Skip to main content

Retrotransposition in tumors and brains

Abstract

LINE-1s (L1s), the only currently active autonomous mobile DNA in humans, occupy at least 17% of human DNA. Throughout evolution, the L1 has also been responsible for genomic insertion of thousands of processed pseudogenes and over one million nonautonomous retrotransposons called SINEs (mainly Alus and SVAs). The 6-kb human L1 has a 5′- untranslated region (UTR) that functions as an internal promoter, two open reading frames—ORF1, which encodes an RNA-binding protein, and ORF2, which expresses endonuclease and reverse transcriptase activities—and a 3′-UTR which ends in a poly(A) signal and tail. Most L1s are molecular fossils: truncated, rearranged or mutated. However, 80 to 100 remain potentially active in any human individual, and to date 101 de novo disease-causing germline retrotransposon insertions have been characterized. It is now clear that significant levels of retrotransposition occur not only in the human germline but also in some somatic cell types. Recent publications and new investigations under way suggest that this may especially be the case for cancers and neuronal cells. This commentary offers a few points to consider to aid in avoiding misinterpretation of data as these studies move forward.

Retrotransposition of non-long terminal repeat (non-LTR) long interspersed nuclear elements (LINE-1 s, or L1s), as well as the mobilization in trans of non-autonomous short interspersed nuclear elements (SINEs) and processed pseudogenes, has built at least 50% of the human genome and remains an ongoing source of gene mutation [1, 2]. As a type of “selfish DNA”, L1 activity was formerly thought to occur predominantly in germ cells, where insertions would pass to the next generation. However, in addition to the massive germline expansion of L1s that occurred during mammalian evolution, recent investigations have documented ongoing retrotransposition in selected somatic cell types, including neural progenitor cells, stem cells, early embryos, tumors and induced pluripotent stem cells [37]. More than 20 years ago, Miki et al. [8] reported the first instance of somatic retrotransposition, an L1 insertion into the adenomatous polyposis coli tumor suppressor gene of a colorectal cancer. The advent of high-throughput sequencing has made it possible to identify numerous non-germline de novo insertions in various kinds of cancer, as recently described in several high-profile papers [6, 913] (Table 1).

Table 1 Summary of published evidence for tumor-specific somatic retrotransposition a

Cancer-associated hypomethylation and elevated transcription of L1s predicted increased retrotransposition in tumors long before new insertions were detected by next-generation sequencing [16]. It is also reasonable to assume that insertions proliferate preferentially in tumors because cancer cells divide more rapidly than their normal cells of origin. Indeed, cell cycling, though not strictly required for retrotransposition, may increase its frequency [1719]. Interestingly, investigations to date have detected new insertion events almost exclusively in tumors of epithelial cell types, some of which proliferate and turn over quickly. Carreira et al. [20] speculated that increased retrotransposon insertions in epithelial tumors may relate to a greater “plasticity” of epithelial cells, which are more easily reprogrammed to yield cancer or pluripotent stem cells.

Recent high-throughput sequence studies have reported tumor-specific insertions that vary greatly in number between different tumors of the same type, ranging from 0 in most instances to 106 in a single colorectal tumor identified by Lee et al. [9]. The application of different methodologies clearly accounts for some of this variation. Furthermore, current sequence analysis pipelines lack sensitivity to detect rare insertion events that occur late in tumor development, thus underestimating the total number of tumor-specific insertions. On the other hand, according to the scenario of Figure 1A, claims for the tumor specificity of much de novo retrotransposition and its absence in matched normal cells might be illusory. This scenario assumes that retrotransposition occurs at significant rates in normal somatic cells. However, in non-tumor tissue sampled in bulk, an individual new somatic insertion may be present in only a single cell, or at most several cells, among the large total population of cells sampled and consequently exist in too small a copy number to be detected. An insertion initially present within a normal cell is more easily detected once that cell by chance clonally expands as a tumor, which, upon sampling, high-throughput sequencing and PCR validation, would falsely appear to possess a tumor-only event. Thus, although tumor-specific de novo retrotransposition events may be underestimated, retrotransposition in normal cells is likely grossly underestimated.

Figure 1
figure 1

Bulk tissue vs. single-cell detection of somatic retrotransposition. (A) Bulk tissue sampling can underestimate the number of normal tissue retrotransposition events. (B) Single-cell sampling can provide truer estimates of rates of normal somatic cell retrotransposition. The numbers of unique de novo retrotransposon events in each cell are indicated. A minor percentage of normal epithelial cells (light brown) contain new unique insertions. Occasionally, a cancer stem cell (dark brown) gives rise to a tumor (green). Every cell of the tumor will contain the de novo retrotransposon insertions present in the originating stem cell. New tumor-specific events may also subsequently occur. Red circles indicate bulk (A) or single-cell (B) sampling for downstream sequencing analyses.

Solyom et al. [10] adopted three indirect approaches to conclude that most, if not all, of the insertions they studied occurred post-tumor initiation: (1) by finding an L1 insertion to be absent in a second section of the same tumor, (2) by detecting empty-site X chromosome alleles in males with an X chromosome tumor insertion, and (3) by querying the heterozygosity of single-nucleotide polymorphisms (SNPs) flanking an L1 insertion site (with the presence of both SNP alleles in the empty-site chromosome implying that the insertion occurred after the one-cell stage of the tumor). Because the tumor samples assayed in this study were not microdissected, however, all three pieces of evidence cited may be subject to misinterpretation because of normal tissue present within the tumor sample. Interestingly, Shukla et al. [11] validated by PCR a single de novo L1 insertion in preneoplasmic liver tissue that was absent in the corresponding hepatocellular carcinoma, although they also considered the possibility that chromosomal loss within the tumor could have deleted the insertion.

Thus, tumor-specific retrotransposon insertions occur, perhaps frequently in some tumors, but apparently vary greatly in number between different types of tumors and between individual tumors of the same cancer type. Although 80 to 100 L1s are estimated to be potentially active in any given human diploid genome [21], each particular genome can harbor its own unique, active L1s, or L1s shared between different individuals may vary in activity. My “hot” L1 may not be your hot L1; perhaps my mobilome has greater “mutational power” than yours [22, 23]. Add to this fact the variability in the epigenetic state of individual active L1s, plus unknown genetic variations in the many cellular factors that associate with the L1 to affect its life cycle [2426], and the task of assessing rates of cancer retrotransposition becomes complicated indeed.

High-throughput sequencing methods can produce false-positive results, underlining the need for validation of a significant number of the total putative somatic insertions found. As much as possible, it is important to identify 5′ as well as 3′ junctions of an insertion to confirm the presence of a poly(A) tail and a target site duplication (TSD), the hallmarks of a true retrotransposition event. A minor but significant number of L1s insert not by target-primed reverse transcription (TPRT), the standard model for L1 retrotransposition [27], but rather by an endonuclease-independent mechanism. These insertion events likely occur at preexisting DNA lesions and generate integrants lacking TSDs [2830]. Interestingly, apparent endonuclease-independent insertions have been detected in significantly increased numbers in some tumors (8 of 35 colorectal cancer insertions reported by Solyom et al. [10]).

Obtaining unbiased estimates of de novo retrotransposition in normal as well as tumor cells is critical for understanding somatic mosaicism, cancer induction, tumor heterogeneity, and the etiology of some neurological diseases [31, 32]. Evrony et al. [33] recently used multiple displacement amplification of single neurons isolated from the frontal cortex and caudate nucleus of three normal individuals to confirm somatic neuronal retrotransposition, but at a rate much lower (0.04 to fewer than 0.6 unique insertions per neuron) than estimates previously suggested by quantitative PCR (qPCR) analyses (a startling “theoretical” increase of about 80 L1 copies per hippocampal neuron compared with heart and liver samples [34]). Using a high-throughput method called RC-seq to analyze bulk DNA, Baillie et al. [15] found almost 8,000 putative somatic L1 insertions in the hippocampus and caudate nucleus of three individuals, an insertion rate much lower than Coufal et al.[34]a.

qPCR techniques are increasingly being used to conclude elevated L1 genomic copy numbers in some cell types or under some cellular conditions [3439]. Typically, the apparent increase in retrotransposon insertions is not verified by downstream sequencing. Apart from the fact that small changes in protocol can alter the results of such sensitive qPCR analyses, an additional but untested source of bias conceivably exists: that is, the promiscuous reverse transcription (RT) of retrotransposon RNAs “free-floating” in the cell and not engaged in TPRT at a site of chromatin integration. Such ectopic RT reactions might be primed by random complementary nucleic acids or perhaps by fold-back and annealing of the L1 poly(A) tail to one of the ten homopolymeric stretches of four to seven U residues that occur across the length of the L1 RNA molecule. Fold-back self-priming of RNA has been observed for a number of RNA-dependent RNA polymerases, including reverse transcriptases [4042]. It has also been shown that RT can initiate from internal sites within L1 RNA during endonuclease-independent insertion [28]. One would expect copy numbers of orphan L1 cDNAs generated by promiscuous RT to be higher in cells with elevated expression of L1 ORF2 or perhaps endogenous retroviral reverse transcriptases; indeed, increased RT activity in some types of neuronal and tumor cells has been reported [4345]. Because these cDNAs are amenable to PCR amplification, qPCR-based estimates of genomic L1 insertion copy numbers in these cells using L1-specific primers would be erroneously high. This source of bias, if true, is not in disagreement with the cis-preference model for L1 retrotransposition. This model states that a retrotransposition-competent integration intermediate consists of L1 RNA bound in cis by its own encoded ORF1 and ORF2 proteins [46, 47]. However, the total number of L1 protein [48] and RNA molecules present in the cell likely greatly exceed in number those bound in cis within bona fide insertion intermediates. These molecules can bind to each other in trans and perhaps engage in RT.

Conclusions

The surprising discovery of frequent somatic retrotransposition has important implications for human health. To fully and accurately assess its extent, concerted efforts are needed to sequence many single cells from a large number of cell types (normal and cancerous) from many individuals, with thorough validation by PCR of individual retrotransposon inserts and capillary sequencing of the PCR products to confirm their identity (Figure 1B). Single-cell, whole-genome amplification is a recent technology and is not without biases, including allelic and locus dropouts, chimeric molecules, uneven amplification due to local variations in G-C content, and incorrect nucleotide insertions [33, 49, 50]. Such artifacts complicate the identification of somatic retrotransposon insertions. However, the ongoing development of new protocols, such as MALBAC (multiple annealing looping-based amplification cycles [51]), promises to reduce bias. Next-generation, single-cell sequencing protocols should help to establish the impact that ongoing L1 retrotransposition manifests in brain biology, cancer, and likely other human diseases where L1 mobilization has not yet been explored. It should also become possible to trace the history of a single somatic retrotransposon back to its origin in a specific cell type or at a specific stage of development.

Endnote

aSix RC-Seq libraries, each prepared from 2.5 μg of DNA, yielded 7743 L1 insertions [15]. Since a human diploid cell contains 6.6 pg of DNA, there were on average .003 unique L1 insertions per brain cell. If most insertions were in neurons, the rate was 1 insertion per 30 to 150 neurons (the brain being variously estimated to be 10-50% neurons [52]).

Abbreviations

ERV:

Endogenous retrovirus

LINE:

Long interspersed nuclear element

LTR:

Long-terminal repeat

ORF:

Open reading frame

RC-seq:

Retrotransposon capture sequencing

SINE:

Short interspersed nuclear element

SVA:

SINE-R, VNTR and Alu

TSD:

Target site duplication

UTR:

Untranslated region

TPRT:

Target-primed reverse transcription

WGS:

Whole-genome sequencing.

References

  1. Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, Devon K, Dewar K, Doyle M, FitzHugh W, Funke R, Gage D, Harris K, Heaford A, Howland J, Kann L, Lehoczky J, LeVine R, McEwan P, McKernan K, Meldrim J, Mesirov JP, Miranda C, Morris W, Naylor J, Raymond C, Rosetti M, Santos R, Sheridan A, Sougnez C, International Human Genome Sequencing Consortium: Initial sequencing and analysis of the human genome. Nature 2001, 409: 860-921. A published erratum appears in Nature 2001, 412: 565 10.1038/35057062

    Article  CAS  PubMed  Google Scholar 

  2. Hancks DC, Kazazian HH Jr: Active human retrotransposons: variation and disease. Review. Curr Opin Genet Dev 2012, 22: 191-203. 10.1016/j.gde.2012.02.006

    Article  CAS  PubMed  Google Scholar 

  3. Muotri AR, Chu VT, Marchetto MCN, Deng W, Moran JV, Gage FH: Somatic mosaicism in neuronal precursor cells mediated by L1 retrotransposition. Nature 2005, 435: 903-910. 10.1038/nature03663

    Article  CAS  PubMed  Google Scholar 

  4. Garcia-Perez JL, Marchetto MC, Muotri AR, Coufal NG, Gage FH, O'Shea KS, Moran JV: LINE-1 retrotransposition in human embryonic stem cells. Hum Mol Genet 2007, 16: 1569-1577. 10.1093/hmg/ddm105

    Article  CAS  PubMed  Google Scholar 

  5. Kano H, Godoy I, Courtney C, Vetter MR, Gerton GL, Ostertag EM, Kazazian HH Jr: L1 retrotransposition occurs mainly in embryogenesis and creates somatic mosaicism. Genes Dev 2009, 23: 1303-1312. 10.1101/gad.1803909

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  6. Iskow RC, McCabe MT, Mills RE, Torene S, Pittard WS, Neuwald AF, Van Meir EG, Vertino PM, Devine SE: Natural mutagenesis of human genomes by endogenous retrotransposons. Cell 2010, 141: 1253-1261. 10.1016/j.cell.2010.05.020

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  7. Wissing S, Muñoz-Lopez M, Macia A, Yang Z, Montano M, Collins W, Garcia-Perez JL, Moran JV, Greene WC: Reprogramming somatic cells into iPS cells activates LINE-1 retroelement mobility. Hum Mol Genet 2012, 21: 208-218. 10.1093/hmg/ddr455

    Article  PubMed Central  PubMed  Google Scholar 

  8. Miki Y, Nishisho I, Horii A, Miyoshi Y, Utsunomiya J, Kinzler KW, Vogelstein B, Nakamura Y: Disruption of the APC gene by a retrotransposal insertion of L1 sequence in a colon cancer. Cancer Res 1992, 52: 643-645.

    CAS  PubMed  Google Scholar 

  9. Lee E, Iskow R, Yang L, Gokcumen O, Haseley P, Luquette LJ 3rd, Lohr JG, Harris CC, Ding L, Wilson RK, Wheeler DA, Gibbs RA, Kucherlapati R, Lee C, Kharchenko PV, Park PJ, Cancer Genome Atlas Research Network: Landscape of somatic retrotransposition in human cancers. Science 2012, 337: 967-971. 10.1126/science.1222077

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  10. Solyom S, Ewing AD, Rahrmann E, Doucet T, Nelson H, Burns B, Harris R, Sigmon D, Casella A, Erlanger B, Wheelan S, Upton KR, Shukla R, Faulkner GJ, Largaespada DA, Kazazian HH Jr: Extensive somatic L1 retrotransposition in colorectal tumors. Genome Res 2012, 22: 2328-2238. 10.1101/gr.145235.112

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Shukla R, Upton KR, Muñoz-Lopez M, Gerhardt DJ, Fisher ME, Nguyen T, Brennan PM, Baillie JK, Collino A, Ghisletti S, Sinha S, Iannelli F, Radaelli E, Dos Santos A, Rapoud D, Guettier C, Samuel D, Natoli G, Carninci P, Ciccarelli FD, Garcia-Perez JL, Faivre J, Faulkner GJ: Endogenous retrotransposition activates oncogenic pathways in hepatocellular carcinoma. Cell 2013, 153: 101-111. 10.1016/j.cell.2013.02.032

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  12. Ewing AD, Ballinger TJ, Earl D, Harris CC, Ding L, Wilson RK, Haussler D, Broad Institute Genome Sequencing and Analysis Program and Platform: Retrotransposition of gene transcripts leads to structural variation in mammalian genomes. Genome Biol 2013, 14: R22. 10.1186/gb-2013-14-3-r22

    Article  PubMed Central  PubMed  Google Scholar 

  13. Helman E, Lawrence M, Stewart C, Getz G, Meyerson M: Identification of somatic retrotransposon insertions across cancer types using RetroSeq. Cancer Res 2012,72(8 Suppl 1):5020. In 103rd Annual Meeting of the American Society for Cancer Research (2012 March 31–2012 April 4, Chicago, IL, USA) Abstract 5060.

    Google Scholar 

  14. Ewing AD, Kazazian HH Jr: High-throughput sequencing reveals extensive variation in human-specific L1 content in individual human genomes. Genome Res 2010, 20: 1262-1270. 10.1101/gr.106419.110

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  15. Baillie JK, Barnett MW, Upton KR, Gerhardt DJ, Richmond TA, De Sapio F, Brennan PM, Rizzu P, Smith S, Fell M, Talbot RT, Gustincich S, Freeman TC, Mattick JS, Hume DA, Heutink P, Carninci P, Jeddeloh JA, Faulkner GJ: Somatic retrotransposition alters the genetic landscape of the human brain. Nature 2011, 479: 534-537. 10.1038/nature10531

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Ehrlich M: DNA methylation in cancer: too much, but also too little. Oncogene 2002, 21: 5400-5413. 10.1038/sj.onc.1205651

    Article  CAS  PubMed  Google Scholar 

  17. Kubo S, Seleme MC, Soifer HS, Perez JL, Moran JV, Kazazian HH Jr, Kasahara N: L1 retrotransposition in nondividing and primary human somatic cells. Proc Natl Acad Sci USA 2006, 103: 8036-8041. 10.1073/pnas.0601954103

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  18. Shi X, Seluanov A, Gorbunova V: Cell divisions are required for L1 retrotransposition. Mol Cell Biol 2007, 27: 1264-1270. 10.1128/MCB.01888-06

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Xie Y, Mates L, Ivics Z, Izsvák Z, Martin SL, An W: Cell division promotes efficient retrotransposition in a stable L1 reporter cell line. Mobile DNA 2013, 4: 10. 10.1186/1759-8753-4-10

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  20. Carreira PE, Richardson SR, Faulkner GJ: L1 retrotransposons, cancer stem cells and oncogenesis. FEBS J 2014, 281: 63-73. 10.1111/febs.12601

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  21. Brouha B, Schustak J, Badge RM, Lutz-Prigge S, Farley AH, Moran JV, Kazazian HH Jr: Hot L1s account for the bulk of retrotransposition in the human population. Proc Natl Acad Sci USA 2003, 100: 5280-5285. 10.1073/pnas.0831042100

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  22. Seleme MC, Vetter MR, Cordaux R, Bastone L, Batzer MA, Kazazian HH Jr: Extensive individual variation in L1 retrotransposition capability contributes to human genetic diversity. Proc Natl Acad Sci USA 2006, 103: 6611-6616. 10.1073/pnas.0601324103

    Article  CAS  PubMed  Google Scholar 

  23. Beck CR, Collier P, Macfarlane C, Malig M, Kidd JM, Eichler EE, Badge RM, Moran JV: LINE-1 retrotransposition activity in human genomes. Cell 2010, 141: 1159-1170. 10.1016/j.cell.2010.05.021

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  24. Reiss D, Mager DL: Stochastic epigenetic silencing of retrotransposons: does stability come with age? Gene 2007, 390: 130-135. 10.1016/j.gene.2006.07.032

    Article  CAS  PubMed  Google Scholar 

  25. Goodier JL, Cheung LE, Kazazian HH Jr: Mapping the LINE1 ORF1 protein interactome reveals associated inhibitors of human retrotransposition. Nucleic Acids Res 2013, 41: 7401-7419. 10.1093/nar/gkt512

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  26. Taylor M, LaCava J, Mita P, Molloy KR, Huang CRL, Li D, Adney EM, Jiang H, Burns KH, Chait BT, Rout MP, Boeke JD, Dai L: Affinity proteomics reveals human host factors implicated in discrete stages of LINE-1 retrotransposition. Cell 2013, 155: 1034-1048. 10.1016/j.cell.2013.10.021

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  27. Luan DD, Korman MH, Jakubczak JL, Eickbush TH: Reverse transcription of R2Bm RNA is primed by a nick at the chromosomal target site: a mechanism for non-LTR retrotransposition. Cell 1993, 72: 595-605. 10.1016/0092-8674(93)90078-5

    Article  CAS  PubMed  Google Scholar 

  28. Morrish TA, Gilbert N, Myers JS, Vincent BJ, Stamato TD, Taccioli GE, Batzer MA, Moran JV: DNA repair mediated by endonuclease-independent LINE-1 retrotransposition. Nat Genet 2002, 31: 159-165. 10.1038/ng898

    Article  CAS  PubMed  Google Scholar 

  29. Morrish TA, Garcia-Perez JL, Stamato TD, Taccioli GE, Sekiguchi J, Moran JV: Endonuclease-independent LINE-1 retrotransposition at mammalian telomeres. Nature 2007, 446: 208-212. 10.1038/nature05560

    Article  CAS  PubMed  Google Scholar 

  30. Sen SK, Huang CT, Han K, Batzer MA: Endonuclease-independent insertion provides an alternative pathway for L1 retrotransposition in the human genome. Nucleic Acids Res 2007, 35: 3741-3751. 10.1093/nar/gkm317

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  31. Erickson RP: Somatic gene mutation and human disease other than cancer: an update. Mutat Res 2010, 705: 96-106. 10.1016/j.mrrev.2010.04.002

    Article  CAS  PubMed  Google Scholar 

  32. Poduri A, Evrony GD, Cai X, Walsh CA: Somatic mutation, genomic variation, and neurological disease. Science 2013, 341: 1237758. 10.1126/science.1237758

    Article  PubMed Central  PubMed  Google Scholar 

  33. Evrony GD, Cai X, Lee E, Hills LB, Elhosary PC, Lehmann HS, Parker JJ, Atabay KD, Gilmore EC, Poduri A, Park PJ, Walsh CA: Single-neuron sequencing analysis of L1 retrotransposition and somatic mutation in the human brain. Cell 2012, 151: 483-496. 10.1016/j.cell.2012.09.035

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Coufal NG, Garcia-Perez JL, Peng GE, Yeo GW, Mu Y, Lovci MT, Morell M, O'Shea KS, Moran JV, Gage FH: L1 retrotransposition in human neural progenitor cells. Nature 2009, 460: 1127-1131. 10.1038/nature08248

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  35. Muotri AR, Marchetto MC, Coufal NG, Oefner R, Yeo G, Nakashima K, Gage FH: L1 retrotransposition in neurons is modulated by MeCP2. Nature 2010, 468: 443-446. 10.1038/nature09544

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  36. Coufal NG, Garcia-Perez JL, Peng GE, Marchetto MC, Muotri AR, Mu Y, Carson CT, Macia A, Moran JV, Gage FH: Ataxia telangiectasia mutated (ATM) modulates long interspersed element-1 (L1) retrotransposition in human neural stem cells. Proc Natl Acad Sci USA 2011, 108: 20382-20387. 10.1073/pnas.1100273108

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  37. Jones RB, Song H, Xu Y, Garrison KE, Buzdin AA, Anwar N, Hunter DV, Mujib S, Mihajlovic V, Martin E, Lee E, Kuciak M, Raposo RA, Bozorgzad A, Meiklejohn DA, Ndhlovu LC, Nixon DF, Ostrowski MA: LINE-1 retrotransposable element DNA accumulates in HIV-1-infected cells. J Virol 2013, 87: 13307-13320. 10.1128/JVI.02257-13

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  38. Gualtieri A, Andreola F, Sciamanna I, Sinibaldi-Vallebona P, Serafino A, Spadafora C: Increased expression and copy number amplification of LINE-1 and SINE B1 retrotransposable elements in murine mammary carcinoma progression. Oncotarget 1882–1893, 2013: 4.

    Google Scholar 

  39. Bundo M, Toyoshima M, Okada Y, Akamatsu W, Ueda J, Nemoto-Miyauchi T, Sunaga F, Toritsuka M, Ikawa D, Kakita A, Kato M, Kasai K, Kishimoto T, Nawa H, Okano H, Yoshikawa T, Kato T, Iwamoto K: Increased L1 retrotransposition in the neuronal genome in schizophrenia. Neuron 2014, 81: 306-313. 10.1016/j.neuron.2013.10.053

    Article  CAS  PubMed  Google Scholar 

  40. Levin HL: An unusual mechanism of self-primed reverse transcription requires the RNase H domain of reverse transcriptase to cleave an RNA duplex. Mol Cell Biol 1996, 16: 5645-5654.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  41. Laurila MR, Salgado PS, Stuart DI, Grimes JM, Bamford DH: Back-priming mode of Φ6 RNA-dependent RNA polymerase. J Gen Virol 2005, 86: 521-526. 10.1099/vir.0.80492-0

    Article  CAS  PubMed  Google Scholar 

  42. Tuiskunen A, Leparc-Goffart I, Boubis L, Monteil V, Klingström J, Tolou HJ, Lundkvist A, Plumet S: Self-priming of reverse transcriptase impairs strand-specific detection of dengue virus RNA. J Gen Virol 2010, 91: 1019-1027. 10.1099/vir.0.016667-0

    Article  CAS  PubMed  Google Scholar 

  43. Douville R, Liu J, Rothstein J, Nath A: Identification of active loci of a human endogenous retrovirus in neurons of patients with amyotrophic lateral sclerosis. Ann Neurol 2011, 69: 141-151. 10.1002/ana.22149

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  44. Sciamanna I, Gualtieri A, Cossetti C, Osimo EF, Ferracin M, Macchia G, Aricò E, Prosseda G, Vitullo P, Misteli T, Spadafora C: A tumor-promoting mechanism mediated by retrotransposon-encoded reverse transcriptase is active in human transformed cell lines. Oncotarget 2013, 4: 2271-2287.

    Article  PubMed Central  PubMed  Google Scholar 

  45. Patnala R, Lee SH, Dahlstrom JE, Ohms S, Chen L, Dheen ST, Rangasamy D: Inhibition of LINE-1 retrotransposon-encoded reverse transcriptase modulates the expression of cell differentiation genes in breast cancer cells. Breast Cancer Res Treat 2013, 143: 239-253.

    Article  PubMed Central  PubMed  Google Scholar 

  46. Esnault C, Maestre J, Heidmann T: Human LINE retrotransposons generate processed pseudogenes. Nat Genet 2000, 24: 363-367. 10.1038/74184

    Article  CAS  PubMed  Google Scholar 

  47. Wei W, Gilbert N, Ooi SL, Lawler JF, Ostertag EM, Kazazian HH, Boeke JD, Moran JV: Human L1 retrotransposition: cis preference versus trans complementation. Mol Cell Biol 2001, 21: 1429-1439. 10.1128/MCB.21.4.1429-1439.2001

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Mandal PK, Ewing AD, Hancks DC, Kazazian HH Jr: Enrichment of processed pseudogene transcripts in L1-ribonucleoprotein particles. Hum Mol Genet 2013, 22: 3730-3748. 10.1093/hmg/ddt225

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  49. Lasken RS: Single-cell sequencing in its prime. Nat Biotechnol 2013, 31: 211-212. 10.1038/nbt.2523

    Article  CAS  PubMed  Google Scholar 

  50. Macaulay IC, Voet T: Single cell genomics: advances and future perspectives. PLoS Genet 2014, 10: e1004126. 10.1371/journal.pgen.1004126

    Article  PubMed Central  PubMed  Google Scholar 

  51. Zong C, Lu S, Chapman AR, Xie XS: Genome-wide detection of single-nucleotide and copy-number variations of a single human cell. Science 2012, 338: 1622-1626. 10.1126/science.1229164

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  52. Hilgetag C, Barbas H: Are there ten times more glia than neurons in the brain? Brain Struct Funct 2009, 213: 365-366. 10.1007/s00429-009-0202-z

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

The author thanks Drs Jose García Pérez, Szilvia Solyom, Jens Mayer, Haig H Kazazian, Geoffrey Faulkner and John Moran for critical reading of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to John L Goodier.

Additional information

Competing interests

The author declares that he has no competing interests.

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly credited. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Cite this article

Goodier, J.L. Retrotransposition in tumors and brains. Mobile DNA 5, 11 (2014). https://doi.org/10.1186/1759-8753-5-11

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1759-8753-5-11

Keywords