Fact-checked by Grok 2 weeks ago

Intron

An intron is a non-coding sequence of DNA located within a gene in eukaryotic organisms, interrupting the coding regions known as exons, and is transcribed into pre-messenger RNA (pre-mRNA) but subsequently removed during RNA splicing to form mature mRNA for translation into proteins. Introns were discovered in 1977 through independent studies by Phillip Sharp and Richard Roberts, who showed that eukaryotic genes, including those in viruses like adenovirus, are discontinuous with non-coding intervening sequences that are spliced out of the primary RNA transcript. This groundbreaking finding, which challenged the prevailing view of genes as continuous coding units, earned Sharp and Roberts the Nobel Prize in Physiology or Medicine in 1993. Spliceosomal introns, the predominant type in eukaryotes, are absent in prokaryotes but present in the vast majority of eukaryotic genes, often numbering in the dozens or hundreds per gene and varying widely in length from tens to tens of thousands of base pairs. Their positions show remarkable conservation across distant species, suggesting evolutionary and functional importance. Beyond being removed during splicing, introns fulfill diverse roles in gene regulation, such as enhancing transcriptional efficiency (sometimes by over 100-fold), enabling alternative splicing that generates multiple protein isoforms from a single gene (affecting up to 95% of human multi-exon genes), promoting mRNA nuclear export and translation, and encoding functional non-coding RNAs like microRNAs and small nucleolar RNAs.

Definition and Discovery

Definition and Characteristics

Introns are non-coding sequences of DNA located within genes, primarily in eukaryotes but also present in some prokaryotes and archaea, that are transcribed into precursor messenger RNA (pre-mRNA) yet excised during RNA splicing to yield mature mRNA for translation into proteins. These sequences interrupt the coding regions of genes and do not contribute to the final protein product in the standard case. Structurally, introns are positioned between exons, the segments that are retained in mature mRNA, and exhibit a wide range of lengths, typically spanning from 50 to more than 6,000 nucleotides, though extremes can exceed 100,000 nucleotides in some cases. Their boundaries are defined by highly conserved sequence motifs essential for recognition by the splicing machinery: the 5' splice site begins with a GU dinucleotide, the 3' splice site ends with an AG dinucleotide, and an internal branch point sequence features a critical adenine residue approximately 20–50 nucleotides upstream of the 3' site. Additionally, many introns contain a polypyrimidine tract, a stretch of pyrimidine-rich nucleotides, near the 3' splice site that aids in spliceosome assembly. In contrast to exons, which encode sequences or regulatory in the mature transcript, introns are generally non-coding and removed to . Exceptions exist where introns harbor genes for functional RNAs, such as small nucleolar RNAs (snoRNAs) that modifications. In eukaryotic genomes, introns ; for instance, in humans, they comprise over 95% of the of many protein-coding genes, with approximately 210,000 introns distributed across the approximately protein-coding genes.

Historical Discovery and Etymology

The discovery of introns marked a paradigm shift in understanding eukaryotic gene structure, emerging from investigations into adenovirus transcription in the mid-1970s. In 1977, Phillip A. Sharp and colleagues at the Massachusetts Institute of Technology hybridized late mRNA from adenovirus type 2 with its genomic DNA and used electron microscopy to observe the resulting structures. These revealed regions where the DNA looped out, unpaired with the mRNA, indicating that the gene contained non-coding intervening sequences separating expressed segments. This work, published in the Proceedings of the National Academy of Sciences, provided the first direct evidence of discontinuous genes in eukaryotes. Independently, in the same year, Richard J. Roberts and his team at Cold Spring Harbor Laboratory analyzed adenovirus type 2 transcripts using similar RNA-DNA hybridization and electron microscopy techniques. Their mapping showed an "amazing sequence arrangement" at the 5' ends of the mRNAs, confirming the presence of spliced intervening sequences that were removed during RNA processing to form mature mRNA. These findings, detailed in Cell, demonstrated that eukaryotic genes are composed of coding exons interrupted by non-coding introns, challenging the long-held assumption of gene-protein colinearity observed in prokaryotes. The 1977 discoveries by Sharp and Roberts were initially met with skepticism, as they contradicted the prevailing view that genes were continuous stretches of DNA directly encoding proteins, a model well-established from bacterial studies. Many scientists questioned whether split genes were artifacts of viral genomes or unique to eukaryotes, given the absence of introns in prokaryotic systems at the time. Confirmation came swiftly through additional electron microscopy studies, which consistently visualized the looped-out intron regions in RNA-DNA hybrids from various eukaryotic genes, solidifying the reality of this "gene-in-pieces" architecture. Earlier hints of RNA processing complexity had appeared in the 1970s from studies on heterogeneous nuclear RNA (hnRNA) in eukaryotic cells, which showed that large precursor transcripts were trimmed to smaller mRNAs, though these did not yet reveal the splicing mechanism. The groundbreaking 1977 papers in Cell and PNAS earned Sharp and Roberts the 1993 Nobel Prize in Physiology or Medicine for their "discoveries concerning split genes." The term "intron," short for intragenic region or intervening sequence, was coined by biochemist Walter Gilbert in 1978 to describe these non-coding elements, while "exon" denoted the expressed sequences joined during splicing. Gilbert introduced these terms in his Nature article "Why genes in pieces?," proposing that introns facilitated evolutionary flexibility by allowing exon shuffling. Before Gilbert's nomenclature, the sequences were commonly called intervening sequences (IVS) in the original 1977 publications. By the early 1980s, further milestones included the elucidation of splicing mechanisms, with the discovery of self-splicing introns in ribosomal RNA precursors, such as the group I intron in Tetrahymena thermophila reported by Thomas Cech in 1982, highlighting the catalytic potential of RNA itself.

Distribution and Occurrence

In Eukaryotic Genomes

Introns are highly abundant in eukaryotic genomes, particularly in more complex organisms. In the human genome, protein-coding genes contain an average of approximately 8 introns per gene, resulting in a total of around 180,000 to 200,000 introns across roughly 20,000 genes. Introns constitute about 24-25% of the total genomic DNA in mammals, significantly contributing to genome size despite not being translated into proteins. Intron sizes exhibit considerable variation across eukaryotic lineages, reflecting differences in genome architecture. In the budding yeast Saccharomyces cerevisiae, introns are rare and small, with only about 5% of genes containing a single intron on average, and typical lengths ranging from 50 to 400 nucleotides. In contrast, vertebrate genomes feature much larger introns; the average human intron is approximately 3.4 kilobases (kb), though sizes can extend up to 100 kb or more in some cases. This expansion contributes to the overall bloat in mammalian genomes, where introns often dwarf exon lengths. Patterns of intron distribution correlate with organismal complexity and gene features. Simpler eukaryotes like yeast have few introns, primarily in ribosomal protein genes, while multicellular organisms show increased numbers; for instance, the fruit fly Drosophila melanogaster genome harbors over 48,000 introns, averaging about 4 per gene with lengths around 487 base pairs. In plants, such as Arabidopsis thaliana, genes average nearly 5 introns with short lengths of about 165 nucleotides, but many plant species exhibit exceptionally long introns exceeding several kilobases, correlating with larger overall genome sizes. Intron number and length also tend to increase with gene length and are more prevalent in housekeeping versus tissue-specific genes in vertebrates.

In Prokaryotic and Archaeal Genomes

Introns are exceedingly rare in bacterial genomes, with the vast majority of prokaryotic genes lacking them entirely. Among the known cases, self-splicing group I and group II introns predominate, often functioning as mobile genetic elements that insert into host genes such as those encoding ribosomal RNAs or surface proteins. For instance, in Clostridium tetani, a group II intron interrupts a surface layer protein gene and undergoes alternative splicing in vivo, representing one of the few documented examples of introns in bacterial protein-coding sequences. Comprehensive analyses indicate that while group II introns occur in approximately 25% of surveyed bacterial genomes, they average only about 5.3 per affected genome, underscoring their scarcity relative to eukaryotic counterparts. In archaeal genomes, introns are more prevalent than in bacteria, though still limited in scope and primarily confined to transfer RNA (tRNA) and ribosomal RNA (rRNA) genes. These introns are typically processed by archaeal tRNA splicing endonucleases that recognize bulge-helix-bulge motifs at the exon-intron boundaries. Notable examples include multiple introns in the 23S rRNA genes of hyperthermophilic species like Pyrobaculum aerophilum and Pyrobaculum islandicum, where a 713-nucleotide intron interrupts the 16S rRNA gene in the former. Additionally, group I introns are widespread in archaeal rRNA and tRNA loci, and according to a 2024 preprint, group II introns have been identified in certain lineages, including members of the Asgard superphylum such as Lokiarchaeota, which exhibit structural and mechanistic parallels to eukaryotic spliceosomal introns. Archaeal and bacterial introns are generally short, ranging from 15 to around 600 nucleotides, in stark contrast to the often kilobase-scale introns in eukaryotes, and genomes harbor only a handful—typically 1 to 10 in total—rather than the thousands found in eukaryotic nuclear genes. This paucity suggests that introns were likely acquired relatively late in the evolution of prokaryote-like ancestors, possibly through horizontal transfer or independent insertions, rather than being ancestral features retained from a common origin.

Classification and Types

Spliceosomal Introns

Spliceosomal introns are non-coding sequences within eukaryotic pre-mRNA transcripts that are excised by the spliceosome, a large ribonucleoprotein complex composed of small nuclear RNAs (snRNAs) and associated proteins. This process is essential for generating mature messenger RNA (mRNA) in the nucleus, where spliceosomal introns predominate as the primary type interrupting protein-coding genes. Unlike other intron classes, spliceosomal introns require the coordinated action of multiple spliceosomal components for accurate removal, distinguishing them as a hallmark of eukaryotic gene architecture. Key structural features of spliceosomal introns include conserved consensus sequences that guide spliceosome recognition and assembly. The 5' splice site typically begins with a GT dinucleotide, while the 3' splice site ends with an AG dinucleotide, and an internal branch point sequence, often featuring an adenine residue, facilitates the splicing reaction. These motifs are recognized by specific small nuclear ribonucleoproteins (snRNPs): U1 snRNP binds the 5' splice site, U2 snRNP interacts with the branch point, and the U4/U5/U6 tri-snRNP complex contributes to catalysis and exon ligation. Introns vary widely in length, from tens to thousands of base pairs, but these conserved elements ensure splicing fidelity across diverse eukaryotic lineages. Spliceosomal introns constitute approximately 99% of all introns in eukaryotic genomes and are entirely absent in prokaryotes, reflecting their role in the complex regulation of eukaryotic gene expression. Their prevalence underscores the evolutionary expansion of nuclear pre-mRNA processing machinery, with densities varying from fewer than one per gene in some unicellular eukaryotes to approximately 8-9 per gene in vertebrates such as humans. This abundance enables alternative splicing, which generates proteomic diversity without expanding gene number. A of spliceosomal introns, known as U12-type introns, deviate from the GT-AG and instead AT-AC termini; these are processed by a distinct spliceosome comprising U11, U12, U4atac, U5, and U6atac snRNPs. In humans, U12-type introns represent about 0.5% of introns, occurring in roughly 700-800 genes, often in clusters within the same transcript. These highlight the spliceosome's adaptability while maintaining mechanistic principles shared with the pathway.

Self-Splicing Introns

Self-splicing introns represent a class of capable of catalyzing their own excision from precursor RNA transcripts through intrinsic activity, of protein enzymes for the core splicing steps. These introns are primarily classified into Group I and Group II based on their distinct secondary structures and catalytic mechanisms, with both types facilitating two sequential transesterification reactions to remove the intron and ligate the flanking exons. Unlike spliceosomal introns, self-splicing introns rely on RNA folding to form active sites, highlighting their role as ancient in organellar and prokaryotic genomes. Group I introns were first identified in the ribosomal RNA (rRNA) precursor of Tetrahymena thermophila, where their self-splicing activity was demonstrated in vitro without added proteins. These introns occur in diverse genes, including rRNA, transfer RNA (tRNA), and mitochondrial protein-coding genes across eukaryotes, bacteria, and organelles. Their splicing mechanism begins with the exogenous guanosine cofactor, whose 3'-hydroxyl group attacks the 5' splice site in the first transesterification step, cleaving the 5' exon and attaching the guanosine to the intron's 5' end; the freed 3'-hydroxyl of the 5' exon then attacks the 3' splice site in the second step, joining the exons and releasing the linear intron. Structurally, Group I introns feature a conserved core of nine helical elements (P1 through P9), where P1 pairs the 5' exon with the internal guide sequence (IGS) to position the splice sites, and the P4-P6 domain forms a key catalytic scaffold, as revealed by crystallographic studies. This ribozyme architecture enables precise recognition and catalysis, with the UGU triplet in P7 coordinating a guanosine-binding pocket. Group II introns, initially characterized in yeast mitochondrial genes such as the cox1 locus, are prevalent in organellar genomes of plants, fungi, and algae, as well as in bacterial chromosomes and plasmids.90264-3.pdf) Their splicing mirrors the lariat-forming pathway of spliceosomal introns, starting with the 2'-hydroxyl of an adenosine bulge in domain VI attacking the 5' splice site to form a lariat intermediate and release the 5' exon; the subsequent attack by the 5' exon's 3'-hydroxyl on the 3' splice site ligates the exons and excises the branched intron. The conserved secondary structure comprises six double-helical domains (I through VI), with domain I serving as the catalytic core that scaffolds the active site through tertiary interactions, including coordination of two Mg²⁺ ions for phosphodiester bond hydrolysis. This ribozyme function is often enhanced by an intron-encoded maturase protein, though core self-splicing occurs in vitro without it. Group II introns exhibit mobility through retrohoming, where a reverse transcriptase-maturase fusion protein facilitates target-primed reverse transcription into homologous DNA sites. In terms of distribution, self-splicing introns are rare in eukaryotic nuclear genomes, where Group I examples are limited to specific fungal and protist rRNA or protein genes, but they are abundant in organelles. For instance, the mitochondrial genome of Saccharomyces cerevisiae contains at least 10 Group I and several Group II introns interrupting genes like cox1, cob, and rRNAs, contributing to genome complexity and enabling independent splicing in isolated transcripts. This organellar prevalence underscores their adaptation to compact, maternally inherited genomes, contrasting with the protein-dependent splicing dominant in nuclear pre-mRNAs.

tRNA and Other Specialized Introns

Introns in transfer RNA (tRNA) genes represent a specialized class distinct from those in messenger RNA (mRNA), as they occur in non-coding RNAs essential for translation and are processed through unique enzymatic mechanisms. In eukaryotes, tRNA introns are invariably positioned at a conserved site within the anticodon loop, specifically one nucleotide downstream of the anticodon between positions 37 and 38 of the mature tRNA sequence. These introns vary in length from 6 to over 100 nucleotides but do not interact extensively with the splicing machinery, allowing accommodation of diverse sequences by the processing enzymes. Unlike spliceosomal introns, tRNA introns are excised by the heterotetrameric tRNA splicing endonuclease (TSEN) complex, composed of subunits TSEN2, TSEN15, TSEN34, and TSEN54 in humans, which employs a molecular ruler mechanism to recognize the pre-tRNA structure and cleave at the exon-intron boundaries without requiring guanosine cofactors or lariat formation. Following cleavage, the exons are ligated by tRNA ligase (RTL or CGI-99 in mammals), ensuring precise restoration of the tRNA's functional cloverleaf structure. The of tRNA introns in eukaryotic genomes is relatively low and across , reflecting evolutionary rather than a for tRNA maturation. For instance, in the yeast Saccharomyces , approximately % of tRNA genes— out of 274—contain introns, distributed across 10 isodecoder families, with all introns at the . In about 7% of the roughly tRNA genes harbor introns, totaling around 28 intron-containing genes, primarily in tRNA-Arg and tRNA-Tyr . Across broader eukaryotic , the proportion ranges from 5% to 25%, with higher incidences in lower eukaryotes like yeast compared to vertebrates, and the introns often serving no essential role in tRNA function, as evidenced by viable intronless mutants in yeast. In archaea, introns in tRNA and (rRNA) genes exhibit distinct structural features adapted to the domain's splicing machinery, emphasizing RNA motifs over extensive secondary structures. These introns are typically recognized by the archaeal splicing endonuclease (aSen) through a conserved bulge-helix-bulge (BHB) motif at the exon-intron boundaries, consisting of two 2- to 3-nucleotide bulges flanking a 4-base-pair helix. The BHB structure facilitates precise cleavage, after which ligation occurs via ATP-dependent RNA ligase, distinguishing this from eukaryotic TSEN-mediated splicing despite superficial similarities in endonuclease recognition. While most archaeal tRNA and rRNA introns rely on this protein-assisted mechanism, some rare group I introns in archaeal rRNA can undergo self-splicing, paralleling organellar variants but remaining infrequent. Examples include BHB-motif introns in the pre-tRNA^{Ile} of Haloferax volcanii and rRNA precursors of Desulfurococcus mobilis, where the motif ensures fidelity in harsh environmental conditions typical of archaeal habitats. Other specialized introns include twintrons, which are nested arrangements where an internal intron is embedded within an external one, requiring sequential splicing for resolution. Twintrons occur rarely in tRNA and rRNA contexts, primarily in organellar genomes, such as group II twintrons in algal chloroplast rRNA genes or mitochondrial tRNA precursors in lycophytes, where the internal intron must be excised first to expose the external splice sites. Another rare variant involves permuted exons associated with group I introns in ciliate rRNA, as seen in Tetrahymena thermophila, where the linear order of exon segments is rearranged, yet self-splicing proceeds via trans-esterification to yield functional circular or linear RNAs. These configurations, including permuted intron-exon (PIE) structures, highlight evolutionary innovations in intron architecture, with the permuted group I introns in ciliates demonstrating autocatalytic activity despite disrupted sequential order. Such specialized forms underscore the adaptability of introns in non-mRNA RNAs, maintaining low overall prevalence to minimize processing burdens.

Splicing Mechanism

Splicing Process Overview

In eukaryotic cells, the splicing process begins with the transcription of pre-mRNA by RNA polymerase II, producing a primary transcript that contains both exons and introns. This pre-mRNA is then subject to splicing, a co-transcriptional process where introns are precisely removed and exons are joined to form mature mRNA. The spliceosome, a large ribonucleoprotein complex, assembles dynamically on the pre-mRNA to catalyze this removal through two sequential transesterification reactions. Recent cryo-electron microscopy (cryo-EM) studies, as of 2024, have provided the first atomic-level blueprint of the human spliceosome, revealing intricate details of its assembly and conformational changes. Spliceosome occurs in a stepwise manner, starting with the E () complex, where U1 binds the 5' , and additional factors recognize the and polypyrimidine tract near the 3' . This progresses to the A (pre-spliceosome) complex with U2 binding the , forming base-pairing interactions that position . The B forms upon of the U4/U6.U5 tri-snRNP, bringing all five major snRNPs (U1, U2, U4, U5, U6) together, followed by structural rearrangements driven by ATP-dependent DExD/H-box helicases like Prp28 and Brr2, which release U1 and activate the catalytic core. The process culminates in the B* and C , where the spliceosome becomes catalytically active, with further ATPase activity (e.g., Prp2) facilitating the first reaction. This is ATP-dependent throughout, relying on helicases for conformational changes, and occurs co-transcriptionally in eukaryotes, coupling splicing to nascent RNA production. The splicing reactions involve two transesterification steps without net consumption of chemical energy. In the first step, the 2'-OH group of an adenosine at the branch point acts as a nucleophile, attacking the phosphate at the 5' splice site, cleaving the 5'-exon and forming a lariat structure with the intron via a 2'-5' phosphodiester bond. The second step follows, where the newly freed 3'-OH of the 5' exon attacks the phosphate at the 3' splice site, ligating the exons with a standard 3'-5' phosphodiester bond and releasing the intron lariat. These reactions are facilitated by the snRNAs in U2, U5, and U6, which mimic ribozyme-like catalysis in the active site. While most introns undergo constitutive splicing, where all introns are removed in a fixed manner, variations arise through , allowing a single pre-mRNA to produce multiple isoforms; for example, excludes specific exons from the mRNA, regulated by splicing factors that modulate snRNP or . This is highly conserved across eukaryotes but can be tuned for regulated splicing in response to cellular signals.

Fidelity and Error Correction

The spliceosome achieves remarkably high fidelity in intron removal, with in vivo splicing error rates typically below 1% per intron, often approaching 0.7% on average across human genes, ensuring that over 99% of splicing events produce accurate exon ligation. This precision is essential given the sequence similarity between splice sites and potential cryptic sites in pre-mRNA, where errors such as exon skipping or intron retention can disrupt coding frames and lead to non-functional transcripts. In specific cases, like the SMN2 gene, splicing errors result in exon 7 skipping in approximately 90% of transcripts, contributing to spinal muscular atrophy pathogenesis by producing unstable SMN protein variants. Cellular proofreading mechanisms enhance this accuracy during spliceosome assembly and catalysis. DEAH-box ATPases, such as Prp16, act as molecular clocks by unwinding suboptimal lariat intermediates formed after the first transesterification step, directing aberrant substrates into a discard pathway that prevents their progression to the second step of splicing. This kinetic proofreading process discriminates against slowly reacting complexes, rejecting those with mismatched branch points or splice sites and thereby reducing error propagation. Recent structural studies (as of 2025) highlight roles for factors like DHX35-GPATCH1 in ensuring splice site fidelity during assembly. Additionally, post-splicing surveillance via nonsense-mediated decay (NMD) degrades aberrant mRNAs containing premature termination codons often introduced by splicing errors like frameshift-inducing exon skips or retained introns. Several factors influence splicing fidelity, including pre-mRNA secondary structure, which can mask or expose splice sites and promote alternative or erroneous pairings, and the concentration of splicing factors like SR proteins that stabilize canonical splice site recognition. Imbalances in these factors, such as reduced SR protein levels, can increase error rates by favoring cryptic sites or inefficient assembly. In disease contexts, mutations altering these elements exacerbate inaccuracies, as seen in SMN2 where a single nucleotide change disrupts an exonic splicing enhancer, leading to predominant exon skipping. Experimental studies highlight differences in between and conditions, with reconstituted spliceosomes showing reduced , such as 10-fold slower under suboptimal conditions, due to the absence of cellular chaperones and pathways. Kinetic models, informed by inhibition assays, demonstrate how energy-dependent branches in the splicing amplify . These models the spliceosome's to speed and accuracy, with Prp16-mediated rejection preventing the accumulation of defective lariats in cellular extracts.

Biological Roles and Evolution

Regulatory and Functional Roles

Introns exert significant influence on gene regulation, primarily through intron-mediated enhancement (IME) and alternative splicing. IME enables specific introns, particularly those located near the 5' end of genes and in their native orientation, to amplify mRNA levels by boosting transcription efficiency, promoting nuclear export, and stabilizing transcripts. This enhancement can increase gene expression by several-fold, with effects observed across diverse organisms from yeast to mammals, underscoring introns' role in fine-tuning protein output without altering coding sequences. Complementing IME, alternative splicing leverages introns to produce multiple mRNA isoforms from a single pre-mRNA, expanding proteomic diversity; in humans, this process affects approximately 95% of multi-exon genes, enabling tissue-specific and developmental regulation of gene function. Beyond direct enhancement, introns serve as reservoirs for non-coding RNAs that regulate cellular processes. A substantial portion—around 50-60%—of human microRNAs (miRNAs) originates from intronic sequences within protein-coding or non-coding host genes, where these miRNAs are excised and mature independently to silence target mRNAs post-transcriptionally. Likewise, the majority of small nucleolar RNAs (snoRNAs), exceeding 95% in vertebrates, are processed from introns, guiding chemical modifications on ribosomal and other RNAs essential for ribosome biogenesis and translation fidelity. Introns also contribute to the formation of circular RNAs (circRNAs) via backsplicing, where flanking intronic repeats or structures facilitate exon circularization, yielding stable RNAs that act as miRNA sponges or modulators of splicing. Introns further mRNA maturation and by facilitating of the (EJC) approximately 20-24 upstream of exon-exon junctions during splicing, which recruits factors to processed transcripts reach the cytoplasm efficiently. They also modulate chromatin , with intronic sequences influencing positioning and to promote or repress transcription in a context-dependent manner. In specific examples, intron retention during stress responses—such as or deprivation—delays of sensor genes by sequestering premature transcripts in the , providing a rapid post-transcriptional brake on protein synthesis. For 40 years, intron retention was often dismissed as splicing noise but is now recognized as a dynamic and evolutionarily conserved mechanism of gene regulation. Similarly, in immune gene diversity, introns within the immunoglobulin heavy chain locus enable V(D)J recombination and class-switch recombination, generating varied antibody specificities and isotypes critical for adaptive immunity.

Evolutionary Origins and Significance

The evolutionary origins of introns remain a subject of debate, encapsulated by the "introns-early" and "introns-late" hypotheses. The introns-early theory proposes that introns were abundant in the last universal common ancestor (LUCA) or even predated the RNA-protein world, facilitating early exon shuffling, with extensive losses occurring in prokaryotic lineages due to streamlining pressures. This view is bolstered by the conservation of intron positions in orthologous genes across eukaryotes, where roughly 25-30% of introns align in sequences from animals, fungi, and plants, often at protosplice sites such as (A/C)AG||G that suggest ancient insertions rather than independent gains. Conversely, the introns-late theory argues that spliceosomal introns emerged as a eukaryotic innovation after the divergence from prokaryotes, driven by the need for complex gene regulation, with evidence from their near-absence in bacterial and archaeal genomes—where introns constitute less than 1% of genes—and the sporadic distribution in eukaryotic paralogs indicating ongoing gains and losses. Recent comparative genomic analyses have identified hundreds of recent intron gain events in human genes, supporting continued intron dynamics in modern eukaryotes. A prevailing compromise reconciles these perspectives by positing that self-splicing group II introns, originally from bacterial endosymbionts like the mitochondrial progenitor, massively invaded the early eukaryotic nuclear genome during eukaryogenesis, creating an intron-rich ancestor before differential losses in descendant lineages. This scenario aligns with the mechanistic parallels between group II intron ribozyme activity and spliceosomal transesterification reactions, where conserved structural domains in group II RNAs mirror spliceosomal snRNAs. Such an invasion likely coincided with the emergence of nuclear-cytoplasmic compartmentalization, enabling intron proliferation without immediate lethality to the host. Introns have profoundly influenced genome evolution by enabling exon shuffling, which promotes the recombination of protein-coding modules to generate functional diversity. In vertebrates, for example, introns in immunoglobulin loci allow V(D)J recombination, shuffling variable exons to produce antibody diversity essential for adaptive immunity. Broader analyses indicate that exon shuffling has assembled modular domains in a significant fraction of eukaryotic multidomain proteins, underscoring introns' role in expanding proteome complexity without de novo sequence invention. Intron proliferation, mediated by gene duplication and retrotransposition-like mobility, correlates strongly with the transition to multicellularity, particularly in metazoans, where intron density surged at the lineage's base to support tissue-specific regulation. Phylogenetic reconstructions reveal that early metazoan ancestors acquired thousands of novel introns, far exceeding those in unicellular relatives, facilitating alternative splicing variants that underpin developmental complexity. This expansion contrasts with the rarity of introns in prokaryotes and archaea, where they appear sporadically, often as mobile group II elements. The fossil record of introns is embodied in ancient preserved in bacterial genomes, such as those in and Sinorhizobium species, indicating their pre-eukaryotic antiquity as retroelements capable of self-propagation. Debates persist on evolution, with structural and phylogenetic supporting its derivation from disassembled components: the intron's catalytic likely fragmented into snRNAs (U2, U6), while maturase proteins evolved into splicing factors like Prp8. This transition highlights introns' significance in driving eukaryotic innovation, from modular protein evolution to the architectural foundations of complex .

Specific Adaptations (e.g., Starvation Response)

One prominent example of intron-mediated adaptation to nutrient stress involves phosphate starvation in Arabidopsis thaliana. Under phosphate (Pi) deficiency, intron retention increases in numerous root transcripts, particularly those associated with phosphate transport and cellular responses, leading to the production of truncated protein isoforms that enhance resource efficiency and stress tolerance. This splicing shift allows plants to fine-tune gene expression without altering transcription levels, promoting survival in low-Pi soils. In heat stress responses, introns facilitate decay mechanisms in plants such as Arabidopsis and tomato, where elevated temperatures induce widespread intron retention, often introducing premature termination codons that trigger nonsense-mediated decay (NMD) of transcripts. This reduces the synthesis of non-essential proteins, conserving energy during thermal stress and contributing to thermotolerance; for instance, retention in heat shock factor genes like HsfA2 modulates isoform production for adaptive protein functions. Similarly, in viral contexts, stable introns from latency-associated transcripts in herpes simplex virus type 1 accumulate post-splicing in infected neurons, suppressing lytic gene expression and maintaining viral latency by interfering with host or viral transcription. Osmotic stress in yeast (Saccharomyces cerevisiae) triggers concerted intron retention in ribosomal protein genes, such as RPS22B, generating bimodal expression patterns that create phenotypic heterogeneity within cell populations. This bet-hedging strategy enables some cells to endure prolonged stress via low protein output while others recover quickly upon relief, enhancing overall population fitness. These adaptations arise through stress-induced shifts in splicing factors, including hnRNP-like proteins in yeast that relocalize to the nucleus under osmotic or thermal stress, altering splice site recognition and favoring retention. In plants, similar changes in SR and hnRNP proteins modulate intron inclusion, often via phosphorylation or binding affinity alterations. Responsive elements, such as weak 5' splice sites or upstream open reading frames in introns, show evolutionary conservation across plant species and even kingdoms, underscoring their adaptive utility. Studies from the 2010s, including genome-wide analyses in Arabidopsis, revealed that approximately 10-20% of intron-containing genes exhibit modulated splicing under various stresses, with intron retention being the dominant event in nutrient and abiotic responses. These findings highlight introns' role in rapid, post-transcriptional adjustments, distinct from broader alternative splicing mechanisms.

Mobility and Genetic Dynamics

Mechanisms of Intron Mobility

Introns exhibit mobility through distinct biochemical mechanisms that enable their spread within and between genomes, primarily observed in self-splicing group I and group II introns, as well as rarer events in spliceosomal introns. Retrohoming is the primary mobility mechanism for group II introns, involving an RNA intermediate that invades a homologous target DNA site. These introns encode a multifunctional intron-encoded protein (IEP) with reverse transcriptase (RT) and endonuclease domains, which assembles with the excised intron RNA to form a ribonucleoprotein (RNP) particle. The RNP targets an intronless allele, where the intron RNA performs reverse splicing directly into one strand of the target DNA, creating a RNA/DNA hybrid; the IEP's endonuclease then nicks the opposite DNA strand, and its RT activity synthesizes the second DNA strand using the intron RNA as a template, resulting in intron insertion. For instance, the Ll.LtrB intron from Lactococcus lactis demonstrates this process with homing efficiencies reaching up to 1.3 × 10^{-3} per recipient cell in vivo, though frequencies can drop to 10^{-5} without full IEP function. In group I introns, mobility occurs through homing endonucleases encoded within the intron open reading frame (ORF). These enzymes, such as I-SceI from the Saccharomyces cerevisiae mitochondrial 21S rRNA intron, recognize and cleave a specific 18-40 base pair sequence in the intronless target DNA, generating a double-strand break. Cellular double-strand break repair via homologous recombination then uses the intron-containing donor allele as a template, copying the intron into the recipient site. This DNA-based mechanism contrasts with the RNA intermediate in retrohoming and is highly site-specific, promoting unidirectional spread. Spliceosomal introns, which rely on the spliceosome for excision, exhibit rare transposition events mediated by DNA intermediates rather than RNA. Experimental evidence from yeast reporter systems has captured intron gain through transposition, where an intron sequence is duplicated and inserted into a new genomic location, potentially via non-long terminal repeat (non-LTR) retrotransposon-like processes or direct DNA copying. Such events are infrequent and contribute to intron proliferation in eukaryotic genomes. Horizontal transfer facilitates intron dissemination across bacterial species, particularly for group I introns, with phylogenetic evidence indicating spread via phage-mediated vectors or direct gene exchange. For example, group I introns in cyanobacterial and α-proteobacterial tRNA genes show patterns inconsistent with vertical inheritance, supporting horizontal transmission events. Mobility rates for such transfers are low, estimated at approximately 10^{-5} per generation in bacterial populations, limiting widespread invasion but enabling occasional colonization of new hosts. Experimental studies of intron mobility often employ in vitro assays to reconstitute these processes. For group II introns, such assays involve assembling RNP particles from purified IEP and intron RNA, then incubating with target DNA to measure reverse splicing and cDNA synthesis efficiencies, as demonstrated for the Ll.LtrB system where insertion occurs preferentially at replication forks. Computational approaches detect ancient "intron fossils"—degenerate or remnant sequences—by scanning genomes for intron-like motifs using sequence homology searches and phylogenetic reconciliation to identify transfer events or losses.

Implications as Mobile Elements

Mobile introns function as selfish genetic elements that promote their own propagation within host genomes, often at the expense of host fitness, thereby driving dynamic patterns of intron gain and loss that shape evolutionary trajectories. In bacteria, mobile group II introns exhibit remarkable abundance and diversity, with many facilitating horizontal gene transfer and contributing to genetic variation across prokaryotic lineages. This mobility enables introns to insert into new genomic sites, influencing gene structure and function over evolutionary time. In eukaryotes, the accumulation of such introns contributes to genome expansion, where non-essential insertions lead to increased genome size through a process of random genetic drift and bloating of non-coding regions. As parasitic entities, mobile introns, particularly those encoding homing endonucleases, can disrupt host genes by inserting into coding sequences, potentially reducing host viability unless counterbalanced by splicing efficiency or host repair mechanisms. These elements exhibit super-Mendelian inheritance, spreading rapidly in populations until fixation, after which endonuclease activity often decays, though persistence in asexual lineages suggests ongoing evolutionary pressures like recombination or rare beneficial roles. Host genomes counteract this parasitism through genetic conflicts, including suppression mechanisms that limit excessive proliferation and maintain genome stability. The broader evolutionary implications of intron mobility include facilitation of speciation by promoting divergence in splicing patterns and gene regulation across populations. In bacteria, mobile introns contribute to the dissemination of genetic material via horizontal transfer, which can indirectly support the spread of adaptive traits such as antibiotic resistance genes embedded in mobile contexts. In modern applications, homing endonucleases derived from group I introns serve as precise tools for genome editing in gene therapy, enabling targeted disruptions or repairs in therapeutic contexts like viral interference and disease correction. Post-2020 advances, as of 2025, have expanded their engineered use in synthetic biology; for example, ARCUS nucleases—derived from the homing endonuclease I-CreI—enable high-efficiency homology-directed insertions in bacterial genomes, while synthetic homing endonuclease gene drives have been developed for applications such as mosquito population control. These developments also support antiviral strategies, including leveraging endonuclease activity for viral interference in phage systems.

References

  1. [1]
    Intron - National Human Genome Research Institute
    An intron is a region that resides within a gene but does not remain in the final mature mRNA molecule following transcription of that gene.
  2. [2]
    intron / introns | Learn Science at Scitable - Nature
    Introns are noncoding sections of an RNA transcript, or the DNA encoding it, that are spliced out before the RNA molecule is translated into a protein.
  3. [3]
    Definition of intron - NCI Dictionary of Genetics Terms
    The sequence of DNA in between exons that is initially copied into RNA, but is cut out of the final, mature messenger RNA transcript.
  4. [4]
    1977: Introns Discovered
    Apr 26, 2013 · Richard Roberts' and Phil Sharp's labs showed that eukaryotic genes contain many interruptions, called introns.Missing: biology | Show results with:biology
  5. [5]
    Introns: The Functional Benefits of Introns in Genomes - PMC - NIH
    First, all of the completely sequenced eukaryotes harbor introns in the genomic structure, whereas no prokaryotes identified so far carry introns. Second, the ...
  6. [6]
    The rise and falls of introns | Heredity - Nature
    Feb 1, 2006 · The discovery of introns and splicing in the 1970s led to two theories of their origin that became known as Introns Early and Introns Late.Introns Early... Or Introns... · The Diversity Of Introns · The Role Of Natural...
  7. [7]
    The Function of Introns - PMC - NIH
    In this review, we show that introns in contemporary species fulfill a broad spectrum of functions, and are involved in virtually every step of mRNA processing.
  8. [8]
    Intron - an overview | ScienceDirect Topics
    An intron is defined as a non-coding segment of RNA that interrupts protein coding or RNA genes, and is involved in processes such as self-catalytic splicing ...
  9. [9]
    Branch Point Identification and Sequence Requirements for Intron ...
    All spliceosomal introns contain 5′ donor and 3′ acceptor splice sites, usually with GU and AG dinucleotides at the respective intron ends and a branch point ...
  10. [10]
    RNA info: Splice site consensus
    These consensus sequences include nearly invariant dinucleotides at each end of the intron, GT at the 5' end of the intron, and AG at the 3' end of the intron.<|separator|>
  11. [11]
    snoRNAs: functions and mechanisms in biological processes, and ...
    May 12, 2022 · snoRNAs are mainly encoded by intronic regions of both protein coding and non-protein coding genes [2]. Normally, snoRNAs can be mainly ...
  12. [12]
    Distributions of exons and introns in the human genome - PubMed
    The 26,564 annotated genes in the human genome (build October, 2003) contain 233,785 exons and 207,344 introns. On average, there are 8.8 exons and 7.8 introns ...
  13. [13]
    Comparative Analysis of the Exon-Intron Structure in Eukaryotic ...
    Many eukaryotic genomic architectures are typified by small exons and flanking introns with variable length. Splice site recognition is more efficient when ...<|control11|><|separator|>
  14. [14]
    Introns: Good Day Junk Is Bad Day Treasure - ScienceDirect
    Introns are ubiquitous in large genomes and most human genes have an average of eight introns. In small genomes, few genes have introns, mostly in highly ...
  15. [15]
    'Junk DNA': Age of Intron - Front Line Genomics
    Feb 23, 2022 · Introns make up a whopping 25% of the human genome, 4 to 5 times higher than protein coding DNA. Introns were first described by Richard Roberts' and Phil ...
  16. [16]
    Intron—exon structures of eukaryotic model organisms
    On average, eukaryotic genes have 3.7 introns per 1kb of coding region, with introns typically 40-125 nt long, but with large variations. Exons are more ...
  17. [17]
    Number of introns in typical gene & average l - Human Homo sapiens
    Number of introns in typical gene & average length of intron. Range, number of introns 8/typical gene: average length of intron 3.4kb. Organism, Human Homo ...
  18. [18]
    Intron Length Evolution in Drosophila - Oxford Academic
    The genome of Drosophila melanogaster contains over 48,000 introns that must be replicated, transcribed, spliced from precursor mRNAs, and enzymatically ...Introduction · Materials and Methods · Results · Discussion
  19. [19]
    Average number of introns per gene - Eukaryotes - BNID 106965
    Average number of introns per gene. Range, 4 to 7 most multicellular ... Fruit fly Drosophila melanogaster ID: 101972 · Number and length of genes and ...
  20. [20]
    Distinct Expansion of Group II Introns During Evolution of ... - Frontiers
    Group II introns (G2Is) are ribozymes that have retroelement characteristics in prokaryotes. Although G2Is are suggested to have been an important evolutionary ...
  21. [21]
    Alternative splicing of a group II intron in a surface layer protein ...
    Nov 8, 2013 · Here we report a group II intron in the human pathogen Clostridium tetani, which undergoes four alternative splicing reactions in vivo.Missing: rarity | Show results with:rarity
  22. [22]
    Evolution of introns in the archaeal world - PNAS
    In Archaea, the introns are also small and often reside in the same location as eukaryal tRNA introns, but not always.Missing: prokaryotes | Show results with:prokaryotes
  23. [23]
    An intron within the 16S ribosomal RNA gene of the archaeon ... - NIH
    However, they differ in that the (lone) 16S rRNA gene of Pyrobaculum aerophilum contains a 713-bp intron not seen in the corresponding gene of Pyrobaculum ...
  24. [24]
    Group I introns are widespread in archaea - PMC - NIH
    May 18, 2018 · They occur most often in ribosomal RNA (rRNA) genes or tRNA genes ... genes, sometimes co-occurring with archaeal BHB introns in the same gene.
  25. [25]
    Group II Introns in Archaeal Genomes and the Evolutionary Origin of ...
    Dec 14, 2024 · A key attribute of eukaryotic genomes is the presence of abundant spliceosomal introns that break up many protein-coding genes into multiple ...
  26. [26]
    Prokaryotic introns and inteins: a panoply of form and function
    It is therefore not surprising that the majority of phage and eubacterial group I introns discovered thus far have been shown to be self-splicing ... 1980.
  27. [27]
    Spliceosome Structure and Function - PMC - PubMed Central - NIH
    The U2-dependent spliceosome is assembled from the U1, U2, U5, and U4/U6 snRNPs and numerous non-snRNP proteins. The main subunits of the U12-dependent ...
  28. [28]
    Origin and evolution of spliceosomal introns | Biology Direct | Full Text
    Apr 16, 2012 · The introns-late concept held that introns emerged only in eukaryotes and new introns have been accumulating continuously throughout eukaryotic evolution.
  29. [29]
    Introns: the “dark matter” of the eukaryotic genome - Frontiers
    ... introns are defined by several extended consensus sequences. These include the 5′ splice site (5′SS), the branch point sequence (BPS), and the 3′ splice ...Missing: prevalence | Show results with:prevalence
  30. [30]
    Intron evolution as a population-genetic process - PMC - NIH
    Moreover, the recent discovery that two types of spliceosomal introns coexist in some eukaryotes ... For example, although >99% of all eukaryotic introns ...
  31. [31]
    The emerging role of minor intron splicing in neurological disorders
    These minor or U12-type introns are found in approximately 700-800 genes in humans and represent approximately 0.5% of all human introns 4.
  32. [32]
    Representation of the secondary and tertiary structure of group I ...
    May 1, 1994 · Group I introns, which are widespread in nature, carry out RNA self–splicing. The secondary structure common to these introns was for the most part established ...
  33. [33]
    Bacterial group I introns: mobile RNA catalysts
    Mar 10, 2014 · The ability of group I introns to self-splice and therefore act as ribozymes was first demonstrated by Cech's group for a group I intron ...
  34. [34]
    Nuclear group I introns in self-splicing and beyond - Mobile DNA
    Jun 5, 2013 · Group I introns are a distinct class of RNA self-splicing introns with an ancient origin. All known group I introns present in eukaryote nuclei interrupt ...Introduction · Group I Intron Mobility At... · Group I Intron Ribozyme That...Missing: 1980s | Show results with:1980s
  35. [35]
    Atomic level architecture of group I introns revealed - PubMed
    Twenty-two years after their discovery as ribozymes, the self-splicing group I introns are finally disclosing their architecture at the atomic level.
  36. [36]
  37. [37]
    Group II intron splicing factors in plant mitochondria - Frontiers
    Typical group II introns are classified as mobile genetic elements, consisting of the self-splicing ribozyme and its intron-encoded maturase protein. A hallmark ...<|control11|><|separator|>
  38. [38]
    Structural insights into the mechanism of group II intron splicing - PMC
    Jun 1, 2018 · Recent advances in structural biology have revealed that all group II introns also share an architecturally conserved tertiary structure, ...
  39. [39]
    Mechanism of maturase‐promoted group II intron splicing
    The group II intron splicing reactions are RNA catalyzed and require a conserved RNA structure formed by six interacting double‐helical domains (DI–VI; Michel ...
  40. [40]
    A maturase-encoding group IIA intron of yeast mitochondria self ...
    Intron 1 of the coxI gene of yeast mitochondrial DNA (aI1) is a group IIA intron that encodes a maturase function required for its splicing in vivo.
  41. [41]
    Mobility of Yeast Mitochondrial Group II Introns: Engineering a New ...
    The mobile group II introns aI1 and aI2 of yeast mtDNA encode endonuclease activities that cleave intronless DNA target sites to initiate mobility by target ...
  42. [42]
    Excised Group II Introns in Yeast Mitochondria Are Lariats ... - PubMed
    Unlike group I introns, this group II intron is not demonstrably dependent on GTP for self-splicing and circularization of the isolated, linear intron is not ...
  43. [43]
    Molecular Mechanisms of pre-mRNA Splicing through Structural ...
    (A) The pre-mRNA splicing cycle. Each cycle includes three phases: assembly and activation of the spliceosome, execution of the splicing reaction, and ...
  44. [44]
    Mechanisms and regulation of spliceosome‐mediated pre‐mRNA ...
    Jul 7, 2024 · Spliceosome assembly and catalysis. The spliceosome assembles on a pre-mRNA, undergoes rearrangements, catalyzes the two reactions that lead to ...
  45. [45]
    Noisy Splicing Drives mRNA Isoform Diversity in Human Cells - NIH
    Dec 9, 2010 · We estimate that the average intron has a splicing error rate of approximately 0.7% and show that introns in highly expressed genes are spliced ...Results · Methods · Associated Data
  46. [46]
    Mechanism of Splicing Regulation of Spinal Muscular Atrophy Genes
    Jun 29, 2018 · SMN2, a nearly identical copy of SMN1, does not compensate for the loss of SMN1 due to predominant skipping of exon 7. However, correction of ...
  47. [47]
    Splicing fidelity: DEAD/H-box ATPases as molecular clocks - PMC
    Given that introns are 10‒100 times longer than exons and are defined by only minimal sequence elements, establishing fidelity during splicing for the accurate ...The Deah-Box Atpase Prp16... · Molecular Events Underlying... · A General Spliceosome...Missing: percentage | Show results with:percentage
  48. [48]
    Splice-site pairing is an intrinsically high fidelity process - PNAS
    Feb 10, 2009 · We demonstrate that the spliceosome pairs exons with an astonishingly high degree of accuracy that may be limited by the quality of pre-mRNAs generated by RNA ...Missing: correction | Show results with:correction
  49. [49]
    Staying on Message: Ensuring Fidelity in Pre-mRNA Splicing - PMC
    Indeed, recent estimates of the error rate for splicing ranges from one in one hundred to one in one hundred thousand [4–6].Dexd/h-Box Atpases... · Figure 3 · Box 4. Spliceosomal Discard...Missing: percentage | Show results with:percentage
  50. [50]
    Full article: Splicing fidelity - Taylor & Francis Online
    In our work, we found evidence for a kinetic proofreading mechanism in splicing in which the DEAH-box ATPase Prp16 discriminates against substrates undergoing ...
  51. [51]
    Proofreading and spellchecking: A two-tier strategy for pre-mRNA ...
    KINETIC PROOFREADING ACTIVITIES MEDIATED BY SPLICEOSOMAL ATPASES PROVIDE A FIRST QUALITY CONTROL MECHANISM TO LIMIT THE RATE OF MISTAKES MADE DURING SPLICING.
  52. [52]
    Intron-Mediated Enhancement: A Tool for Heterologous Gene ...
    This review highlights the complexity of IME on the levels of its regulation ... IME affects all levels of gene expression, but the strongest effect of IME ...
  53. [53]
    Intron-mediated regulation of gene expression - PubMed
    The introns involved in IME must be within transcribed sequences near the start of a gene and in their natural orientation to increase expression. The intron ...
  54. [54]
    Alternative splicing: Human disease and quantitative analysis ... - NIH
    Dec 24, 2020 · It adds another regulation layer of gene expression. Up to 95% of human multi-exon genes undergo alternative splicing to encode proteins with ...
  55. [55]
    From snoRNA to miRNA: Dual function regulatory non-coding RNAs
    In human, approximately 60% of miRNAs are encoded in introns [10], [29], of both protein-coding and non-protein-coding genes [30]. Intronic miRNAs are often co- ...Missing: percentage | Show results with:percentage
  56. [56]
    Annotation of snoRNA abundance across human tissues reveals ...
    Jun 4, 2021 · Indeed, intergenic snoRNAs contribute only to 2% of the total snoRNA abundance, confirming the mostly intronic origin of human snoRNAs [35].
  57. [57]
    Internal Introns Promote Backsplicing to Generate Circular RNAs ...
    Jun 25, 2022 · We observed an enhanced rate of circRNA generation when introns joining exons to be incorporated into circRNAs were present as compared to the intronless ...
  58. [58]
    The exon–exon junction complex provides a binding platform for ...
    This complex is the species responsible for enhancing nucleocytoplasmic export of spliced mRNAs. It does so by providing a strong binding site for the mRNA ...
  59. [59]
    Introns as Gene Regulators: A Brick on the Accelerator - PMC - NIH
    This article focuses on the specific kind of intron that increases mRNA accumulation because these introns seem to play a major role in regulating the gene in ...
  60. [60]
    Intron retention is a stress response in sensor genes and is restored ...
    Jul 1, 2022 · Intron retention (IR) is a regulatory mechanism that can retard protein production by acting at the level of mRNA processing.
  61. [61]
    The generation of diversity in immunoglobulins - NCBI - NIH
    The immunoglobulin gene segments are organized into three clusters or genetic loci—the κ, λ, and heavy-chain loci. These are on different chromosomes and each ...
  62. [62]
    The origin of introns and their role in eukaryogenesis: a compromise ...
    Although, when introns have been discovered, their origin appeared completely mysterious, the discovery of self-splicing introns offered a solution that ...'introns Early' Versus... · Reviewer's Report 1 · Reviewer's Report 4
  63. [63]
    The trouble with (group II) introns - PNAS
    May 6, 2014 · According to this, spliceosomal introns are the descendants of group II introns introduced into eukaryotes via the genome of the α- ...
  64. [64]
    Exon structure conservation despite low sequence similarity: a relic ...
    For example, studies of intron positions within modular extracellular proteins have suggested that exon shuffling played a key role in the evolution of complex ...
  65. [65]
    Signatures of Domain Shuffling in the Human Genome - PMC - NIH
    Phase combinations of flanking introns are useful indicators of exon shuffling. Intron phase is a parameter that determines the intron position relative to the ...Missing: percentage | Show results with:percentage
  66. [66]
    The genome of the choanoflagellate Monosiga brevicollis ... - Nature
    Feb 14, 2008 · This is consistent with a proliferation of introns during the early evolution of the Metazoa. ... multicellularity in modern metazoans.
  67. [67]
    A Detailed History of Intron-rich Eukaryotic Ancestors Inferred from a ...
    Intron-rich ancestors were confidently inferred for each major eukaryotic group including 53% to 74% of the human intron density for the last eukaryotic common ...
  68. [68]
    Mobile Bacterial Group II Introns at the Crux of Eukaryotic Evolution
    Group II introns are remarkable mobile retroelements that use the combined activities of an autocatalytic RNA and an intron-encoded reverse transcriptase (RT) ...
  69. [69]
    Genome-Wide Detection of Condition-Sensitive Alternative Splicing ...
    Similar to samples from Fe-deficient plants, the majority of intron retention events were induced upon Pi starvation, with this trend being more pronounced in ...
  70. [70]
    Relevance and Regulation of Alternative Splicing in Plant Heat ...
    Repression of intron splicing is temperature-dependent for many genes, as increased temperatures are associated with a higher number of genes undergoing IR ( ...
  71. [71]
    The Stable 2.0-Kilobase Intron of the Herpes Simplex Virus Type 1 ...
    During latency, herpes simplex virus expresses a unique set of latency-associated transcripts (LATs). As the 2.0-kb LAT intron is complementary to, ...
  72. [72]
    Intron-mediated induction of phenotypic heterogeneity - Nature
    Apr 20, 2022 · Concerted intron retention in transcripts occurs in response to stress, suggesting that splicing regulation has a functional role in yeast. Such ...
  73. [73]
    Review Alternative splicing: Enhancing ability to cope with stress via ...
    Many alternative splicing events are well conserved among plant species, and also across kingdoms, especially those observed in response to stress, for genes ...Review · Abstract · Introduction
  74. [74]
    Alternative splicing landscapes in Arabidopsis thaliana across ...
    Jan 14, 2021 · A. thaliana shows high levels of AS, similar to fruitflies, and that, compared to animals, disproportionately uses AS for stress responses.
  75. [75]
    A decades-long journey with mobile introns - PMC - NIH
    Because group II introns do so via an RNA intermediate, the process is termed retrohoming. Group II introns can also mobilize into ectopic sites at low ...
  76. [76]
    Bacterial group I introns: mobile RNA catalysts - PubMed Central - NIH
    Mar 10, 2014 · Group I introns are structured self-splicing introns that in part persist in genomes by minimizing the impact of their insertion into host genes ...
  77. [77]
    Retrohoming of a Bacterial Group II Intron: Mobility via Complete ...
    Group II intron homing in bacteria presents unique opportunities for studying mobility mechanisms. A group I–group II twintron construct was instrumental in ...
  78. [78]
    Insertion of group II intron retroelements after intrinsic transcriptional ...
    Apr 17, 2007 · Here we show that a subclass of group II introns avoids host damage by inserting directly after transcriptional terminator motifs in bacterial ...<|separator|>
  79. [79]
    Group II intron mobility using nascent strands at DNA ... - EMBO Press
    LtrB group II intron uses a major retrohoming mechanism in which the excised intron RNA reverse splices into one strand of a DNA target site, while the intron ...
  80. [80]
    Homing endonucleases from mobile group I introns - PubMed Central
    Homing endonucleases are highly specific DNA cleaving enzymes that are encoded within genomes of all forms of microbial life including phage and eukaryotic ...
  81. [81]
    Homing endonucleases: structural and functional insight into the ...
    Homing endonucleases confer mobility to their host intervening sequence, either an intron or intein, by catalyzing a highly specific double-strand break in ...
  82. [82]
    Homing Endonucleases: From Microbial Genetic Invaders to ...
    Jan 12, 2011 · Homing endonucleases are microbial DNA-cleaving enzymes that mobilize their own reading frames by generating double strand breaks at ...
  83. [83]
    Spliceosomal intronogenesis - PNAS
    May 23, 2016 · The most common event in our screen is plasmid-borne intron loss ... Although how the spliceosomal introns emerged and why only eukaryotes ...Spliceosomal Intronogenesis · Sign Up For Pnas Alerts · Results
  84. [84]
    Evidence for Extensive Recent Intron Transposition in Closely ...
    Oct 12, 2011 · Though spliceosomal introns are a major structural component of most eukaryotic genes and intron density varies by more than three orders of ...<|control11|><|separator|>
  85. [85]
    Sporadic Distribution of tRNACCUArg Introns among α-Purple ... - NIH
    This represents an interesting case of intron transposition and horizontal transmission between widely divergent species (cyanobacteria and α-purple bacteria).
  86. [86]
    Horizontal Transfer and Gene Conversion as an Important Driving ...
    Group I introns are highly dynamic and mobile, featuring extensive presence-absence variation and widespread horizontal transfer. Group I introns can invade ...
  87. [87]
    Functionality of In vitro Reconstituted Group II Intron RmInt1-Derived ...
    The functional unit of mobile group II introns is a ribonucleoprotein particle (RNP) consisting of the intron-encoded protein (IEP) and the excised intron RNA.
  88. [88]
    A computational approach for identifying pseudogenes in the ...
    Aug 7, 2006 · We describe a computational pipeline for identifying them, which in contrast to previous work explicitly uses intron-exon structure in parent genes to classify ...
  89. [89]
    Selfish genetic elements, genetic conflict, and evolutionary innovation | PNAS
    ### Summary of Selfish Genetic Elements (SGEs) Including Introns
  90. [90]
    Remarkable Abundance and Evolution of Mobile Group II Introns in ...
    Sep 6, 2010 · ... bacteria. Our results show that bacterial endosymbionts with reduced ... Mobile introns: pathways and proteins. ,. 2002. Mobile DNA II ...
  91. [91]
    The Repatterning of Eukaryotic Genomes by Random Genetic Drift
    This general syndrome of genomic bloating, operating in parallel across eukaryotic ... For example, the addition of every intron to a gene imposes a constraint on ...
  92. [92]
    Inteins, introns, and homing endonucleases: recent revelations ...
    Nov 13, 2006 · Self splicing introns and inteins that rely on a homing endonuclease for propagation are parasitic genetic elements. Their life-cycle and evolutionary fate has ...
  93. [93]
    Generalized bacterial genome editing using mobile group II introns ...
    Sep 3, 2013 · The introns deliver lox sites to specific genomic loci, enabling genomic manipulations. Efficiency is enhanced by adding flexibility to the RNA ...Results · Deletions · Genome Engineering In...<|separator|>
  94. [94]
    An intron endonuclease facilitates interference competition between ...
    Jul 4, 2024 · This work demonstrates how a homing endonuclease can be deployed in interference competition among viruses and provide a relative fitness advantage.