 2019 · We find that the structure of the CDS regulates protein expression through changes in functional mRNA half-life (ie, mRNA being actively translated).
  129. [129]
    Codon optimality is a major determinant of mRNA stability - PMC - NIH
    Specifically, mRNAs with less than 40% optimal codons were typically found to be unstable, with a median half-life of 5.4 minutes. In contrast, mRNAs with 70% ...
  130. [130]
    Nonsense-mediated mRNA decay at the crossroads of many cellular ...
    Nonsense-mediated mRNA decay (NMD) is a surveillance mechanism ensuring the fast decay of mRNAs harboring a premature termination codon (PTC).
  131. [131]
    Mechanism of Nonsense-Mediated mRNA Decay Stimulation ... - NIH
    NMD factors UPF2 and UPF3 bridge UPF1 to the exon junction complex and stimulate its RNA helicase activity. Nat Struct Mol Biol. 2008;15:85–93. doi: 10.1038 ...
  132. [132]
    Evidence that Tristetraprolin Binds to AU-Rich Elements and ... - NIH
    TTP can bind directly to the AU-rich element (ARE) in TNF-α mRNA (E. Carballo, WS Lai, and PJ Blackshear, Science 281:1001–1005, 1998).Transfection Of Hek 293... · Results · Effect Of Ttp On Tnf-α Mrna...
  133. [133]
    Tristetraprolin (TTP): Interactions with mRNA and proteins, and ...
    Adenosine and uridine (AU)-rich elements (ARE), often located in the 3′ untranslated regions (3′UTR) of mRNAs, are known to target transcripts for rapid decay.
  134. [134]
    mRNA degradation by miRNAs and GW182 requires both CCR4 ...
    Our findings indicate that GW182 links the miRNA pathway to mRNA degradation by interacting with AGO1 and promoting decay of at least a subset of miRNA targets.
  135. [135]
    Ago-TNRC6 complex triggers microRNA-mediated mRNA decay by ...
    We show that let-7 miRNA-induced silencing complexes (miRISCs), which contain Argonaute (Ago) and TNRC6 (also known as GW182) proteins, trigger highly rapid ...
  136. [136]
    Antiviral silencing in animals - PMC - NIH
    Antiviral silencing in animals is a small RNA-guided gene silencing mechanism, part of RNAi, that plays a role in responses to virus infection.
  137. [137]
    Antiviral RNA interference in plants: Increasing complexity and ... - NIH
    Aug 26, 2025 · RNA interference (RNAi, also known as RNA silencing) is one of the most important plant defense responses against viral invasion.
  138. [138]
  139. [139]
    Post-transcriptional regulation of inflammation by RNA-binding ...
    These findings suggested that HuR competes with mRNA destabilizing ARE-BPs, such as TTP, and protect target mRNAs from decay. HuR targets mRNAs of several ...
  140. [140]
    RNA-binding protein TTP is a global post-transcriptional regulator of ...
    Our study uncovers a role of TTP as a suppressor of feedback inhibitors of inflammation and highlights the importance of fine-tuned TTP activity-regulation by ...
  141. [141]
  142. [142]
    Modular RNA interactions shape FXR1 condensates involved in ...
    Sep 29, 2025 · We therefore concluded that FXR1 proteins can undergo phase separation in vitro, and form spatially specific mRNP condensates in hESCs. FXR1 ...
  143. [143]
    Auto-regulatory feedback by RNA-binding proteins - PubMed Central
    RBPs engage in an auto-regulatory feedback by directly binding to and influencing the fate of their own mRNAs, exerting control over their own expression.
  144. [144]
    Building Robust Transcriptomes with Master Splicing Factors
    Oct 23, 2014 · RBPs can also feature in positive feedback loops alongside transcription factors, for example, during the regulation of developmental decisions ...
  145. [145]
    Splicing to keep splicing: A feedback system for cellular ...
    Jun 19, 2025 · RBPMS (RNA Binding Protein, MRNA processing factor) is a gene coding the master splicing regulator in mammalian smooth muscle cells (SMCs). It ...
  146. [146]
    Circular RNAs: biogenesis, expression and their potential roles in ...
    Jan 17, 2018 · First, circRNAs can function as a miRNA sponge to regulate the function of miRNAs [19, 27]. Some circRNAs, such as circRNA for miRNA-7 (ciRS-7) ...Missing: paper | Show results with:paper
  147. [147]
    Quick or quality? How mRNA escapes nuclear quality control during ...
    We uncovered a simple mechanism, in which the export block of regular mRNAs and a fast export of heat shock mRNAs is achieved by deactivation of the nuclear ...Pre-Mrna Maturation · Nuclear Mrna Quality Control... · Figure 3Missing: sensor paper
  148. [148]
    The role of RNA in biological phase separations - PubMed Central
    RNA plays a key role in phase separation events that modulate various aspects of RNA metabolism. Here, we review the role that RNA plays in ribonucleoprotein ...
  149. [149]
    mRNA structure determines specificity of a polyQ-driven phase ...
    RNA promotes liquid-liquid phase separation (LLPS) to build membraneless compartments in cells. How distinct molecular compositions are established and ...
  150. [150]
    Differentiating Protein-Coding and Noncoding RNA - NIH
    Nov 28, 2008 · One of the most fundamental criteria used to distinguish long ncRNAs from mRNAs is ORF length. ... Discrimination of non-protein–coding ...
  151. [151]
    The Nobel Prize in Physiology or Medicine 1993 - Press release
    Roberts and Phillip A. Sharp in 1977 independently discovered that genes could be discontinuous, that is, a given gene could be present in the genetic ...