Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Towards safe, non-viral therapeutic gene expression in humans

Key Points

  • The development of a leukaemia-like disorder in three patients in a recent retroviral-based clinical trial underlines the need for safe and efficient approaches for achieving gene restoration.

  • Non-viral vectors, with their safety and manufacturing advantages, have been held back from clinical use because of their poor in vivo efficiency and their inability to confer long-term expression of therapeutic genes. Recent advances in our understanding of viral integrases and episomal replication might be applied to non-viral vectors, offering, for the first time, the possibility of achieving sustained non-viral gene expression.

  • Site-specific recombinases from bacteriophages and other sources have been exploited — in conjunction with non-viral vectors — to catalyze the integration of therapeutic genes into safe, defined genome locations. This approach allows genetic disorders to be corrected but avoids the risk of insertional mutagenesis.

  • So that they can be used in mammalian cells, the activity of site-specific recombinases from prokaryotic origins may be enhanced by the addition of nuclear import sequences, and by using directed evolution to alter the specificity of the recognized site of integration.

  • Rep integrase isolated from the adeno-associated virus is able to mediated efficient site-specific integration into the human genome at a safe site on chromosome 19. Recent studies have focused on overcoming problems of Rep cytotoxicity.

  • By using scaffold/nuclear matrix attachment region sequences, plasmids have been created that can replicate and segregate during mitosis in an extrachromosomal (episomal) form without either requiring viral proteins or interfering with the host DNA.

  • Most engineered human artificial chromosomes are either unstable or are difficult to fully characterize, but neocentromere-based minichromosomes have recently been created that are stable and structurally defined, showing great potential for gene therapy.

  • The various new advances make it possible for the first time to achieve long-term therapeutic gene expression in humans through non-viral approaches.

Abstract

The potential dangers of using viruses to deliver and integrate DNA into host cells in gene therapy have been poignantly highlighted in recent clinical trials. Safer, non-viral gene delivery approaches have been largely ignored in the past because of their inefficient delivery and the resulting transient transgene expression. However, recent advances indicate that efficient, long-term gene expression can be achieved by non-viral means. In particular, integration of DNA can be targeted to specific genomic sites without deleterious consequences and it is possible to maintain transgenes as small episomal plasmids or artificial chromosomes. The application of these approaches to human gene therapy is gradually becoming a reality.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Site-directed integration of foreign DNA.
Figure 2: Site-specific recombination mechanisms.
Figure 3: Site-specific integration mediated by the adeno-associated virus Rep protein.
Figure 4: Structure and visualization of an episomally replicating plasmid (pEPI) and a human artificial chromosome (HAC).

Similar content being viewed by others

References

  1. Hacein-Bey-Abina, S. et al. Sustained correction of X-linked severe combined immunodeficiency by ex vivo gene therapy. N. Engl. J. Med. 346, 1185–1193 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. Roesler, J. et al. Third-generation, self-inactivating gp91phox lentivector corrects the oxidase defect in NOD/SCID mouse-repopulating peripheral blood-mobilized CD34+ cells from patients with X-linked chronic granulomatous disease. Blood 100, 4381–4390 (2002).

    Article  CAS  PubMed  Google Scholar 

  3. Hollon, T. Researchers and regulators reflect on first gene therapy death. Nature Med. 6, 6 (2000).

    Article  CAS  PubMed  Google Scholar 

  4. Schroder, A. R. et al. HIV-1 integration in the human genome favors active genes and local hotspots. Cell 110, 521–529 (2002).

    Article  CAS  PubMed  Google Scholar 

  5. Woods, N. B. et al. Lentiviral vector transduction of NOD/SCID repopulating cells results in multiple vector integrations per transduced cell: risk of insertional mutagenesis. Blood 101, 1284–1289 (2003).

    Article  CAS  PubMed  Google Scholar 

  6. Li, Z. et al. Murine leukemia induced by retroviral gene marking. Science 296, 497 (2002).

    Article  CAS  PubMed  Google Scholar 

  7. Wu, X., Li, Y., Crise, B. & Burgess, S. M. Transcription start regions in the human genome are favored targets for MLV integration. Science 300, 1749–1751 (2003).

    Article  CAS  PubMed  Google Scholar 

  8. Hacein-Bey-Abina, S. et al. A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. N. Engl. J. Med. 348, 255–256 (2003).

    Article  PubMed  Google Scholar 

  9. Hacein-Bey-Abina, S. et al. LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science 302, 415–419 (2003). A report of the development of leukaemia in two patients that had been successfully cured of SCIDX1.

    Article  CAS  PubMed  Google Scholar 

  10. McCormack, M. P., Forster, A., Drynan, L., Pannell, R. & Rabbitts, T. H. The LMO2 T-cell oncogene is activated via chromosomal translocations or retroviral insertion during gene therapy but has no mandatory role in normal T-cell development. Mol. Cell Biol. 23, 9003–9013 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Dave, U. P., Jenkins, N. A. & Copeland, N. G. Gene therapy insertional mutagenesis insights. Science 303, 333 (2004).

    Article  PubMed  Google Scholar 

  12. Thomas, C. E., Ehrhardt, A. & Kay, M. A. Progress and problems with the use of viral vectors for gene therapy. Nature Rev. Genet. 4, 346–358 (2003). A comprehensive overview of the current and future challenges of using viruses for gene delivery.

    Article  CAS  PubMed  Google Scholar 

  13. Johnson-Saliba, M. & Jans, D. A. Gene therapy: optimising DNA delivery to the nucleus. Curr. Drug Targets 2, 371–399 (2001).

    Article  CAS  PubMed  Google Scholar 

  14. Ferber, D. Gene therapy. Safer and virus-free? Science 294, 1638–1642 (2001).

    Article  CAS  PubMed  Google Scholar 

  15. Kircheis, R. et al. Polyethylenimine/DNA complexes shielded by transferrin target gene expression to tumors after systemic application. Gene Ther. 8, 28–40 (2001).

    Article  CAS  PubMed  Google Scholar 

  16. Jans, D. A., Xiao, C. Y. & Lam, M. H. Nuclear targeting signal recognition: a key control point in nuclear transport? Bioessays 22, 532–544 (2000).

    Article  CAS  PubMed  Google Scholar 

  17. Lyman, S. K., Guan, T., Bednenko, J., Wodrich, H. & Gerace, L. Influence of cargo size on Ran and energy requirements for nuclear protein import. J. Cell Biol. 159, 55–67 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Kursa, M. et al. Novel shielded transferrin-polyethylene glycol-polyethylenimine/DNA complexes for systemic tumor-targeted gene transfer. Bioconjug. Chem. 14, 222–231 (2003).

    Article  CAS  PubMed  Google Scholar 

  19. Chan, C. K. & Jans, D. A. Enhancement of MSH receptor- and GAL4-mediated gene transfer by switching the nuclear import pathway. Gene Ther. 8, 166–171 (2001).

    Article  CAS  PubMed  Google Scholar 

  20. Kakudo, T. et al. Transferrin-modified liposomes equipped with a pH-sensitive fusogenic peptide: an artificial viral-like delivery system. Biochemistry 43, 5618–5628 (2004).

    Article  CAS  PubMed  Google Scholar 

  21. Zanta, M. A., Belguise-Valladier, P. & Behr, J. P. Gene delivery: a single nuclear localization signal peptide is sufficient to carry DNA to the cell nucleus. Proc. Natl Acad. Sci. USA 96, 91–96 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Uherek, C., Fominaya, J. & Wels, W. A modular DNA carrier protein based on the structure of diphtheria toxin mediates target cell-specific gene delivery. J. Biol. Chem. 273, 8835–8841 (1998). A classic paper showing that novel proteins engineered to overcome cellular barriers to gene delivery can efficiently transfect target cells.

    Article  CAS  PubMed  Google Scholar 

  23. Box, M. et al. A multi-domain protein system based on the HC fragment of tetanus toxin for targeting DNA to neuronal cells. J. Drug Target. 11, 333–343 (2003).

    Article  CAS  PubMed  Google Scholar 

  24. Kapsa, R. et al. In vivo and in vitro correction of the mdx Dystrophin gene nonsense mutation by short-fragment homologous replacement. Hum. Gene Ther. 12, 629–642 (2001).

    Article  CAS  PubMed  Google Scholar 

  25. Goncz, K. K., Prokopishyn, N. L., Chow, B. L., Davis, B. R. & Gruenert, D. C. Application of SFHR to gene therapy of monogenic disorders. Gene Ther. 9, 691–694 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Voziyanov, Y., Pathania, S. & Jayaram, M. A general model for site-specific recombination by the integrase family recombinases. Nucleic Acids Res. 27, 930–941 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Kolot, M., Silberstein, N. & Yagil, E. Site-specific recombination in mammalian cells expressing the Int recombinase of bacteriophage HK022. Mol. Biol. Rep. 26, 207–213 (1999).

    Article  CAS  PubMed  Google Scholar 

  28. Thyagarajan, B., Guimaraes, M. J., Groth, A. C. & Calos, M. P. Mammalian genomes contain active recombinase recognition sites. Gene 244, 47–54 (2000).

    Article  CAS  PubMed  Google Scholar 

  29. Ortiz-Urda, S. et al. Stable nonviral genetic correction of inherited human skin disease. Nature Med. 8, 1166–1170 (2002).

    Article  CAS  PubMed  Google Scholar 

  30. Olivares, E. C. et al. Site-specific genomic integration produces therapeutic Factor IX levels in mice. Nature Biotechnol. 20, 1124–1128 (2002). The first in vivo demonstration of the ability of the phage φC31 integrase to integrate a corrective gene into liver cells and express therapeutic levels of coagulation factor IX in mice.

    Article  CAS  Google Scholar 

  31. Dymecki, S. M. Flp recombinase promotes site-specific DNA recombination in embryonic stem cells and transgenic mice. Proc. Natl Acad. Sci. USA 93, 6191–6196 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Buchholz, F., Angrand, P. O. & Stewart, A. F. Improved properties of FLP recombinase evolved by cycling mutagenesis. Nature Biotechnol. 16, 657–662 (1998).

    Article  CAS  Google Scholar 

  33. Christ, N., Corona, T. & Droge, P. Site-specific recombination in eukaryotic cells mediated by mutant λ integrases: implications for synaptic complex formation and the reactivity of episomal DNA segments. J. Mol. Biol. 319, 305–314 (2002).

    Article  CAS  PubMed  Google Scholar 

  34. Esposito, D., Thrower, J. S. & Scocca, J. J. Protein and DNA requirements of the bacteriophage HP1 recombination system: a model for intasome formation. Nucleic Acids Res. 29, 3955–3964 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Olivares, E. C., Hollis, R. P. & Calos, M. P. Phage R4 integrase mediates site-specific integration in human cells. Gene 278, 167–176 (2001).

    Article  CAS  PubMed  Google Scholar 

  36. Hollis, R. P. et al. Phage integrases for the construction and manipulation of transgenic mammals. Reprod. Biol. Endocrinol. 1, 79–89 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Groth, A. C., Olivares, E. C., Thyagarajan, B. & Calos, M. P. A phage integrase directs efficient site-specific integration in human cells. Proc. Natl Acad. Sci. USA 97, 5995–6000 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Thyagarajan, B., Olivares, E. C., Hollis, R. P., Ginsburg, D. S. & Calos, M. P. Site-specific genomic integration in mammalian cells mediated by phage φC31 integrase. Mol. Cell Biol. 21, 3926–3934 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Ortiz-Urda, S. et al. φC31 integrase-mediated nonviral genetic correction of junctional epidermolysis bullosa. Hum. Gene Ther. 14, 923–928 (2003).

    Article  CAS  PubMed  Google Scholar 

  40. Ortiz-Urda, S. et al. Injection of genetically engineered fibroblasts corrects regenerated human epidermolysis bullosa skin tissue. J. Clin. Invest. 111, 251–255 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Quennevillea, S. P. et al. Nucleofection of muscle-derived stem cells and myoblasts with C31 integrase: stable expression of a full-length-dystrophin fusion gene by human myoblasts. Mol. Ther. 10, 679–687 (2004).

    Article  CAS  Google Scholar 

  42. Gagneten, S., Le, Y., Miller, J. & Sauer, B. Brief expression of a GFP cre fusion gene in embryonic stem cells allows rapid retrieval of site-specific genomic deletions. Nucleic Acids Res. 25, 3326–3331 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Peitz, M., Pfannkuche, K., Rajewsky, K. & Edenhofer, F. Ability of the hydrophobic FGF and basic TAT peptides to promote cellular uptake of recombinant Cre recombinase: a tool for efficient genetic engineering of mammalian genomes. Proc. Natl Acad. Sci. USA 99, 4489–4494 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Le, Y., Gagneten, S., Tombaccini, D., Bethke, B. & Sauer, B. Nuclear targeting determinants of the phage P1 cre DNA recombinase. Nucleic Acids Res. 27, 4703–4709 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Andreas, S., Schwenk, F., Kuter-Luks, B., Faust, N. & Kuhn, R. Enhanced efficiency through nuclear localization signal fusion on phage φC31-integrase: activity comparison with Cre and FLPe recombinase in mammalian cells. Nucleic Acids Res. 30, 2299–2306 (2002). This article describes how the addition of a nuclear localization signal improves the ability of the phage φC31 integrase to enter the eukaryotic nucleus and mediate recombination.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Stemmer, W. P. Rapid evolution of a protein in vitro by DNA shuffling. Nature 370, 389–391 (1994).

    Article  CAS  PubMed  Google Scholar 

  47. Sclimenti, C. R., Thyagarajan, B. & Calos, M. P. Directed evolution of a recombinase for improved genomic integration at a native human sequence. Nucleic Acids Res. 29, 5044–5051 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Santoro, S. W. & Schultz, P. G. Directed evolution of the site specificity of Cre recombinase. Proc. Natl Acad. Sci. USA 99, 4185–4190 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Buchholz, F. & Stewart, A. F. Alteration of Cre recombinase site specificity by substrate-linked protein evolution. Nature Biotechnol. 19, 1047–1052 (2001). This study describes an efficient and innovative approach for evolving a recombinase that recognizes new recombination sites.

    Article  CAS  Google Scholar 

  50. Huser, D., Weger, S. & Heilbronn, R. Kinetics and frequency of adeno-associated virus site-specific integration into human chromosome 19 monitored by quantitative real-time PCR. J. Virol. 76, 7554–7559 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Rizzuto, G. et al. Development of animal models for adeno-associated virus site-specific integration. J. Virol. 73, 2517–2526 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Dutheil, N., Shi, F., Dupressoir, T. & Linden, R. M. Adeno-associated virus site-specifically integrates into a muscle-specific DNA region. Proc. Natl Acad. Sci. USA 97, 4862–4866 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Surosky, R. T. et al. Adeno-associated virus Rep proteins target DNA sequences to a unique locus in the human genome. J. Virol. 71, 7951–7959 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Lamartina, S., Roscilli, G., Rinaudo, D., Delmastro, P. & Toniatti, C. Lipofection of purified adeno-associated virus Rep68 protein: toward a chromosome-targeting nonviral particle. J. Virol. 72, 7653–7658 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Young, S. M., McCarty, D. M., Degtyareva, N. & Samulski, R. J. Roles of adeno-associated virus Rep protein and human chromosome 19 in site-specific recombination. J. Virol. 74, 3953–3966 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Manno, C. S. et al. AAV-mediated factor IX gene transfer to skeletal muscle in patients with severe hemophilia B. Blood 101, 2963–2972 (2003).

    Article  CAS  PubMed  Google Scholar 

  57. Moskalenko, M. et al. Epitope mapping of human anti-adeno-associated virus type 2 neutralizing antibodies: implications for gene therapy and virus structure. J. Virol. 74, 1761–1766 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Dong, J. Y., Fan, P. D. & Frizzell, R. A. Quantitative analysis of the packaging capacity of recombinant adeno-associated virus. Hum. Gene Ther. 7, 2101–2112 (1996).

    Article  CAS  PubMed  Google Scholar 

  59. Duan, D. et al. Circular intermediates of recombinant adeno-associated virus have defined structural characteristics responsible for long-term episomal persistence in muscle tissue. J. Virol. 72, 8568–8577 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Miller, D. G., Petek, L. M. & Russell, D. W. Adeno-associated virus vectors integrate at chromosome breakage sites. Nature Genet. 36, 767–773 (2004).

    Article  CAS  PubMed  Google Scholar 

  61. Nakai, H. et al. AAV serotype 2 vectors preferentially integrate into active genes in mice. Nature Genet. 34, 297–302 (2003). This study shows that rAAV vectors (which lack rep ) integrate in a non-random fashion at high frequency into active genes of experimental animals.

    Article  CAS  PubMed  Google Scholar 

  62. Donsante, A. et al. Observed incidence of tumorigenesis in long-term rodent studies of rAAV vectors. Gene Ther. 8, 1343–1346 (2001).

    Article  CAS  PubMed  Google Scholar 

  63. Balague, C., Kalla, M. & Zhang, W. W. Adeno-associated virus Rep78 protein and terminal repeats enhance integration of DNA sequences into the cellular genome. J. Virol. 71, 3299–3306 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Pieroni, L. et al. Targeted integration of adeno-associated virus-derived plasmids in transfected human cells. Virology 249, 249–259 (1998).

    Article  CAS  PubMed  Google Scholar 

  65. Steigerwald, R. et al. Requirements for adeno-associated virus-derived non-viral vectors to achieve stable and site-specific integration of plasmid DNA in liver carcinoma cells. Digestion 68, 13–23 (2003).

    Article  CAS  PubMed  Google Scholar 

  66. Philpott, N. J. et al. Efficient integration of recombinant adeno-associated virus DNA vectors requires a p5-rep sequence in cis. J. Virol. 76, 5411–5421 (2002). This report shows that a 138 bp cis element within the p5 promoter is able to mediate efficient site-specific integration of plasmid in the absence of the AAV inverted terminal-repeat elements.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Philpott, N. J., Gomos, J. & Falck-Pedersen, E. Transgene expression after Rep-mediated site-specific integration into chromosome 19. Hum. Gene Ther. 15, 47–61 (2004).

    Article  CAS  PubMed  Google Scholar 

  68. Kogure, K. et al. Targeted integration of foreign DNA into a defined locus on chromosome 19 in K562 cells using AAV-derived components. Int. J. Hematol. 73, 469–475 (2001).

    Article  CAS  PubMed  Google Scholar 

  69. Urabe, M. et al. Positive and negative effects of adeno-associated virus Rep on AAVS1-targeted integration. J. Gen. Virol. 84, 2127–2132 (2003).

    Article  CAS  PubMed  Google Scholar 

  70. Tsunoda, H., Hayakawa, T., Sakuragawa, N. & Koyama, H. Site-specific integration of adeno-associated virus-based plasmid vectors in lipofected HeLa cells. Virology 268, 391–401 (2000).

    Article  CAS  PubMed  Google Scholar 

  71. Philpott, N. J., Gomos, J., Berns, K. I. & Falck-Pedersen, E. A p5 integration efficiency element mediates Rep-dependent integration into AAVS1 at chromosome 19. Proc. Natl Acad. Sci. USA 99, 12381–12385 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Satoh, W., Hirai, Y., Tamayose, K. & Shimada, T. Site-specific integration of an adeno-associated virus vector plasmid mediated by regulated expression of Rep based on Cre-loxP recombination. J. Virol. 74, 10631–10638 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Rinaudo, D., Lamartina, S., Roscilli, G., Ciliberto, G. & Toniatti, C. Conditional site-specific integration into human chromosome 19 by using a ligand-dependent chimeric adeno-associated virus/Rep protein. J. Virol. 74, 281–294 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Bushman, F. D. Tethering human immunodeficiency virus 1 integrase to a DNA site directs integration to nearby sequences. Proc. Natl Acad. Sci. USA 91, 9233–9237 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Holmes-Son, M. L. & Chow, S. A. Correct integration mediated by integrase-LexA fusion proteins incorporated into HIV-1. Mol. Ther. 5, 360–370 (2002).

    Article  CAS  PubMed  Google Scholar 

  76. Dildine, S. L., Respess, J., Jolly, D. & Sandmeyer, S. B. A chimeric Ty3/Moloney murine leukemia virus integrase protein is active in vivo. J. Virol. 72, 4297–4307 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Tan, W., Zhu, K., Segal, D. J., Barbas, C. F. & Chow, S. A. Fusion proteins consisting of human immunodeficiency virus type 1 integrase and the designed polydactyl zinc finger protein E2C direct integration of viral DNA into specific sites. J. Virol. 78, 1301–1313 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Ali, S. H., Kasper, J. S., Arai, T. & DeCaprio, J. A. Cul7/p185/p193 binding to simian virus 40 large T antigen has a role in cellular transformation. J. Virol. 78, 2749–2757 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Li, D. et al. Structure of the replicative helicase of the oncoprotein SV40 large tumour antigen. Nature 423, 512–518 (2003).

    Article  CAS  PubMed  Google Scholar 

  80. Humme, S. et al. The EBV nuclear antigen 1 (EBNA1) enhances B cell immortalization several thousandfold. Proc. Natl Acad. Sci. USA 100, 10989–10994 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Wu, H., Ceccarelli, D. F. & Frappier, L. The DNA segregation mechanism of Epstein–Barr virus nuclear antigen 1. EMBO Rep. 1, 140–144 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. White, R. E., Wade-Martins, R. & James, M. R. Sequences adjacent to oriP improve the persistence of Epstein–Barr virus-based episomes in B cells. J. Virol. 75, 11249–11252 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Cooper, M. J. et al. Safety-modified episomal vectors for human gene therapy. Proc. Natl Acad. Sci. USA 94, 6450–6455 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Sclimenti, C. R. et al. Epstein–Barr virus vectors provide prolonged robust factor IX expression in mice. Biotechnol. Prog. 19, 144–151 (2003).

    Article  CAS  PubMed  Google Scholar 

  85. Wade-Martins, R., White, R. E., Kimura, H., Cook, P. R. & James, M. R. Stable correction of a genetic deficiency in human cells by an episome carrying a 115 kb genomic transgene. Nature Biotechnol. 18, 1311–1314 (2000).

    Article  CAS  Google Scholar 

  86. Piechaczek, C., Fetzer, C., Baiker, A., Bode, J. & Lipps, H. J. A vector based on the SV40 origin of replication and chromosomal S/MARs replicates episomally in CHO cells. Nucleic Acids Res. 27, 426–428 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Schaarschmidt, D., Baltin, J., Stehle, I. M., Lipps, H. J. & Knippers, R. An episomal mammalian replicon: sequence-independent binding of the origin recognition complex. EMBO J. 23, 191–201 (2004).

    Article  CAS  PubMed  Google Scholar 

  88. Jenke, B. H. et al. An episomally replicating vector binds to the nuclear matrix protein SAF-A in vivo. EMBO Rep. 3, 349–354 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Mearini, G., Nielsen, P. E. & Fackelmayer, F. O. Localization and dynamics of small circular DNA in live mammalian nuclei. Nucleic Acids Res. 32, 2642–2651 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Heng, H. H. et al. Chromatin loops are selectively anchored using scaffold/matrix-attachment regions. J. Cell Sci. 117, 999–1008 (2004).

    Article  CAS  PubMed  Google Scholar 

  91. Goetze, S., Gluch, A., Benham, C. & Bode, J. Computational and in vitro analysis of destabilized DNA regions in the interferon gene cluster: potential of predicting functional gene domains. Biochemistry 42, 154–166 (2003).

    Article  CAS  PubMed  Google Scholar 

  92. Jenuwein, T. et al. Extension of chromatin accessibility by nuclear matrix attachment regions. Nature 385, 269–272 (1997).

    Article  CAS  PubMed  Google Scholar 

  93. Baiker, A. et al. Mitotic stability of an episomal vector containing a human scaffold/matrix-attached region is provided by association with nuclear matrix. Nature Cell Biol. 2, 182–184 (2000). This paper demonstrates that the specific interaction between an S/MAR-containing plasmid and the nuclear matrix can confer episomal replication and retention through cell division.

    Article  CAS  PubMed  Google Scholar 

  94. Kipp, M. et al. SAF-Box, a conserved protein domain that specifically recognizes scaffold attachment region DNA. Mol. Cell Biol. 20, 7480–7489 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Stehle, I. M., Scinteie, M. F., Baiker, A., Jenke, A. C. & Lipps, H. J. Exploiting a minimal system to study the epigenetic control of DNA replication: the interplay between transcription and replication. Chromosome Res. 11, 413–421 (2003). This report shows that active transcription of the S/MAR is required for episomal replication.

    Article  CAS  PubMed  Google Scholar 

  96. Jenke, A. C. et al. Nuclear scaffold/matrix attached region modules linked to a transcription unit are sufficient for replication and maintenance of a mammalian episome. Proc. Natl Acad. Sci. USA 101, 11322–11327 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Ohba, R., Matsumoto, K. & Ishimi, Y. Induction of DNA replication by transcription in the region upstream of the human c-myc gene in a model replication system. Mol. Cell Biol. 16, 5754–5763 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Keller, C., Ladenburger, E. M., Kremer, M. & Knippers, R. The origin recognition complex marks a replication origin in the human TOP1 gene promoter. J Biol. Chem. 277, 31430–31440 (2002).

    Article  CAS  PubMed  Google Scholar 

  99. Murray, A. W. & Szostak, J. W. Construction of artificial chromosomes in yeast. Nature 305, 189–193 (1983).

    Article  CAS  PubMed  Google Scholar 

  100. Alazami, A. M., Mejia, J. E. & Monaco, Z. L. Human artificial chromosomes containing chromosome 17 alphoid DNA maintain an active centromere in murine cells but are not stable. Genomics 83, 844–851 (2004).

    Article  CAS  PubMed  Google Scholar 

  101. Harrington, J. J., Van Bokkelen, G., Mays, R. W., Gustashaw, K. & Willard, H. F. Formation of de novo centromeres and construction of first-generation human artificial microchromosomes. Nature Genet. 15, 345–355 (1997).

    Article  CAS  PubMed  Google Scholar 

  102. Henning, K. A. et al. Human artificial chromosomes generated by modification of a yeast artificial chromosome containing both human alpha satellite and single-copy DNA sequences. Proc. Natl Acad. Sci. USA 96, 592–597 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Ebersole, T. A. et al. Mammalian artificial chromosome formation from circular alphoid input DNA does not require telomere repeats. Hum. Mol. Genet. 9, 1623–1631 (2000).

    Article  CAS  PubMed  Google Scholar 

  104. Farr, C. J. et al. Generation of a human X-derived minichromosome using telomere-associated chromosome fragmentation. EMBO J. 14, 5444–5454 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Heller, R., Brown, K. E., Burgtorf, C. & Brown, W. R. Mini-chromosomes derived from the human Y chromosome by telomere directed chromosome breakage. Proc. Natl Acad. Sci. USA 93, 7125–7130 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. du Sart, D. et al. A functional neo-centromere formed through activation of a latent human centromere and consisting of non-alpha-satellite DNA. Nature Genet 16, 144–153 (1997).

    Article  CAS  PubMed  Google Scholar 

  107. Saffery, R. et al. Construction of neocentromere-based human minichromosomes by telomere-associated chromosomal truncation. Proc. Natl Acad. Sci. USA 98, 5705–5710 (2001). This paper describes the use of telomere-directed fragmentation of a neocentomere-containing chromosome to produce a minichromsome that is mitotically stable and able to interact with several centromere-associated proteins.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Wong, L. H., Saffery, R. & Choo, K. H. Construction of neocentromere-based human minichromosomes for gene delivery and centromere studies. Gene Ther. 9, 724–726 (2002).

    Article  CAS  PubMed  Google Scholar 

  109. Wong, L. H. et al. Analysis of mitotic and expression properties of human neocentromere-based transchromosomes in mice. J. Biol. Chem. 280, 3954–3962 (2005).

    Article  CAS  PubMed  Google Scholar 

  110. Grimes, B. R. et al. Stable gene expression from a mammalian artificial chromosome. EMBO Rep. 2, 910–914 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Mejia, J. E., Willmott, A., Levy, E., Earnshaw, W. C. & Larin, Z. Functional complementation of a genetic deficiency with human artificial chromosomes. Am. J. Hum. Genet. 69, 315–326 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Auriche, C. et al. Functional human CFTR produced by a stable minichromosome. EMBO Rep. 3, 862–868 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Shen, M. H. et al. A structurally defined mini-chromosome vector for the mouse germ line. Curr. Biol. 10, 31–34 (2000).

    Article  CAS  PubMed  Google Scholar 

  114. Telenius, H. et al. Stability of a functional murine satellite DNA-based artificial chromosome across mammalian species. Chromosome Res. 7, 3–7 (1999).

    Article  CAS  PubMed  Google Scholar 

  115. Lamb, B. T. et al. Introduction and expression of the 400 kilobase amyloid precursor protein gene in transgenic mice. Nature Genet. 5, 22–30 (1993).

    Article  CAS  PubMed  Google Scholar 

  116. Lee, J. T. & Jaenisch, R. A method for high efficiency YAC lipofection into murine embryonic stem cells. Nucleic Acids Res. 24, 5054–5055 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Magin-Lachmann, C. et al. In vitro and in vivo delivery of intact BAC DNA — comparison of different methods. J. Gene Med. 6, 195–209 (2004).

    Article  CAS  PubMed  Google Scholar 

  118. Marschall, P., Malik, N. & Larin, Z. Transfer of YACs up to 2.3 Mb intact into human cells with polyethylenimine. Gene Ther. 6, 1634–1637 (1999).

    Article  CAS  PubMed  Google Scholar 

  119. White, R. E. et al. Functional delivery of large genomic DNA to human cells with a peptide-lipid vector. J. Gene Med. 5, 883–892 (2003).

    Article  CAS  PubMed  Google Scholar 

  120. Oberle, V., de Jong, G., Drayer, J. I. & Hoekstra, D. Efficient transfer of chromosome-based DNA constructs into mammalian cells. Biochim. Biophys. Acta. 1676, 223–230 (2004).

    Article  CAS  PubMed  Google Scholar 

  121. Chan, C. K. & Jans, D. A. Using nuclear targeting signals to enhance non-viral gene transfer. Immunol. Cell Biol. 80, 119–130 (2002).

    Article  CAS  PubMed  Google Scholar 

  122. Jones, P. A. & Baylin, S. B. The fundamental role of epigenetic events in cancer. Nature Rev. Genet. 3, 415–428 (2002).

    Article  CAS  PubMed  Google Scholar 

  123. Grassi, G. et al. Inhibitors of DNA methylation and histone deacetylation activate cytomegalovirus promoter-controlled reporter gene expression in human glioblastoma cell line U87. Carcinogenesis 24, 1625–1635 (2003).

    Article  CAS  PubMed  Google Scholar 

  124. Hong, K., Sherley, J. & Lauffenburger, D. A. Methylation of episomal plasmids as a barrier to transient gene expression via a synthetic delivery vector. Biomol. Eng. 18, 185–192 (2001).

    Article  CAS  PubMed  Google Scholar 

  125. Hsieh, C. L. In vivo activity of murine de novo methyltransferases, Dnmt3a and Dnmt3b. Mol. Cell Biol. 19, 8211–8218 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Scharfmann, R., Axelrod, J. H. & Verma, I. M. Long-term in vivo expression of retrovirus-mediated gene transfer in mouse fibroblast implants. Proc. Natl Acad. Sci. USA 88, 4626–4630 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Li, X., Eastman, E. M., Schwartz, R. J. & Draghia-Akli, R. Synthetic muscle promoters: activities exceeding naturally occurring regulatory sequences. Nature Biotechnol. 17, 241–245 (1999).

    Article  CAS  Google Scholar 

  128. Martinelli, R. & De Simone, V. Short and highly efficient synthetic promoters for melanoma-specific gene expression. FEBS Lett. 579, 153–156 (2005).

    Article  CAS  PubMed  Google Scholar 

  129. Villemure, J. F., Savard, N. & Belmaaza, A. Promoter suppression in cultured mammalian cells can be blocked by the chicken β-globin chromatin insulator 5'HS4 and matrix/scaffold attachment regions. J. Mol. Biol. 312, 963–974 (2001).

    Article  CAS  PubMed  Google Scholar 

  130. Ramezani, A., Hawley, T. S. & Hawley, R. G. Performance- and safety-enhanced lentiviral vectors containing the human interferon-β scaffold attachment region and the chicken β-globin insulator. Blood 101, 4717–4724 (2003).

    Article  CAS  PubMed  Google Scholar 

  131. Jenke, A. C., Scinteie, M. F., Stehle, I. M. & Lipps, H. J. Expression of a transgene encoded on a non-viral episomal vector is not subject to epigenetic silencing by cytosine methylation. Mol. Biol. Rep. 31, 85–90 (2004).

    Article  CAS  PubMed  Google Scholar 

  132. Chen, Z. Y., He, C. Y., Meuse, L. & Kay, M. A. Silencing of episomal transgene expression by plasmid bacterial DNA elements in vivo. Gene Ther. 11, 856–864 (2004).

    Article  CAS  PubMed  Google Scholar 

  133. Bigger, B. W. et al. An araC-controlled bacterial cre expression system to produce DNA minicircle vectors for nuclear and mitochondrial gene therapy. J. Biol. Chem. 276, 23018–23027 (2001).

    Article  CAS  PubMed  Google Scholar 

  134. Chen, Z. Y., He, C. Y., Ehrhardt, A. & Kay, M. A. Minicircle DNA vectors devoid of bacterial DNA result in persistent and high-level transgene expression in vivo. Mol. Ther. 8, 495–500 (2003).

    Article  CAS  PubMed  Google Scholar 

  135. Haseloff, J., Siemering, K. R., Prasher, D. C. & Hodge, S. Removal of a cryptic intron and subcellular localization of green fluorescent protein are required to mark transgenic Arabidopsis plants brightly. Proc. Natl Acad. Sci. USA 94, 2122–2127 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Sarukhan, A. et al. Successful interference with cellular immune responses to immunogenic proteins encoded by recombinant viral vectors. J. Virol. 75, 269–277 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Gross, D. A., Leboeuf, M., Gjata, B., Danos, O. & Davoust, J. CD4+CD25+ regulatory T cells inhibit immune-mediated transgene rejection. Blood 102, 4326–4328 (2003).

    Article  CAS  PubMed  Google Scholar 

  138. Dobrzynski, E. et al. Induction of antigen-specific CD4+ T-cell anergy and deletion by in vivo viral gene transfer. Blood 104, 969–977 (2004).

    Article  CAS  PubMed  Google Scholar 

  139. Honigman, A. et al. Imaging transgene expression in live animals. Mol. Ther. 4, 239–249 (2001).

    Article  CAS  PubMed  Google Scholar 

  140. Zhang, G., Budker, V., Williams, P., Subbotin, V. & Wolff, J. A. Efficient expression of naked DNA delivered intraarterially to limb muscles of nonhuman primates. Hum. Gene. Ther. 12, 427–438 (2001).

    Article  CAS  PubMed  Google Scholar 

  141. Andre, F. & Mir, L. M. DNA electrotransfer: its principles and an updated review of its therapeutic applications. Gene Ther. 11 (Suppl. 1), 33–42 (2004).

    Article  CAS  Google Scholar 

  142. Matsui, H., Johnson, L. G., Randell, S. H. & Boucher, R. C. Loss of binding and entry of liposome-DNA complexes decreases transfection efficiency in differentiated airway epithelial cells. J. Biol. Chem. 272, 1117–1126 (1997).

    Article  CAS  PubMed  Google Scholar 

  143. Filion, M. C. & Phillips, N. C. Toxicity and immunomodulatory activity of liposomal vectors formulated with cationic lipids toward immune effector cells. Biochim. Biophys. Acta 1329, 345–356 (1997).

    Article  CAS  PubMed  Google Scholar 

  144. Mislick, K. A. & Baldeschwieler, J. D. Evidence for the role of proteoglycans in cation-mediated gene transfer. Proc. Natl Acad. Sci. USA 93, 12349–12354 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Sonawane, N. D., Szoka, F. C. & Verkman, A. S. Chloride accumulation and swelling in endosomes enhances DNA transfer by polyamine–DNA polyplexes. J. Biol. Chem. 278, 44826–44831 (2003).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors acknowledge the support of the National Health and Medical Research Council to D.A.J., and of the Deutsche Forschungsgemeinschaft and the European Union to H.-J.L. The authors are also indebted to Michele Calos for critical reading of the manuscript, and to Isa M. Stehle, Lee Wong and Richard Saffery for the FISH illustrations shown in Figure 4.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David A. Jans.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

DATABASES

OMIM

cystic fibrosis

Duchenne muscular dystrophy

Lesch-Nyhan syndrome

Nijmegen breakage syndrome

RDEB

SCIDX1

FURTHER INFORMATION

Agence Française de Sécurité Sanitaire des Produits de Santé press release

David Jans' laboratory

Hans J. Lipps' laboratory

Journal of Gene Medicine Clinical Trials web site (Wiley Database)

Glossary

RETROVIRUS

A family of RNA viruses that replicate by reverse transcription and then integrate into the host genome.

ADENOVIRUS

A DNA-virus that replicates in the nucleus in an extra-chromosomal form.

X-LINKED SEVERE COMBINED IMMUNE DEFICIENCY

(SCIDX1). Disorder characterised by the absence of an immune system, caused by mutations that result in a failure to make a protein essential for T- and B-cell function.

LONG TERMINAL REPEAT

(LTR). A long DNA sequence, repeated at each end of retroviral DNA. An LTR contains regulatory sequences that are required to initiate transcription of the viral DNA into an RNA form that is inserted into new viral particles.

PROVIRUS

A stage in the life cycle of some viruses in which viral DNA has been inserted into the chromosome of the host cell.

INTERLEUKIN 2 RECEPTOR γ

The IL2RG protein is located on the surface of blood-forming cells and is a receptor for the ligand IL2, crucial for directing the growth and activation of cells of the immune system.

LIPOSOME

A spherical lipid-bilayer vesicle that can enclose DNA for transport into cells.

DNA COMPACTION

The reduction in the volume occupied by a DNA molecule caused by the addition of multivalent cations such as polyamine to the DNA.

EPISOME

A stable DNA molecule that persists in the nucleus without integrating into the cellular genome.

LYSOGENIC

A phase of the virus life cycle during which the virus integrates into the host chromosome of the infected cell, often remaining dormant for some period of time.

PARTIAL HEPATECTOMY

Removal of two-thirds of the liver, which stimulates liver cell division resulting in the removal of almost all unintergrated DNA from the liver.

RECESSIVE DYSTROPHIC EPIDERMOLYSIS BULLOSA

A blistering skin condition in which the filaments that anchor the epidermis to the underlying dermis are either absent or do not function. This is due to defects in the gene for type VII collagen, a fibrous protein that is the main component of the anchoring filaments.

FLUORESCENCE-ACTIVATED CELL SORTING

A cell-separation technique that is based on differences in physical and chemical properties of cells, such as the degree of fluorescence conferred by cell surface labelling.

INVERTED TERMINAL REPEATS

Repeat sequences in reverse orientation that serve as the viral origins of replication. ITRs are essential for virus packaging and rescue of the integrated viral genome.

CYTOSTATIC

Inhibiting or suppressing cellular growth and multiplication.

NUCLEAR MATRIX

The dense fibrillar network within the nucleus to which loops of chromatin attach.

ALPHA-SATELLITE DNA

A class of 170 bp repeating sequences of nucleotide pairs, found at centromeres.

MICROCELL-MEDIATED CHROMOSOME TRANSFER

A technique to deliver chromosomes into cells, whereby the chromosomes of donor cells are partitioned into discrete subnuclear packets. These microcells can be fused with recipient cells, resulting in transfer of chromosomes.

CYTOSINE METHYLATION

The addition of methyl groups (-CH3) to cytosine residues; This has a role in transcriptional silencing in higher organisms.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Glover, D., Lipps, H. & Jans, D. Towards safe, non-viral therapeutic gene expression in humans. Nat Rev Genet 6, 299–310 (2005). https://doi.org/10.1038/nrg1577

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg1577

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing