Abstract
Transposable elements (TEs), once considered genomic “junk”, are now recognized as critical regulators of genome function and human disease. These mobile genetic elements—including retrotransposons (long interspersed nuclear elements [LINE-1], Alu, short interspersed nuclear element-variable numbers of tandem repeats-Alu [SVA], and human endogenous retrovirus [HERV]) and DNA transposons—are tightly regulated by multilayered mechanisms that operate from transcription through to genomic integration. Although typically silenced in somatic cells, TEs are transiently activated during key developmental stages—such as zygotic genome activation and cell fate determination—where they influence chromatin architecture, transcriptional networks, RNA processing, and innate immune responses. Dysregulation of TEs, however, can lead to genomic instability, chronic inflammation, and various pathologies, including cancer, neurodegeneration, and aging. Paradoxically, their reactivation also presents new opportunities for clinical applications, particularly as diagnostic biomarkers and therapeutic targets. Understanding the dual role of TEs—and balancing their contributions to normal development and disease—is essential for advancing novel therapies and precision medicine.
Keywords: Transposable elements, Transposon regulation, Regulatory DNA elements, Long interspersed nuclear elements, Long terminal repeats
Introduction and Classification of Transposable Elements
Transposable elements (TEs), also known as “jumping genes”, are dynamic DNA sequences capable of relocating within the genome. First discovered in maize by Barbara McClintock in the 1940s, TEs have revolutionized our understanding of genomes by revealing their mobile and plastic nature. Ubiquitous across all life forms, TEs exhibit diverse structures and mobilization mechanisms and act as major drivers of genetic diversity and genome evolution.[1,2,3]
In humans, TEs constitute approximately 50% of the genome, far exceeding the 1.5% comprised of protein-coding genes [Figure 1A]. Based on their transposition mechanisms, TEs are classified into two major categories: Class I (retrotransposons) and Class II (DNA transposons) [Figure 1A and B]. Retrotransposons move through a “copy-and-paste” mechanism that involves an RNA intermediate, which is then reverse-transcribed into DNA. In contrast, DNA transposons use a “cut-and-paste” mechanism and do not involve an RNA intermediate [Figure 1B–E].[1,2,3] Retrotransposons are further divided into long terminal repeats (LTRs) and non-LTR retrotransposons. The latter includes long interspersed nuclear elements (LINEs) and short interspersed nuclear elements (SINEs) [Figure 1A and B].
Figure 1.
Structure, abundance and life cycle of TEs in the human genome. (A) Pie chart showing the distribution of various TEs in the human genome. (B) Structural representation of different types of TEs. The frequencies of new insertion events for L1, Alu and SVA in newborn babies are indicated, while DNA transposons and ERVs are considered molecular fossils that are incapable of mobilization in modern humans. (C) Lifecycle of DNA transposon, which encodes the DNA transposase that cuts itself and integrate into another genomic location. (D) Lifecycle of the LTR retrotransposon or ERV. ERV regulators and therapies targeting ERVs are marked as indicated. (E) Schematic of L1 retrotransposition. L1 regulators and therapies targeting L1s are marked as indicated. (F) Phylogenetic tree showing the evolution of L1 subfamilies (L1PA6 to L1Hs) in primates, including rhesus, orangutan, gorilla, chimpanzee, and human. L1Hs, L1 homo sapiens or L1PA1, is the youngest L1 subfamily in the human genome. A: A boxes; AR: A-rich linker; ASO: Antisense oligo; B: B boxes; C: Cysteine-rich domains; EN: Endonuclease; Env: Envelope protein; ERVs: Endogenous retrovirus; Gag: Group-specific antigen protein; ITR: Inverted terminal repeat; LTR: Long terminal repeat; L1: LINE-1; LINE-1: Long interspersed nuclear elements; mya; Million years; ORF: Open reading frame; Pol: Pol protein; RT: Reverse transcription; SINE: Short interspersed nuclear element; SVA: Short interspersed nuclear element-variable numbers of tandem repeats-Alu; TEs: Transposable elements; UTR: Untranslated region; VNTR: Variable numbers of tandem repeats.
DNA transposon
A typical DNA transposon consists of a coding sequence that encodes DNA transposase, flanked by two inverted terminal repeat (ITR) sequences. The transposase binds and cleaves the ITRs, excises the transposon, and facilitates its reintegration into a new genomic location[1,2,3] [Figure 1B and C]. Although DNA transposons are relatively rare in humans compared to other eukaryotes and are now largely inactive (“molecular fossils”), they have played a significant role in genome evolution[1,2,3]. Notably, they contributed to the emergence of essential genes, such as RAG1 and RAG2, which mediate V(D)J recombination[4] and gave rise to host-TE fusion proteins that function as key epigenetic regulators.[5]
LINE-1 (or L1)
As the only active autonomous TE in humans, L1 encodes all necessary components for its own mobilization. RNA polymerase II (Pol II) transcribes full-length L1 RNA, which is then translated into two key proteins: Open reading frame (ORF) 1 (RNA-binding chaperone) and ORF2 (a bifunctional enzyme with endonuclease activity at AT-rich DNA and reverse transcriptase activity). Retrotransposition proceeds via target-primed reverse transcription (TPRT) initiated at genomic nicks, followed by complementary DNA (cDNA) synthesis and integration [Figure 1E].[1,2,3] Recent structural studies have resolved the biochemical architecture of ORF2, identifying its template and target sites.[6,7]
L1s are thought to have originated from Class II introns and have persisted in vertebrate genomes for over 150 million years, diversifying through mutations and amplifications.[8] Human L1 elements include the L1Hs (L1 Homo sapiens, the youngest and most active subfamily), L1PA (primate-specific and older than L1Hs, e.g., L1PA2 and L1PA3), and L1M (mammalian-specific and ancient, predating primate divergence) [Figure 1F]. Although L1s makes up approximately 17% of the human genome with approximately 500,000 L1 copies, most are truncated or mutated, with only about 100 copies remaining capable of retrotransposition.[9] Nonetheless, hundreds to thousands of L1 elements per cell remain transcriptionally active and can serve as DNA regulatory elements, influencing gene expression.[1,2,3,10,11]
Alu
Alu is a subfamily of SINEs, approximately 300 bp long, derived from the 7SL RNA gene.[12] Alu elements are characterized by two internal repeats separated by a short spacer region. Among the Alu subfamilies, AluY is the most recent and active, followed by AluS and AluJ. AluY plays a significant role in shaping human genome evolution and gene regulation.[12]
Alu does not encode the proteins required for its own mobilization. Instead, Alu relies on L1-encoded ORF2 for reverse transcription (RT) and genomic integration [Figure 1E]. Alu RNAs (transcribed by Pol III) hijack L1’s enzymatic machinery in trans, enabling their proliferation throughout the genome. Despite their dependence on L1, Alu mobilizes at frequencies comparable to L1, with estimated insertion rates of 1:200–1:20 (L1) and 1:40–1:20 (Alu) per birth [Figure 1B].[13]
SINE-variable numbers of tandem repeats (VNTR)-Alu (SVA)
SVA is specific to hominids and is the youngest transposon family in the human genome. It consists of Alu-like and SINE-R sequences separated by VNTRs [Figure 1B]. SVAs constitute approximately 0.2% of the human genome and are grouped into subfamilies SVA-A through SVA-F [Figure 1A].[14] Similar as Alu, SVA belongs to the nonautonomous retrotransposon. SVA RNAs (transcribed by Pol II) hijack L1-encoded ORF2 for RT and genomic integration, with estimated insertion rates of 1:500–1:63 per birth [Figure 1B and 1E].[13]
Long terminal repeats (LTR)
LTR retrotransposons are flanked by LTRs and are considered remnants of ancient RNA viruses that infected ancestral genomes—examples include human endogenous retroviruses (HERVs) [Figure 1D]. These retrotransposons are classified based on the type of transfer RNA (tRNA) that binds to their primer-binding site to initiate RT. For instance, HERVK designates a group of proviruses that use lysine (K) tRNA as a primer, whereas HERVH uses histidine (H) tRNA.[15]
Over evolutionary time, these elements have accumulated mutations, internal deletions and recombinations, often leading to the loss of protein-coding sequences essential for replication. These genetic alterations prevent LTR retrotransposons from forming functional viral-like particles or producing reverse transcriptase, rendering them incapable of reintegration. Consequently, these inactive LTRs become fixed in the genome, where they may still modulate gene regulation or chromatin architecture but have lost their ability to mobilize.[1,2,3]
Molecular Mechanisms of TE Activity
Multilayered regulation of TEs
TEs present a genomic paradox: While they drive genetic innovation, their unchecked activity can threaten genomic integrity.[1,2,3] To maintain this balance, cells utilize a sophisticated, multitiered regulatory system [Table 1, Figure 1D and E].
Table 1.
Key regulatory pathways that control human transposable elements (TEs).
Regulatory pathways | Description | Key regulators | TEs affected |
---|---|---|---|
DNA methylation | m5C methylation of CpG sites in TE promoters for silencing; 6mA DNA methylation | DNMTs, TETs | L1, Alu, ERV |
Histone modification | Repressive (H3K9me3, H3K27me3, H2A/H4R3me2, H4K20me3); Activating (H3K27ac, H3K4me3, H4K16ac) | KAP1, HUSH, HUSH2, SETDB1, KRAB-ZFP, NuRD, HDACs, KMT2D, SUV39H1/2, MSL3, PRMT5 | L1, SVA, ERV |
Chromatin remodeling | Chromatin remodelers regulate TE accessibility | SWI/SNF, INO80, MORC1/2 | L1, Alu, HERVK |
Transcription factor | Transcription factors activate or repress TE transcription | YY1, RUNX3, MYC, FOXA1, P53, KRAB-ZFP, CTCF, MeCP2, SOX, DPPA2/4 | L1, Alu, ERV |
RNA interference | Small RNAs (siRNAs, miRNAs, piRNAs) degrade TE RNAs | PIWI, RIG-I, miR-128, let-7 | L1, Alu, ERV |
RNA modification | m6A and m5C modification regulates the TE RNA fate, including chromatin epigenetics | METTL3/14, YTHDC1, MBD6 | L1, Alu, ERV |
RNA binding protein | RNA binding protein regulates TE RNA processing and stability | SAFB, PTBP1, MATR3, hnRNPM, DBR1, SRSF10, DIS3, EXOSC6, NEXT, HELZ2, TUT4/7, MOV10, DDX42, SAMHD1, UPF1 | L1, ERV |
Translation and protein homeostasis | L1 proteins are controlled at the level of translation, protein assembly and degradation | SAMHD1, Condensin I/II, TDP-43, cGAS, PML, PABPC1, PABPN1 | L1 |
DNA damage & repair | DNA damage and repair factors regulate reverse transcription and DNA integration | APOBEC3, SETX, PCNA, Fanconi anemia factor, BRCA1, Homologous recombination | L1, ERV |
APOBEC3: Apolipoprotein B mRNA editing enzyme catalytic subunit 3; BRCA1: Breast cancer type 1 susceptibility protein; cGAS: Cyclic GMP-AMP synthase; CTCF: CCCTC-binding factor; DBR1: Debranching RNA lariats 1; DDX42: DEAD-box helicase 42; DIS3: DIS3 homolog, exosome endoribonuclease and 3′-5′ exoribonuclease; DNMTs: DNA methyltransferases; DPPA2/4: developmental pluripotency associated 2/4; ERV: Endogenous retrovirus; EXOSC6: Exosome component 6; FOXA1: Forkhead box A1; HDACs: Histone deacetylases; HELZ2: Helicase with zinc finger 2; HERVK: Human endogenous retrovirus K; HUSH: human silencing hub; INO80: INO80 complex ATPase subunit; KAP1: Krüppel-associated box domain–associated protein 1; KMT2D: Lysine (K)-specific methyltransferase 2D; KRAB-ZFP: Krüppel-associated box domain zinc finger proteins; L1: Long interspersed nuclear elements 1; let-7: Lethal-7; MATR3: Matrin 3; hnRNPM: Heterogeneous nuclear ribonucleoprotein M; MBD6: Methyl-CpG binding domain protein 6; MECP2: Methyl-CpG binding protein 2; METTL3/14: Methyltransferase 3/14, N6-adenosine-methyltransferase complex catalytic subunit; miRNA: MicroRNA; miR-128: MicroRNA-128; MORC1/2: MORC family CW-type zinc finger 1/2; MOV10: Mov10 RNA helicase; MSL3: MSL complex subunit 3; MYC: MYC proto-oncogene, bHLH transcription factor; NEXT: Nuclear exosome targeting complex; NuRD: Nucleosome remodeling and deacetylase; PABPC1: Poly(A) binding protein cytoplasmic 1; PABPN1: Poly(A) binding protein nuclear 1; piRNA: PIWI-interacting RNA; PIWI: P-element induced wimpy testis; PML: Progressive multifocal leukoencephalopathy; PRMT5: Protein arginine methyltransferase 5; PTBP1: Polypyrimidine tract binding protein 1; RIG-I: Retinoic acid-inducible gene I; RUNX3: Runt-related transcription factor 3; SAFB: Scaffold attachment factor B; SAMHD1: SAM and HD domain containing deoxynucleoside triphosphate triphosphohydrolase 1; SETDB1: SET domain bifurcated histone lysine methyltransferase 1; SETX: Senataxin; siRNA: Small interfering RNA; PCNA: Proliferating cell nuclear antigen; SOX: SRY-related high-mobility group; SRSF10: Serine and arginine rich splicing factor 10; SUV39H1/2: SUV39H1/2 histone lysine methyltransferase; SWI/SNF: Switch/sucrose-nonfermentable; TDP-43: Transactive response DNA-binding protein 43; TE: Transposable element; TETs: Ten-eleven translocation family proteins; TUT4/7: Terminal uridylyl transferase 4/7; UPF1: UPF1 RNA helicase and ATPase; YTHDC1: YTH N6-methyladenosine RNA binding protein C1; YY1: Yin Yang 1.
DNA methylation
The first layer of TE regulation involves epigenetic modifications [Table 1]. At the DNA methylation level, DNA methyltransferases (DNMTs) add methyl groups to CpG islands in TE promoters, thereby inhibiting transcription.[16] This mechanism is especially important in germline cells to prevent the transgenerational transmission of active TE.[17] Interestingly, the Ten-eleven translocation protein family can demethylate 5-methylcytosine (m5C) to promote TE transcription and can also repress L1 expression independently of its enzymatic activity.[18]
Histone modification
Repressive histone marks, such as H3K9me3 and H3K27me3, establish heterochromatin at LINEs, SINEs, and LTRs to silence their transcription.[19] Several silencing complexes have been identified: (1) The KAP1/TRIM28 complex recruits SETDB1 to deposit H3K9me3, particularly targeting young L1s, SVAs, and endogenous retroviruses (ERVs).[20] (2) The HUSH complex—MPP8, TASOR, and PPHLN1—recognizes intronless cDNAs and reinforces H3K9me3 deposition on active retrotransposons.[21,22] Interestingly, the HUSH complex is antagonized by an alternative complex, HUSH2, which incorporates TASOR2 (FAM208B) instead of TASOR. Unlike HUSH, TASOR2 promotes L1 expression, demonstrating how paralog switching can reverse regulatory outcomes.[23,24] (3) H3K9me3-generating methyltransferase, SUV39H, is involved in suppressing retrotransposons in mouse embryonic stem cells.[25] (4) The NuRD complex partners with retinoblastoma proteins to recruit histone deacetylases (HDACs), thereby removing activating acetylation marks.[26]
In contrast, activation marks—such as H3K4me1 and H3K27ac—are associated with TE-derived enhancers.[27] Knockout of CTBP1 and KMT2D reduces H3K9me3 levels and increases H3K4me1/H3K27ac, thereby de-repressing multiple TE families.[10] Multiple additional factors that erase or suppress activating histone modifications have been recently reviewed elsewhere, and therefore will not be discussed in detail here.[28]
RNA interference (RNAi)
In germline cells, the PIWI-interacting RNA (piRNA) pathway serves as the primary defense mechanism against TEs. The piRNA-PIWI complex mediates TE suppression through dual mechanisms: (1) post-transcriptional cleavage of TE-derived RNAs (including L1, Alu, and ERV transcripts) and (2) transcriptional silencing at the chromatin level, thereby safeguarding genomic integrity during germ cell development.[29,30,31] In somatic cells, TE regulation shifts to small RNA-mediated pathways, where endogenous small interfering RNAs (siRNAs) and microRNAs (miRNAs) target TE-derived RNAs for degradation.[32,33] For a more comprehensive discussion of RNAi in transposon control, we refer readers to a recent in-depth review on this subject.[34]
Transcription factors (TFs)
TFs such as Yin Yang 1 and Runt-related transcription factor 3 bind to L1 promoters to enhance transcription.[10,21] Similarly, p53 regulates the expression of Alu and ERV elements in response to cellular stress and DNA damage.[35] In addition, TEs have evolved strategies to evade epigenetic suppression—for instance, some L1s acquire mutations or 5′-untranslated region (UTR) deletions to escape Zn finger (ZNF)-mediated repression.[36]
RNA-binding protein (RBP)
TE-derived RNAs are regulated by various RBPs that control their stability, splicing, and retrotransposition.[10,11,37,38,39] For example, PTBP1, MATR3, hnRNPM, SRSF10, and SAFB bind L1 RNAs to suppress aberrant splicing into adjacent exons.[10,11,37,38,39] Remarkably, genome-wide clustered regularly interspaced short palindromic repeats (CRISPR) screens targeting both endogenous and codon-optimized L1 elements identified multiple RBPs—including HUSH and SAFB,[21] confirmed in subsequent independent studies[11,38,40,41,42]—that regulate L1 activity in a sequence-dependent manner. Additional RBPs, such as condensin I/II complexes, PABPC1, and PABPN1, regulate L1 translation and RNP assembly.[43] Furthermore, DIS3, EXOSC6, PML bodies, and nuclear cyclic GMP-AMP synthase (cGAS) contribute to depleting TE-derived RNAs and proteins.[10,44,45]
RNA modification
N6-methyladenosine (m6A) RNA methylation can directly recruit YTH domain-containing proteins or indirectly influence RNA-protein interactions by altering RNA structure, thereby affecting multiple aspects of RNA fate, including TE-derived transcripts.[10,40,46] Recent findings suggest crosstalk between m6A RNA modifications and histone/DNA modifications in regulating TE transcription, supporting the concept of m6A-dependent chromatin epigenetics.[10,28,47,48,49,50,51,52] In addition to m6A, m5C and its reader protein MBD6 play important roles in regulating chromatin-associated TE RNAs and, consequently, influence chromatin compaction.[53] Moreover, adenosine-to-inosine (A-to-I) RNA editing in Alu RNAs can reduce complementarity between inverted Alu repeats, unwind Alu double-stranded RNA (dsRNA) and mitigate the potential risks associated with its accumulation.[54] Thus, TE RNAs harbor diverse RNA modifications that influence their fates.
DNA damage and repair
DNA damage response pathways are also involved in regulating TE activity. APOBEC3 cytidine deaminases hypermutate the cDNA of L1 and ERV elements to prevent successful RT and genomic integration.[55] In addition, DNA repair factors—including PCNA, non-homologous end-joining genes, Fanconi anemia-related factors, BRCA1-mediated homologous recombination pathways, and DNA/RNA helix-resolving proteins—process L1 retrotransposition intermediates to either promote or restrict their integration.[21,56,57] However, the precise mechanisms by which these repair factors influence retrotransposition remain unclear and warrant further investigation.
These multilayered regulatory mechanisms—including DNA methylation, histone modifications, RNA interference, TFs, RNA modifications, and DNA damage response pathways—work together to control TE activity. Yet the fundamental question remains: How are these regulatory factors coordinated to achieve spatiotemporal control of individual TE elements? Are there species- or cell-type-specific hierarchies among these pathways? Which TE subsets are most relevant to disease development? Addressing these questions will offer critical insights into the regulation of TEs across various cell types and developmental stages.
Multifaceted regulatory functions of TEs
Comprising nearly half of the human genome, TEs are dynamic regulators of chromatin architecture, transcriptional networks, RNA processing, and innate immunity [Figure 2]. Through various mechanisms, their activity has significantly influenced genomic complexity and the evolution of gene regulation.
Figure 2.
Molecular basis and clinical implications for human TEs in development and disease. TEs are originated from functional RNAs or retrovirus, continue to shape the human genome through transposition, and have been co-opted as diverse regulatory DNA elements, which are particularly active and essential for multiple key biological contexts, including embryonic development, neurodevelopmental disorders, ageing, immune responses, and cancer progression. TEs have emerged as diagnostic biomarkers and therapeutic targets due to their disease-specific activation patterns and immunogenic properties. ASO: Antisense oligo; DNMTi: DNA methyltransferase inhibitor; ERV: Endogenous retroviruse; HDACi: Histone deacetylase inhibitor; RT: Reverse transcription; SVA: SINE-variable numbers of tandem repeats (VNTR)-Alu; TEs: Transposable elements.
Exon insertion
TE insertions into coding regions or exons can have major consequences, including loss-of-function mutations, altered protein products, and increased genomic instability.[58,59] In breast and colon cancers, L1 insertion into tumor suppressor genes or oncogenes—such as adenomatous polyposis coli—can contribute to carcinogenesis by disrupting normal gene expression and cellular regulation.[58,59] Moreover, Alu insertions can affect gene regulation and have been implicated in diseases such as multiple sclerosis (MS).[60] TE insertions into non-coding regions, including regulatory elements, may also lead to long-term effects on gene expression and cellular functions.[61,62]
Chromatin compartment
TEs can act as nucleation sites for heterochromatin formation, typically within transcriptionally inactive regions of the genome.[63,64] L1 elements frequently localize to heterochromatic domains, such as lamina-associated domains and nucleolus-associated domains, to organize the 3D genome structure.[64] Both L1 RNA and DNA can trigger phase separation of heterochromatin protein 1α, promoting heterochromatin compartmentalization and silencing of L1-enriched genomic regions.[64]
Topologically associating domain (TAD)
TEs also function as TAD boundaries—regions of the genome within which genes and regulatory elements interact more frequently than with regions outside the domain.[3] HERVH and L1 elements can establish TAD boundaries independently of CTCF, and their functions are closely linked to transcriptional activity.[11,65] Active L1s can recruit RNA Pol II, generate L1 chimeric RNAs, and help form domain boundaries within euchromatin.[11,38] The nuclear matrix protein SAFB interacts with L1-enriched regions, modulates Pol II occupancy, and weakens L1-associated boundaries, illustrating how TE-related transcription and RNA processing shape genome architecture.[11] Over time, L1 insertions generate lineage-specific TAD boundaries, contributing to species-specific chromatin organization and gene regulatory networks.[11]
Enhancer
TEs also function as enhancers.[10,66,67,68,69] The 5′-UTRs of younger L1 subfamilies, such as L1PA1–L1PA6, show stronger STARR-seq enrichments than older L1 subfamilies, suggesting greater enhancer potential.[10,67] This activity is supported by Hi-C and 4C analysis, which demonstrate that L1 elements can physically interact with distal target genes and modulate their expression.[10] In human cells, various factors, including KMT2D, DIS3, ACTL6A, and m6A RNA modifications, are involved in modulating L1 enhancer activity.[10,67] Alu repeats can also facilitate enhancer-promoter interactions by forming complementary RNA duplexes that bridge these genomic regions.[70] In addition, ERVs function as enhancers in modulating immune response and tumorigenesis.[71,72] And SVAs act as RNA-dependent enhancers to regulate both erythropoiesis and myelopoiesis.[73] However, further research is needed to elucidate the precise mechanisms by which TEs selectively activate the transcription of one or more genes across long genomic distances.
Trans-regulation
L1-derived RNAs can serve as scaffolds for transcriptional complexes that regulate gene expression, demonstrating trans-regulatory roles in ESCs.[74] During cellular stress—such as heat shock and viral infection—Alu transcripts are upregulated and can globally inhibit RNA Pol II-mediated transcription, acting as a regulatory brake on gene expression.[12] Moreover, TE-derived RNAs or their RT products can activate RNA sensors (RIG-1, MAVS) or DNA sensors (the cGAS–STING pathway) to trigger type I interferon (IFN) responses.[7,44,75,76]
Alternative initiation, splicing, and termination
TEs contribute initiation, splicing, and termination sites that give rise to TE chimeric transcripts.[77] Evolutionarily, young L1 elements typically have repressive effects by recruiting splicing factors such as SAFB, MATR3, PTBP1, and HNRNPM, which help prevent cryptic splicing in intronic regions.[11,37,38,39] In contrast, older L1s are more likely to lose binding sites for these repressive proteins, allowing them to be spliced into exons in a tissue-specific manner.[39] In addition, the 5′-UTR of L1 exhibits bidirectional promoter activity—both sense and antisense—and can generate fusion proteins with nearby exons.[11,78]
Overall, TEs are integral components of the genome, influencing key biological processes such as genome stability, 3D genome folding, TAD boundary formation, enhancer activity, RNA splicing, and innate immunity. Further research is needed to dissect the molecular features, mechanisms, and context-specific functions of individual TE elements, as well as how they interact with other regulatory DNA elements to ensure precise gene expression during development and disease.
Physiological Roles of TEs in Development and Disease
TEs serve as pivotal genomic architects, contributing to both evolutionary innovation and disease pathogenesis.[1,2,3,79] Their capacity to reshape genomes and modulate gene expression makes them critical regulators in various biological contexts [Table 2].
Table 2.
Impact of representative TEs in human development and disease.
Development and disease | TE types | Impact |
---|---|---|
Embryonic development | ||
ZGA | L1, SVA, ERVs (HERVH, LTR7) | Activates ZGA and facilitates ZGA exit; Facilitates chromatin remodeling; Activates pluripotency genes. |
Placental development | MER50, RLTR13 | Promotes invasive placentation. |
Neuro-development disorders | ||
Neurodevelopment | L1, HERV, Alu, SVA | Regulates neural differentiation, synaptic plasticity, and brain circuitry. |
Schizophrenia | L1 | Disrupts synaptic genes and neuronal signaling. |
Rett syndrome | L1 | Genome destabilization by retrotransposition. |
Autism pectrum disorder | L1, Alu | Alters synaptic connectivity and gene expression. |
Ataxia-telangiectasia | L1 | Purkinje cell dysfunction and degeneration. |
Alzheimer’s disease | L1, Alu, HERV | Induced accumulation of G-quadruplex structures; triggers inflammatory pathways and apoptosis in non-neural cells |
Amyotrophic lateral sclerosis | HERVK | Triggers neuroinflammation, neuron spreading, and neuron apoptosis |
Aicardi-Goutières syndrome | L1 | cGAS–STING pathway; neuroinflammatory activation. |
Immune-related disease | ||
Systemic lupus erythematosus | L1, HERV | Triggers chronic inflammation and autoantibody production. |
Rheumatoid arthritis | HERVK | Drives synovial inflammation and joint damage. |
Cancer | ||
Breast, colon, liver, prostate, skin cancer | L1, ERV, HERVK, SVA | Promotes tumorigenesis and immune evasion. |
cGAS-STING: Cyclic GMP-AMP synthase - stimulator of interferon genes; ERV: Endogenous retroviruse; HERVH: Human endogenous retrovirus H; HERVK: Human endogenous retrovirus K; L1: Long interspersed nuclear elements 1; LTR: Long terminal repeat; SLE: Systemic lupus erythematosus; SVA: SINE-variable numbers of tandem repeats (VNTR)-Alu; TE: Transposable element; ZGA: Zygotic genome activation.
Embryonic development
TEs exhibit dynamic, stage-specific regulatory roles in embryonic development.[19,80] Distinct TE families show precise temporal activation patterns: the LTR7Y (HERVH) subfamily is active during blastocyst formation, while MLT2A1 (HERVL) peaks at the eight-cell stage.[81] These stage-specific activities highlight TEs as critical developmental regulators.
During zygotic genome activation (ZGA)—the transition from maternal to zygotic control of gene expression—LTRs and L1 elements are reactivated and function as enhancers or alternative promoters to modulate the expression of genes essential for embryogenesis.[80] L1 is highly expressed in preimplantation embryos and is subsequently downregulated during differentiation; its repression impairs chromatin accessibility and blastocyst formation.[82] Percharde et al[59] demonstrated that L1 RNA acts as a nuclear scaffold in mouse embryonic stem cells (ESCs), facilitating ribosomal DNA expression and repressing DUX, a master regulator of the ZGA program. In humans, L1 RNAs repress gene activation in 8-cell-like cells, which correspond to the ZGA stage.[83] Recently, Li et al[10] found that transcriptionally active L1s function as enhancers that selectively contact and activate minor ZGA genes in early mouse embryos. Depletion of L1 RNA abolished the activation of nearly half of these genes, resulting in embryo arrest at the 2–4-cell stage. Collectively, these findings suggest that L1 transcription and RNA initially act in cis to promote ZGA gene expression and initiate ZGA and subsequently act in trans to repress ZGA genes and facilitate ZGA exit.
Beyond ZGA, TEs also play essential roles in placental development.[84,85] In mammals, MER50 and RLTR13 ERVs provide key enhancers that regulate genes involved in invasive placentation—a defining trait of eutherian mammals.[86] The ERV-derived envelope (env) gene encodes proteins, such as syncytin, which are crucial for trophoblast fusion, proliferation, and antiviral defense.[87] Besides, a mouse-specific retrotransposon, MT2B2, provides an alternative promoter that drives expression of an N-terminally truncated isoform of Cdk2ap1, which plays a key role in embryo implantation.[88] Moreover, a systematic scan has identified 1507 human env-derived ORFs, including truncated forms, such as Suppressyn and Refrex-1, which protect placental cells via receptor interference.[89]
While TEs clearly contribute to developmental regulation, their evolutionary importance requires nuanced interpretation. Recent studies challenge the widespread assumption that TE-derived regulatory elements are universally essential. Instead, many may represent lineage-specific adaptations with limited functions or evolutionarily neutral “genomic noise” lacking conserved biological impact.[90] For example, the MER50 endogenous retrovirus-derived enhancer plays a critical role in placental development, specifically in eutherian mammals, displaying strict trophoblast-specific activity and being entirely absent in non-placental species.[91] These findings underscore the need for a more nuanced framework to assess TE contributions to gene regulation—one that considers both their biological relevance and the frequent lack of conservation across species.
Neural development and disorders
TEs play crucial regulatory roles in brain development through multiple mechanisms: (1) Somatic L1 retrotransposition occurs frequently in neural precursor cells, potentially contributing to genome plasticity and gene regulatory programs.[92,93] (2) L1 elements regulate genes involved in neurogenesis, shaping the transcriptional landscape of neural progenitors and stem cells.[94] (3) Alu elements modulate the expression of neurotransmitter receptors and growth factors, thereby influencing synaptic plasticity.[95] These findings suggest that TEs actively participate in regulating key genes involved in neural differentiation, synaptic plasticity, and neurogenesis.
Dysregulation or over-activation of TEs contributes to the pathogenesis of neurological disorders through multiple pathways: (1) Somatic L1 insertions in the prefrontal cortex and hippocampal neurons can disrupt synaptic genes (e.g., DISC1) and alter neuronal signaling pathways in patients with schizophrenia.[96] (2) Alu elements regulate neuroligins and neurexins, whose dysregulation is implicated in autism spectrum disorders (ASDs), characterized by impaired social communication and repetitive behavior.[97] In addition, TE-induced genomic instability may cause copy number variations (CNVs), which are frequently observed in patients with ASD and are believed to contribute to disease etiology.[98] (3) In Rett syndrome, mutations in MeCP2 lead to L1 hypomethylation and increased L1 retrotransposition, resulting in genomic instability and disease progression.[99] (4) L1 activation contributes to Purkinje cell dysfunction and degeneration in the mouse cerebellum, driving the onset of ataxia.[100] This finding is supported by increased L1 mobility observed in patients with ataxia-telangiectasia.[101] (5) In Alzheimer’s disease (AD), active L1 elements promote G4 structure accumulation, potentially disrupting gene expression in neurons.[102] Alu RNAs have also been shown to activate inflammatory pathways and induce apoptosis in non-neural cells, including retinal pigment epithelial cells.[103] In addition, cytoplasmic nucleic acids derived from HERVs stimulate innate immune sensors, leading to type I IFN production and neuroinflammation in AD.[104] (6) In amyotrophic lateral sclerosis (ALS), elevated TDP-43 expression enhances HERVK activity, triggering neuroinflammation and motor neuron apoptosis through neuronal spreading.[105] (7) In Aicardi–Goutières syndrome, the accumulation of L1-derived single-stranded DNA activates the cGAS–STING pathway, leading to neuroinflammation.[106] Overall, these findings highlight the conserved role of TE dysregulation in neurological disorders and support the potential for therapeutic targeting of L1 activity, including epigenetic modulation, anti-inflammatory strategies, and genome stabilization therapies.
Immune response
TEs influence both innate and adaptive immunity, as well as autoimmune disease pathogenesis, highlighting their complex and multifaceted roles in human health.[62,107,108]
Innate immunity, the body’s first line of defense against pathogens, is activated in part by TEs.[108] For example, L1-derived dsRNA activates RIG-I and MDA5, triggering a type I IFN response.[75] In parallel, L1 cDNA can stimulate the cGAS pathway.[109] HERVK also engages RNA sensors, such as TLR7 and TLR8, and DNA sensors, including cGAS and AIM2.[110,111] Beyond these trans-effects, ERVs and L1 elements can act in cis as IFN-inducible enhancers that activate IFN-stimulated genes and modulate the IFN pathway.[69,107] In addition, ERVs recruit TFs such as IRF3, STAT1, and NF-κB, promoting immune activation and inflammation.[71] Although these interactions remain context-dependent and mechanistically underexplored, further research is required to clarify the specific pathways involved and their implications in various disease settings.
TEs also influence adaptive immunity by regulating T-cell development and function.[112] L1, HERV, and SVA elements contribute to T-cell receptor signaling and differentiation by serving as regulatory sequences for genes essential to T-cell activation.[83,113] HERV-derived sequences have been shown to regulate the expression of key cytokines and TFs such as FOXP3, which is critical for developing regulatory T cells that maintain immune tolerance and prevent autoimmunity.[114]
Although TEs are crucial for normal immune function, their dysregulation can contribute to immune-related diseases, including autoimmunity and chronic inflammation.[108] For instance, TEs have been linked to T-cell exhaustion in chronic infections, impairing immune surveillance and response.[58,115] TE overactivation also induces genomic instability and persistent inflammation, contributing to systemic lupus erythematosus (SLE), rheumatoid arthritis (RA), and MS.[116] In SLE, aberrant activation of L1s and HERVs leads to the production of dsRNA and retrotransposon-derived transcripts that amplify type I IFN response.[117] Dysregulated TE expression in immune cells, including plasma and dendritic cells, exacerbates autoimmunity and drives tissue damage and disease progression.[118] HERVK has been implicated in RA-associated synovial inflammation, where its expression in synovial fibroblasts promotes cytokine production and immune cell infiltration into joints.[118] Interestingly, TE activation is also observed during acute viral infections,[119] although its functional significance remains unclear. Is TE activation a programmed component of antiviral defense or merely a byproduct of infection-induced transcription? Further mechanistic studies are needed to resolve these context-dependent roles and assess their therapeutic potential.
Aging and aging-related disease
As organisms age, TEs become more active and contribute to genomic instability.[120] Somatic mutations arising from TE-induced genomic instability accumulate over time and are associated with age-related diseases, such as neurodegeneration and cancer.[121] In aging neurons, L1 activation can lead to increased DNA damage and the formation of double-strand breaks, which are inefficiently repaired in aging cells.[122] This accumulation of DNA damage contributes to cellular dysfunction and the aging process.[122]
Chronic, low-grade inflammation is another well-documented feature of aging.[121] TEs play a significant role in driving this inflammation by activating immune responses. For example, derepressed HERVK elements promote cellular senescence through activating the cGAS-STING pathway in premature aging models.[123] Moreover, cytoplasmic accumulation of L1 DNA during aging can trigger type I IFN responses, thereby promoting age-related inflammation.[121] Notably, this effect can be reduced by nucleoside reverse transcriptase inhibitors, suggesting that L1 represents a potential target for mitigating age-associated inflammatory conditions.[124]
A recent study revealed a mechanism by which derepressed TEs contribute to aging: the SVA composite transposon promotes myeloid-biased hematopoiesis via its enhancer activity.[73] This evidence suggests that the age-associated increase in TE activity and their accompanying enhancer functions constitute a potent force driving the aging process.
Cancer
TEs are frequently reactivated in cancer cells, leading to somatic retrotransposition and genomic instability.[58,59] TE retrotransposition can insert new genetic materials into coding or regulatory regions, resulting in mutations that disrupt oncogenes or tumor suppressor genes.[1,59,125,126] TE activation in tumor cells is also associated with CNVs, which are prevalent in cancer and linked to poor prognosis.[59]
Despite their tumor-promoting effects, reactivated L1s can also induce genome instability, activate DNA damage response, and exert tumor-suppressive roles, particularly in myeloid leukemia.[115] In addition, TEs influence tumor progression by modulating immune responses that may suppress cancer development.[7,44,76,127] TE-derived RNAs and their RT products (complementary single-stranded DNA or DNA/RNA hybrids) can activate nucleic acid-sensing pathways, including RNA receptors like RIG-I-like receptors, Toll-like receptor 3, and the DNA sensor cGAS enzyme.[115,125] Moreover, the translation of TE-host chimeric RNAs can generate novel peptides, broadening the range of tumor antigens.[58,128,129,130] Thus, TEs’ ability to stimulate innate immune sensors and generate tumor-specific antigens makes them promising targets for cancer immunotherapy.[131]
TEs in Diagnostics and Therapy
TEs have emerged as valuable tools in precision medicine, serving as diagnostic biomarkers and therapeutic targets due to their disease-specific activation patterns and immunogenic properties [Figure 1D, 1E and Figure 2].
TEs as biomarkers
TEs are often reactivated in cancer and other diseases, making them promising diagnostic biomarkers.[58,79] Hypomethylation of L1, and L1 ORF1, are associated with worse clinical outcomes in cancers, indicating its potential as an early biomarker for cancer detection.[126] Somatic retrotransposition of L1 and other TEs is linked to tumorigenesis in breast, colon, and liver cancers.[59] L1-derived sequences can be detected in circulating tumor DNA (ctDNA), exosomes, and plasma, offering a noninvasive approach for tumor detection and monitoring.[132] In addition, Alu elements are upregulated in specific cancers and may serve as cancer-specific biomarkers.[133]
Beyond oncology, TEs show diagnostic potential for neurodegenerative diseases, including Alzheimer’s disease (AD) and Parkinson’s disease (PD).[134,135] Reactivation of L1 and HERVs in neurons correlates with neuroinflammation and neurodegeneration, suggesting that TE activity in cerebrospinal fluid or serum could track disease progression and therapeutic response.[136]
Tumor-specific antigens and immunotherapy
TE reactivation in cancer cells generates immunogenic neoantigens, which are absent in healthy cells due to tight epigenetic silencing.[58,115,128,129,130] These TE-derived peptides are presented via MHC molecules, triggering T-cell and antibody responses against tumors.[58,115,128,129,130] For example, L1-encoded proteins (ORF1 and ORF2) are highly expressed in cancers (e.g., breast, colon) and recognized by cytotoxic T lymphocytes.[58,115,125] The L1 antisense promoter-derived L1PA2–GABRG2 fusion peptides, presented on the surface of cancer cells, represent new potential targets for cancer therapy.[128] Besides, HERVK and Alu produce tumor-specific peptides, expanding the pool of targets for cancer vaccines and personalized immunotherapy.[58,115,125] Moreover, a recent study showed that HIF2α upregulates intact ERVs (e.g., ERVE-4, ERV3.2) in clear cell renal carcinoma, whose peptides elicit T-cell responses post-immunotherapy or transplantation.[137] Pharmacological HIF2α stabilization induces similar ERV expression in other cancers, highlighting broader therapeutic potential.[137]
Epigenetic modulation of TEs for therapy
Epigenetic drugs can reactivate silenced TEs, enhancing type I IFN responses, and expose TE-derived neoantigens to improve immune recognition.[79,129] For example, DNMT/HDAC inhibitors promote TE expression, boost type I IFN responses, and expose TE neoantigens for immune recognition.[79,129] Besides, H3K27M mutations in pediatric gliomas redistribute H3K27 acetylation, derepress ERVs, and sensitize tumors to combined treatment with DNMT/HDAC inhibitors.[138] Moreover, depleting or inhibiting histone modifiers (e.g., KDM1A, KDM5B, SETDB1) reactivates TEs, triggers interferon signaling, and synergizes with immune checkpoint blockade.[139,140,141] These approaches hold promise for advancing epigenetic therapies and improving cancer treatment outcomes. However, challenges such as tumor heterogeneity in TE expression and immune evasion mechanisms remain barriers to efficacy.
TE depletion via genome or RNA therapy
Strategies aimed at silencing TE can reduce genomic instability and impede disease progression.[58,142] For example, L1 knockdown has been shown to suppress tumor growth, highlighting the therapeutic potential of TE silencing in cancer and other disorders associated with genomic instability.[143,144] Moreover, antisense oligonucleotides targeting L1 have been shown to extend the lifespan of mice by approximately 25%, suggesting their potential as antiaging interventions.[145]
RT inhibition
RT inhibitors—such as 3TC, originally developed for the human immunodeficiency virus—have been repurposed to block retrotransposition in clinical trials for AD, PD, ALS, cancer, lupus, and aging.[146,147,148] Although these drugs can slow tumor progression, they typically fail to eliminate malignant cells entirely, likely because TEs are not the sole drivers of these complex diseases.[147] Consequently, combination therapies are often required. In addition, given the physiological roles of TEs, there are concerns about off-target effects from RT inhibitors.[148]
Harnessing technological innovations for TE-targeted therapies
TE research has long been hindered by their repetitive nature and tight epigenetic silencing, which obscure detection of their mobilization, expression, and function. Recent advances in single-cell RNA sequencing (scRNA-seq), long-read sequencing and machine learning are now overcoming these barriers, revealing unprecedented insights into TE biology and their contributions to health and disease.
scRNA-seq: mapping TE dynamics at cellular resolution
scRNA-seq enables high-resolution profiling of gene expression, revealing how TEs influence critical biological processes such as cell differentiation, immune responses, and tumor progression.[149,150] Complementary single-cell epigenomic techniques, such as single-cell ATAC-seq, further elucidate the chromatin dynamics governing TE activation or silencing.[151] These approaches offer valuable perspectives on the contribution of TEs to aging, cancer, and immune dysfunction.
Long-read sequencing: overcoming the repetitive challenge
Conventional short-read sequencing struggles to resolve repetitive TE sequences, whereas long-read platforms (e.g., PacBio and Oxford Nanopore) can accurately assemble full-length TE sequences and detect mobilization events and genomic rearrangements.[152,153] For example, long-read sequencing of ctDNA could enable non-invasive detection of TE-driven mutations in cancer. However, long-read sequencing is often non-quantitative, expensive, and difficult to resolve full-length L1 insertions. Overcoming these challenges will be essential to fully leverage its potential for TE research and precision medicine.
Machine learning for TE analysis
The development of computational tools for TE analysis from long-read data represents an emerging frontier. machine learning approaches trained on multi-omic datasets show particular promise for identifying immunogenic TE-derived epitopes, which have been notoriously difficult to predict due to the complexity of immune recognition.[154] Tools such as NeoFuse and nextNEOpi utilize machine learning algorithms, including linear regression and artificial neural networks, to predict neoantigens from TE-containing RNAs.[155,156] While machine learning enhances high-throughput screening for TE-based immunotherapies, further validation, such as cross-referencing with experimental data, is needed to refine these models for clinical use.
Expanding the therapeutic toolkit
The recent identification of 40 novel integration-competent DNA TEs in the human genome, along with development of associated molecular toolkits, highlights the growing potential for TE-based therapeutic applications.[157] Thus, as we deepen our understanding of fundamental TE biology, these discoveries are catalyzing innovative approaches to genome engineering and targeted therapies, presenting exciting opportunities for novel prognostic markers and targeted therapies.
Conclusions
TEs, once dismissed as “junk DNA”, are now recognized as dynamic regulators of human biology. Their tightly regulated activity shapes neural development, immune responses, and aging, while their dysregulation contributes to diverse diseases, including cancer, autoimmune disorders, and neurodegenerative diseases. Advances in single-cell omics, long-read sequencing and machine learning have enabled unprecedented resolution in mapping TE activity and interactions, paving the way for novel TE-based biomarkers and targeted therapies. While clinical translation faces challenges, such as cell-type specificity and off-target effects, TEs offer immense potential for advancing precision medicine by bridging genetic variation, gene regulation, and disease mechanisms.
Funding
This work was supported by grants from Beijing Natural Science Foundation (No. JQ24033), the National Key Research and Development Program (No. 2022YFA1302700), the National Natural Science Foundation of China (Nos. 32350011, 32270593, 32070631, and 32400441), and the China Postdoctoral Science Foundation (No. 2024T170499). We thank the support from the Benyuan Fund-Young Investigator Exploration Grant in Life Sciences, Tsinghua University Initiative Scientific Research Program and Pillars of the Nation Funding for Life Sciences, Tsinghua university.
Conflicts of interest
None.
Footnotes
How to cite this article: Hong YQ, Liu N. Transposable elements in health and disease: Molecular basis and clinical implications. Chin Med J 2025;138:2220–2233. doi: 10.1097/CM9.0000000000003775
References
- 1.Kazazian HH, Jr, Moran JV. Mobile DNA in Health and Disease. N Engl J Med 2017;377:361–370. doi: 10.1056/NEJMra1510092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Feschotte C. Transposable elements: McClintock’s legacy revisited. Nat Rev Genet 2023;24:797–800. doi: 10.1038/s41576-023-00652-3. [DOI] [PubMed] [Google Scholar]
- 3.Lawson HA, Liang Y, Wang T. Transposable elements in mammalian chromatin organization. Nat Rev Genet 2023;24:712–723. doi: 10.1038/s41576-023-00609-6. [DOI] [PubMed] [Google Scholar]
- 4.Kapitonov VV, Koonin EV. Evolution of the RAG1-RAG2 locus: both proteins came from the same transposon. Biol Direct 2015;10:20. doi: 10.1186/s13062-015-0055-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Cosby RL Judd J Zhang R Zhong A Garry N Pritham EJ, et al. Recurrent evolution of vertebrate transcription factors by transposase capture. Science 2021;371:eabc6405. doi: 10.1126/science.abc6405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Thawani A, Florez Ariza AJ, Nogales E, Collins K. Template and target-site recognition by human LINE-1 in retrotransposition. Nature 2024;626:186–193. doi: 10.1038/s41586-023-06933-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Baldwin ET Van Eeuwen T Hoyos D Zalevsky A Tchesnokov EP Sánchez R, et al. Structures, functions and adaptations of the human LINE-1 ORF2 protein. Nature 2024;626:194–206. doi: 10.1038/s41586-023-06947-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Ivancevic AM, Kortschak RD, Bertozzi T, Adelson DL. LINEs between Species: Evolutionary Dynamics of LINE-1 Retrotransposons across the Eukaryotic Tree of Life. Genome Biol Evol 2016;8:3301–3322. doi: 10.1093/gbe/evw243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Brouha B Schustak J Badge RM Lutz-Prigge S Farley AH Moran JV, et al. Hot L1s account for the bulk of retrotransposition in the human population. Proc Natl Acad Sci USA 2003;100:5280–5285. doi: 10.1073/pnas.0831042100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Li X Bie L Wang Y Hong Y Zhou Z Fan Y, et al. LINE-1 transcription activates long-range gene expression. Nat Genet 2024;56:1494–1502. doi: 10.1038/s41588-024-01789-5. [DOI] [PubMed] [Google Scholar]
- 11.Hong Y Bie L Zhang T Yan X Jin G Chen Z, et al. SAFB restricts contact domain boundaries associated with L1 chimeric transcription. Mol Cell 2024;84:1637–1650.e10. doi: 10.1016/j.molcel.2024.03.021. [DOI] [PubMed] [Google Scholar]
- 12.Zhang X-O, Pratt H, Weng Z. Investigating the Potential Roles of SINEs in the Human Genome. Annu Rev Genomics Hum Genet 2021;22:199–218. doi: 10.1146/annurev-genom-111620-100736. [DOI] [PubMed] [Google Scholar]
- 13.Feusier J Watkins WS Thomas J Farrell A Witherspoon DJ Baird L, et al. Pedigree-based estimation of human mobile element retrotransposition rates. Genome Res 2019;29:1567–1577. doi: 10.1101/gr.247965.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Chu C Lin EW Tran A Jin H Ho NI Veit A, et al. The landscape of human SVA retrotransposons. Nucleic Acids Res 2023;51:11453–11465. doi: 10.1093/nar/gkad821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Dopkins N, Nixon DF. Activation of human endogenous retroviruses and its physiological consequences. Nat Rev Mol Cell Biol 2024;25:212–222. doi: 10.1038/s41580-023-00674-z. [DOI] [PubMed] [Google Scholar]
- 16.Smith ZD, Hetzel S, Meissner A. DNA methylation in mammalian development and disease. Nat Rev Genet 2025;26:7–30. doi: 10.1038/s41576-024-00760-8. [DOI] [PubMed] [Google Scholar]
- 17.Zoch A Konieczny G Auchynnikava T Stallmeyer B Rotte N Heep M, et al. C19ORF84 connects piRNA and DNA methylation machineries to defend the mammalian germ line. Mol Cell 2024;84:1021–1035.e11. doi: 10.1016/j.molcel.2024.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Arroyo M Hastert FD Zhadan A Schelter F Zimbelmann S Rausch C, et al. Isoform-specific and ubiquitination dependent recruitment of Tet1 to replicating heterochromatin modulates methylcytosine oxidation. Nat Commun 2022;13:5173. doi: 10.1038/s41467-022-32799-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Xu R Li S Wu Q Li C Jiang M Guo L, et al. Stage-specific H3K9me3 occupancy ensures retrotransposon silencing in human pre-implantation embryos. Cell Stem Cell 2022;29:1051–1066.e8. doi: 10.1016/j.stem.2022.06.001. [DOI] [PubMed] [Google Scholar]
- 20.Senft AD, Macfarlan TS. Transposable elements shape the evolution of mammalian development. Nat Rev Genet 2021;22:691–711. doi: 10.1038/s41576-021-00385-1. [DOI] [PubMed] [Google Scholar]
- 21.Liu N Lee CH Swigut T Grow E Gu B Bassik MC, et al. Selective silencing of euchromatic L1s revealed by genome-wide screens for L1 regulators. Nature 2018;553:228–232. doi: 10.1038/nature25179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Seczynska M, Bloor S, Cuesta SM, Lehner PJ. Genome surveillance by HUSH-mediated silencing of intronless mobile elements. Nature 2022;601:440–445. doi: 10.1038/s41586-021-04228-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Danac J Matthews RE Gungi A Qin C Parsons H Antrobus R, et al. Competition between two HUSH complexes orchestrates the immune response to retroelement invasion. Mol Cell 2024;84:2870–2881.e5. doi: 10.1016/j.molcel.2024.06.020. [DOI] [PubMed] [Google Scholar]
- 24.Jensvold ZD, Flood JR, Christenson AE, Lewis PW. Interplay between Two Paralogous Human Silencing Hub (HuSH) Complexes in Regulating LINE-1 Element Silencing. Nat Commun 2024;15:9492. doi: 10.1038/s41467-024-53761-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Bulut-Karslioglu A De La Rosa-Velázquez IA Ramirez F Barenboim M Onishi-Seebacher M Arand J, et al. Suv39h-dependent H3K9me3 marks intact retrotransposons and silences LINE elements in mouse embryonic stem cells. Molr Cell 2014;55:277–290. doi: 10.1016/j.molcel.2014.05.029. [DOI] [PubMed] [Google Scholar]
- 26.Montoya-Durango DE, Ramos KA, Bojang P, Ruiz L, Ramos IN, Ramos KS. LINE-1 silencing by retinoblastoma proteins is effected through the nucleosomal and remodeling deacetylase multiprotein complex. BMC Cancer 2016;16:38. doi: 10.1186/s12885-016-2068-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Karttunen K Patel D Xia J Fei L Palin K Aaltonen L, et al. Transposable elements as tissue-specific enhancers in cancers of endodermal lineage. Nat Commun 2023;14:5313. doi: 10.1038/s41467-023-41081-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Li X, Liu N. Advances in understanding LINE-1 regulation and function in the human genome. Trends Genet 2025. doi: 10.1016/j.tig.2025.04.011. [DOI] [PubMed] [Google Scholar]
- 29.Dias Mirandela M Zoch A Leismann J Webb S Berrens RV Valsakumar D, et al. Two-factor authentication underpins the precision of the piRNA pathway. Nature 2024;634:979–985. doi: 10.1038/s41586-024-07963-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Wang X, Ramat A, Simonelig M, Liu M-F. Emerging roles and functional mechanisms of PIWI-interacting RNAs. Nat Rev Mol Cell Biol 2023;24:123–141. doi: 10.1038/s41580-022-00528-0. [DOI] [PubMed] [Google Scholar]
- 31.Wang L, Dou K, Moon S, Tan FJ, Zhang ZZ. Hijacking Oogenesis Enables Massive Propagation of LINE and Retroviral Transposons. Cell 2018;174:1082–1094.e12. doi: 10.1016/j.cell.2018.06.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Svoboda P. Renaissance of mammalian endogenous RNAi. FEBS Letters 2014;588:2550–2556. doi: 10.1016/j.febslet.2014.05.030. [DOI] [PubMed] [Google Scholar]
- 33.Tristán-Ramos P Rubio-Roldan A Peris G Sánchez L Amador-Cubero S Viollet S, et al. The tumor suppressor microRNA let-7 inhibits human LINE-1 retrotransposition. Nat Commun 2020;11:5712. doi: 10.1038/s41467-020-19430-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Iwasaki YW, Shoji K, Nakagwa S, Miyoshi T, Tomari Y. Transposon–host arms race: a saga of genome evolution. Trends Genet 2025;41:369–389. doi: 10.1016/j.tig.2025.01.009. [DOI] [PubMed] [Google Scholar]
- 35.Tiwari B, Jones AE, Abrams JM. Transposons, p53 and Genome Security. Trend Genet 2018;34:846–855. doi: 10.1016/j.tig.2018.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Imbeault M, Helleboid P-Y, Trono D. KRAB zinc-finger proteins contribute to the evolution of gene regulatory networks. Nature 2017;543:550–554. doi: 10.1038/nature21683. [DOI] [PubMed] [Google Scholar]
- 37.Zheng R Dunlap M Bobkov G Gonzalez-Figueroa C Patel KJ Lyu J, et al. hnRNPM protects against the dsRNA-mediated interferon response by repressing LINE-associated cryptic splicing. Mol Cell 2024;84:2087–2103.e8. doi: 10.1016/j.molcel.2024.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ilık İA Glažar P Tse K Brändl B Meierhofer D Müller F-J, et al. Autonomous transposons tune their sequences to ensure somatic suppression. Nature 2024;626:1116–1124. doi: 10.1038/s41586-024-07081-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Attig J Agostini F Gooding C Chakrabarti AM Singh A Haberman N, et al. Heteromeric RNP Assembly at LINEs Controls Lineage-Specific RNA Processing. Cell 2018;174:1067–1081. e17. doi: 10.1016/j.cell.2018.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Xiong F Wang R Lee J-H Li S Chen S-F Liao Z, et al. RNA m6A modification orchestrates a LINE-1–host interaction that facilitates retrotransposition and contributes to long gene vulnerability. Cell Res 2021;31:861–885. doi: 10.1038/s41422-021-00515-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Spencley AL Bar S Swigut T Flynn RA Lee CH Chen L-F, et al. Co-transcriptional genome surveillance by HUSH is coupled to termination machinery. Mol Cell 2023;83:1623–1639.e8. doi: 10.1016/j.molcel.2023.04.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Garland W Müller I Wu M Schmid M Imamura K Rib L, et al. Chromatin modifier HUSH co-operates with RNA decay factor NEXT to restrict transposable element expression. Molecular Cell 2022;82:1691–1707.e8. doi: 10.1016/j.molcel.2022.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Ward JR Khan A Torres S Crawford B Nock S Frisbie T, et al. Condensin I and condensin II proteins form a LINE-1 dependent super condensin complex and cooperate to repress LINE-1. Nucleic Acids Res 2022;50:10680–10694. doi: 10.1093/nar/gkac802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Zhen Z Chen Y Wang H Tang H Zhang H Liu H, et al. Nuclear cGAS restricts L1 retrotransposition by promoting TRIM41-mediated ORF2p ubiquitination and degradation. Nat Commun 2023;14:8217. doi: 10.1038/s41467-023-43001-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Mathavarajah S Vergunst KL Habib EB Williams SK He R Maliougina M, et al. PML and PML-like exonucleases restrict retrotransposons in jawed vertebrates. Nucleic Acids Research 2023;51:3185–3204. doi: 10.1093/nar/gkad152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Liu N, Pan T. N6-methyladenosine–encoded epitranscriptomics. Nat Struct Mol Biol 2016;23:98–102. doi: 10.1038/nsmb.3162. [DOI] [PubMed] [Google Scholar]
- 47.Chelmicki T Roger E Teissandier A Dura M Bonneville L Rucli S, et al. m6A RNA methylation regulates the fate of endogenous retroviruses. Nature 2021;591:312–316. doi: 10.1038/s41586-020-03135-1. [DOI] [PubMed] [Google Scholar]
- 48.Wei J Yu X Yang L Liu X Gao B Huang B, et al. FTO mediates LINE1 m6A demethylation and chromatin regulation in mESCs and mouse development. Science 2022;376:968–973. doi: 10.1126/science.abe9582. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Liu J Gao M He J Wu K Lin S Jin L, et al. The RNA m6A reader YTHDC1 silences retrotransposons and guards ES cell identity. Nature 2021;591:322–326. doi: 10.1038/s41586-021-03313-9. [DOI] [PubMed] [Google Scholar]
- 50.Liu J Dou X Chen C Chen C Liu C Xu MM, et al. N6-methyladenosine of chromosome-associated regulatory RNA regulates chromatin state and transcription. Science 2020;367:580–586. doi: 10.1126/science.aay6018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Li Z Fang F Zafar MI Wu X Liu X Tan X, et al. RNA m6A modification regulates L1 retrotransposons in human spermatogonial stem cell differentiation in vitro and in vivo. Cell Mol Life Sci 2024;81:92. doi: 10.1007/s00018-024-05119-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Sun T, Xu Y, Xiang Y, Ou J, Soderblom EJ, Diao Y. Crosstalk between RNA m6A and DNA methylation regulates transposable element chromatin activation and cell fate in human pluripotent stem cells. Nat Genet 2023;55:1324–1335. doi: 10.1038/s41588-023-01452-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Zou Z Dou X Li Y Zhang Z Wang J Gao B, et al. RNA m5C oxidation by TET2 regulates chromatin state and leukaemogenesis. Nature 2024;634:986–994. doi: 10.1038/s41586-024-07969-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Kaneko H Dridi S Tarallo V Gelfand BD Fowler BJ Cho WG, et al. DICER1 deficit induces Alu RNA toxicity in age-related macular degeneration. Nature 2011;471:325–330. doi: 10.1038/nature09830. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Modenini G, Abondio P, Boattini A. The coevolution between APOBEC3 and retrotransposons in primates. Mob DNA 2022;13:27. doi: 10.1186/s13100-022-00283-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Bona N, Crossan GP. Fanconi anemia DNA crosslink repair factors protect against LINE-1 retrotransposition during mouse development. Nat Struct Mol Biol 2023;30:1434–1445. doi: 10.1038/s41594-023-01067-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Mita P Sun X Fenyo D Kahler DJ Li D Agmon N, et al. BRCA1 and S phase DNA repair pathways restrict LINE-1 retrotransposition in human cells. Nat Struct Mol Biol 2020;27:179–191. doi: 10.1038/s41594-020-0374-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Liang Y, Qu X, Shah NM, Wang T. Towards targeting transposable elements for cancer therapy. Nat Rev Cancer 2024;24:123–140. doi: 10.1038/s41568-023-00653-8. [DOI] [PubMed] [Google Scholar]
- 59.Rodriguez-Martin B Alvarez EG Baez-Ortega A Zamora J Supek F Demeulemeester J, et al. Pan-cancer analysis of whole genomes identifies driver rearrangements promoted by LINE-1 retrotransposition. Nat Genet 2020;52:306–319. doi: 10.1038/s41588-019-0562-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.IMSGC, Beecham AH Patsopoulos NA Xifara DK Davis MF Kemppinen A, et al. Analysis of immune-related loci identifies 48 new susceptibility variants for multiple sclerosis. Nat Genet 2013;45:1353–1360. doi: 10.1038/ng.2770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Gebrie A. Transposable elements as essential elements in the control of gene expression. Mob DNA 2023;14:9. doi: 10.1186/s13100-023-00297-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Fueyo R, Judd J, Feschotte C, Wysocka J. Roles of transposable elements in the regulation of mammalian transcription. Nat Rev Mol Cell Biol 2022;23:481–497. doi: 10.1038/s41580-022-00457-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Huo X Ji L Zhang Y Lv P Cao X Wang Q, et al. The Nuclear Matrix Protein SAFB Cooperates with Major Satellite RNAs to Stabilize Heterochromatin Architecture Partially through Phase Separation. Mol Cell 2020;77:368–383. e7. doi: 10.1016/j.molcel.2019.10.001. [DOI] [PubMed] [Google Scholar]
- 64.Lu JY Chang L Li T Wang T Yin Y Zhan G, et al. Homotypic clustering of L1 and B1/Alu repeats compartmentalizes the 3D genome. Cell Res 2021;31:613–630. doi: 10.1038/s41422-020-00466-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Zhang Y Li T Preissl S Amaral ML Grinstein JD Farah EN, et al. Transcriptionally active HERV-H retrotransposons demarcate topologically associating domains in human pluripotent stem cells. Nat Genet 2019;51:1380–1388. doi: 10.1038/s41588-019-0479-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Du AY, Chobirko JD, Zhuo X, Feschotte C, Wang T. Regulatory transposable elements in the encyclopedia of DNA elements. Nat Commun 2024;15:7594. doi: 10.1038/s41467-024-51921-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Pal D Patel M Boulet F Sundarraj J Grant OA Branco MR, et al. H4K16ac activates the transcription of transposable elements and contributes to their cis-regulatory function. Nat Struct Mol Biol 2023;30:935–947. doi: 10.1038/s41594-023-01016-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Meng S Liu X Zhu S Xie P Fang H Pan Q, et al. Young LINE-1 transposon 5′ UTRs marked by elongation factor ELL3 function as enhancers to regulate naïve pluripotency in embryonic stem cells. Nat Cell Biol 2023;25:1319–1331. doi: 10.1038/s41556-023-01211-y. [DOI] [PubMed] [Google Scholar]
- 69.Buttler CA, Ramirez D, Dowell RD, Chuong EB. An intronic LINE-1 regulates IFNAR1 expression in human immune cells. Mob DNA 2023;14:20. doi: 10.1186/s13100-023-00308-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Liang L Cao C Ji L Cai Z Wang D Ye R, et al. Complementary Alu sequences mediate enhancer–promoter selectivity. Nature 2023;619:868–875. doi: 10.1038/s41586-023-06323-x. [DOI] [PubMed] [Google Scholar]
- 71.Chuong EB, Elde NC, Feschotte C. Regulatory evolution of innate immunity through co-option of endogenous retroviruses. Science 2016;351:1083–1087. doi: 10.1126/science.aad5497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Deniz Ö, Ahmed M, Todd CD, Rio-Machin A, Dawson MA, Branco MR. Endogenous retroviruses are a source of enhancers with oncogenic potential in acute myeloid leukaemia. Nat Commun 2020;11:3506. doi: 10.1038/s41467-020-17206-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Zhou Z Zhu S Hong Y Jin G Ma R Lin F, et al. Composite transposons with bivalent histone marks function as RNA-dependent enhancers in cell fate regulation. Cell 2025:S0092-8674(25)00803-7. doi: 10.1016/j.cell.2025.07.014. Epub ahead of print. [DOI] [PubMed] [Google Scholar]
- 74.Percharde M Lin CJ Yin Y Guan J Peixoto GA Bulut-Karslioglu A, et al. A LINE1-Nucleolin Partnership Regulates Early Development and ESC Identity. Cell 2018;174:391–405 e19. doi: 10.1016/j.cell.2018.05.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Chen YG, Hur S. Cellular origins of dsRNA, their recognition and consequences. Nat Rev Mol Cell Biol 2022;23:286–301. doi: 10.1038/s41580-021-00430-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Tunbak H Enriquez-Gasca R Tie C Gould PA Mlcochova P Gupta RK, et al. The HUSH complex is a gatekeeper of type I interferon through epigenetic regulation of LINE-1s. Nat Commun 2020;11:5387. doi: 10.1038/s41467-020-19170-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Oomen ME Rodriguez-Terrones D Kurome M Zakhartchenko V Mottes L Simmet K, et al. An atlas of transcription initiation reveals regulatory principles of gene and transposable element expression in early mammalian development. Cell 2025;188:1156–1174.e20. doi: 10.1016/j.cell.2024.12.013. [DOI] [PubMed] [Google Scholar]
- 78.Denli AM Narvaiza I Kerman BE Pena M Benner C Marchetto MCN, et al. Primate-specific ORF0 contributes to retrotransposon-mediated diversity. Cell 2015;163:583–593. doi: 10.1016/j.cell.2015.09.025. [DOI] [PubMed] [Google Scholar]
- 79.Lee M, Jr, Ahmad SF, Xu J. Regulation and function of transposable elements in cancer genomes. Cell Mol Life Sci 2024;81:157. doi: 10.1007/s00018-024-05195-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Guo Y, Li TD, Modzelewski AJ, Siomi H. Retrotransposon renaissance in early embryos. Trend Genet 2024;40:39–51. doi: 10.1016/j.tig.2023.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Goke J Lu X Chan YS Ng HH Ly LH Sachs F, et al. Dynamic transcription of distinct classes of endogenous retroviral elements marks specific populations of early human embryonic cells. Cell Stem Cell 2015;16:135–141. doi: 10.1016/j.stem.2015.01.005. [DOI] [PubMed] [Google Scholar]
- 82.Jachowicz JW, Bing X, Pontabry J, Boskovic A, Rando OJ, Torres-Padilla ME. LINE-1 activation after fertilization regulates global chromatin accessibility in the early mouse embryo. Nat Genet 2017;49:1502–1510. doi: 10.1038/ng.3945. [DOI] [PubMed] [Google Scholar]
- 83.Zhang J Ataei L Mittal K Wu L Caldwell L Huynh L, et al. LINE1 and PRC2 control nucleolar organization and repression of the 8C state in human ESCs. Dev Cell 2025;60:186–203.e13. doi: 10.1016/j.devcel.2024.09.024. [DOI] [PubMed] [Google Scholar]
- 84.Frost JM Amante SM Okae H Jones EM Ashley B Lewis RM, et al. Regulation of human trophoblast gene expression by endogenous retroviruses. Nat Struct Mol Biol 2023;30:527–538. doi: 10.1038/s41594-023-00960-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Grow EJ Flynn RA Chavez SL Bayless NL Wossidlo M Wesche DJ, et al. Intrinsic retroviral reactivation in human preimplantation embryos and pluripotent cells. Nature 2015;522:221–225. doi: 10.1038/nature14308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Chuong EB, Rumi MAK, Soares MJ, Baker JC. Endogenous retroviruses function as species-specific enhancer elements in the placenta. Nat Genet 2013;45:325–329. doi: 10.1038/ng.2553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.C.West R, Ezashi T, B.Schoolcraft W, Yuan Y. Beyond fusion: A novel role for ERVW-1 in trophoblast proliferation and type I interferon receptor expression. Placenta 2022;126:150–159. doi: 10.1016/j.placenta.2022.06.012. [DOI] [PubMed] [Google Scholar]
- 88.Modzelewski AJ Shao W Chen J Lee A Qi X Noon M, et al. A mouse-specific retrotransposon drives a conserved Cdk2ap1 isoform essential for development. Cell 2021;184:5541–5558.e22. doi: 10.1016/j.cell.2021.09.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Frank JA Singh M Cullen HB Kirou RA Benkaddour-Boumzaouad M Cortes JL, et al. Evolution and antiviral activity of a human protein of retroviral origin. Science 2022;378:422–428. doi: 10.1126/science.abq7871. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Choudhary M, Quaid K, Xing X, Schmidt H, Wang T. Widespread contribution of transposable elements to the rewiring of mammalian 3D genomes. Nat Commun 2023;14:634. doi: 10.1038/s41467-023-36364-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Yu M Hu X Pan Z Du C Jiang J Zheng W, et al. Endogenous retrovirus-derived enhancers confer the transcriptional regulation of human trophoblast syncytialization. Nucleic Acids Res 2023;51:4745–4759. doi: 10.1093/nar/gkad109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Bodea GO Botto JM Ferreiro ME Sanchez-Luque FJ de Los Rios Barreda J Rasmussen J, et al. LINE-1 retrotransposons contribute to mouse PV interneuron development. Nat Neurosci 2024;27:1274–1284. doi: 10.1038/s41593-024-01650-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Pascarella G Hon CC Hashimoto K Busch A Luginbühl J Parr C, et al. Recombination of repeat elements generates somatic complexity in human genomes. Cell 2022;185:3025–3040.e6. doi: 10.1016/j.cell.2022.06.032. [DOI] [PubMed] [Google Scholar]
- 94.Mangoni D Simi A Lau P Armaos A Ansaloni F Codino A, et al. LINE-1 regulates cortical development by acting as long non-coding RNAs. Nat Commun 2023;14:4974. doi: 10.1038/s41467-023-40743-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Larsen PA, Hunnicutt KE, Larsen RJ, Yoder AD, Saunders AM. Warning SINEs: Alu elements, evolution of the human brain, and the spectrum of neurological disease. Chromosome Res 2018;26:93–111. doi: 10.1007/s10577-018-9573-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Erwin JA, Marchetto MC, Gage FH. Mobile DNA elements in the generation of diversity and complexity in the brain. Nat Rev Neurosci 2014;15:497–506. doi: 10.1038/nrn3730. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Saeliw T Kanlayaprasit S Thongkorn S Songsritaya K Sanannam B Sae-Lee C, et al. Epigenetic Gene-Regulatory Loci in Alu Elements Associated with Autism Susceptibility in the Prefrontal Cortex of ASD. Int J Mol Sci 2023;24. doi: 10.3390/ijms24087518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Suarez NA, Macia A, Muotri AR. LINE-1 retrotransposons in healthy and diseased human brain. Dev Neurobiol 2018;78:434–455. doi: 10.1002/dneu.22567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Muotri AR Marchetto MCN Coufal NG Oefner R Yeo G Nakashima K, et al. L1 retrotransposition in neurons is modulated by MeCP2. Nature 2010;468:443–446. doi: 10.1038/nature09544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Takahashi T Stoiljkovic M Song E Gao X-B Yasumoto Y Kudo E, et al. LINE-1 activation in the cerebellum drives ataxia. Neuron 2022;110:3278–3287.e8. doi: 10.1016/j.neuron.2022.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Coufal NG Garcia-Perez JL Peng GE Marchetto MC Muotri AR Mu Y, et al. Ataxia telangiectasia mutated (ATM) modulates long interspersed element-1 (L1) retrotransposition in human neural stem cells. Proc Natl Acad Sci USA 2011;108:20382–20387. doi: 10.1073/pnas.1100273108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Hanna R, Flamier A, Barabino A, Bernier G. G-quadruplexes originating from evolutionary conserved L1 elements interfere with neuronal gene expression in Alzheimer’s disease. Nat Commun 2021;12:1828. doi: 10.1038/s41467-021-22129-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Polesskaya O Kananykhina E Roy-Engel AM Nazarenko O Kulemzina I Baranova A, et al. The role of Alu-derived RNAs in Alzheimer’s and other neurodegenerative conditions. Med Hypotheses 2018;115:29–34. doi: 10.1016/j.mehy.2018.03.008. [DOI] [PubMed] [Google Scholar]
- 104.Ochoa E Ramirez P Gonzalez E De Mange J Ray WJ Bieniek KF, et al. Pathogenic tau–induced transposable element–derived dsRNA drives neuroinflammation. Sci Adv 2023;9:eabq5423. doi: 10.1126/sciadv.abq5423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Steiner JP Bachani M Malik N DeMarino C Li W Sampson K, et al. Human Endogenous Retrovirus K Envelope in Spinal Fluid of Amyotrophic Lateral Sclerosis Is Toxic. Ann Neurol 2022;92:545–561. doi: 10.1002/ana.26452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Thomas CA Tejwani L Trujillo CA Negraes PD Herai RH Mesci P, et al. Modeling of TREX1-Dependent Autoimmune Disease using Human Stem Cells Highlights L1 Accumulation as a Source of Neuroinflammation. Cell Stem Cell 2017;21:319–331.e8. doi: 10.1016/j.stem.2017.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Pasquesi GIM Allen H Ivancevic A Barbachano-Guerrero A Joyner O Guo K, et al. Regulation of human interferon signaling by transposon exonization. Cell 2024;187:7621–7636.e19. doi: 10.1016/j.cell.2024.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Kassiotis G, Stoye JP. Immune responses to endogenous retroelements: taking the bad with the good. Nat Rev Immunol 2016;16:207–219. doi: 10.1038/nri.2016.27. [DOI] [PubMed] [Google Scholar]
- 109.De Cecco M Ito T Petrashen AP Elias AE Skvir NJ Criscione SW, et al. L1 drives IFN in senescent cells and promotes age-associated inflammation. Nature 2019;566:73–78. doi: 10.1038/s41586-018-0784-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Lima-Junior DS Krishnamurthy SR Bouladoux N Collins N Han S-J Chen EY, et al. Endogenous retroviruses promote homeostatic and inflammatory responses to the microbiota. Cell 2021;184:3794–3811.e19. doi: 10.1016/j.cell.2021.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Dembny P Newman AG Singh M Hinz M Szczepek M Krüger C, et al. Human endogenous retrovirus HERV-K(HML-2) RNA causes neurodegeneration through Toll-like receptors. JCI Insight 2020;5:e131093. doi: 10.1172/jci.insight.131093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Chi H, Pepper M, Thomas PG. Principles and therapeutic applications of adaptive immunity. Cell 2024;187:2052–2078. doi: 10.1016/j.cell.2024.03.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Marasca F Sinha S Vadalà R Polimeni B Ranzani V Paraboschi EM, et al. LINE1 are spliced in non-canonical transcript variants to regulate T cell quiescence and exhaustion. Nat Genet 2022;54:180–193. doi: 10.1038/s41588-021-00989-7. [DOI] [PubMed] [Google Scholar]
- 114.Costa ALD Prieto-Oliveira P Duarte-Barbosa M Andreata-Santos R Peter CM Prolo de Brito T, et al. The Relationship between HERV, Interleukin, and Transcription Factor Expression in ZIKV Infected versus Uninfected Trophoblastic Cells. Cells 2024;13:1491. doi: 10.3390/cells13171491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Shah NM Jang HJ Liang Y Maeng JH Tzeng S-C Wu A, et al. Pan-cancer analysis identifies tumor-specific antigens derived from transposable elements. Nat Genet 2023;55:631–639. doi: 10.1038/s41588-023-01349-3. [DOI] [PubMed] [Google Scholar]
- 116.Dendrou CA, Fugger L, Friese MA. Immunopathology of multiple sclerosis. Nat Rev Immunol 2015;15:545–558. doi: 10.1038/nri3871. [DOI] [PubMed] [Google Scholar]
- 117.Treger RS, Pope SD, Kong Y, Tokuyama M, Taura M, Iwasaki A. The Lupus Susceptibility Locus Sgp3 Encodes the Suppressor of Endogenous Retrovirus Expression SNERV. Immunity 2019;50:334–347.e9. doi: 10.1016/j.immuni.2018.12.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Ko EJ, Cha HJ. The Roles of Human Endogenous Retroviruses (HERVs) in Inflammation. Kosin Med J 2021;36:69–78. doi: 10.7180/kmj.2021.36.2.69. [Google Scholar]
- 119.Macchietto MG, Langlois RA, Shen SS. Virus-induced transposable element expression up-regulation in human and mouse host cells. Life Sci Alliance 2020;3:e201900536. doi: 10.26508/lsa.201900536. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Yushkova E, Moskalev A. Transposable elements and their role in aging. Ageing Res Rev 2023;86:101881. doi: 10.1016/j.arr.2023.101881. [DOI] [PubMed] [Google Scholar]
- 121.Gorbunova V Seluanov A Mita P McKerrow W Fenyö D Boeke JD, et al. The role of retrotransposable elements in ageing and age-associated diseases. Nature 2021;596:43–53. doi: 10.1038/s41586-021-03542-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Zhao B Wu Q Ye AY Guo J Zheng X Yang X, et al. Somatic LINE-1 retrotransposition in cortical neurons and non-brain tissues of Rett patients and healthy individuals. PLoS Genet 2019;15:e1008043. doi: 10.1371/journal.pgen.1008043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Liu X Liu Z Wu Z Ren J Fan Y Sun L, et al. Resurrection of endogenous retroviruses during aging reinforces senescence. Cell 2023;186:287–304.e26. doi: 10.1016/j.cell.2022.12.017. [DOI] [PubMed] [Google Scholar]
- 124.Simon M Van Meter M Ablaeva J Ke Z Gonzalez RS Taguchi T, et al. LINE1 Derepression in Aged Wild-Type and SIRT6-Deficient Mice Drives Inflammation. Cell Metabolism 2019;29:871–885.e5. doi: 10.1016/j.cmet.2019.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Jang HS Shah NM Du AY Dailey ZZ Pehrsson EC Godoy PM, et al. Transposable elements drive widespread expression of oncogenes in human cancers. Nat Genet 2019;51:611–617. doi: 10.1038/s41588-019-0373-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Burns KH. Transposable elements in cancer. Nat Rev Cancer 2017;17:415–424. doi: 10.1038/nrc.2017.35. [DOI] [PubMed] [Google Scholar]
- 127.Gu Z Liu Y Zhang Y Cao H Lyu J Wang X, et al. Silencing of LINE-1 retrotransposons is a selective dependency of myeloid leukemia. Nat Genet 2021;53:672–682. doi: 10.1038/s41588-021-00829-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Szymanski JJ, Chauhan PS, Chaudhuri AA. LINE-1 Retrotransposons as Cell-free DNA Biomarkers for Multicancer Early Detection. Clin Cancer Res 2025;31:1179–1181. doi: 10.1158/1078-0432.CCR-24-4051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Jang HJ Shah NM Maeng JH Liang Y Basri NL Ge J, et al. Epigenetic therapy potentiates transposable element transcription to create tumor-enriched antigens in glioblastoma cells. Nat Genet 2024;56:1903–1913. doi: 10.1038/s41588-024-01880-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Taylor MS Wu C Fridy PC Zhang SJ Senussi Y Wolters JC, et al. Ultrasensitive detection of circulating LINE-1 ORF1p as a specific multi-cancer biomarker. Cancer Discovery 2023;13:2532–2547. doi: 10.1158/2159-8290.CD-23-0313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Attermann AS, Bjerregaard A-M, Saini SK, Grønbæk K, Hadrup SR. Human endogenous retroviruses and their implication for immunotherapeutics of cancer. Ann Oncology 2018;29:2183–2191. doi: 10.1093/annonc/mdy413. [DOI] [PubMed] [Google Scholar]
- 132.Tang Z, Li L, Shen L, Shen X, Ju S, Cong H. Diagnostic Value of Serum Concentration and Integrity of Circulating Cell-Free DNA in Breast Cancer: A Comparative Study With CEA and CA15-3. Lab Med 2018;49:323–328. doi: 10.1093/labmed/lmy019. [DOI] [PubMed] [Google Scholar]
- 133.Cheng J Cuk K Heil J Golatta M Schott S Sohn C, et al. Cell-free circulating DNA integrity is an independent predictor of impending breast cancer recurrence. Oncotarget 2017;8:54537–54547. doi: 10.18632/oncotarget.17384. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Zhang W Huang C Yao H Yang S Jiapaer Z Song J, et al. Retrotransposon: an insight into neurological disorders from perspectives of neurodevelopment and aging. Transl Neurodegener 2025;14:14. doi: 10.1186/s40035-025-00471-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Yan Z, Zhou Z, Wu Q, Chen ZB, Koo EH, Zhong S. Presymptomatic Increase of an Extracellular RNA in Blood Plasma Associates with the Development of Alzheimer’s Disease. Curr Biol 2020;30:1771–1782.e3. doi: 10.1016/j.cub.2020.02.084. [DOI] [PubMed] [Google Scholar]
- 136.Tam OH, Ostrow LW, Gale Hammell M. Diseases of the nERVous system: retrotransposon activity in neurodegenerative disease. Mob DNA 2019;10:32. doi: 10.1186/s13100-019-0176-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Jiang Q Braun DA Clauser KR Ramesh V Shirole NH Duke-Cohan JE, et al. HIF regulates multiple translated endogenous retroviruses: Implications for cancer immunotherapy. Cell 2025;188:1807–1827.e34. doi: 10.1016/j.cell.2025.01.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Krug B De Jay N Harutyunyan AS Deshmukh S Marchione DM Guilhamon P, et al. Pervasive H3K27 Acetylation Leads to ERV Expression and a Therapeutic Vulnerability in H3K27M Gliomas. Cancer Cell 2019;35:782–797.e8. doi: 10.1016/j.ccell.2019.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Griffin GK Wu J Iracheta-Vellve A Patti JC Hsu J Davis T, et al. Epigenetic silencing by SETDB1 suppresses tumour intrinsic immunogenicity. Nature 2021;595:309–314. doi: 10.1038/s41586-021-03520-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Zhang S-M Cai WL Liu X Thakral D Luo J Chan LH, et al. KDM5B promotes immune evasion by recruiting SETDB1 to silence retroelements. Nature 2021;598:682–687. doi: 10.1038/s41586-021-03994-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Sheng W LaFleur MW Nguyen TH Chen S Chakravarthy A Conway JR, et al. LSD1 Ablation Stimulates Anti-tumor Immunity and Enables Checkpoint Blockade. Cell 2018;174:549–563.e19. doi: 10.1016/j.cell.2018.05.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Lanciano S, Cristofari G. Cancer Immunotherapy: How to Exploit Transposable Elements? Clin Chem 2024;70:17–20. doi: 10.1093/clinchem/hvad091. [DOI] [PubMed] [Google Scholar]
- 143.Sun S You E Hong J Hoyos D Del Priore I Tsanov KM, et al. Cancer cells restrict immunogenicity of retrotransposon expression via distinct mechanisms. Immunity 2024;57:2879–2894.e11. doi: 10.1016/j.immuni.2024.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Zhang R Zhang F Sun Z Liu P Zhang X Ye Y, et al. LINE-1 Retrotransposition Promotes the Development and Progression of Lung Squamous Cell Carcinoma by Disrupting the Tumor-Suppressor Gene FGGY. Cancer Res 2019;79:4453–4465. doi: 10.1158/0008-5472.CAN-19-0076. [DOI] [PubMed] [Google Scholar]
- 145.Della Valle F Reddy P Yamamoto M Liu P Saera-Vila A Bensaddek D, et al. LINE-1 RNA causes heterochromatin erosion and is a target for amelioration of senescent phenotypes in progeroid syndromes. Sci Transl Med 2022;14:eabl6057. doi: 10.1126/scitranslmed.abl6057. [DOI] [PubMed] [Google Scholar]
- 146.Sullivan AC Zuniga G Ramirez P Fernandez R Wang C-P Li J, et al. A Phase IIa clinical trial to evaluate the effects of anti-retroviral therapy in Alzheimer’s disease (ART-AD). NPJ Dement 2025;1:2. doi: 10.1038/s44400-024-00001-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Rajurkar M Parikh AR Solovyov A You E Kulkarni AS Chu C, et al. Reverse Transcriptase Inhibition Disrupts Repeat Element Life Cycle in Colorectal Cancer. Cancer Discovery 2022;12:1462–1481. doi: 10.1158/2159-8290.CD-21-1117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Leslie M. Risk moves. Science 2025;388:244–247. doi: 10.1126/science.ady2285. [DOI] [PubMed] [Google Scholar]
- 149.Wang C Shi Z Huang Q Liu R Su D Chang L, et al. Single-cell analysis of isoform switching and transposable element expression during preimplantation embryonic development. PLoS Biol 2024;22:e3002505. doi: 10.1371/journal.pbio.3002505. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Wang R Zheng Y Zhang Z Song K Wu E Zhu X, et al. MATES: a deep learning-based model for locus-specific quantification of transposable elements in single cell. Nat Commun 2024;15:8798. doi: 10.1038/s41467-024-53114-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Lanciano S, Cristofari G. Measuring and interpreting transposable element expression. Nat Rev Genet 2020;21:721–736. doi: 10.1038/s41576-020-0251-y. [DOI] [PubMed] [Google Scholar]
- 152.Bilgrav Saether K, Eisfeldt J. Detecting transposable elements in long-read genomes using sTELLeR. Bioinformatics 2024;40:btae686. doi: 10.1093/bioinformatics/btae686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Ewing AD Smits N Sanchez-Luque FJ Faivre J Brennan PM Richardson SR, et al. Nanopore Sequencing Enables Comprehensive Transposable Element Epigenomic Profiling. Mol Cell 2020;80:915–928.e5. doi: 10.1016/j.molcel.2020.10.024. [DOI] [PubMed] [Google Scholar]
- 154.Xie N, Shen G, Gao W, Huang Z, Huang C, Fu L. Neoantigens: promising targets for cancer therapy. Signal Transduct Target Ther 2023;8:9. doi: 10.1038/s41392-022-01270-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Rieder D Fotakis G Ausserhofer M René G Paster W Trajanoski Z, et al. nextNEOpi: a comprehensive pipeline for computational neoantigen prediction. Bioinformatics 2022;38:1131–1132. doi: 10.1093/bioinformatics/btab759. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Fotakis G, Rieder D, Haider M, Trajanoski Z, Finotello F. NeoFuse: predicting fusion neoantigens from RNA sequencing data. Bioinformatics 2020;36:2260–2261. doi: 10.1093/bioinformatics/btz879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Zhang T Tan S Tang N Li Y Zhang C Sun J, et al. Heterologous survey of 130 DNA transposons in human cells highlights their functional divergence and expands the genome engineering toolbox. Cell 2024;187:3741–3760.e30. doi: 10.1016/j.cell.2024.05.007. [DOI] [PubMed] [Google Scholar]