Skip to main content
Molecular Biology and Evolution logoLink to Molecular Biology and Evolution
. 2024 Feb 2;41(3):msae025. doi: 10.1093/molbev/msae025

The Dynamic Interplay Between Ribosomal DNA and Transposable Elements: A Perspective From Genomics and Cytogenetics

Sònia Garcia 1,, Ales Kovarik 2, Sophie Maiwald 3, Ludwig Mann 4, Nicola Schmidt 5, Joan Pere Pascual-Díaz 6, Daniel Vitales 7,8, Beatrice Weber 9, Tony Heitkam 10,11,
Editor: Miriam Barlow
PMCID: PMC10946416  PMID: 38306580

Abstract

Although both are salient features of genomes, at first glance ribosomal DNAs and transposable elements are genetic elements with not much in common: whereas ribosomal DNAs are mainly viewed as housekeeping genes that uphold all prime genome functions, transposable elements are generally portrayed as selfish and disruptive. These opposing characteristics are also mirrored in other attributes: organization in tandem (ribosomal DNAs) versus organization in a dispersed manner (transposable elements); evolution in a concerted manner (ribosomal DNAs) versus evolution by diversification (transposable elements); and activity that prolongs genomic stability (ribosomal DNAs) versus activity that shortens it (transposable elements). Re-visiting relevant instances in which ribosomal DNA–transposable element interactions have been reported, we note that both repeat types share at least four structural and functional hallmarks: (1) they are repetitive DNAs that shape genomes in evolutionary timescales, (2) they exchange structural motifs and can enter co-evolution processes, (3) they are tightly controlled genomic stress sensors playing key roles in senescence/aging, and (4) they share common epigenetic marks such as DNA methylation and histone modification. Here, we give an overview of the structural, functional, and evolutionary characteristics of both ribosomal DNAs and transposable elements, discuss their roles and interactions, and highlight trends and future directions as we move forward in understanding ribosomal DNA–transposable element associations.

Keywords: repetitive DNA, ribosomal DNA, transposable elements, concerted evolution, homogenization, transposition, recombination, housekeeping genes, genome stability, genome size, molecular cytogenetics, long-read sequencing

Ribosomal DNAs and Transposable Elements: Two Opposing Faces of Repetitive DNAs

Ribosomal DNAs: Conserved and Heavily Transcribed, These Housekeepers Are In Charge of Cell Maintenance

Ribosomal RNA genes (rDNAs) play fundamental key roles in cellular processes. They are repetitive and encode the RNA components of ribosomes, the most ancient and complex of all molecular machines (Moss et al. 2006). Despite rDNAs not representing a large proportion of an organism's genome size, they produce around 80–90% of the RNAs found in most cells (Eaves et al. 2020) and ribosomal RNAs (rRNAs) accounting for 60% of the ribosomal mass (O’Connor and Adams 2010). Ribosomal DNA repeats are crucial players in maintaining genome stability (Kobayashi 2006) and any disturbance at rDNA loci may have a great impact on cellular processes, including the response to DNA damage and overall cell longevity (Ganley and Kobayashi 2014). Because of their universality, sequence conservation, and usually high copy numbers, rDNAs and their spacers have been widely used to resolve evolutionary relationships among organisms (Nieto-Feliner and Rosselló 2007) and deployed as molecular markers for breeding purposes or addressing hybridization processes (Garcia et al. 2020, 2023). Although highly conserved due to their housekeeping nature, there are some differences between prokaryotic and eukaryotic rRNAs (Fig. 1). Whereas prokaryotic ribosomes contain three rRNAs, the 16S, 23S, and 5S rRNAs, eukaryotic ribosomes carry four: the corresponding 18S, 26S (or 28S, depending on the organism group), the 5S rRNAs, and additionally, the eukaryote-specific 5.8S rRNA. Prokaryotes typically encode all rRNAs in a single operon, an arrangement thought to facilitate gene regulation. Nevertheless, in some bacteria and archaea, the 16S and 23S rRNA genes can occur in an unlinked manner (Brewer et al. 2020). In eukaryotes, 18S, 5.8S, and 26S/28S rRNA genes are usually encoded in a single operon, called 35S in plants and 45S in animals (Hemleben and Zentgraf 1994; Hemleben et al. 2021). Although it is generally assumed that the 18S–5.8S–26/28S operon spreads as a whole (Bueno et al. 2013), independent mobility of each of these genes has been occasionally observed, e.g. in fish (Symonová et al. 2013) and in grasshoppers (Ferretti et al. 2019). The fourth gene, the 5S rRNA, is the only one ubiquitous to prokaryotes and eukaryotes; it can be either linked (Garcia et al. 2009) to the other rRNA genes or separated from these, located in other genomic regions. Much less frequently, the 5S rRNA gene can also be linked to other multigene families, such as histone genes or the trans-spliced leader, or even, very occasionally, can be dispersed across the genome (Drouin and de Sá 1995).

Fig. 1.

Fig. 1.

Structure of a prokaryotic (left) and a eukaryotic (right) ribosome. The different rRNA components of both ribosome types are highlighted, focusing on prokaryotic (blue, left side of the ribosome), as well as the eukaryotic (violet, right side of the ribosome) variants. For the 26S/28S rDNA, the 26S variant is found in plants and the 28S variant in animals. The indicated rRNAs are encoded in the tandemly arranged rRNA genes.

Another distinction between prokaryotic and eukaryotic rDNAs is their copy number, ranging from a few operons in Escherichia coli (e.g. seven in strain K12) (Maeda et al. 2015) to hundreds to thousands of copies in plants and animals (Ingle et al. 1975; Prokopowich et al. 2003). Moreover, rDNA copy number is dynamic and can experience large interspecific and even intraspecific variation. Variation that can also be driven by environmental changes (Gibbons et al. 2014; Lavirinenko et al. 2021). Copy number of each of the rDNAs may be influenced by the need to maintain a balanced amount of 5S and 45S arrays. In this regard, Gibbons et al. (2015) demonstrated that 5S and 45S rDNA arrays of human and mouse exhibit a tightly coupled variation in copy numbers (a concerted copy number variation, cCNV), despite their location in different chromosomes. Although the exact mechanism(s) on how cCNV of rRNA genes is coordinated is unclear, the discovery of cCNV appears as a new way to achieve gene and genome balance (Malone 2015). Besides, rRNA transcription is also precisely regulated, as demonstrated by Condon et al. (1993) in E. coli, where a depletion of the number of functional rDNA copies lead to an increased expression of the remaining ones.

As other tandemly arranged multigene families, rDNAs generally evolve in concert. Concerted evolution is one of the molecular evolution models put forward for multigene families that keeps sequence integrity across all gene copies (Brown et al. 1972). Other options have been proposed to explain the evolution of rDNAs in certain cases such as the birth-and-death model (e.g. Pinhal et al. 2011; Zhang et al. 2021). In this model, gene variants arise by gene duplication with some staying for a long time in the genome whereas others become pseudogenized and ultimately deleted (Nei and Hughes 1992; Nei and Rooney 2005). Nevertheless, diversity in non-transcribed spacer (NTS) regions of rDNAs within and between species is common (Coen et al. 1982; Williams et al. 1990). Also, in the rRNA genes, intragenomic diversity has been found in species as distant as yeast (Sultanov and Hochwagen 2022) and human (Fan et al. 2022). Despite this, the concerted model explains the evolution of rRNA genes better than all other models proposed so far. A recent review, which addresses the intragenomic rDNA variation across a wide range of organisms, highlights that rDNA evolution is complex and still a subject of debate, even more than 50 years after the concerted evolution model for multigene families evolution was first proposed (Wang et al. 2023).

The rDNAs represent heavily transcribed units along the chromosomes. These actively transcribed 35S/45S rDNA loci constitute the nucleolus organizer regions (NORs), the site where ribosome biogenesis takes place. The nucleolus is the most visible component of the interphase nucleus, and its physical relationship with the rDNA locus was first recognized by Barbara McClintock in maize (1934). But the rDNA's presumed stability in chromosomal locations is untrue in some cases: already 40 years ago, mobile NORs were observed in Allium species (Schubert 1984; Schubert and Wolbus 1985). The mobile NOR hypothesis is based on the variable sizes, numbers, or chromosomal positions of rDNA loci after silver staining and in situ hybridization of rRNA genes. Over the last years, this observation was also shown for other taxa (e.g. Pedrosa-Harand et al. 2006; Schmidt et al. 2019). Mobility of rDNAs is a sporadic event and likely results from recombination rather than being a transposable element (TE)-mediated process. Yet, in the context of this review, it is interesting to note that rDNAs can sometimes be mobile.

Despite the long history of research into rDNAs, they still remain elusive (Hall et al. 2022). Due to their repetitiveness, they are usually absent from most of todays genome assemblies, and, despite the rDNA's significance for cellular maintenance, we still know relatively little about rDNA evolution, copy number preservation, and impact on genome integrity.

TEs: Hyperdiverse and Mostly Repressed, These Evolutionary Drivers Bring Genomic Novelty

In sharp contrast with rDNAs, TEs are dispersed genomic repeats, often termed “mobile DNAs” or “jumping genes”. These sequences can change their position in the genome or generate copies of themselves in a process termed (retro-) transposition. Depending on the presence of an mRNA intermediate, TEs are classified as Class I - retrotransposons and Class II - DNA transposons (Finnegan 1989; Wells and Feschotte 2020). As such, TEs are the embodiment of the mobile genome and serve as agents of fast genomic change.

TEs are ubiquitous. They have been detected across eukaryotes and prokaryotes including animals, plants, fungi, and bacteria. Unlike rRNA genes, making up only a tiny fraction of the genomes, TEs can account for up to 80–90% in some cases (Schnable et al. 2009) and they exist in a huge diversity, coming in many sizes, shapes, and proportions. Among TEs, long and short interspersed nuclear elements (LINEs and SINEs) are, by far, the most abundant in mammalian genomes (Platt et al. 2018), whereas long terminal repeat (LTR) retrotransposons dominate plant genomes (Gao et al. 2016).

Regarding the position of TEs in eukaryotic genomes, TEs are considered to be dispersed along the chromosomes. Despite this generalist view, there are many instances of TEs with non-random distribution: for example, Arabidopsis thaliana centromeres are enriched in certain Athila retrotransposons (Naish et al. 2021) and fruit fly telomeres are made up of HeT-A and TART (telomere-associated) retrotransposons (Silva-Sousa et al. 2012). It is now widely accepted that the observed patterns result from the interplay of TE insertion and TE removal mechanisms. To understand these complex TE integration dynamics, many TE mutagenesis and comparative genomics studies have been performed in the recent years (reviewed, for example, in Sultana et al. 2017).

Historically, it took not only the discovery of TEs in maize, by McClintock (1950), but additional years to accept that genomes “were not static, stable, and immobile” (as later highlighted by Kazazian 2011). Even then, TEs were mainly recognized as “junk” or “parasitic DNAs” contaminating the genome (Orgel and Crick 1980). In our current times, the perceived relevance and impact of TEs are again shifting. There is a growing conviction that TE disruption could indeed serve as evolutionary driver. Thus, TEs provide a reservoir of genetic and epigenetic variability, with some TEs even having adaptive potential (Chuong et al. 2017; Schrader and Schmitz 2019; Almeida et al. 2022). Despite the vast majority of studied TEs being deleterious, the advance of genomics brings forward more cases in which TEs take over new, unsuspected, and sometimes beneficial roles, such as gene master regulators, evolutionary drivers, and structural genomic components, among others.

In terms of molecular evolution, TEs also differ significantly from rDNAs. Since they are mostly free of selection pressures, they are hyperdiverse, and their mode of evolution is usually explained by the neutral theory (Kimura 1968; Arkhipova 2018). With neutrality, it is usually understood that insertional TE mutagenesis is mostly neutral or slightly deleterious (Arkhipova 2018). Of course, with the advance of genetics and genomics, more cases come to light where TE insertions cause disruption, but sometimes also produce phenotypic and/or regulatory variability. Two of the most well-known examples are Mendel's wrinkled peas and the industrial melanism of peppered moths (Bhattacharyya et al. 1990; van’t Hof et al. 2016). Regarding TE evolution within the host, we outline the typical life cycle of a TE according to the birth-and-death model, the currently favored model explaining TE evolution in the host (Blumenstiel 2019):

  1. Birth and initial amplification: typically, point mutations or modular reshuffling can lead to enough sequence variation that a TE arises, which is not yet silenced by the epigenetic machinery, but still contains all necessary components for activation. This is usually followed by an initial amplification (“burst of amplification”) that increases the copy number of the new variant. Along these lines, a TE family is typically considered active or mobilizable, if its members (at least one member) are capable of producing TE copies under favorable circumstances. Depending on the TE type, this often involves harboring continuous, undisrupted open reading frames and intact promoters. The hallmarks of an active TE family are much more difficult to define, if non-autonomous TE families are concerned, which often do not encode any protein domains.

  2. Silencing: After amplification, the host's epigenetic silencing machinery will silence the TEs. Depending on the genomic circumstances, re-activation is possible, if silencing was released.

  3. Decay and death: Over evolutionary timescales, by accumulation of mutations, the TE's potential to be mobilized will decrease, until the TE is either decayed or deleted, e.g. by recombination (Devos et al. 2002). Of course, the process of mutation can also lead to the emergence of new TE variants, thus starting anew the TE life cycle.

rDNA and TEs: Bridging the Differences and Coming Together?

Summarizing, while rDNAs make up a small but essential genomic fraction, TEs make up the largest but mostly dispensable part of the genome, but are essentially disregarded in their functions; while the rDNA's organization is in tandem, localized in specific chromosomal loci, TEs are typically dispersed across genomes; while at least some rDNA loci are always active, TEs are silenced in their majority; while rDNAs tend to homogenize their sequences, diversification is the rule for TEs; and while rDNAs follow preferentially a concerted mode of evolution (Wang et al. 2023), the evolution of TEs may better fit the birth-and-death model.

Yet, both rDNA and TEs also have several things in common: (1) while TEs and their derivatives are certainly major contributors to genome size, recent evidence suggests that rDNA by-products (such as pseudogenized rDNA copies and/or fragments) can also contribute to the “junk DNA” accumulating over evolutionary time (Robicheau et al. 2017). (2) In terms of mobility and chromosomal positioning, TEs occasionally jump into ribosomal DNA loci and can integrate more-or-less stably in some rDNA arrays; upon re-activation, TEs may carry rDNA fragments (Pérez-González and Eickbush 2001) and distribute them across the genome in their path (see “Ribosomal DNA in TEs” of this review). Hence, TEs have the potential to structurally embed rDNA sequence units, to mobilize rDNA copies, and to diffuse the rDNA's restriction to distinct chromosomal sites (although such processes may also be explained by recombination). (3) Regarding activity, TE-mediated silencing can sometimes spread to silence rDNA chromatin. Hence, some TE insertions have been shown to prevent rDNA transcription (Long and Dawid 1979; Jamrich and Miller 1984; Fefelova et al. 2022). Interestingly, if TEs are transcriptionally activated, rDNA can follow (Fefelova et al. 2022). Similarly, TEs embedded in the rDNA or using rDNA promoters (see “TEs in Ribosomal DNA” and “Ribosomal DNA in TEs” of this review) may circumvent silencing and be actively transcribed.

Evolutionarily speaking, the mechanistic differences between rDNA and TEs are sometimes overridden: occasionally, similarly to TEs, rDNA copies emerge and decay according to the birth-and-death model (see above). Complimentarily, in a few instances, TEs may also evolve by homogenization, reminiscent of rDNAs: first, they may act as entry points for illegitimate recombination processes, leading to copy number expansion or contraction (Devos et al. 2002). Second, TEs may also be homogenized across the genome by non-allelic gene conversion events, thereby spreading mutations from one copy to another (Kejnovsky et al. 2007; Ellison and Bachtrog 2015; Fawcett and Innan 2019). Third, occasionally, TEs can form tandemly repeated structures undergoing homogenization (Paço et al. 2019; Maiwald et al. 2021). Nevertheless, TE-driven homogenization is usually not considered a dominant force: instead, recent large sequencing works show TEs as the titular antagonists that disrupt ongoing homogenization processes (Naish et al. 2021).

Here, we address the complex relationship between rDNAs and TEs (Fig. 2), two of the most salient figures of genomes, by analyzing the evolution of methodological approaches to reveal rDNAs, TEs, and their interactions and also by reviewing relevant instances in which an interaction between both has been described. We also aim to find connections between both, aiming to understand how their interactions may contribute to their mutual evolution and genomic distribution and how this can provide material for genomic innovation. We finally address their molecular evolution, mobility, and trajectory across the cellular lifespan in these interactions, given that both rDNAs and TEs play major roles in genome evolution, chromosomal stability, and gene regulation.

Fig. 2.

Fig. 2.

Association of rDNAs with TEs. A) rDNA unit (monomer) composition follows a strict blueprint as they always contain the genetic information for the actual ribosomal genes (highlighted in gray), which are separated by spacer regions (left, red: 5S rDNA, right, purple: 35S/45S rDNA). Spacers are named according to their positions. As the 5S rDNA spacer is usually not transcribed, it is called NTS. The 35S/45S rDNA monomer is composed of the 5′ ETS, a several kb-long coding region, a 3′ ETS, and an IGS. The spacing units in between coding regions are called ITS. For 5S rDNA, we also observe an internal control region, with highly conserved promoter box motifs (A-Box, intermediate element, and C-Box) for Polymerase III transcription. The 35S/45S monomer consists of multiple rDNA genes for 18S, 5.8S, and 26S/28S rDNA. B) rDNA monomers appear in large numbers and form large arrays with a tandem arrangement. Insertions of TEs are highlighted in blue. C) Usually, the 35S/45S rDNA resides distally on the chromosomes, whereas the 5S rDNA occupies interstitial chromosomal loci. D) Some TEs are associated with these rDNA sequences and/or locations. This has been described for various types of TEs, such as the LTR retrotransposon lineage Galadriel and the LINEs R1 and R2 that target the 26S/28S rDNA; as well as many non-autonomous TEs (including SINEs, Cassandra TRIMs, and DNA transposons) that harbor rDNA-derived sequences such as the 5S promoter.

Methods to Detect rDNA, TEs, and Their Associations

Method-wise, due to their repetitiveness, the most striking rDNA and TE similarity is their absence from genome assemblies. TEs and rDNAs usually clutter up the “chromosome 0” or “random chromosome”, where the unassembled bits and pieces are compiled. Nevertheless, during the past 50 yr, many methods were brought forward that have the potential to identify rDNA–TE associations.

In the 1970s, reannealing kinetics studies by Flavell et al. (1974) already established a clear relationship between genome size and the proportion of repetitive DNAs—a positive and usually significant correlation. In other words, most genome size variation is due to the variation in repeated DNA. This still holds true today, with caveats being polyploidy-related genome growth and the accumulation of sequence remnants of formerly repetitive DNAs in mega-genomes (Novák et al. 2020). The rise of molecular biology methods in the 1980s led to the identification of the most abundant repeats by Southern blot hybridization of genomic DNA with radioactively labeled probes, usually with short exposition times. The first cloning and sequencing experiments revealed all kinds of repeats, among them were ribosomal DNAs and fragments of the most abundant TEs. Already in the late 1970s, the presence of insertion sequences (a kind of short, simple TE) in ribosomal DNA was detected by Northern blot hybridization in E. coli (Nisen and Shapiro 1979). Insertions into many rDNA monomers of Bombyx mori, detected by restriction analysis (Lecanidou et al. 1984), were later identified as R1/R2 elements (Xiong and Eickbush 1988). Likewise, PCR, Southern blot, and similarity searches suggested that proximal rDNA-flanking sequences may consist of retrotransposons in rice (Fujisawa et al. 2006).

Cytological methods have also been a classical tool to show rDNA and TE (co)localization along chromosomes, usually focusing on TE integrations in rDNA (see “Ribosomal DNA in TEs” of this review). One of the first techniques to be developed was silver staining of interphase nuclei and chromosomes: silver nitrate reacts strongly with proteins and allows for visualization of actively transcribed rDNA loci (Bloom and Goodpasture 1976; Blum et al. 1987); however, a real surge in understanding arose when molecular methods were more regularly combined with cytology in the 1990s. This coincided with the spread of cytogenetic and fluorescent in situ hybridization (FISH) techniques. Due to the universal conservation of the ribosomal genes, rDNAs were not only among the first repetitive sequences to be identified for most organisms but also the first repeat probes to be hybridized. As even rDNAs from distantly related organisms can be used as FISH probes for many species, they are still among the most used probes for karyotyping (Heitkam and Garcia 2023). More than 2,000 papers have been published reporting the number and position of rDNA loci in chromosomes of plants and animals, and this information has been organized and summarized in www.plantrdnadatabase.com (version 4.0 February 2023; Rodríguez-González et al. 2023) and in www.animalrdnadatabase.com (Sochorová et al. 2018), respectively. TEs, however, are far more diverse and require identification and characterization experiments prior to cytogenetic use. Hence, their adoption for cytogenetics has been lagging. In most experiments, TEs were mapped dispersedly along the chromosomes, with some notable exceptions: for example, TEs of the Chromovirus lineage frequently reside in specific chromosomal positions: CRM-type chromoviruses are often centromeric (Neumann et al. 2011; Weber et al. 2013), whereas Galadriel-type chromoviruses are often detected within 35S rDNA copies (Balint-Kurti et al. 2000; Weber et al. 2013). These kinds of rDNA–TE associations can be visualized using two-colour FISH onto mitotic, meiotic, and interphase nuclei as well as onto stretched DNA fibers (Fig. 3A; also see “TEs in Ribosomal DNA” for further biological details on this specific rDNA–TE association).

Fig. 3.

Fig. 3.

Methods to identify rDNA–TE associations. A) Using cytogenetics, an association of rDNAs and TEs becomes visible by co-localization of the respective signals. In this example, sequences corresponding to the Galadriel-type retrotransposon (red) and the 18S rDNA (green) are hybridized to sugar beet nuclei. Along the mitotic metaphase chromosomes, hybridization of Galadriel clearly shows two major sites covering the typical rDNA regions on chromosome 1 of sugar beet (left). At higher resolution, on interphase nuclei, clear co-hybridization of the rDNA and the TE probe indicate their association (middle). Benefitting from the higher resolution of meiotic pachytene chromosomes, clear interspersed and overlapping (yellow) signals support the co-localization of the 35S rDNA and the Galadriel retrotransposon in sugar beet (right). Reprinted from Weber et al. (2013), under CC-BY. B) To visualize the rDNA–TE association at the sequence level, error-prone genome assemblies are not suitable. Instead, long reads can be screened for co-occurrence of the rDNA and a specific TE on the same locus. Here, the association already indicated in panel (A), between the Galadriel retrotransposon and the 18S rDNA, was used. Two self-dotplots of sugar beet PacBio long reads were produced with FlexiDot (Seibt et al. 2018) and shaded to indicate the integration of the retrotransposon (pastel tones) into the rDNA (green tones). In the left instance, a canonical 18S–5.8S–26S rDNA locus is shown, in which the 18S rRNA gene is interrupted by a Galadriel retrotransposon. To the right, two rDNA repeats are arranged in an inverted orientation. Both are interrupted by Galadriel in the 18S rRNA gene. C) The 35S rDNA cluster graphs contain information about the locus: here, three RepeatExplorer-derived graphs from sugar beet are shown. On the right, a smooth, uninterrupted graph indicates that many uninterrupted, homogenized 18S–5.8S–26S rDNA monomers exist. In the middle, a disorganized, fuzzy graph with some similarities to TEs indicates TE disruption and rearrangement. To the right, the 18S–5.8S–26S rDNA locus produced a single, circular cluster. The association with TEs (in this case with the Galadriel retrotransposon also indicated in panels A and B) becomes visible as a branch emerging from the 18S rDNA region. D) An association of the rDNA and TEs often becomes visible after graphical read clustering. A shared RepeatExplorer graph indicates the sequence similarity between the 5S rDNA and retrotransposons of the Cassandra type in two species (left: Tragopogon porrifolius; middle: Senecio vulgaris). A more complex graph (right: Musa acuminata) indicates multiple 5S rDNA variants and sequence similarities with a Ty1/copia retrotransposon. Reprinted from Garcia et al. (2020), under CC-BY.

The genomics era brought a new dimension to our understanding of the associations between rDNA and TEs. Conserved and tandemly arranged genes such as rDNAs are straightforward to identify along newly generated sequence data. In contrast, the high diversity of TEs complicates their identification and requires a range of diverse, often not integrated, tools (TE Hub Consortium et al. 2021). Due to the collapsing of repetitive regions in genome assemblies, most insights are gained from assembly-free approaches, using both short and long reads. The high sequencing throughput coupled with fast software development have now revealed rDNA and TE features and their possible interactions, which only a few years ago would have been difficult to imagine. Long reads are especially ideal to gain insights into the fine organization of rDNA and TEs, including their genomic environments. Automatized dot plotting tools like Flexidot (Seibt et al. 2018) allow the visualization of their respective positions. For example, self-dotplots of the Galadriel-type chromovirus mentioned above reveal the embedment into the 18S gene of the rDNA of sugar beet (Fig. 3B). Other tools, such as the computational pipeline RepeatExplorer (Novak et al. 2010, 2020), allow to identify and characterize repetitive DNAs in next-generation sequencing data, using graph-based clustering of sequence reads to identify repetitive elements. The resulting repeat clusters portray, for example, the homogenization of the 35S rDNA locus by smooth, uninterrupted 18S–5.8S–26S rDNA graphs (Fig. 3C, left), and may also show TE-disrupted loci as fuzzy or branched graphs (Fig. 3C, middle and right). Sequence similarities between rDNA and TEs (see “Future Perspectives” of this review) can also be revealed, as for example by shared graphs between the 5S rDNA and the 5S promoter-containing Cassandra TEs (Fig. 3D; Garcia et al. 2020). For the future, as genomics becomes more accessible for research on all organisms, we predict that more sequence-based methods to resolve rDNA–TE associations will be developed. Nevertheless, due to the repetitive nature of both, we recommend corroboration by molecular biology and/or cytogenetics to exclude being misled by sequence artifacts.

TEs in Ribosomal DNA

Many studies have reported the presence of TEs or TE fragments within rRNA genes, in their spacers or their close vicinity; a summary of the most relevant cases in animals, plants, fungi, protozoans, and bacteria can be consulted in Table 1. While in the earlier works the precise localization of TEs was not possible, with cytogenetic and genomic technological advances newer studies exactly positioned these TEs in the rDNA. Some TEs occur exclusively within rDNA arrays, whereas others are more widespread and dispersed with only some copies in the rDNA. The latter seems to be more prominent and here we lay out some plant-specific specific examples: in sugar beet, some TEs such as the LINE BNR1 (Heitkam and Schmidt 2009) and the endogenous pararetrovirus beetEPRV3 (Schmidt et al. 2021) have some copies in the rDNA, but many more in other genomic regions. This observation is not limited to beets, but represents a general phenomenon (e.g. Raskina et al. 2008). In contrast, specific TEs targeting or accumulating in the rDNA have also been detected, such as Ty3-gypsy LTR retrotransposons of the Galadriel clade; one example is the Monkey element, found in the banana genome (Balint-Kurti et al. 2000; Hríbova et al. 2010), which has been recently confirmed on long reads (Eva Hríbova, personal communication). Galadriel-type TEs were also found in some of the 18S genes in sugar beet and other Beta species (Fig. 3A–C; Weber et al. 2013). Nevertheless, the sugar beet genome contains about 1,200 potentially functional 18S rDNA copies lacking TE insertions (Fig. 3C). Similarly, full-length Galadriel elements have been found integrated into the 18S gene in tomato [also called tomato rDNA-related retrotransposons (TRRTs); Jo et al. 2009], which was later sorted into the Galadriel lineage (Weber et al. 2013). This observation was later repeated in the mango genome, suggesting that this could be a common phenomenon (Nusrat Sultana and Tony Heitkam, personal communication). The chromodomains in the Galadriel lineage likely allow targeting of the open chromatin, of the rDNA, to allow integration of Galadriel-type TEs (Novikova 2009; Weber et al. 2013).

Table 1.

Evidence of TE found in rDNA sequences of animals, plants, fungi, protozoans, and bacteria

Organismal group Organismal subgroup Organism rDNA integration TE type TE superfamily TE order/lineage TE name Type of evidence Reference
Animals Arachnids Hosarius adansoni Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Ascidians Ciona intestinalis, C. savignyi Into 28S rRNA gene Retrotransposon Non-LTR LINE R2, R4 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Birds Coturnix japonica Not defined Not defined Not defined Not defined Not defined Chimeric rDNA–TE amplicons after PCR Saifitdinova et al. (2019)
Collembola Anurida maritima Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 PCR, cloning and sequencing Burke et al. (1998)
Chelicerata Limulus polyphemus Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 PCR, cloning and sequencing Burke et al. (1998)
Crustaceans Daphnia obtusa Into 28S rRNA gene DNA transposon piggyBac Full DNA transposons and MITEs PokeyA and PokeyB Quantitative PCR (qPCR) to estimate rRNA gene and Pokey number in isolates from natural populations of D. obtusa, and in clonally propagated mutation accumulation lines LeRiche et al. (2014)
Daphnia pulex Into 28S rRNA gene DNA transposon piggyBac Full DNA transposons and MITEs Pokey Subcloning and sequencing Penton et al. (2002)
D. pulex Into 28S rRNA gene DNA transposon piggyBac Full DNA transposons and MITEs Pokey PCR, cloning, and sequencing upstream of Pokey insertions Glass et al. (2008)
D. pulex Into 28S rRNA gene DNA transposon piggyBac Full DNA transposons and MITEs Pokey and mPok Computational analysis of the genome assembly Elliott et al. (2013)
D. pulex, D. pulicaria Into 28S rRNA gene DNA transposon piggyBac Full DNA transposons and MITEs Pokey qPCR to estimate the number of 18S and 28S ribosomal RNA genes and Pokey elements in rDNA (rPokey), as well as other genomic locations (gPokey) Eagle and Crease (2012)
Daphnia species Into 28S rRNA gene DNA transposon piggyBac Full DNA transposons and MITEs Pokey PCR and sequencing of the Pokey integration site using a Poley and a 28S rDNA primer Penton and Crease (2004)
Procambarus clarkii Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Triops longicaudatus Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Echinoderms Metacrinus rotundus Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Fish Coregonus albula Into 45S Retrotransposon Non-LTR LINE Rex1 Co-localization observed after FISH Symonová et al. (2013)
Danio rerio Into 5S Retrotransposon Non-LTR LINE Mutsu RNAseq, Northern blotting, qPCR Locati et al. (2017)
D. rerio Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Diplodus sargus Into the spacer of the 5S rDNA Several retrotransposons LTR and non-LTR Ty3-gypsy and LINE Not defined Cloning, sequencing, co-localization observed after FISH Merlo et al. (2013)
Eptatretus burgeri Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Erythrinus erythrinus into 5S Retrotransposon Non-LTR LINE Rex3 Co-localization observed after FISH Cioffi et al. (2010)
Gymnotus mamiraua Into the spacer of the 5S rDNA DNA transposon Tc1/mariner Not defined Not defined Comparative cytogenetics da Silva et al. (2016)
Oreochromis niloticus Into 28S rRNA gene Retrotransposon Non-LTR LINE R4 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
O. niloticus and others related Not defined with precision Retrotransposon LTR and non-LTR SINE, Copia SINE2-1AFC, SINE_FR2, Copia-53_MLp-1 Sequencing data Nakajima et al. (2012)
Oryzias lotipes Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005), Kuroki-Kami et al. (2019)
Rineloricaria latirostris Into the spacer of the 5S rDNA DNA transposon hAT Not defined Not defined PCR, cloning, sequencing; co-localization in FISH Glukosky et al. (2018)
Tanichthys albonubes Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Triportheus species Into the spacer of the 5S rDNA Different TEs, linkage with U1 snRNA multigene family Not defined Not defined Not defined PCR, cloning, sequencing; co-localization in FISH Yano et al. (2020)
Insects Anopheles gambiae Not defined Retrotransposon Non-LTR Not defined Not defined Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Bombyx mori Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 Genomic screening, DNA hybridization, subcloning and sequencing Fujiwara et al. (1984), Burke et al. (1987, 1993), Xiong and Eickbush (1988)
Drosophila melanogaster Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 Cloning, sequencing, genomic DNA blot analysis Jakubczak et al. (1990, 1992)
D. melanogaster Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 rDNA transgenics, RNA-FISH, ChIP, RNA-seq Luo et al. (2020)
D. simulans Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 RNA analysis by RT-PCR (Reverse Transcription PCR), Northern blot of different crosses Eickbush et al. (2008)
Diadromus pulchellus, Eupelmus vuilleti, Apis mellifera Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 Cloning, sequencing, Southern blot hybridization Bigot et al. (1992)
Forficula auricularia Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 PCR, cloning and sequencing Burke et al. (1998)
Hepialus humuli Into the 28S–18S spacer Retrotransposon LTR Gypsy Hh Ty3/gypsyA Sequencing, RepeatExplorer analysis, FISH Daliková et al. (2023)
Inachis io Into the 28S–18S spacer Retrotransposon Non-LTR LINE liR2 Sequencing, RepeatExplorer analysis, FISH Daliková et al. (2023)
Nasonia vitripennis Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 Genomic screening, degenerate PCR, cloning, sequencing Burke et al. (1993)
Popillia japonica Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 Genomic screening, degenerate PCR, cloning, sequencing Burke et al. (1993)
Sciara coprophila Into 28S rRNA gene Retrotransposon Non-LTR LINE R1, R2 Genomic screening, degenerate PCR, cloning, sequencing Renkwatz-Pohl et al. (1981), Kerrebrock et al. (1989)
Tribolium castaneum Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 PCR of 5′ junctions Eickbush et al. (2013)
Nematodes Ascaris lumbricoides Into the 28S rRNA gene Retrotransposon LTR Ty1-copia Ty1 Sequencing of rDNA region, Southern and northern blot hybridization Neuhaus et al. (1987)
Strongyloides rattii Into 28S rRNA gene Retrotransposon Non-LTR LINE R4 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Mammals Primates (human, chimpanzee, gorilla, orangutan, gibbon, rhesus monkey) Into IGS: fragment C (upstream rDNA promoter) Retrotransposon Non-LTR SINE Alu Sequencing of rDNA region Gonzalez et al. (1993)
Mollusc Biomphalaria glabrata Into 28S rRNA gene Retrotransposon Non-LTR LINE R4 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Reptiles Chinemys reevesii Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Degenerate PCR, cloning and sequencing Kojima and Fujiwara (2005)
Trematodes Schistosoma mansoni Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Plants Eudicots Beta vulgaris Into 18S rRNA gene Retrotransposon LTR Ty3-gypsy, chromovirus, Galadriel-type Beon Detected in genome assembly, co-localization of 35S rDNA and TE by FISH on mitotic and meiotic chromosomes and fibers Weber et al. (2013)
Armeria macrophylla, A. pungens Into 26S–18S rRNA spacer (IGS) Transposon and retrotransposon fragments LTR and DNA transposons Sabrina and EnSpm (CACTA) PCR, cloning and sequencing Nieto-Feliner et al. (2019)
Cucumis melo, C. sativus Into ITS and ETS rDNA regions Retrotransposon Non-LTR LINE Menolird18 Co-localization of 35S rDNA and TE by FISH Setiawan et al. (2020)
Solanum lycopersicum Into 18S gene Retrotransposon LTR Ty3-gypsy, chromovirus, Galadriel-type TRRTs Co-localization of 35S rDNA and TE by fiber FISH, BAC sequencing and sequencing assembly of the rDNA units Jo et al. (2009)
Monocots Allium cernuum Between 26S and 18S rRNA genes Retrotransposon LTR Ty1-copia Not defined Detected during IGS sequencing, confirmed by comparative Southern hybridization and co-localization of 35S rDNA and TE by FISH Chester et al. (2010)
Musa acuminata Not defined Retrotransposon LTR Ty3-gypsy, chromovirus, Galadriel-type Monkey Subcloning and sequencing, co-localization of 35S rDNA and TE by FISH on mitotic chromosomes Balint-Kurti et al. (2000)
Oryza sativa Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Zea mays Into 28S rRNA gene Retrotransposon Non-LTR LINE R2 Cross genome screening, blast searches in genomic databases Kojima and Fujiwara (2004)
Fungi Ascomycota Ascobolus immersus In the ETS, separating the 26S and the 18S rRNA genes Retrotransposon LTR Not defined Hideaway Subcloning and sequencing, comparative Sothern hybridization Kempken (2001)
Saccharomyces cerevisiae Between the 3′ ends of the 37S and 5S rRNAs Retrotransposon Non-LTR Ty1-copia Ty1 Subcloning and sequencing, comparative Sothern hybridization Vincent and Petes (1986)
Microsporidia Nosema bombycis Particular rDNA unit: LSU-ITS-SSU-5S; integration in LSU (large subunit), ITS and SSU (small subunit) DNA transposon Not defined further MITE Four families: LSUME1, ITSME1, SSUME1, SSUME2 Whole genome sequencing (WGS) and genome-wide survey using LSU rRNA, SSU rRNA and 5S rRNA as queries; Northern blotting Liu et al. (2013)
Protozoans Euglenozoa Trypanosoma brucei Not defined Not defined Not defined Not defined Not defined Ordered restriction map (BgI II fragments) Hassan et al. (1992)
Protists Amoebozoa Physarium polycephalum Within the coding region of the 26S rRNA gene Not defined Not defined Not defined Strain-specific group I intron Southern blot hybridization, restriction analysis Muscarella and Vogt (1989)
Bacteria Corynebacterium aurimucosum, Thermus scotoductius Into 16S rRNA gene Not defined Not defined Not defined Not defined Analysis of genomic sequences searching for imperfect rRNA genes Lim et al. (2012)
Escherichia coli Between the 16S and 23S rRNA genes DNA transposon Not defined Not defined Tn10 Lambda-specialized transducing phage, analysis of expression of tRNA and rRNA coded from genes on the phages Morgan (1980)
E. coli 16S rRNA and 23S rRNA Simple TEs Not defined Not defined Insertion sequences IS1 and IS2 Northern blot hybridization of probes of IS1 and IS2 to E. coli RNA Nisen and Shapiro (1979)
Mycoplasma imitans In 16S–23S rRNA IGS DNA transposon? Not defined Not defined A transposase gene from an undetermined element Size comparison of PCR products of the ITS region between M. imitans and M. gallisepticum Harasawa et al. (2004)

In animals, there are many examples of TEs in the 28S rDNA: especially the piggyBac-type DNA transposon Pokey has colonized the rDNA of several Daphnia species (Penton et al. 2002; Penton and Crease 2004; Glass et al. 2008; Eagle and Crease 2012; Elliott et al. 2013). There are also several reports of R1, R2, and R4 elements, classified as LINEs, integrating into the 28S rDNA. These LINEs are nearly ubiquitous in arthropods, such as Drosophila melanogaster (Jakubczak et al. 1992; Luo et al. 2020), B. mori (Xiong and Eickbush 1988), some parasitoid wasps and honeybees (Bigot et al. 1992), and also in beetles (Eickbush et al. 2013), reptiles, echinoderms, arachnids, crustaceans, and fish, among others (Kojima and Fujiwara 2005); see Table 1 for more examples. Insertions of these R elements often occur in non-essential regions of the 28S gene, thus favoring the conservation of insertion sites (Kojima and Fujiwara 2005). Eickbush et al. (2013) hypothesized that R2 elements lacked their own promoters, relying on the rDNA transcription machinery for their expression. Considering that rDNA arrays are prone to copy number loss due to their repetitiveness, Nelson et al. (2023) recently proposed the challenging idea that R2 would indeed be essential for maintaining rDNA copy numbers (see “Future Perspectives”), defying the notion that retrotransposons are solely self-serving elements.

As for the intergenic spacer (IGS) of the 35S/45S rDNA, there are also several works reporting TE integration; indeed, the IGS is highly variable and repetitive, and hence, prone to TE insertion polymorphisms. Among the most relevant TEs integrating in the IGS, we highlight the Alu SINEs, found in primates, including humans (González et al. 1993), and the LTR retrotransposon Hideaway detected in the fungus Ascobolus immersus (Kempken 2001). In Nosema bombycis, members of four miniature inverted-repeat transposable element (MITE) families were detected in the genic and intergenic rDNA regions, with the rDNA remaining fully functional (Liu et al. 2013). In some lepidopterans (here: Inachis io), the recent work by Daliková et al. (2023) reported the association of the IGS with an R2 element and a satellite DNA. Also, in the same organism group (here: Hepialus humuli), a non-functional Ty3/gypsy retrotransposon was detected at the very end of the rDNA unit [at the junction between the 28S rDNA and the external transcribed spacer (ETS)]. The authors concluded that mobile elements would have hardly contributed to mediating the spread of rDNA, while conversely, the satellite DNA arrays found in the IGS could have promoted the homology-mediated spread of rDNA through ectopic recombination or by integrating extrachromosomal circular rDNA. As for plants, integration of TEs in the IGS has been described for Allium cernuum, where a Ty1/copia retrotransposon, which occurs in and outside of the rDNA arrays (Chester et al. 2010). Also, in species from the genus Cucumis, LINEs were detected in the internal transcribed spacers (ITS) and ETS rDNA regions (Setiawan et al. 2020). As the mentioned examples encompass animals, fungi, and plants, we conclude that TE integration in the IGS is a widespread phenomenon.

While, to our knowledge, no TE insertion has been reported within the 5S gene, TE insertions in the adjacent NTS are relatively frequent. In Allium schoenoprasum, for example, the insertion of a TE into the 5S rDNA spacer has led to the emergence of a second 5S rDNA family. Both 5S rDNA variants now co-exist and differ in length of the spacer (Shibata and Hizume 2002). In other plants, e.g. for the grains Aegilops speltoides and Hordeum spontaneum, En/Spm-like transposons also reside in the 5S rDNA NTS (Altinkut et al. 2006); in several banana species, Ty1/copia-like TE fragments also constitute a part of the 5S rDNA NTS (Hribova et al. 2010 and Fig. 3D). As for animals, we can only report examples for fish: among the TEs in the 5S spacer, DNA transposons, LINEs, SINEs, and Ty3/gypsy elements are the most commonly found. Remarkably, TE-flanking rRNA genes in cichlid fish have been regarded as a source of rRNA gene movement, not only for the 5S rRNA gene but also for 18S rRNA genes—generally, TE movements seem to play a large role in the generation of cichlid fish diversity (Carleton et al. 2020).

Finally, several early studies describe close relationships between a TEs and rDNAs, but without defining specifically the insertion sites or the kind of element. Examples are an unspecified mobile element in the rRNA genes of Trypanosoma brucei (Hassan et al. 1992), a Ty1/copia element in the rDNA of yeast (Vincent and Petes 1986), Ty1/copia intervening sequences in the rDNA of the nematode Ascaris lumbricoides (Neuhaus et al. 1987), or even large variations in bacterial rDNAs attributed to recent TE insertions into the 16S rRNA genes (Lim et al. 2012). Summarizing all incidents (Table 1), we conclude that TEs in rDNA occur in all major species groups—animals, fungi, plants, and even bacteria. But how do rDNAs retain the integrated TEs, are rDNAs preferred by TEs, and is an rDNA–TE association solely deleterious or can there sometimes be benefits?

Are rDNAs TE Insertion Hotspots?

Based on the observations of plant genomes, some researchers assert that the infrequent occurrences of mobile elements in rDNA point toward a rare targeting of these genes by TEs (Chester et al. 2010). Others claim that rDNA is indeed a TE insertion hotspot; in other words, TEs insert with a higher frequency in rDNA than in other genomic regions (Nieto-Feliner et al. 2019; Bendich and Rogers 2023). But which is more likely the case? First, there may be a study intensity bias, as rDNA is one of the most studied genomic loci in molecular biology. Hence, the many reported cases of TEs in rDNA may be a result of the many research efforts spent on the rDNA. Second, as most genome assemblies are far from being complete and as the rDNA regions are difficult to assemble, we cannot yet measure universal TE integration rates in the rDNA. Nevertheless, some TE-specific rates are already available: Perez-Gonzalez and Eickbush (2001) provided an estimate of R1/R2 retrotransposition into rDNA and found it to be similar to other TEs. Hence, the rate of R1/R2 transposition into rDNA could be as similar as the rate in other genomic locations. In contrast, Redd et al. (2023) introduced the rice-derived TE mPING into yeast and found preferable mPING insertion into yeast rDNA.

Going back to the examples of R1/R2 LINEs, we discuss the impact of TEs in the rDNA on rRNA functionality. In arthropod genomes, the rRNA genes lose the capability to synthesize functional 28S rRNA genes after the invasion of rDNA by these TEs (Long and Dawid 1979; Jamrich and Miller 1984; Eickbush and Eickbush 2003). However, despite their deleterious effect, R1/R2 insertions have been maintained in arthropods by vertical transmission since the origin of this lineage (Lathe et al. 1995; Burke et al. 1998; Malik and Eickbush 1999). This does not affect the host viability as intact rDNA copies are still present—the levels of R1 or R2 insertion can vary from a small percentage to over 70% in Drosophila (Malik and Eickbush 1999). Interestingly, in arthropods these elements exhibit conserved insertion sites (Bigot et al. 1992; Zhang and Eickbush 2005; Eickbush and Eickbush 2012; Eickbush et al. 2013). Their precise location in the rRNA genes becomes relevant for the TE after transcription: it allows folding of the RNA in such a way that the TE-derived part is autocatalytically cut, retained, and then can enter the retrotransposition process (Oyun et al. 2018). Precise integration sites are also documented for other TEs, including LTR retrotransposons and DNA transposons (Penton et al. 2002; Jo et al. 2009). As with R1/R2 insertions, the viability of the host genome is not compromised in these cases because of the high number of (still functional) rDNA copies.

Two scenarios that are not mutually exclusive may explain these multiple precise locations: (1) the rDNA–TE associations may arise from the homogenization of a single TE insertion across the whole rDNA array, due to unequal and illegitimate recombination. As a result, many rDNA monomers with identical TE insertions would arise from a single TE integration event. (2) The rDNA–TE associations may also arise from a targeting mechanism of the TE, e.g. facilitated by a chromodomain (Gao et al. 2008; Abascal-Palacios et al. 2021). If a certain rDNA locus is targeted by a TE family with an insertion preference, multiple times, we would also observe identical TE insertions across the rDNA monomers (Jo et al. 2009; Weber et al. 2013).

Do TE Insertions in rDNA Simply Occupy a Niche for Their Own Profit or May They Even Be Beneficial to the Host?

Before discussing potential benefits of TEs being integrated into rDNA, it should be noted that there is still too little information to completely ascertain whether rDNA–TE associations are coincidental or may indeed provide a benefit onto which selection can act upon. Most TEs can also integrate into other locations (not just into the rDNA), hence the impact on the host fitness may be an important aspect to evaluate if an rDNA–TE association will be sustained across generations. That said, as rRNA genes are multicopy, actively transcribed and highly conserved, it is clear that a TE insertion in rDNA ensures its transcription and propagation. Hence, the rDNA can be considered a “ideal niche” for TE insertion. This niche would provide a safe haven for TEs, from which they propagate (Penton et al. 2002). In turn, as a repetitive multigene family, there are many more copies of rDNA than those that are essential. Thus, rDNA-integrated TEs can exist up to a certain threshold without important phenotypic effects which may hinder the host fitness (Eickbush and Eickbush 1995; Malik and Eickbush 1999), as intact rDNA copies would still be available. Besides, TE insertions into conserved, but non-essential regions, ensure the integrity and functionality of rDNA genes of the respective monomer. This preference likely allows the TE to “survive” (and also thrive) in rDNA loci, as strong selection against disrupted and inactive rDNA units would eliminate them from the genome. However, there is also a danger for TEs in integrating into the rDNA. Due to rDNA's homogenization processes (concerted evolution), TE insertions in rDNA may also be lost more quickly than in other genomic regions. Hence, the rDNAs capacity for quick homogenization is a double-edged sword for TEs. It allows quick spread and also quick removal, as proposed earlier by Perez-González et al. (2003).

Are there also benefits for rDNA regulation of the host brought by TEs? First, TEs may serve as a vehicle to transpose rDNA from one genomic location to the other. This may be advantageous in some karyotypes, which suffer from frequent chromosome rearrangements. Second, transposition can inactivate supernumerary rDNA copies, which may be helpful to better regulate a gene dosage, e.g. after polyploidization or horizontal transfer. A close association of inactive heavily mutated rDNA units with TEs has been reported in plant allopolyploids (Handa et al. 2018; Tulpová et al. 2022) and in alien rDNA following horizontal transfer (Mahelka et al. 2017). In these instances, TEs may accumulate in those rDNA loci (units) which are already inactive and have lost homogenization capacity. Beneficial roles of TEs may also be indirect. For example, in Drosophila, the rDNA-specific R2 endonuclease introduces double-strand breaks within the rDNA, which immediately serve as a starting point for rDNA homogenization, maintenance, and copy number upholding (see above). Similarly, a regulatory function of TEs (McClintock 1951) that allows the neighboring rDNA chromatin to benefit from the epigenetic machinery of TEs cannot be excluded.

Considering all the points above, it becomes clear that only TEs in conserved but non-essential rDNA sites stand a chance of long-term survival, proliferation, and spread in the rDNA. This leads to few rDNA landing sites that may have a chance of becoming preferred TE hotspots. All these aspects point to tightly linked co-evolution between TEs and rDNA, in which the TEs evolve to select the most appropriate regions to insert into, whereas the rDNA's homogenization capacity removes the potentially harmful insertions from the genome. Hence, we speculate that more complete sequencing and assembly of rDNA loci may bring to light more TE integration hotspots at distinct sites within the rDNA monomers. Whether the TE ratio in rDNA outshines other iconic chromosomal regions, such as the telomeres and centromeres, remains to be seen (and is likely organism-specific).

Ribosomal DNA in TEs

Several studies have detected ribosomal DNA fragments in TEs (Table 2). In general, TEs have a tendency to acquire sequence modules and protein domains by reshuffling (e.g. Jiang et al. 2004; Seibt et al. 2020). In some of these cases, TEs have likely co-opted and domesticated certain domains (reviewed by Cosby et al. 2019; Wang and Han 2021). The most common rDNA genes or fragments found associated with TEs are 5S ribosomal DNA sequences. TEs containing 5S sequences usually keep, more or less intact, the internal 5S RNA promoter for Pol III for the TE's proliferation benefits. Escaping from the boundaries of Pol II transcription (typical for TEs and genes) and instead opting for the Pol III promoter enables the TE to unlock a completely different proliferation strategy, at different time points and likely in different tissues as opposed to the typical Pol II-transcribed TEs. In the first release of SINEBase, a database of short interspersed elements (SINEs) (Vassetzky and Kramerov 2013), 5S-derived SINEs accounted for 2.3% of its content and to date, they are known in reptiles, insects, fish, and mammals.

Table 2.

Evidence of rDNA-derived TE sequences

Organismal group Organismal subgroup Organism Similarity to which rDNA TE type TE superfamily TE name Reference
Animals Crustaceans Daphnia pulex Region of the IGS in internal region DNA transposon/MITE piggyBac Pokey and mPok Elliot et al. (2013)
Fish Salmonoidei 5S rDNA promoter Retrotransposon SINE OS Matveev and Okada (2009)
Danio rerio 5S rDNA promoter Retrotransposon SINE SINE3/AmnSINE1 Kapitonov and Jurka (2003)
Insects Several species 5S rDNA promoter Retrotransposon SINE HaSE3 Wang et al. (2012)
Mammals Several species 5S rDNA promoter Retrotransposon SINE OS Nishihara et al. (2006)
Megachiroptera 5S rDNA promoter Retrotransposon SINE MEG-RL, MEG-RS, MEG-TR Gogolevsky et al. (2009)
Pedetes capensis 5S rDNA promoter Retrotransposon SINE Ped-1 Gogolevsky et al. (2009)
Homo sapiens 28S rDNA Retrotransposon SINE SINE28 Longo et al. (2015)
Nemerteans (ribbon worms) rDNA: 5S rDNA promoter Retrotransposon LTR/TRIM Ajax Kojima (2024)
Reptiles Anolis carolinensis 5S rDNA promoter Retrotransposon SINE 5S-Sauria Piskurek et al. (2009)
Plants Angiosperms-eudicots Amaranthaceae 5S rDNA promoter Retrotransposon LTR/TRIM Ama_Cassandra Maiwald et al. (2021)
Asteraceae 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Maiwald et al. (2024)
Several species 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Gao et al. (2016)
Several species 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Kalendar et al. (2008)
Rosaceae 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Yin et al. (2014)
Several species 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Kalendar et al. (2020)
Brassicaceae 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Sampath and Yang (2014)
Angiosperms-monocots Agavaceae 5S rDNA promoter Retrotransposon LTR/TRIM Cassandra Tamayo-Ordóñez et al. (2018)

To our knowledge, there are no 5S-derived SINEs in plants. Instead, plants harbor non-autonomous LTR retrotransposons with 5S rDNA promoter sequences in their LTRs, named Cassandra (Kalendar et al. 2008, 2020; Maiwald et al. 2021, 2024; a recent study also found analogous retrotranspons in ribbon worms, named Ajax (Kojima 2024)). Cassandras were first described in 50 plant species including ferns, monocots, and eudicots and later (re)confirmed in several angiosperm lineages (Yin et al 2014; Gao et al. 2016; Kalendar et al. 2020), including Arabidopsis (Sampath and Yang 2014), Agave (Tamayo-Ordóñez et al. 2018), and beets (Maiwald et al. 2021). At present, there are no reports of Cassandras in gymnosperms. Looking closely at the Asteraceae, the plant family with most variability in the 5S rDNA, including rDNA arrangements (Garcia et al. 2010) and promoter shifts (Garcia et al. 2012), Maiwald et al. (2024) found that Cassandra retrotransposons closely mimic the promoter motif changes, thus providing a recent example of rDNA–TE co-evolution.

In some cases, Cassandra TEs show a tandem-array pattern, reminiscent of cellular 5S (Kalendar et al. 2008; Maiwald et al. 2021). This arrangement has been used to explain the very variable distribution of 5S rDNA loci along plant chromosomes, in which Cassandra signals would have been confused with authentic 5S rDNA loci in FISH experiments. Indeed, cases with extraordinarily high numbers of 5S loci have been detected (for example, up to 71 loci in tulip genomes; Mizouchi et al. 2007) that could likely be Cassandra elements accounting for these possibly “fake” rDNA loci numbers. A similar situation might apply for the 22–38 5S rDNA loci for Alstroemeria (Kamstra et al. 1997; Baeza et al. 2007) or 10–38 for Paphiopedilum (Lan and Albert 2011), considering that the median for angiosperms is two 5S sites (one locus; Garcia et al. 2017). For the latter genus, the authors already proposed that TE activity may underlie the surprisingly high 5S loci number, i.e. some of the 5S signals may result from pseudogenes that were mobilized by TEs (Lan and Albert 2011). TEs also harbor other rRNA genes or fragments such as the mammalian SINE28 that contains 28S rDNA fragments (Longo et al. 2015). Similarly, Pokey TEs possess a series of 200 bp repeats upstream of the transposase domain, derived from the rDNA IGS (Elliott et al. 2013).

The many cases of rDNA-related acquisitions inside TEs are likely the result of modular reshuffling as has been often observed in TE evolution (e.g. Wollrab et al. 2012; Seibt et al. 2020), especially for non-autonomous elements. While we speculate that most rDNA-related sequence modules do not benefit the TE, some may make a difference toward the TE's evolutionary success—such as the 5S rDNA promoter. These successful acquisitions are seen several times across the tree of life, for instance in animals and plants, benefitting different TE families, including the iconic Cassandra retrotransposons.

Future Perspectives

We have shown that rDNAs and TEs are two important actors in genomes and have both similarities and differences in terms of their organization, function, evolution, and mobility. Interactions between both have been described and these can contribute to their mutual evolution and genomic distribution. This landscape of rDNAs and TEs in the genome is a mixed one, with the interplay between their organization, evolution, and function leading to a diverse range of outcomes.

What to Expect From Current and Next Technical Advancements

The advancements of the long-read technology have the potential to provide even more insights into the complex interplay between rDNA, TEs, and the genome as a whole. Long reads capture longer stretches of repetitive DNAs in a single read, hence reducing the need for assembly and improving the accuracy of analysis (Marx 2023). This approach has meant a huge step forward since the repetitiveness and complexity characterizing both rDNA and TE-rich regions were a big handicap for previous sequencing technologies based on much shorter reads. Long reads can reveal the precise and (if present) the higher-order organization of rRNA genes, as done by Havlová et al. (2016), Symonová et al. (2017) and Heitkam et al. (2020), and the respective positions of rDNA and TE (Fig. 3B). Recently, McKinlay et al. (2021) went one step further using a target enrichment method to enrich for rDNA loci in ultra-long Oxford Nanopore Technology sequencing reads, given that these loci usually represent a small proportion of the genome; a similar approach, but for TEs, was followed by Merkulov et al. (2023). Research in TEs may be perhaps the most affected field by the availability of long reads (Shahid and Slotkin 2020), given their intrinsic and enormous diversity, although advances in long-read sequencing may also be the key to finally and precisely assessing sequence, structure, and copy number variation of rDNA (Hall et al. 2022). Overall, the dropping cost of even the most sophisticated omics approaches and the new analytical tools that are continuously arising will surely also allow a more detailed examination of the relationships between rDNAs and TEs.

Apart from resolving repetitive loci by long-read sequencing, optical mapping technologies have the capability to provide high-resolution images of entire chromosomes including the highly repetitive ribosomal loci. Thus, optical mapping can inform about the structure and organization of ribosomal DNA and TEs, by allowing the visualization of their distribution and orientation (Tulpová et al. 2022). Genome editing based on CRISPR/Cas9 may also be used in the study of repeats such as rDNA and TEs, by enabling their targeted modification and maybe even deletion from distinct sites (Smith et al. 2020; Lopez et al. 2021). Changing the overall location of rDNA and TEs may be also possible in the near future using chromosome modification techniques (Schmidt et al. 2020). Summarizing, we predict that in the near future, genome sequencing and mapping, as well as editing technologies, will likely produce the largest gain of knowledge in unraveling the rDNA–TE associations.

Biological Roles and Consequences of the rDNA–TE Interplay

Beyond genome structure and evolution, research on ribosomal DNA and TEs also contributes to the understanding of many biological processes, such as rDNA maintenance, control of copy numbers, and even more complex processes such as aging and senescence. Regarding rDNA maintenance, a retrotransposon-induced mechanism was recently reported in Drosophila (Nelson et al. 2023) and perhaps other insects. It remains to be determined whether similar mechanisms operate in other groups, such as plants, whose genomes are inherently known to be thronged with TEs. If so, this would pave a way to studies of potential of beneficial effects of TEs on host fitness as a driving force behind their success across different organisms.

Similarly, new structural links between rDNAs and TEs are arising. Both components also can exist beyond the chromosomes, for example as extrachromosomal circular DNAs (eccDNAs; Flavell and Ish-Horowicz 1981; Pont et al. 1987; Cohen et al. 2003, 2008). These DNA rings mostly originate from recombinational processes and play roles in upholding rDNA copy number (Mansisidor et al. 2018). Regarding TEs, eccDNAs can be side-products (Lanciano et al. 2017; Mann et al. 2022; Peng et al. 2022) or even necessary intermediates for the mobilization of TEs (Yang et al. 2023). In the framework of this review, eccDNAs are interesting to explore, as both—some rDNAs and some TEs—may acquire mobility, if present in circular form. Further eccDNA study may likely bring to light more insight into rDNA–TE associations.

On a physiological level, rDNAs are associated with cellular aging, as the accumulation of rDNA mutations over time can affect cellular function and contribute to age-related decline (Sinclair and Guarente 1997; Kasselimi et al. 2022). Loss of rDNA repeats, possibly through rDNA circle formation and accumulation, is associated with lifespan in both yeast and humans (Goffová and Fajkus 2021). Despite its importance for maintaining all cellular functions, rDNAs become increasingly fragile with age, and are prone to copy number and DNA methylation changes (Watada et al. 2020). TEs also influence aging processes, as their epigenetic control becomes more unstable with increasing age, resulting in increasing TE transcript levels, gene regulatory change, and more transposition (Gorbunova et al. 2021; Yang et al. 2022; Yushkova and Moskalev 2023). Concluding, aging profoundly affects both rDNA and TEs. Meanwhile, also rDNA and TEs affect aging, on at least two levels: DNA methylation and regulatory control as well as mobility and copy number variation. Hence, with aging, these seemingly opposite genomic components can both serve as starting points of chromosomal instability, thus speeding up the cellular trajectory toward mortality. We expect that the next decade will provide molecular insights into the role of TEs and rDNAs in aging, and that regulatory control of both repetitive genome components likely plays a role in identifying targets for intervention and treatment of age-related diseases.

Conclusion

Despite the perceived disparity between rDNAs and TEs, both genome components share more than both being discovered in maize by Barbara McClintock. We outline the many examples, where TEs and rDNAs co-occur, interact, benefit, and even evolve together. We outline what to expect from the latest technical advances and tap into the shared biological roles of rDNAs and TEs. Starting out as seemingly antagonistic forces—with rDNAs as housekeepers upholding cell maintenance and TEs as silenced disruptive agents—rDNAs and TEs often cross paths. However, with passage through the cell's lifespan, their genomic effects converge, both leading toward genomic fragility. We conclude that the often-overlooked interplay of rDNAs and TEs is a major force not only in genome evolution but also in cellular maintenance, gene regulation, and chromosomal stability.

Acknowledgments

Open access funding enabled and organized by the Programa de Apoyo a la Publicación en Acceso Abierto para autores CSIC (PROA). This research was supported by the Agencia Estatal de Investigación, Government of Spain (PID2020-119163GB-I00 and PRE2021-097873, granted to S.G. and J.P.P.D., respectively, all funded by MCIN/AEI/10.13039/501100011033). T.H. receives grants from the Bundesministerium für Bildung und Forschung (BMBF) for the EpicBeet project (FKZ 031B1221A) and the German Research Foundation (DFG) for the project 433081887/HE7194/2-1. S.M. is funded by a Back-to-Research Grant financed by the Ministry of Research, Culture and Tourism of the Free State of Saxony (SMWK). Interactions between the Dresden, Graz, and Barcelona labs are enabled by an EMBO short-term fellowship (Ref. 8989) and a Dresden Senior Fellowship to S.G. The research of A.K. is supported by the Czech Science Foundation (grant no. 22-16826S).

Contributor Information

Sònia Garcia, Institut Botànic de Barcelona (IBB), CSIC-CMCNB, 08038 Barcelona, Catalonia, Spain.

Ales Kovarik, Institute of Biophysics, Academy of Sciences of the Czech Republic, 61265 Brno, Czech Republic.

Sophie Maiwald, Faculty of Biology, Technische Universität Dresden, D-01069 Dresden, Germany.

Ludwig Mann, Faculty of Biology, Technische Universität Dresden, D-01069 Dresden, Germany.

Nicola Schmidt, Faculty of Biology, Technische Universität Dresden, D-01069 Dresden, Germany.

Joan Pere Pascual-Díaz, Institut Botànic de Barcelona (IBB), CSIC-CMCNB, 08038 Barcelona, Catalonia, Spain.

Daniel Vitales, Institut Botànic de Barcelona (IBB), CSIC-CMCNB, 08038 Barcelona, Catalonia, Spain; Laboratori de Botànica–Unitat Associada CSIC, Facultat de Farmàcia i Ciències de l’Alimentació, Universitat de Barcelona, 08028 Barcelona, Catalonia, Spain.

Beatrice Weber, Faculty of Biology, Technische Universität Dresden, D-01069 Dresden, Germany.

Tony Heitkam, Faculty of Biology, Technische Universität Dresden, D-01069 Dresden, Germany; Institute of Biology, NAWI Graz, Karl-Franzens-Universität, A-8010 Graz, Austria.

Data Availability

Not applicable.

References

  1. Abascal-Palacios  G, Jochem  L, Pla-Prats  C, Beuron  F, Vannini  A. Structural basis of Ty3 retrotransposon integration at RNA polymerase III-transcribed genes. Nat Commun. 2021:12(1):6992. 10.1038/s41467-021-27338-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Almeida  MV, Vernaz  G, Putman  ALK, Miska  EA. Taming transposable elements in vertebrates: from epigenetic silencing to domestication. Trends Genet. 2022:38(6):529–553. 10.1016/j.tig.2022.02.009. [DOI] [PubMed] [Google Scholar]
  3. Altinkut  A, Raskina  O, Nevo  E, Belyayev  A. En/Spm-like transposons in Poaceae species: transposase sequence variability and chromosomal distribution. Cell Mol Biol Lett. 2006:11(2):214–229. 10.2478/s11658-006-0017-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Arkhipova  IR. Neutral theory, transposable elements, and eukaryotic genome evolution. Mol Biol Evol. 2018:35(6):1332–1337. 10.1093/molbev/msy083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Baeza  C, Schrader  O, Budahn  H. Characterization of geographically isolated accessions in five Alstroemeria L. species (Chile) using FISH of tandemly repeated DNA sequences and RAPD analysis. Plant Syst Evol. 2007:269(1–2):1–14. 10.1007/s00606-007-0591-5. [DOI] [Google Scholar]
  6. Balint-Kurti  PJ, Clendennen  SK, Doleželová  M, Valárik  M, Doležel  J, Beetham  PR, May  GD. Identification and chromosomal localization of the monkey retrotransposon in Musa sp. Mol Gen Genet. 2000:263(6):908–915. 10.1007/s004380000265. [DOI] [PubMed] [Google Scholar]
  7. Bendich  AJ, Rogers  SO. Ribosomal intergenic spacers are filled with transposon remnants. Genome Biol Evol. 2023:15(7):evad114. 10.1093/gbe/evad114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bhattacharyya  MK, Smith  AM, Ellis  THN, Hedley  C, Martin  C. The wrinkled-seed character of pea described by Mendel is caused by a transposon-like insertion in a gene encoding starch-branching enzyme. Cell. 1990:60(1):115–122. 10.1016/0092-8674(90)90721-P. [DOI] [PubMed] [Google Scholar]
  9. Bigot  Y, Lutcher  F, Hamelin  MH, Périquet  G. The 28S ribosomal RNA-encoding gene of Hymenoptera: inserted sequences in the retrotransposon-rich regions. Gene. 1992:121(2):347–352. 10.1016/0378-1119(92)90142-C. [DOI] [PubMed] [Google Scholar]
  10. Bloom  SE, Goodpasture  C. An improved technique for selective silver staining of nucleolar organizer regions in human chromosomes. Hum Genet. 1976:34(2):199–206. 10.1007/BF00278889. [DOI] [PubMed] [Google Scholar]
  11. Blum  H, Beier  H, Gross  HJ. Improved silver staining of plant proteins, RNA and DNA in polyacrylamide gels. Electrophoresis. 1987:8(2):93–99. 10.1002/elps.1150080203. [DOI] [Google Scholar]
  12. Blumenstiel  JP. Birth, school, work, death, and resurrection: the life stages and dynamics of transposable element proliferation. Genes (Basel).  2019:10(5):336. 10.3390/genes10050336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Brewer  TE, Albertsen  M, Edwards  A, Kirkegaard  RH, Rocha  EP, Fierer  N. Unlinked rRNA genes are widespread among bacteria and archaea. ISME J. 2020:14(2):597–608. 10.1038/s41396-019-0552-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Brown  DD, Wensink  PC, Jordan  E. A comparison of the ribosomal DNA's of Xenopus laevis and Xenopus mulleri: the evolution of tandem genes. J Mol Biol. 1972:63(1):57–73. 10.1016/0022-2836(72)90521-9. [DOI] [PubMed] [Google Scholar]
  15. Bueno  D, Palacios-Gimenez  OM, Cabral-de-Mello  DC. Chromosomal mapping of repetitive DNAs in the grasshopper Abracris flavolineata reveal possible ancestry of the B chromosome and H3 histone spreading. PLoS One. 2013:8(6):e66532. 10.1371/journal.pone.0066532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Burke  WD, Calalang  CC, Eickbush  TH. The site-specific ribosomal insertion element type II of Bombyx mori (R2Bm) contains the coding sequence for a reverse transcriptase-like enzyme. Mol Cell Biol. 1987:7(6):2221–2230. 10.1128/mcb.7.6.2221-2230.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Burke  WD, Eickbush  DG, Xiong  Y, Jakubczak  J, Eickbush  TH. Sequence relationship of retrotransposable elements R1 and R2 within and between divergent insect species. Mol Biol Evol. 1993:10(1):163–185. 10.1093/oxfordjournals.molbev.a039990. [DOI] [PubMed] [Google Scholar]
  18. Burke  WD, Malik  HS, Lathe  WC  III, Eickbush  TH. Are retrotransposons long-term hitchhikers?  Nature. 1998:392(6672):141–142. 10.1038/32330. [DOI] [PubMed] [Google Scholar]
  19. Carleton  KL, Conte  MA, Malinsky  M, Nandamuri  SP, Sandkam  BA, Meier  JI, Mwaiko  S, Seehausen  O, Kocher  TD. Movement of transposable elements contributes to cichlid diversity. Mol Ecol. 2020:29(24):4956–4969. 10.1111/mec.15685. [DOI] [PubMed] [Google Scholar]
  20. Chester  M, Sykorova  E, Fajkus  J, Leitch  AR. Single integration and spread of a Copia-like sequence nested in rDNA intergenic spacers of Allium cernuum (Alliaceae). Cytogenet Genome Res. 2010:129(1–3):35–46. 10.1159/000312959. [DOI] [PubMed] [Google Scholar]
  21. Chuong  EB, Elde  NC, Feschotte  C. Regulatory activities of transposable elements: from conflicts to benefits. Nat Rev Genet. 2017:18(2):71–86. 10.1038/nrg.2016.139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cioffi  MB, Martins  C, Bertollo  LAC. Chromosome spreading of associated transposable elements and ribosomal DNA in the fish Erythrinus erythrinus. Implications for genome change and karyoevolution in fish. BMC Evol Biol. 2010:10:271. 10.1186/1471-2148-10-271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Coen  E, Strachan  T, Dover  G. Dynamics of concerted evolution of ribosomal DNA and histone gene families in the melanogaster species subgroup of Drosophila. J Mol Biol. 1982:158(1):17–35. 10.1016/0022-2836(82)90448-X. [DOI] [PubMed] [Google Scholar]
  24. Cohen  S, Houben  A, Segal  D. Extrachromosomal circular DNA derived from tandemly repeated genomic sequences in plants. Plant J. 2008:53(6):1027–1034. 10.1111/j.1365-313X.2007.03394.x. [DOI] [PubMed] [Google Scholar]
  25. Cohen  S, Yacobi  K, Segal  D. Extrachromosomal circular DNA of tandemly repeated genomic sequences in Drosophila. Genome Res. 2003:13(6a):1133–1145. 10.1101/gr.907603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Condon  C, French  S, Squires  C, Squires  CL. Depletion of functional ribosomal RNA operons in Escherichia coli causes increased expression of the remaining intact copies. EMBO J. 1993:12(11):4305–4315. 10.1002/j.1460-2075.1993.tb06115.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Cosby  RL, Chang  NC, Feschotte  C. Host–transposon interactions: conflict, cooperation, and cooption. Genes Dev. 2019:33(17–18):1098–1116. 10.1101/gad.327312.119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Dalíková  M, Provazníková  I, Provazník  J, Grof-Tisza  P, Pepi  A, Nguyen  P. The role of repetitive sequences in repatterning of major ribosomal DNA clusters in Lepidoptera. Gen Biol Evol. 2023:15(6):evad090. 10.1093/gbe/evad090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. da Silva  M, Barbosa  P, Artoni  RF, Feldberg  E. Evolutionary dynamics of 5S rDNA and recurrent association of transposable elements in electric fish of the family Gymnotidae (Gymnotiformes): the case of Gymnotus mamiraua. Cytogenet Genome Res. 2016:149(4):297–303. 10.1159/000449431. [DOI] [PubMed] [Google Scholar]
  30. Devos  KM, Brown  JKM, Bennetzen  JL. Genome size reduction through illegitimate recombination counteracts genome expansion in Arabidopsis. Genome Res. 2002:12(7):1075–1079. 10.1101/gr.132102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Drouin  G, De Sa  MM. The concerted evolution of 5S ribosomal genes linked to the repeat units of other multigene families. Mol Biol Evol. 1995:12(3):481–493. 10.1093/oxfordjournals.molbev.a040223. [DOI] [PubMed] [Google Scholar]
  32. Eagle  SHC, Crease  TJ. Copy number variation of ribosomal DNA and Pokey transposons in natural populations of Daphnia. Mob DNA. 2012:3(1):4. 10.1186/1759-8753-3-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Eaves  LA, Gardner  AJ, Fry  RC. Tools for the assessment of epigenetic regulation. In: Fry  RC, editor. Environmental epigenetics in toxicology and public health. Amsterdam: Elsevier; 2020. p. 33–64. [Google Scholar]
  34. Eickbush  DG, Burke  WD, Eickbush  TH. Evolution of the R2 retrotransposon ribozyme and its self-cleavage site. PLoS One. 2013:8(9):e66441. 10.1371/journal.pone.0066441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Eickbush  DG, Eickbush  TH. Vertical transmission of the retrotransposable elements R1 and R2 during the evolution of the Drosophila melanogaster species subgroup. Genetics. 1995:139(2):671–684. 10.1093/genetics/139.2.671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Eickbush  DG, Eickbush  TH. Transcription of endogenous and exogenous R2 elements in the rRNA gene locus of Drosophila melanogaster. Mol Cell Biol. 2003:23(11):3825–3836. 10.1128/MCB.23.11.3825-3836.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Eickbush  DG, Eickbush  TH. R2 and R2/R1 hybrid non-autonomous retrotransposons derived by internal deletions of full-length elements. Mob DNA. 2012:3(1):10. 10.1186/1759-8753-3-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Eickbush  DG, Ye  J, Zhang  X, Burke  WD, Eickbush  TH. Epigenetic regulation of retrotransposons within the nucleolus of Drosophila. Mol Cell Biol. 2008:28(20):6452–6461. 10.1128/MCB.01015-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Elliott  TA, Stage  DE, Crease  TJ, Eickbush  TH. In and out of the rRNA genes: characterization of Pokey elements in the sequenced Daphnia genome. Mob DNA. 2013:4(1):20. 10.1186/1759-8753-4-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Ellison  CE, Bachtrog  D. Non-allelic gene conversion enables rapid evolutionary change at multiple regulatory sites encoded by transposable elements. Elife. 2015:4:e05899. 10.7554/eLife.05899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Fan  W, Eklund  E, Sherman  RM, Liu  H, Pitts  S, Ford  B, Rajeshkumar  NV, Laiho  M. Widespread genetic heterogeneity of human ribosomal RNA genes. RNA. 2022:28(4):478–492. 10.1261/rna.078925.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Fawcett  JA, Innan  H. The role of gene conversion between transposable elements in rewiring regulatory networks. Genome Biol Evol. 2019:11(7):1723–1729. 10.1093/gbe/evz124. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Fefelova  EA, Pleshakova  IM, Mikhaleva  EA, Pirogov  SA, Poltorachenko  VA, Abramov  YA, Romashin  DD, Shatskikh  AS, Blokh  RS, Gvozdev  VA, et al.  Impaired function of rDNA transcription initiation machinery leads to derepression of ribosomal genes with insertions of R2 retrotransposon. Nucleic Acids Res. 2022:50(2):867–884. 10.1093/nar/gkab1276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Ferretti  ABSM, Ruiz-Ruano  FJ, Milani  D, Loreto  V, Martí  DA, Ramos  E, Martins  C, Cabral-de-Mello  DC. How dynamic could be the 45S rDNA cistron? An intriguing variability in a grasshopper species revealed by integration of chromosomal and genomic data. Chromosoma. 2019:128(2):165–175. 10.1007/s00412-019-00706-8. [DOI] [PubMed] [Google Scholar]
  45. Finnegan  DJ. Eukaryotic transposable elements and genome evolution. Trends Genet. 1989:5(4):103–107. 10.1016/0168-9525(89)90039-5. [DOI] [PubMed] [Google Scholar]
  46. Flavell  AJ, Ish-Horowicz  D. Extrachromosomal circular copies of the eukaryotic transposable element Copia in cultured Drosophila cells. Nature. 1981:292(5824):591–595. 10.1038/292591a0. [DOI] [PubMed] [Google Scholar]
  47. Flavell  RB, Bennett  MD, Smith  JB, Smith  DB. Genome size and the proportion of repeated nucleotide sequence DNA in plants. Biochem Genet. 1974:12(4):257–269. 10.1007/BF00485947. [DOI] [PubMed] [Google Scholar]
  48. Fujisawa  M, Yamagata  H, Kamiya  K, Nakamura  M, Saji  S, Kanamori  H, Wu  J, Matsumoto  T, Sasaki  T. Sequence comparison of distal and proximal ribosomal DNA arrays in rice (Oryza sativa L.) chromosome 9S and analysis of their flanking regions. Theor Appl Genet. 2006:113(3):419–428. 10.1007/s00122-006-0307-1. [DOI] [PubMed] [Google Scholar]
  49. Fujiwara  H, Ogura  T, Takada  N, Miyajima  N, Ishikawa  H, Maekawa  H. Introns and their flanking sequences of Bombyx mori rDNA. Nucleic Acids Res. 1984:12(17):6861–6869. 10.1093/nar/12.17.6861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Ganley  ARD, Kobayashi  T. Ribosomal DNA and cellular senescence: new evidence supporting the connection between rDNA and aging. FEMS Yeast Res. 2014:14(1):49–59. 10.1111/1567-1364.12133. [DOI] [PubMed] [Google Scholar]
  51. Gao  D, Li  Y, Kim  KD, Abernathy  B, Jackson  SA. Landscape and evolutionary dynamics of terminal repeat retrotransposons in miniature in plant genomes. Genome Biol. 2016:17(1):7. 10.1186/s13059-015-0867-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Gao  X, Hou  Y, Ebina  H, Levin  H, Voytas  DF. Chromodomains direct integration of retrotransposons to heterochromatin. Genome Res. 2008:18(3):359–369. 10.1101/gr.7146408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Garcia  S, Crhák Khaitová  L, Kovařík  A. Expression of 5 S rRNA genes linked to 35 S rDNA in plants, their epigenetic modification and regulatory element divergence. BMC Plant Biol. 2012:12:95. 10.1186/1471-2229-12-95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Garcia  S, Kovařík  A, Leitch  A, Garnatje  T. Cytogenetic features of rRNA genes across land plants: analysis of the plant rDNA database. Plant J. 2017:89(5):1020–1030. 10.1111/tpj.13442. [DOI] [PubMed] [Google Scholar]
  55. Garcia  S, Lim  KY, Chester  M, Garnatje  T, Pellicer  J, Vallès  J, Leitch  AR, Kovařík  A. Linkage of 35S and 5S rRNA genes in Artemisia (family Asteraceae): first evidence from angiosperms. Chromosoma. 2009:118(1):85–97. 10.1007/s00412-008-0179-z. [DOI] [PubMed] [Google Scholar]
  56. Garcia  S, Panero  JL, Siroky  J, Kovarik  A. Repeated reunions and splits feature the highly dynamic evolution of 5S and 35S ribosomal RNA genes (rDNA) in the Asteraceae family. BMC Plant Biol. 2010:10:176. 10.1186/1471-2229-10-176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Garcia  S, Pascual-Díaz  JP, Krumpolcová  A, Kovařík  A. Analysis of 5S rDNA genomic organization through the RepeatExplorer2 pipeline: a simplified protocol. In: Heitkam  T, Garcia  S, editors. Plant cytogenetics and cytogenomics. Methods in molecular biology. 1st ed. Vol. 2672. New York (NY): Humana; 2023. p. 501–512. 10.1007/978-1-0716-3226-0_30. [DOI] [PubMed] [Google Scholar]
  58. Garcia  S, Wendel  JF, Borowska-Zuchowska  N, Aïnouche  M, Kuderova  A, Kovarik  A. The utility of graph clustering of 5S ribosomal DNA homoeologs in plant allopolyploids, homoploid hybrids, and cryptic introgressants. Front Plant Sci. 2020:11:41. 10.3389/fpls.2020.00041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Gibbons  JG, Branco  AT, Godinho  SA, Yu  S, Lemos  B. Concerted copy number variation balances ribosomal DNA dosage in human and mouse genomes. Proc Natl Acad Sci U S A. 2015:112(8):2485–2490. 10.1073/pnas.1416878112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Gibbons  JG, Branco  AT, Yu  S, Lemos  B. Ribosomal DNA copy number is coupled with gene expression variation and mitochondrial abundance in humans. Nat Commun. 2014:5:4850. 10.1038/ncomms5850. [DOI] [PubMed] [Google Scholar]
  61. Glass  SK, Moszczynska  A, Crease  TJ. The effect of transposon Pokey insertions on sequence variation in the 28S rRNA gene of Daphnia pulex. Genome. 2008:51(12):988–1000. 10.1139/G08-092. [DOI] [PubMed] [Google Scholar]
  62. Glugoski  L, Giuliano-Caetano  L, Moreira-Filho  O, Vicari  MR, Nogaroto  V. Co-located hAT transposable element and 5S rDNA in an interstitial telomeric sequence suggest the formation of Robertsonian fusion in armored catfish. Gene. 2018:650:49–54. 10.1016/j.gene.2018.01.099. [DOI] [PubMed] [Google Scholar]
  63. Goffová  I, Fajkus  J. The rDNA loci—intersections of replication, transcription, and repair pathways. Int J Mol Sci. 2021:22(3):1302. 10.3390/ijms22031302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Gogolevsky  KP, Vassetzky  NS, Kramerov  DA. 5S rRNA-derived and tRNA-derived SINEs in fruit bats. Genomics. 2009:93(5):494–500. 10.1016/j.ygeno.2009.02.001. [DOI] [PubMed] [Google Scholar]
  65. Gonzalez  IL, Tugendreich  S, Hieter  P, Sylvester  JE. Fixation times of retroposons in the ribosomal DNA spacer of human and other primates. Genomics. 1993:18(1):29–36. 10.1006/geno.1993.1423. [DOI] [PubMed] [Google Scholar]
  66. Gorbunova  V, Seluanov  A, Mita  P, McKerrow  W, Fenyö  D, Boeke  JD, Linker  SB, Gage  FH, Kreiling  JA, Petrashen  AP, et al.  The role of retrotransposable elements in ageing and age-associated diseases. Nature. 2021:596(7870):43–53. 10.1038/s41586-021-03542-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Hall  AN, Morton  E, Queitsch  C. First discovered, long out of sight, finally visible: ribosomal DNA. Trends Genet. 2022:38(6):587–597. 10.1016/j.tig.2022.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Handa  H, Kanamori  H, Tanaka  T, Murata  K, Kobayashi  F, Robinson  SJ, Koh  CS, Pozniak  CJ, Sharpe  AG, Paux  E, et al.  Structural features of two major nucleolar organizer regions (NORs), Nor-B1 and Nor-B2, and chromosome-specific rRNA gene expression in wheat. Plant J. 2018:96(6):1148–1159. 10.1111/tpj.14094. [DOI] [PubMed] [Google Scholar]
  69. Harasawa  R, Pitcher  DG, Ramírez  AS, Bradbury  JM. A putative transposase gene in the 16S–23S rRNA intergenic spacer region of Mycoplasma imitans. Microbiology (Reading). 2004:150(Pt 4):1023–1029. 10.1099/mic.0.26629-0. [DOI] [PubMed] [Google Scholar]
  70. Hassan  M, Das  S, Adhya  S. Mini-exon derived RNA gene of Leishmania donovani: structure, organization and expression. J Biosci. 1992:17(1):55–66. 10.1007/BF02716774. [DOI] [Google Scholar]
  71. Havlová  K, Dvořáčková  M, Peiro  R, Abia  D, Mozgová  I, Vansáčová  L, Gutierrez  C, Fajkus  J. Variation of 45S rDNA intergenic spacers in Arabidopsis thaliana. Plant Mol Biol. 2016:92(4–5):457–471. 10.1007/s11103-016-0524-1. [DOI] [PubMed] [Google Scholar]
  72. Heitkam  T, Garcia  S. Plant cytogenetics and cytogenomics: methods in molecular biology. 1st ed. Vol. 2672. New York (NY): Humana; 2023. p. 1–568. 10.1007/978-1-0716-3226-0. [DOI] [PubMed] [Google Scholar]
  73. Heitkam  T, Schmidt  T. BNR—a LINE family from Beta vulgaris—contains a RRM domain in open reading frame 1 and defines a L1 sub-clade present in diverse plant genomes. Plant J. 2009:59(6):872–882. 10.1111/j.1365-313X.2009.03923.x. [DOI] [PubMed] [Google Scholar]
  74. Heitkam  T, Weber  B, Walter  I, Liedtke  S, Ost  C, Schmidt  T. Satellite DNA landscapes after allotetraploidization of quinoa (Chenopodium quinoa) reveal unique A and B subgenomes. Plant J. 2020:103(1):32–52. 10.1111/tpj.14705. [DOI] [PubMed] [Google Scholar]
  75. Hemleben  V, Grierson  D, Borisjuk  N, Volkov  RA, Kovarik  A. Personal perspectives on plant ribosomal RNA genes research: from precursor-rRNA to molecular evolution. Front Plant Sci. 2021:12:797348. 10.3389/fpls.2021.797348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Hemleben  V, Zentgraf  U. Structural organization and regulation of transcription by RNA polymerase I of plant nuclear ribosomal RNA genes. Results Probl Cell Differ. 1994:20:3–24. 10.1007/978-3-540-48037-2_1. [DOI] [PubMed] [Google Scholar]
  77. Hřibová  E, Neumann  P, Matsumoto  T, Roux  N, Macas  J, Doležel  J. Repetitive part of the banana (Musa acuminata) genome investigated by low-depth 454 sequencing. BMC Plant Biol. 2010:10:204. 10.1186/1471-2229-10-204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Ingle  J, Timmis  JN, Sinclair  J. The relationship between satellite deoxyribonucleic acid, ribosomal ribonucleic acid gene redundancy, and genome size in plants. Plant Physiol. 1975:55(3):496–501. 10.1104/pp.55.3.496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Jakubczak  JL, Xiong  Y, Eickbush  TH. Type I (R1) and type II (R2) ribosomal DNA insertions of Drosophila melanogaster are retrotransposable elements closely related to those of Bombyx mori. J Mol Biol. 1990:212(1):37–52. 10.1016/0022-2836(90)90303-4. [DOI] [PubMed] [Google Scholar]
  80. Jakubczak  JL, Zenni  MK, Woodruff  RC, Eickbush  TH. Turnover of R1 (type I) and R2 (type II) retrotransposable elements in the ribosomal DNA of Drosophila melanogaster. Genetics. 1992:131(1):129–142. 10.1093/genetics/131.1.129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Jamrich  M, Miller  OL. The rare transcripts of interrupted rRNA genes in Drosophila melanogaster are processed or degraded during synthesis. EMBO J. 1984:3(7):1541–1545. 10.1002/j.1460-2075.1984.tb02008.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Jiang  N, Bao  Z, Zhang  X, Eddy  SR, Wessler  SR. Pack-MULE transposable elements mediate gene evolution in plants. Nature. 2004:431(7008):569–573. 10.1038/nature02953. [DOI] [PubMed] [Google Scholar]
  83. Jo  S-H, Koo  D-H, Kim  JF, Hur  C-G, Lee  S, Yang  TJ, Kwon  SY, Choi  D. Evolution of ribosomal DNA-derived satellite repeat in tomato genome. BMC Plant Biol. 2009:9(1):1–14. 10.1186/1471-2229-9-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Kalendar  R, Raskina  O, Belyayev  A, Schulman  AH. Long tandem arrays of Cassandra retroelements and their role in genome dynamics in plants. Int J Mol Sci. 2020:21(8):2931. 10.3390/ijms21082931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Kalendar  R, Tanskanen  J, Chang  W, Antonius  K, Sela  H, Peleg  O, Schulman  AH. Cassandra retrotransposons carry independently transcribed 5S RNA. Proc Natl Acad Sci U S A. 2008:105(15):5833–5838. 10.1073/pnas.0709698105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Kamstra  SA, Kuipers  AGJ, De Jeu  MJ, Ramanna  MS, Jacobsen  E. Physical localisation of repetitive DNA sequences in Alstroemeria: karyotyping of two species with species-specific and ribosomal DNA. Genome. 1997:40(5):652–658. 10.1139/g97-086. [DOI] [PubMed] [Google Scholar]
  87. Kapitonov  VV, Jurka  J. A novel class of SINE elements derived from 5S rRNA. Mol Biol Evol. 2003:20(5):694–702. 10.1093/molbev/msg075. [DOI] [PubMed] [Google Scholar]
  88. Kasselimi  E, Pefani  DE, Taraviras  S, Lygerou  Z. Ribosomal DNA and the nucleolus at the heart of aging. Trends Biochem Sci. 2022:47(4):328–341. 10.1016/j.tibs.2021.12.007. [DOI] [PubMed] [Google Scholar]
  89. Kazazian  HH. Mobile DNA transposition in somatic cells. BMC Biol. 2011:9(1):1–4. 10.1186/1741-7007-9-62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Kejnovsky  E, Hobza  R, Kubat  Z, Widmer  A, Marais  GAB, Vyskot  B. High intrachromosomal similarity of retrotransposon long terminal repeats: evidence for homogenization by gene conversion on plant sex chromosomes?  Gene. 2007:390(1–2):92–97. 10.1016/j.gene.2006.10.007. [DOI] [PubMed] [Google Scholar]
  91. Kempken  F. Hideaway, a repeated element from Ascobolus immersus, is rDNA-associated and may resemble a retrotransposon. Curr Genet. 2001:40(3):179–185. 10.1007/s002940100253. [DOI] [PubMed] [Google Scholar]
  92. Kerrebrock  AW, Srivastava  R, Gerbi  SA. Isolation and characterization of ribosomal DNA variants from Sciara coprophila. J Mol Biol. 1989:210(1):1–13. 10.1016/0022-2836(89)90286-6. [DOI] [PubMed] [Google Scholar]
  93. Kimura  M. Evolutionary rate at the molecular level. Nature. 1968:217(5129):624–626. 10.1038/217624a0. [DOI] [PubMed] [Google Scholar]
  94. Kobayashi  T. Strategies to maintain the stability of the ribosomal RNA gene repeats—collaboration of recombination, cohesion, and condensation. Genes Genet Syst. 2006:81(3):155–161. 10.1266/ggs.81.155. [DOI] [PubMed] [Google Scholar]
  95. Kojima  KK, Fujiwara  H. Cross-genome screening of novel sequence specific non-LTR retrotransposons: various multicopy RNA genes and microsatellites are selected as targets. Mol Biol Evol. 2004:21(2):207–217. 10.1093/molbev/msg235. [DOI] [PubMed] [Google Scholar]
  96. Kojima  KK, Fujiwara  H. Long-term inheritance of the 28S rDNA-specific retrotransposon R2. Mol Biol Evol. 2005:22(11):2157–2165. 10.1093/molbev/msi210. [DOI] [PubMed] [Google Scholar]
  97. Kojima  KK. Helenus and Ajax, Two groups of non-autonomous LTR retrotransposons, represent a new type of small rna gene-derived mobile elements. Biology. 2024:13(2):119. 10.3390/biology13020119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Kuroki-Kami  A, Nichuguti  N, Yatabe  H, Mizuno  S, Kawamura  S, Fujiwara  H. Targeted gene knockin in zebrafish using the 28S rDNA-specific non-LTR-retrotransposon R2Ol. Mob DNA. 2019:10(1):23. 10.1186/s13100-019-0167-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Lan  T, Albert  VA. Dynamic distribution patterns of ribosomal DNA and chromosomal evolution in Paphiopedilum, a lady's slipper orchid. BMC Plant Biol. 2011:11(1):126. 10.1186/1471-2229-11-126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Lanciano  S, Carpentier  MC, Llauro  C, Jobet  E, Robakowska-Hyzorek  D, Lasserre  E, Ghesquière  A, Panaud  O, Mirouze  M. Sequencing the extrachromosomal circular mobilome reveals retrotransposon activity in plants. PLoS Genet. 2017:13(2):e1006630. 10.1371/journal.pgen.1006630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lathe  WC, Burke  WD, Eickbush  DG, Eickbush  TH. Evolutionary stability of the R1 retrotransposable element in the genus Drosophila. Mol Biol Evol. 1995:12(6):1094–1105. 10.1093/oxfordjournals.molbev.a040283. [DOI] [PubMed] [Google Scholar]
  102. Lavrinienko  A, Jernfors  T, Koskimäki  JJ, Pirttilä  AM, Watts  PC. Does intraspecific variation in rDNA copy number affect analysis of microbial communities?  Trends Microbiol. 2021:29(1):19–27. 10.1016/j.tim.2020.05.019. [DOI] [PubMed] [Google Scholar]
  103. Lecanidou  R, Eickbush  TH, Kafatos  FC. Ribosomal DNA genes of Bombyx mori: a minor fraction of the repeating units contain insertions. Nucleic Acids Res. 1984:12(11):4703–4713. 10.1093/nar/12.11.4703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. LeRiche  K, Eagle  SHC, Crease  TJ. Copy number of the transposon, Pokey, in rDNA is positively correlated with rDNA copy number in Daphnia obtuse. PLoS ONE. 2014:9(12):e114773. 10.1371/journal.pone.0114773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Lim  K, Furuta  Y, Kobayashi  I. Large variations in bacterial ribosomal RNA genes. Mol Biol Evol. 2012:29(10):2937–2948. 10.1093/molbev/mss101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Liu  H, Pan  G, Dang  X, Li  T, Zhou  Z. Characterization of active ribosomal RNA harboring MITEs insertion in microsporidian Nosema bombycis genome. Parasitol Res. 2013:112(3):1011–1020. 10.1007/s00436-012-3223-0. [DOI] [PubMed] [Google Scholar]
  107. Locati  MD, Pagano  JFB, Girard  G, Ensink  WA, van Olst  M, van Leeuwen  S, Nehrdich  U, Spaink  HP, Rauwerda  H, Jonker  MJ, et al.  Expression of distinct maternal and somatic 5.8S, 18S, and 28S rRNA types during zebrafish development. RNA. 2017:23(8):1188–1199. 10.1261/rna.061515.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Long  EO, Dawid  IB. Expression of ribosomal DNA insertions in Drosophila melanogaster. Cell. 1979:18(4):1185–1196. 10.1016/0092-8674(79)90231-9. [DOI] [PubMed] [Google Scholar]
  109. Longo  MS, Brown  JD, Zhang  C, O’Neill  MJ, O’Neill  RJ. Identification of a recently active mammalian SINE derived from ribosomal RNA. Genome Biol Evol. 2015:7(3):775–788. 10.1093/gbe/evv015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Lopez  FB, Fort  A, Tadini  L, Probst  AV, McHale  M, Friel  J, Ryder  P, Pontvianne  F, Pesaresi  P, Sulpice  R, et al.  Gene dosage compensation of rRNA transcript levels in Arabidopsis thaliana lines with reduced ribosomal gene copy number. Plant Cell. 2021:33(4):1135–1150. 10.1093/plcell/koab020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Luo  Y, Fefelova  E, Ninova  M, Chen  YCA, Aravin  AA. Repression of interrupted and intact rDNA by the sumo pathway in Drosophila melanogaster. Elife. 2020:9:e52416. 10.7554/eLife.52416. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Maeda  M, Shimada  T, Ishihama  A. Strength and regulation of seven rRNA promoters in Escherichia coli. PLoS One. 2015:10(12):e0144697. 10.1371/journal.pone.0144697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Mahelka  V, Krak  K, Kopecký  D, Fehrer  J, Šafář  J, Bartoš  J, Hobza  R, Blavet  N, Blattner  FR. Multiple horizontal transfers of nuclear ribosomal genes between phylogenetically distinct grass lineages. Proc Natl Acad Sci U S A. 2017:114(7):1726–1731. 10.1073/pnas.1613375114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Maiwald  S, Mann  L, Garcia  S, Heitkam  T. Evolving together: Cassandra retrotransposons gradually mirror promoter mutations of the 5S rRNA genes. Mol Biol Evol. 2024:41(2):msae010. 10.1093/molbev/msae010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Maiwald  S, Weber  B, Seibt  KM, Schmidt  T, Heitkam  T. The Cassandra retrotransposon landscape in sugar beet (Beta vulgaris) and related Amaranthaceae: recombination and re-shuffling lead to a high structural variability. Ann Bot. 2021:127(1):91–109. 10.1093/aob/mcaa176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Malik  HS, Eickbush  TH. Modular evolution of the integrase domain in the Ty3/Gypsy class of LTR retrotransposons. J Virol. 1999:73(6):5186–5190. 10.1128/JVI.73.6.5186-5190.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Malone  JH. Balancing copy number in ribosomal DNA. Proc Natl Acad Sci U S A. 2015:112(9):2635–2636. 10.1073/pnas.1500054112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Mann  L, Seibt  KM, Weber  B, Heitkam  T. ECCsplorer: a pipeline to detect extrachromosomal circular DNA (eccDNA) from next-generation sequencing data. BMC Bioinformatics. 2022:23(1):1–15. 10.1186/s12859-021-04545-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Mansisidor  A, Molinar  T, Srivastava  P, Dartis  DD, Pino Delgado  A, Blitzblau  HG, Klein  H, Hochwagen  A. Genomic copy-number loss is rescued by self-limiting production of DNA circles. Mol Cell. 2018:72(3):583–593.e4. 10.1016/j.molcel.2018.08.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Marx  V. Method of the year: long-read sequencing. Nat Methods. 2023:20(1):6–11. 10.1038/s41592-022-01730-w. [DOI] [PubMed] [Google Scholar]
  121. Matveev  V, Okada  N. Retroposons of salmonoid fishes (Actinopterygii: Salmonoidei) and their evolution. Gene. 2009:434(1-2):16–28. 10.1016/j.gene.2008.04.022. [DOI] [PubMed] [Google Scholar]
  122. McClintock  B. The relation of a particular chromosomal element to the development of the nucleoli in Zea mays. Zeitschrift für Zellforschung und Mikroskopische Anatomie. 1934:21(2):294–326. 10.1007/BF00374060. [DOI] [Google Scholar]
  123. McClintock  B. The origin and behavior of mutable loci in maize. Proc Natl Acad Sci U S A. 1950:36(6):344–355. 10.1073/pnas.36.6.344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. McClintock  B. Chromosome organization and genic expression. Cold Spring Harb Symp Quant Biol. 1951:16:13–47. 10.1101/SQB.1951.016.01.004. [DOI] [PubMed] [Google Scholar]
  125. McKinlay  A, Fultz  D, Wang  F, Pikaard  CS. Targeted enrichment of rRNA gene tandem arrays for ultra-long sequencing by selective restriction endonuclease digestion. Front Plant Sci. 2021:12:762. 10.3389/fpls.2021.656049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Merkulov  P, Egorova  E, Kirov  I. Composition and structure of Arabidopsis thaliana extrachromosomal circular DNAs revealed by nanopore sequencing. Plants (Basel). 2023:12(11):2178. 10.3390/plants12112178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Merlo  MA, Cross  I, Rodríguez-Rúa  A, Manchado  M, Rebordinos  L. First approach to studying the genetics of the meagre (Argyrosomus regius; Asso, 1801) using three multigene families. Aquacult Res. 2013:44(6):974–984. 10.1111/j.1365-2109.2012.03103.x. [DOI] [Google Scholar]
  128. Mizuochi  H, Marasek  A, Okazaki  K. Molecular cloning of Tulipa fosteriana rDNA and subsequent FISH analysis yields cytogenetic organization of 5S rDNA and 45S rDNA in T. gesneriana and T. fosteriana. Euphytica. 2007:155(1–2):235–248. 10.1007/s10681-006-9325-y. [DOI] [Google Scholar]
  129. Morgan  EA. Insertions of Tn10 into an E. coli ribosomal RNA operon are incompletely polar. Cell. 1980:21(1):257–265. 10.1016/0092-8674(80)90133-6. [DOI] [PubMed] [Google Scholar]
  130. Moss  T, Stefanovsky  V, Langlois  F, Gagnon-Kugler  T. A new paradigm for the regulation of the mammalian ribosomal RNA genes. Biochem Soc Trans. 2006:34(6):1079–1081. 10.1042/BST0341079. [DOI] [PubMed] [Google Scholar]
  131. Muscarella  DE, Vogt  VM. A mobile group I intron in the nuclear rDNA of Physarum polycephalum. Cell. 1989:56(3):443–454. 10.1016/0092-8674(89)90247-x. [DOI] [PubMed] [Google Scholar]
  132. Naish  M, Alonge  M, Wlodzimierz  P, Tock  AJ, Abramson  BW, Schmücker  A, Mandáková  T, Jamge  B, Lambing  C, Kuo  P, et al.  The genetic and epigenetic landscape of the Arabidopsis centromeres. Science. 2021:374(6569):eabi7489. 10.1126/science.abi7489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Nakajima  RT, Cabral-de-Mello  DC, Valente  GT, Venere  PC, Martins  C. Evolutionary dynamics of rRNA gene clusters in cichlid fish. BMC Evol Biol. 2012:12:198. 10.1186/1471-2148-12-198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Nei  M, Hughes  AL. Balanced polymorphism and evolution by the birth-and-death process in the MHC loci. In: Tsuji  K, Aizawa  M, Sasazuki  T, editors. 11th Histocompatibility workshop and conference. Oxford: Oxford University Press; 1992. p. 27–38. [Google Scholar]
  135. Nei  M, Rooney  AP. Concerted and birth-and-death evolution of multigene families. Ann Rev Genet. 2005:39(1):121–152. 10.1146/annurev.genet.39.073003.112240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Nelson  JO, Slicko  A, Yamashita  YM. The retrotransposon R2 maintains Drosophila ribosomal DNA repeats. Proc Natl Acad Sci U S A. 2023:120(23):e2221613120. 10.1073/pnas.2221613120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Neuhaus  H, Müller  F, Etter  A, Tobler  H. Type I-like Intervening sequences are found in the rDNA of the nematode Ascaris lumbricoides. Nucleic Acids Res. 1987:15(19):7689–7707. 10.1093/nar/15.19.7689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Neumann  P, Navrátilová  A, Koblížková  A, Kejnovsk  E, Hřibová  E, Hobza  R, Widmer  A, Doležel  J, Macas  J. Plant centromeric retrotransposons: a structural and cytogenetic perspective. Mob DNA. 2011:2(1):4. 10.1186/1759-8753-2-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Nieto-Feliner  G, Rosato  M, Alegre  G, San Segundo  P, Rosselló  JA, Garnatje  T, Garcia  S. Dissimilar molecular and morphological patterns in an introgressed peripheral population of a sand dune species (Armeria pungens, Plumbaginaceae). Plant Biol. 2019:21(6):1072–1082. 10.1111/plb.13035. [DOI] [PubMed] [Google Scholar]
  140. Nieto-Feliner  G, Rosselló  JA. Better the devil you know? Guidelines for insightful utilization of nrDNA ITS in species-level evolutionary studies in plants. Mol Phylogenet Evol. 2007:44(2):911–919. 10.1016/j.ympev.2007.01.013. [DOI] [PubMed] [Google Scholar]
  141. Nisen  P, Shapiro  L. E. coli ribosomal RNA contains sequences homologous to insertion sequences IS1 and IS2. Nature. 1979:282(5741):872–874. 10.1038/282872a0. [DOI] [PubMed] [Google Scholar]
  142. Nishihara  H, Smit  AFA, Okada  N. Functional noncoding sequences derived from SINEs in the mammalian genome. Genome Res. 2006:16(7):864–874. 10.1101/gr.5255506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Novák  P, Guignard  MS, Neumann  P, Kelly  LJ, Mlinarec  J, Koblížková  A, Dodsworth  S, Kovařík  A, Pellicer  J, Wang  W, et al.  Repeat-sequence turnover shifts fundamentally in species with large genomes. Nat Plants. 2020:6(11):1325–1329. 10.1038/s41477-020-00785-x. [DOI] [PubMed] [Google Scholar]
  144. Novák  P, Neumann  P, Macas  J. Graph-based clustering and characterization of repetitive sequences in next-generation sequencing data. BMC Bioinformatics. 2010:11:378. 10.1186/1471-2105-11-378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Novikova  O. Chromodomains and LTR retrotransposons in plants. Commun Integr Biol. 2009:2(2):158–162. 10.4161/cib.7702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. O’Connor  C, Adams  JU. Essentials of cell biology. Cambridge (MA): NPG Education; 2010. [Google Scholar]
  147. Orgel  LE, Crick  FHC. Selfish DNA: the ultimate parasite. Nature. 1980:284(5757):604–607. 10.1038/284604a0. [DOI] [PubMed] [Google Scholar]
  148. Oyun  NY, Zagoskina  AS, Mukha  DV. Inheritance of 5′-truncated copies of R2 retrotransposon in a series of generations of German cockroach, Blattella germanica. Russ J Genet. 2018:54(12):1438–1444. 10.1134/S1022795418120116. [DOI] [Google Scholar]
  149. Paço  A, Freitas  R, Vieira-da-Silva  A. Conversion of DNA sequences: from a transposable element to a tandem repeat or to a gene. Genes (Basel).  2019:10(12):1014. 10.3390/genes10121014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Pedrosa-Harand  A, de Almeida  CCS, Mosiolek  M, Blair  MW, Schweizer  D, Guerra  M. Extensive ribosomal DNA amplification during Andean common bean (Phaseolus vulgaris L.) evolution. Theor Appl Genet. 2006:112(5):924–933. 10.1007/s00122-005-0196-8. [DOI] [PubMed] [Google Scholar]
  151. Peng  H, Mirouze  M, Bucher  E. Extrachromosomal circular DNA: a neglected nucleic acid molecule in plants. Curr Opin Plant Biol. 2022:69:102263. 10.1016/j.pbi.2022.102263. [DOI] [PubMed] [Google Scholar]
  152. Penton  EH, Crease  TJ. Evolution of the transposable element Pokey in the ribosomal DNA of species in the subgenus Daphnia (Crustacea: Cladocera). Mol Biol Evol. 2004:21(9):1727–1739. 10.1093/molbev/msh189. [DOI] [PubMed] [Google Scholar]
  153. Penton  EH, Sullender  BW, Crease  TJ. Pokey, a new DNA transposon in Daphnia (Cladocera: Crustacea). J Mol Evol. 2002:55(6):664–673. 10.1007/s00239-002-2362-9. [DOI] [PubMed] [Google Scholar]
  154. Pérez-González  CE, Burke  WD, Eickbush  TH. R1 and R2 retrotransposition and deletion in the rDNA loci on the X and Y chromosomes of Drosophila melanogaster. Genetics. 2003:165(2):675–685. 10.1093/genetics/165.2.675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Pérez-González  CE, Eickbush  TH. Dynamics of R1 and R2 elements in the rDNA locus of Drosophila simulans. Genetics. 2001:158(4):1557–1567. 10.1093/genetics/158.4.1557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Pinhal  D, Yoshimura  TS, Araki  CS, Martins  C. The 5S rDNA family evolves through concerted and birth-and-death evolution in fish genomes: an example from freshwater stingrays. BMC Evol Biol. 2011:11:151. 10.1186/1471-2148-11-151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Piskurek  O, Nishihara  H, Okada  N. The evolution of two partner LINE/SINE families and a full-length chromodomain-containing Ty3/Gypsy LTR element in the first reptilian genome of Anolis carolinensis. Gene. 2009:441(1-2):111–118. 10.1016/j.gene.2008.11.030. [DOI] [PubMed] [Google Scholar]
  158. Platt  RN, Vandewege  MW, Ray  DA. Mammalian transposable elements and their impacts on genome evolution. Chromosome Res. 2018:26(1–2):25–43. 10.1007/s10577-017-9570-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Pont  G, Degroote  F, Picard  G. Some extrachromosomal circular DNAs from Drosophila embryos are homologous to tandemly repeated genes. J Mol Biol. 1987:195(2):447–451. 10.1016/0022-2836(87)90665-6. [DOI] [PubMed] [Google Scholar]
  160. Prokopowich  CD, Gregory  TR, Crease  TJ. The correlation between rDNA copy number and genome size in eukaryotes. Genome. 2003:46(1):48–50. 10.1139/g02-103. [DOI] [PubMed] [Google Scholar]
  161. Raskina  O, Barber  JC, Nevo  E, Belyayev  A. Repetitive DNA and chromosomal rearrangements: speciation-related events in plant genomes. Cytogenet Genome Res. 2008:120(3–4):351–357. 10.1159/000121084. [DOI] [PubMed] [Google Scholar]
  162. Redd  PS, Payero  L, Gilbert  DM, Page  CA, King  R, McAssey  EV, Bodie  D, Diaz  S, Hancock  CN. Transposase expression, element abundance, element size, and DNA repair determine the mobility and heritability of PIF/Pong/Harbinger transposable elements. Front Cell Dev Biol. 2023:11:1184046. 10.3389/fcell.2023.1184046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Renkawitz-Pohl  R, Matsumoto  L, Gerbi  SA. Structure of the ribosomal DNA repeat of Sciara coprophila. Nucleic Acids Res. 1981:9(15):3747–3764. 10.1093/nar/9.15.3747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Robicheau  BM, Susko  E, Harrigan  AM, Snyder  M. Ribosomal RNA genes contribute to the formation of pseudogenes and junk DNA in the human genome. Genome Biol Evol. 2017:9(2):380–397. 10.1093/gbe/evw307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Rodríguez-González  R, Gutiérrez  ML, Fuentes  I, Gálvez-Prada  F, Sochorová  J, Kovařík  A, Garcia  S. Release 4.0 of the plant rDNA database: a database on plant ribosomal DNA loci number, their position, and organization: an information source for comparative cytogenetics. In: Garcia  S, Nualart  N, editors. Plant genomic and cytogenetic databases: methods in molecular biology. Vol. 2703. New York (NY): Humana; 2023. p. 237–245. 10.1007/978-1-0716-3389-2_18. [DOI] [PubMed] [Google Scholar]
  166. Saifitdinova  A, Galkina  S, Kulak  M, Fillon  V, Volodkina  V, Pavlova  O, Gaginskaya  E. The dispersal of ribosomal gene sequences in the karyotype of Coturnix japonica. Biopolym Cell. 2019:35(3):229–230. 10.7124/bc.0009F5. [DOI] [Google Scholar]
  167. Sampath  P, Yang  T-J. Miniature inverted-repeat transposable elements (MITEs) as valuable genomic resources for the evolution and breeding of Brassica crops. Plant Breed Biotechnol. 2014:2(4):322–333. 10.9787/PBB.2014.2.4.322. [DOI] [Google Scholar]
  168. Schmidt  C, Fransz  P, Rönspies  M, Dreissig  S, Fuchs  J, Heckmann  S, Houben  A, Puchta  H. Changing local recombination patterns in Arabidopsis by CRISPR/Cas mediated chromosome engineering. Nat Commun. 2020:11(1):4418. 10.1038/s41467-020-18277-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Schmidt  N, Seibt  KM, Weber  B, Schwarzacher  T, Schmidt  T, Heitkam  T. Broken, silent, and in hiding: tamed endogenous pararetroviruses escape elimination from the genome of sugar beet (Beta vulgaris). Ann Bot. 2021:128(3):281–299. 10.1093/aob/mcab042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Schmidt  T, Heitkam  T, Liedtke  S, Schubert  V, Menzel  G. Adding color to a century-old enigma: multi-color chromosome identification unravels the autotriploid nature of saffron (Crocus sativus) as a hybrid of wild Crocus cartwrightianus cytotypes. New Phytol. 2019:222(4):1965–1980. 10.1111/nph.15715. [DOI] [PubMed] [Google Scholar]
  171. Schnable  PS, Ware  D, Fulton  RS, Stein  JC, Wei  F, Pasternak  S, Liang  C, Zhang  J, Fulton  L, Graves  TA, et al.  The B73 maize genome: complexity, diversity, and dynamics. Science. 2009:326(5956):1112–1115. 10.1126/science.1178534. [DOI] [PubMed] [Google Scholar]
  172. Schrader  L, Schmitz  J. The impact of transposable elements in adaptive evolution. Mol Ecol. 2019:28(6):1537–1549. 10.1111/mec.14794. [DOI] [PubMed] [Google Scholar]
  173. Schubert  I. Mobile nucleolus organizing regions (NORs) in Allium (Liliaceae s. lat.)?—inferences from the specifity of silver staining. Plant Syst Evol. 1984:144(3–4):291–305. 10.1007/BF00984139. [DOI] [Google Scholar]
  174. Schubert  I, Wobus  U. In situ hybridization confirms jumping nucleolus organizing regions in Allium. Chromosoma. 1985:92(2):143–148. 10.1007/BF00328466. [DOI] [Google Scholar]
  175. Seibt  KM, Schmidt  T, Heitkam  T. FlexiDot: highly customizable, ambiguity-aware dotplots for visual sequence analyses. Bioinformatics. 2018:34(20):3575–3577. 10.1093/bioinformatics/bty395. [DOI] [PubMed] [Google Scholar]
  176. Seibt  KM, Schmidt  T, Heitkam  T. The conserved 3′ Angio-domain defines a superfamily of short interspersed nuclear elements (SINEs) in higher plants. Plant J. 2020:101(3):681–699. 10.1111/tpj.14567. [DOI] [PubMed] [Google Scholar]
  177. Setiawan  AB, Teo  CH, Kikuchi  S, Sassa  H, Kato  K, Koba  T. Chromosomal locations of a non-LTR retrotransposon, Menolird18, in Cucumis melo and Cucumis sativus, and its implication on genome evolution of Cucumis species. Cytogenet Genome Res. 2020:160(9):554–564. 10.1159/000511119. [DOI] [PubMed] [Google Scholar]
  178. Shahid  S, Slotkin  RK. The current revolution in transposable element biology enabled by long reads. Curr Opin Plant Biol. 2020:54:49–56. 10.1016/j.pbi.2019.12.012. [DOI] [PubMed] [Google Scholar]
  179. Shibata  F, Hizume  M. Evolution of 5S rDNA units and their chromosomal localization in Allium cepa and Allium schoenoprasum revealed by microdissection and FISH. Theor Appl Genet. 2002:105(2):167–172. 10.1007/s00122-002-0950-0. [DOI] [PubMed] [Google Scholar]
  180. Silva-Sousa  R, López-Panadès  E, Casacuberta  E. Drosophila telomeres: an example of co-evolution with transposable elements. Genome Dyn. 2012:7:46–67. 10.1159/000337127. [DOI] [PubMed] [Google Scholar]
  181. Sinclair  DA, Guarente  L. Extrachromosomal rDNA circles—a cause of aging in yeast. Cell. 1997:91(7):1033–1042. 10.1016/S0092-8674(00)80493-6. [DOI] [PubMed] [Google Scholar]
  182. Smith  CJ, Castanon  O, Said  K, Volf  V, Khoshakhlagh  P, Hornick  A, Ferreira  R, Wu  CT, Güell  M, Garg  S, et al.  Enabling large-scale genome editing at repetitive elements by reducing DNA nicking. Nucleic Acids Res. 2020:48(9):5183–5195. 10.1093/nar/gkaa239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Sochorová  J, Garcia  S, Gálvez  F, Symonová  R, Kovařík  A. Evolutionary trends in animal ribosomal DNA loci: introduction to a new online database. Chromosoma. 2018:127(1):141–150. 10.1007/s00412-017-0651-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Sultana  T, Zamborlini  A, Cristofari  G, Lesage  P. Integration site selection by retroviruses and transposable elements in eukaryotes. Nat Rev Genet. 2017:18(5):292–308. 10.1038/nrg.2017.7. [DOI] [PubMed] [Google Scholar]
  185. Sultanov  D, Hochwagen  A. Varying strength of selection contributes to the intragenomic diversity of rRNA genes. Nat Commun. 2022:13(1):7245. 10.1038/s41467-022-34989-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Symonová  R, Majtánová  Z, Sember  A, Staaks  GBO, Bohlen  J, Freyhof  J, Rábová  M, Ráb  P. Genome differentiation in a species pair of coregonine fishes: an extremely rapid speciation driven by stress-activated retrotransposons mediating extensive ribosomal DNA multiplications. BMC Evol Biol. 2013:13:42. 10.1186/1471-2148-13-42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Symonová  R, Ocalewicz  K, Kirtiklis  L, Delmastro  GB, Pelikánová  Š, Garcia  S, Kovařík  A. Higher-order organisation of extremely amplified, potentially functional and massively methylated 5S rDNA in European pikes (Esox sp.). BMC Genomics. 2017:18(1):391. 10.1186/s12864-017-3774-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Tamayo-Ordóñez  YJ, Narváez-Zapata  JA, Tamayo-Ordóñez  MC, Sánchez-Teyer  LF. Retroelements and DNA methylation could contribute to diversity of 5S rDNA in Agave L. J Mol Evol. 2018:86(6):404–423. 10.1007/s00239-018-9856-6. [DOI] [PubMed] [Google Scholar]
  189. TE Hub Consortium, Elliot  TA, Heitkam  T, Hubley  R, Quesneville  H, Suh  A, Wheeler  TJ. TE Hub: a community-oriented space for sharing and connecting tools, data, resources, and methods for transposable element annotation. Mob DNA. 2021:12(1):16. 10.1186/s13100-021-00244-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Tulpová  Z, Kovařík  A, Toegelová  H, Navrátilová  P, Kapustová  V, Hřibová  E, Vrána  J, Macas  J, Doležel  J, Šimková  H. Fine structure and transcription dynamics of bread wheat ribosomal DNA loci deciphered by a multi-omics approach. Plant Genome. 2022:15(1):e20191. 10.1002/tpg2.20191. [DOI] [PubMed] [Google Scholar]
  191. van’t Hof  AE, Campagne  P, Rigden  DJ, Yung  CJ, Lingley  J, Quail  MA, Hall  N, Darby  AC, Saccheri  IJ. The industrial melanism mutation in British peppered moths is a transposable element. Nature. 2016:534(7605):102–105. 10.1038/nature17951. [DOI] [PubMed] [Google Scholar]
  192. Vassetzky  NS, Kramerov  DA. SINEBase: a database and tool for SINE analysis. Nucleic Acids Res. 2013:41(D1):83–89. 10.1093/nar/gks1263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Vincent  A, Petes  TD. Isolation and characterization of a Ty element inserted into the ribosomal DNA of the yeast Saccharomyces cerevisiae. Nucleic Acids Res. 1986:14(7):2939–2949. 10.1093/nar/14.7.2939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Wang  J, Han  G-Z. Unearthing LTR retrotransposon gag genes co-opted in the deep evolution of eukaryotes. Mol Biol Evol. 2021:38(8):3267–3278. 10.1093/molbev/msab101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Wang  J, Wang  A, Han  Z, Zhang  Z, Li  F, Li  X. Characterization of three novel SINE families with unusual features in Helicoverpa armigera. PLoS ONE. 2012:7(2):e31355. 10.1371/journal.pone.0031355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Wang  W, Zhang  X, Garcia  S, Leitch  AR, Kovařík  A. Intragenomic rDNA variation-the product of concerted evolution, mutation, or something in between?  Heredity (Edinb).  2023:131(3):179–188. 10.1038/s41437-023-00634-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Watada  E, Li  S, Hori  Y, Fujiki  K, Shirahige  K, Inada  T, Kobayashi  T. Age-dependent ribosomal DNA variations in mice. Mol Cell Biol. 2020:40(22):e00368-20. 10.1128/MCB.00368-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Weber  B, Heitkam  T, Holtgräwe  D, Weisshaar  B, Minoche  AE, Dohm  JC, Himmelbauer  H, Schmidt  T. Highly diverse chromoviruses of Beta vulgaris are classified by chromodomains and chromosomal integration. Mob DNA. 2013:4(1):8. 10.1186/1759-8753-4-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Wells  JN, Feschotte  C. A field guide to eukaryotic transposable elements. Annu Rev Genet. 2020:54(1):539–561. 10.1146/annurev-genet-040620-022145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Williams  SM, Robbins  LG, Cluster  PD, Allard  RW, Strobeck  C. Superstructure of the Drosophila ribosomal gene family. Proc Natl Acad Sci U S A. 1990:87(8):3156–3160. 10.1073/pnas.87.8.3156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Wollrab  C, Heitkam  T, Holtgräwe  D, Weisshaar  B, Minoche  AE, Dohm  JC, Himmelbauer  H, Schmidt  T. Evolutionary reshuffling in the Errantivirus lineage Elbe within the Beta vulgaris genome. Plant J. 2012:72(4):636–651. 10.1111/j.1365-313X.2012.05107.x. [DOI] [PubMed] [Google Scholar]
  202. Xiong  Y, Eickbush  TH. The site-specific ribosomal DNA insertion element R1Bm belongs to a class of non-long-terminal-repeat retrotransposons. Mol Cell Biol. 1988:8(1):114–123. 10.1128/mcb.8.1.114-123.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Yang  N, Srivastav  SP, Rahman  R, Ma  Q, Dayama  G, Li  S, Chinen  M, Lei  EP, Rosbash  M, Lau  NC. Transposable element landscapes in aging Drosophila. PLoS Genet. 2022:18(3):e1010024. 10.1371/journal.pgen.1010024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Yang  F, Su  W, Chung  OW, Tracy  L, Wang  L, Ramsden  DA, Zhang  ZZ. Retrotransposons hijack alt-EJ for DNA replication and eccDNA biogenesis. Nature. 2023:620(7972):218–225. 10.1038/s41586-023-06327-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Yano  CF, Merlo  MA, Portela-Bens  S, Cioffi  MDB, Bertollo  LAC, Santos-Júnior  CD, Rebordinos  L. Evolutionary dynamics of multigene families in Triportheus (Characiformes, Triportheidae): a transposon mediated mechanism?  Front Mar Sci. 2020:7:6. 10.3389/fmars.2020.00006. [DOI] [Google Scholar]
  206. Yin  H, Du  J, Li  L, Jin  C, Fan  L, Li  M, Wu  J, Zhang  S. Comparative genomic analysis reveals multiple long terminal repeats, lineage-specific amplification, and frequent interelement recombination for Cassandra retrotransposon in pear (Pyrus bretschneideri Rehd.). Genome Biol Evol. 2014:6(6):1423–1436. 10.1093/gbe/evu114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Yushkova  E, Moskalev  A. Transposable elements and their role in aging. Ageing Res Rev. 2023:86:101881. 10.1016/j.arr.2023.101881. [DOI] [PubMed] [Google Scholar]
  208. Zhang  X, Eickbush  TH. Characterization of active R2 retrotransposition in the rDNA locus of Drosophila simulans. Genetics. 2005:170(1):195–205. 10.1534/genetics.104.038703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Zhang  M, Tang  Y-W, Xu  Y, Yonezawa  T, Shao  Y, Wang  Y-G, Song  Z-P, Yang  J, Zhang  W-J. Concerted and birth-and-death evolution of 26S ribosomal DNA in Camellia L. Ann Bot. 2021:127(1):63–73. 10.1093/aob/mcaa169. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Not applicable.


Articles from Molecular Biology and Evolution are provided here courtesy of Oxford University Press

RESOURCES