Skip to main content

Insertional Mutagenesis in Hematopoietic Cells: Lessons Learned from Adverse Events in Clinical Gene Therapy Trials

  • Chapter
  • First Online:

Abstract

From an early stage in the development of retroviral vectors for gene therapy applications, there has been a concern that recombinant vectors could elicit cellular transformation by altering expression of either cellular proto-oncogenes or tumor suppressor genes that are proximal to the genomic integration site (Fig. 6.1). This phenomenon, referred to as insertional mutagenesis, was characterized as a property of wild type gammaretroviral viruses (previously known as murine oncoretroviral viruses) which are the prototype virus from which the recombinant vectors used in the majority of the clinical trials described in this chapter were derived (reviewed in [178]). Wild type retroviruses fall into two categories with regards to their transforming properties: Slow transforming and acute transforming retroviruses.

Lars U. Müller and Michael D. Milsom have equally contributed to this work.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD   109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

References

  1. Aiuti, A., et al. (2007). Multilineage hematopoietic reconstitution without clonal selection in ADA-SCID patients treated with stem cell gene therapy. The Journal of Clinical Investigation, 117, 2233–2240.

    PubMed  CAS  Google Scholar 

  2. Aiuti, A., et al. (2002). Correction of ADA-SCID by stem cell gene therapy combined with nonmyeloablative conditioning. Science, 296, 2410–2413.

    PubMed  CAS  Google Scholar 

  3. Aiuti, A., et al. (2009). Gene therapy for immunodeficiency due to adenosine deaminase deficiency. The New England Journal of Medicine, 360, 447–458.

    PubMed  CAS  Google Scholar 

  4. Alexander, B. L., et al. (2007). Progress and prospects: Gene therapy clinical trials (part 1). Gene Therapy, 14, 1439–1447.

    PubMed  CAS  Google Scholar 

  5. Antoine, C., et al. (2003). Long-term survival and transplantation of haemopoietic stem cells for immunodeficiencies: Report of the European experience 1968–1999. Lancet, 361, 553–560.

    PubMed  Google Scholar 

  6. Arumugam, P. I., Scholes, J., Perelman, N., Xia, P., Yee, J. K., & Malik, P. (2007). Improved human beta-globin expression from self-inactivating lentiviral vectors carrying the chicken hypersensitive site-4 (cHS4) insulator element. Molecular Therapy, 15, 1863–1871.

    PubMed  CAS  Google Scholar 

  7. Arumugam, P. I., et al. (2009). Genotoxic potential of lineage-specific lentivirus vectors carrying the beta-globin locus control region. Molecular Therapy, 17, 1929–1937.

    PubMed  CAS  Google Scholar 

  8. Barjesteh van Waalwijk van Doorn-Khosrovani, S., et al. (2003). High EVI1 expression predicts poor survival in acute myeloid leukemia: A study of 319 de novo AML patients. Blood, 101, 837–845.

    Google Scholar 

  9. Baum, C., Hegewisch-Becker, S., Eckert, H. G., Stocking, C., & Ostertag, W. (1995). Novel retroviral vectors for efficient expression of the multidrug resistance (MDR-1) gene in early hematopoietic cells. Journal of Virology, 69, 7541–7547.

    PubMed  CAS  Google Scholar 

  10. Beard, B. C., et al. (2007). Comparison of HIV-derived lentiviral and MLV-based gammaretroviral vector integration sites in primate repopulating cells. Molecular Therapy, 15, 1356–1365.

    PubMed  CAS  Google Scholar 

  11. Berenson, R. J., et al. (1988). Antigen CD34+ marrow cells engraft lethally irradiated baboons. The Journal of Clinical Investigation, 81, 951–955.

    PubMed  CAS  Google Scholar 

  12. Berenson, R. J., et al. (1991). Engraftment after infusion of CD34+ marrow cells in patients with breast cancer or neuroblastoma. Blood, 77, 1717–1722.

    PubMed  CAS  Google Scholar 

  13. Berger, F., et al. (2001). Efficient retrovirus-mediated transduction of primitive human peripheral blood progenitor cells in stroma-free suspension culture. Gene Therapy, 8, 687–696.

    PubMed  CAS  Google Scholar 

  14. Bertrand, Y., et al. (1999). Influence of severe combined immunodeficiency phenotype on the outcome of HLA non-identical, T-cell-depleted bone marrow transplantation: A retrospective European survey from the European group for bone marrow transplantation and the European society for immunodeficiency. The Journal of Pediatrics, 134, 740–748.

    PubMed  CAS  Google Scholar 

  15. Beug, H., von Kirchbach, A., Doderlein, G., Conscience, J. F., & Graf, T. (1979). Chicken hematopoietic cells transformed by seven strains of defective avian leukemia viruses display three distinct phenotypes of differentiation. Cell, 18, 375–390.

    PubMed  CAS  Google Scholar 

  16. Blaese, R. M., et al. (1995). T lymphocyte-directed gene therapy for ADA-SCID: Initial trial results after 4 years. Science, 270, 475–480.

    PubMed  CAS  Google Scholar 

  17. Blumenthal, M., Skelton, D., Pepper, K. A., Jahn, T., Methangkool, E., & Kohn, D. B. (2007). Effective suicide gene therapy for leukemia in a model of insertional oncogenesis in mice. Molecular Therapy, 15, 183–192.

    PubMed  CAS  Google Scholar 

  18. Bokhoven, M., et al. (2009). Insertional gene activation by lentiviral and gammaretroviral vectors. Journal of Virology, 83, 283–294.

    PubMed  CAS  Google Scholar 

  19. Booth, C., et al. (2007). Management options for adenosine deaminase deficiency; proceedings of the EBMT satellite workshop (Hamburg, March 2006. Clinical Immunology, 123, 139–147.

    PubMed  CAS  Google Scholar 

  20. Bordignon, C., et al. (1995). Gene therapy in peripheral blood lymphocytes and bone marrow for ADA-immunodeficient patients. Science, 270, 470–475.

    PubMed  CAS  Google Scholar 

  21. Bosticardo, M., Ghosh, A., Du, Y., Jenkins, N., Copeland, N., & Candotti, F. (2009). Self-inactivating retroviral vector-mediated gene transfer induces oncogene activation and immortilization of primary murine bone marrow cells. Molecular Therapy, 17, 1910–1918.

    PubMed  CAS  Google Scholar 

  22. Bosticardo, M., Marangoni, F., Aiuti, A., Villa, A., & Grazia Roncarolo, M. (2009). Recent advances in understanding the pathophysiology of Wiskott-Aldrich syndrome. Blood, 113, 6288–6295.

    PubMed  CAS  Google Scholar 

  23. Bousso, P., et al. (2000). Diversity, functionality, and stability of the T cell repertoire derived in vivo from a single human T cell precursor. Proceedings of the National Academy of Sciences of the USA, 97, 274–278.

    PubMed  CAS  Google Scholar 

  24. Bozorgmehr, F., et al. (2007). No evidence of clonal dominance in primates up to 4 years following transplantation of multidrug resistance 1 retrovirally transduced long-term repopulating cells. Stem Cells, 25, 2610–2618.

    PubMed  CAS  Google Scholar 

  25. Boztug, K., et al. (2007). Hematopoietic stem cell gene therapy for Wiskott-Aldrich Syndrome. p. 502.

    Google Scholar 

  26. Buckley, R. H., et al. (1999). Hematopoietic stem-cell transplantation for the treatment of severe combined immunodeficiency. The New England Journal of Medicine, 340, 508–516.

    PubMed  CAS  Google Scholar 

  27. Buonamici, S., Chakraborty, S., Senyuk, V., & Nucifora, G. (2003). The role of EVI1 in normal and leukemic cells. Blood Cells, Molecules and Diseases, 31, 206–212.

    CAS  Google Scholar 

  28. Buonamici, S., et al. (2004). EVI1 induces myelodysplastic syndrome in mice. The Journal of Clinical Investigation, 114, 713–719.

    PubMed  CAS  Google Scholar 

  29. Calmels, B., et al. (2005). Recurrent retroviral vector integration at the Mds1/Evi1 locus in nonhuman primate hematopoietic cells. Blood, 106, 2530–2533.

    PubMed  CAS  Google Scholar 

  30. Cassani, B., et al. (2009). Integration of retroviral vectors induces minor changes in the transcriptional activity of T cells from ADA-SCID patients treated with gene therapy. Blood, 114, 3546–3556.

    PubMed  CAS  Google Scholar 

  31. Cattoglio, C., et al. (2007). Hot spots of retroviral integration in human CD34+ hematopoietic cells. Blood, 110, 1770–1778.

    PubMed  CAS  Google Scholar 

  32. Cavazzana-Calvo, M., & Fischer, A. (2007). Gene therapy for severe combined immunodeficiency: Are we there yet? The Journal of Clinical Investigation, 117, 1456–1465.

    PubMed  CAS  Google Scholar 

  33. Cavazzana-Calvo, M., Lagresle, C., Hacein-Bey-Abina, S., & Fischer, A. (2005). Gene therapy for severe combined immunodeficiency. Annual Review of Medicine, 56, 585–602.

    PubMed  CAS  Google Scholar 

  34. Cavazzana-Calvo, M., et al. (2000). Gene therapy of human severe combined immunodeficiency (SCID)-X1 disease [see comments]. Science, 288, 669–672.

    PubMed  CAS  Google Scholar 

  35. Cavazzana-Calvo, M., et al. (2008). Hematopoietic Stem Cell Gene Therapy Trial with Lentiviral Vector in X-Linked Adrenoleukodystrophy. American Society of Hematology, ISI:000262104701045, 304–305.

    Google Scholar 

  36. Chong, H., & Vile, R. G. (1996). Replication-competent retrovirus produced by a ‘split-function’ third generation amphotropic packaging cell line. Gene Therapy, 3, 624–629.

    PubMed  CAS  Google Scholar 

  37. Cleynen, I., & Van de Ven, W. J. (2008). The HMGA proteins: A myriad of functions [Review]. International Journal of Oncology, 32, 289–305.

    PubMed  CAS  Google Scholar 

  38. Cornetta, K., & Anderson, W. F. (1989). Protamine sulfate as an effective alternative to polybrene in retroviral-mediated gene-transfer: Implications for human gene therapy. Journal of Virological Methods, 23, 187–194.

    PubMed  CAS  Google Scholar 

  39. Cornils, K., et al. (2009). Stem cell marking with promotor-deprived self-inactivating retroviral vectors does not lead to induced clonal imbalance. Molecular Therapy, 17, 131–143.

    PubMed  CAS  Google Scholar 

  40. Couture, L. A., Mullen, C. A., & Morgan, R. A. (1994). Retroviral vectors containing chimeric promoter/enhancer elements exhibit cell-type-specific gene expression. Human Gene Therapy, 5, 667–677.

    PubMed  CAS  Google Scholar 

  41. Crooks, G. M., & Kohn, D. B. (1993). Growth factors increase amphotropic retrovirus binding to human CD34+ bone marrow progenitor cells. Blood, 82, 3290–3297.

    PubMed  CAS  Google Scholar 

  42. Daniel, R., & Smith, J. A. (2008). Integration site selection by retroviral vectors: Molecular mechanism and clinical consequences. Human Gene Therapy, 19, 557–568.

    PubMed  CAS  Google Scholar 

  43. Dave, U. P., et al. (2009). Murine leukemias with retroviral insertions at Lmo2 are predictive of the leukemias induced in SCID-X1 patients following retroviral gene therapy. PLoS Genetics, 5, e1000491.

    PubMed  Google Scholar 

  44. De Palma, M., et al. (2005). Promoter trapping reveals significant differences in integration site selection between MLV and HIV vectors in primary hematopoietic cells. Blood, 105, 2307–2315.

    PubMed  Google Scholar 

  45. Deichmann, A., et al. (2007). Vector integration is nonrandom and clustered and influences the fate of lymphopoiesis in SCID-X1 gene therapy. The Journal of Clinical Investigation, 117, 2225–2232.

    PubMed  CAS  Google Scholar 

  46. Dilber, M. S., Bjorkstrand, B., Li, K. J., Smith, C. I., Xanthopoulos, K. G., & Gahrton, G. (1994). Basic fibroblast growth factor increases retroviral-mediated gene transfer into human hematopoietic peripheral blood progenitor cells. Experimental Hematology, 22, 1129–1133.

    PubMed  CAS  Google Scholar 

  47. Donahue, R. E., et al. (1992). Helper virus induced T cell lymphoma in nonhuman primates after retroviral mediated gene transfer. The Journal of Experimental Medicine, 176, 1125.

    PubMed  CAS  Google Scholar 

  48. Emery, D. W., Yannaki, E., Tubb, J., Nishino, T., Li, Q., & Stamatoyannopoulos, G. (2002). Development of virus vectors for gene therapy of beta chain hemoglobinopathies: Flanking with a chromatin insulator reduces gamma-globin gene silencing in vivo. Blood, 100, 2012–2019.

    PubMed  CAS  Google Scholar 

  49. Evans-Galea, M. V., Wielgosz, M. M., Hanawa, H., Srivastava, D. K., & Nienhuis, A. W. (2007). Suppression of clonal dominance in cultured human lymphoid cells by addition of the cHS4 insulator to a lentiviral vector. Molecular Therapy, 15, 801–809.

    PubMed  CAS  Google Scholar 

  50. Felice, B., et al. (2009). Transcription factor binding sites are genetic determinants of retroviral integration in the human genome. PloS One, 4, e4571.

    PubMed  Google Scholar 

  51. Feron, C., et al. (1993). Purification, expansion, and multiple fluorochrome labeling of cord blood hemopoietic precursors: Preliminary results. Journal of Hematotherapy, 2, 259–261.

    PubMed  CAS  Google Scholar 

  52. Fischer, A., et al. (1997). Naturally occurring primary deficiencies of the immune system. Annual Review of Immunology, 15, 93–124.

    PubMed  CAS  Google Scholar 

  53. Foster, H., et al. (2008). Codon and mRNA sequence optimization of microdystrophin transgenes improves expression and physiological outcome in dystrophic mdx mice following AAV2/8 gene transfer. Molecular Therapy, 16, 1825–1832.

    PubMed  CAS  Google Scholar 

  54. Freas-Lutz, D. L., Correll, P. H., Dougherty, S. F., Xu, L., Pluznik, D. H., & Karlsson, S. (1994). Expression of human glucocerebrosidase in murine macrophages: Identification of efficient retroviral vectors. Experimental Hematology, 22, 857–865.

    PubMed  CAS  Google Scholar 

  55. Garcia, I. S., et al. (1991). A study of chromosome 11p13 translocations involving TCR beta and TCR delta in human T cell leukaemia. Oncogene, 6, 577–582.

    PubMed  CAS  Google Scholar 

  56. Garcia-Higuera, I., et al. (2001). Interaction of the Fanconi anemia proteins and BRCA1 in a common pathway. Molecular Cell, 7, 249–262.

    PubMed  CAS  Google Scholar 

  57. Garrett, E., Miller, A. R., Goldman, J. M., Apperley, J. F., & Melo, J. V. (2000). Characterization of recombination events leading to the production of an ecotropic replication-competent retrovirus in a GP+envAM12-derived producer cell line. Virology, 266, 170–179.

    PubMed  CAS  Google Scholar 

  58. Gaspar, H. B., et al. (2004). Gene therapy of X-linked severe combined immunodeficiency by use of a pseudotyped gammaretroviral vector. Lancet, 364, 2181–2187.

    PubMed  CAS  Google Scholar 

  59. Gaspar, H. B., et al. (2006). Successful reconstitution of immunity in ADA-SCID by stem cell gene therapy following cessation of PEG-ADA and use of mild preconditioning. Molecular Therapy, 14, 505–513.

    PubMed  CAS  Google Scholar 

  60. Glimm, H., et al. (1997). Efficient gene transfer in primitive CD34+/CD38lo human bone marrow cells reselected after long-term exposure to GALV-pseudotyped retroviral vector. Human Gene Therapy, 8, 2079–2086.

    PubMed  CAS  Google Scholar 

  61. Goerner, M., Bruno, B., McSweeney, P. A., Buron, G., Storb, R., & Kiem, H. P. (1999). The use of granulocyte colony-stimulating factor during retroviral transduction on fibronectin fragment CH-296 enhances gene transfer into hematopoietic repopulating cells in dogs. Blood, 94, 2287–2292.

    PubMed  CAS  Google Scholar 

  62. Gonda, T. J., Sheiness, D. K., & Bishop, J. M. (1982). Transcripts from the cellular homologs of retroviral oncogenes: Distribution among chicken tissues. Molecular and Cellular Biology, 2, 617–624.

    PubMed  CAS  Google Scholar 

  63. Grunebaum, E., et al. (2006). Bone marrow transplantation for severe combined immune deficiency. Jama, 295, 508–518.

    PubMed  CAS  Google Scholar 

  64. Hacein-Bey-Abina, S., et al. (2002). Sustained correction of X-linked severe combined immunodeficiency by ex vivo gene therapy. The New England Journal of Medicine, 346, 1185–1193.

    PubMed  CAS  Google Scholar 

  65. Hacein-Bey-Abina, S., et al. (2003). LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science, 302, 415–419.

    PubMed  CAS  Google Scholar 

  66. Hacein-Bey-Abina, S., et al. (2008). Insertional oncogenesis in 4 patients after retrovirus-mediated gene therapy of SCID-X1. The Journal of Clinical Investigation, 118, 3132–3142.

    PubMed  CAS  Google Scholar 

  67. Hacker, C. V., et al. (2006). The integration profile of EIAV-based vectors. Molecular Therapy, 14, 536–545.

    PubMed  CAS  Google Scholar 

  68. Haddad, E., et al. (1998). Long-term immune reconstitution and outcome after HLA-nonidentical T-cell-depleted bone marrow transplantation for severe combined immunodeficiency: A European retrospective study of 116 patients. Blood, 91, 3646–3653.

    PubMed  CAS  Google Scholar 

  69. Haneline, L. S., et al. (2003). Retroviral-mediated expression of recombinant Fancc enhances the repopulating ability of Fancc–/– hematopoietic stem cells and decreases the risk of clonal evolution. Blood, 101, 1299–1307.

    PubMed  CAS  Google Scholar 

  70. Hanenberg, H., Xiao, X. L., Dilloo, D., Hashino, K., Kato, I., & Williams, D. A. (1996). Colocalization of retrovirus and target cells on specific fibronectin fragments increases genetic transduction of mammalian cells. Nature Medicine, 2, 876–882.

    PubMed  CAS  Google Scholar 

  71. Hargrove, P. W., et al. (2008). Globin lentiviral vector insertions can perturb the expression of endogenous genes in beta-thalassemic hematopoietic cells. Molecular Therapy, 16, 525–533.

    PubMed  CAS  Google Scholar 

  72. Hematti, P., et al. (2004). Distinct genomic integration of MLV and SIV vectors in primate hematopoietic stem and progenitor cells. PLoS Biology, 2, e423.

    PubMed  Google Scholar 

  73. Hershfield, M. S. (1998). Adenosine deaminase deficiency: Clinical expression, molecular basis, and therapy. Seminars in Hematology, 35, 291–298.

    PubMed  CAS  Google Scholar 

  74. Higashimoto, T., et al. (2007). The woodchuck hepatitis virus post-transcriptional regulatory element reduces readthrough transcription from retroviral vectors. Gene Therapy, 14, 1298–1304.

    PubMed  CAS  Google Scholar 

  75. Hildinger, M., Abel, K. L., Ostertag, W., & Baum, C. (1999). Design of 5 untranslated sequences in retroviral vectors developed for medical use. Journal of Virology, 73, 4083–4089.

    PubMed  CAS  Google Scholar 

  76. Hildinger, M., Eckert, H. G., Schilz, A. J., John, J., Ostertag, W., & Baum, C. (1998). FMEV vectors: Both retroviral long terminal repeat and leader are important for high expression in transduced hematopoietic cells. Gene Therapy, 5, 1575–1579.

    PubMed  CAS  Google Scholar 

  77. Hirschhorn, R., Yang, D. R., Puck, J. M., Huie, M. L., Jiang, C. -K., & Kurlandsky, L. E. (1996). Spontaneous in vivo reversion to normal of an inherited mutation in a patient with adenosine deaminase deficiency. Nature Genetics, 13, 290–295.

    PubMed  CAS  Google Scholar 

  78. Hock, R. A., Miller, A. D., & Osborne, W. R. (1989). Expression of human adenosine deaminase from various strong promoters after gene transfer into human hematopoietic cell lines. Blood, 74, 876–881.

    PubMed  CAS  Google Scholar 

  79. Hoogerbrugge, P. M., et al. (1996). Bone marrow gene transfer in three patients with adenosine deaminase deficiency. Gene Therapy, 3, 179–183.

    PubMed  CAS  Google Scholar 

  80. Howe, S. J., et al. (2008). Insertional mutagenesis combined with acquired somatic mutations causes leukemogenesis following gene therapy of SCID-X1 patients. The Journal of Clinical Investigation, 118, 3143–3150.

    PubMed  CAS  Google Scholar 

  81. Hu, J., et al. (2003). Direct comparison of RD114-pseudotyped versus amphotropic-pseudotyped retroviral vectors for transduction of rhesus macaque long-term repopulating cells. Molecular Therapy, 8, 611–617.

    PubMed  CAS  Google Scholar 

  82. Hu, J., et al. (2008). Reduced genotoxicity of avian sarcoma leukosis virus vectors in rhesus long-term repopulating cells compared to standard murine retrovirus vectors. Molecular Therapy, 16, 1617–1623.

    PubMed  CAS  Google Scholar 

  83. Hudspeth, M. P., & Raymond, G. V. (2007). Immunopathogenesis of adrenoleukodystrophy: Current understanding. Journal of Neuroimmunology, 182, 5–12.

    PubMed  CAS  Google Scholar 

  84. Kiem, H. P., et al. (1997). Gene transfer into marrow repopulating cells: Comparison between amphotropic and gibbon ape leukemia virus pseudotyped retroviral vectors in a competitive repopulation assay in baboons. Blood, 90, 4638–4645.

    PubMed  CAS  Google Scholar 

  85. Kiem, H. -P., et al. (1998). Improved gene transfer into baboon marrow repopulating cells using recombinant human fibronectin fragment CH-296 in combination with IL-6, stem cell factor, FLT-3 ligand, and megakaryocyte growth and development factor. Blood, 92, 1878–1886.

    PubMed  CAS  Google Scholar 

  86. Kiem, H. P., et al. (2004). Long-term clinical and molecular follow-up of large animals receiving retrovirally transduced stem and progenitor cells: No progression to clonal hematopoiesis or leukemia. Molecular Therapy, 9, 389–395.

    PubMed  CAS  Google Scholar 

  87. Kim, W. S., Weickert, C. S., & Garner, B. (2008). Role of ATP-binding cassette transporters in brain lipid transport and neurological disease. Journal of Neurochemistry, 104, 1145–1166.

    PubMed  CAS  Google Scholar 

  88. Kim, Y. J., et al. (2009). Sustained high-level polyclonal hematopoietic marking and transgene expression 4 years after autologous transplantation of rhesus macaques with SIV lentiviral vector-transduced CD34+ cells. Blood, 113, 5434–5443.

    PubMed  CAS  Google Scholar 

  89. Knipper, R., et al. (2001). Improved post-transcriptional processing of an MDR1 retrovirus elevates expression of multidrug resistance in primary human hematopoietic cells. Gene Therapy, 8, 239–246.

    PubMed  CAS  Google Scholar 

  90. Kohn, D. B., et al. (1995). Engraftment of gene-modified umbilical cord blood cells in neonates with adenosine deaminase deficiency. Nature Medicine, 1, 1017–1023.

    PubMed  CAS  Google Scholar 

  91. Kokoris, M. S., Sabo, P., Adman, E. T., & Black, M. E. (1999). Enhancement of tumor ablation by a selected HSV-1 thymidine kinase mutant. Gene Therapy, 6, 1415–1426.

    PubMed  CAS  Google Scholar 

  92. Kurre, P., Morris, J., Horn, P. A., Harkey, M. A., Andrews, R. G., & Kiem, H. P. (2001). Gene transfer into baboon repopulating cells: A comparison of Flt-3 Ligand and megakaryocyte growth and development factor versus IL-3 during ex vivo transduction. Molecular Therapy, 3, 920–927.

    PubMed  CAS  Google Scholar 

  93. Kustikova, O., et al. (2005). Clonal dominance of hematopoietic stem cells triggered by retroviral gene marking. Science, 308, 1171–1174.

    PubMed  CAS  Google Scholar 

  94. Kustikova, O. S., et al. (2007). Retroviral vector insertion sites associated with dominant hematopoietic clones mark “stemness” pathways. Blood, 109, 1897–1907.

    PubMed  CAS  Google Scholar 

  95. Laufs, S., Nagy, K. Z., Giordano, F. A., Hotz-Wagenblatt, A., Zeller, W. J., & Fruehauf, S. (2004). Insertion of retroviral vectors in NOD/SCID repopulating human peripheral blood progenitor cells occurs preferentially in the vicinity of transcription start regions and in introns. Molecular Therapy, 10, 874–881.

    PubMed  CAS  Google Scholar 

  96. Lewinski, M. K., & Bushman, F. D. (2005). Retroviral DNA integration–mechanism and consequences. Advances in Genetics, 55, 147–181.

    PubMed  CAS  Google Scholar 

  97. Lewinski, M. K., et al. (2006). Retroviral DNA integration: Viral and cellular determinants of target-site selection. PLoS Pathogens, 2, e60.

    PubMed  Google Scholar 

  98. Li, C. L., & Emery, D. W. (2008). The cHS4 chromatin insulator reduces gammaretroviral vector silencing by epigenetic modifications of integrated provirus. Gene Therapy, 15, 49–53.

    PubMed  Google Scholar 

  99. Li, C. L., Xiong, D., Stamatoyannopoulos, G., & Emery, D. W. (2009). Genomic and functional assays demonstrate reduced gammaretroviral vector genotoxicity associated with use of the cHS4 chromatin insulator. Molecular Therapy, 17, 716–724.

    PubMed  CAS  Google Scholar 

  100. Li, X., et al. (2005). Ex vivo culture of Fancc–/– stem/progenitor cells predisposes cells to undergo apoptosis, and surviving stem/progenitor cells display cytogenetic abnormalities and an increased risk of malignancy. Blood, 105, 3465–3471.

    PubMed  CAS  Google Scholar 

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

    PubMed  CAS  Google Scholar 

  102. Likar, Y., et al. (2008). A new acycloguanosine-specific supermutant of herpes simplex virus type 1 thymidine kinase suitable for PET imaging and suicide gene therapy for potential use in patients treated with pyrimidine-based cytotoxic drugs. Journal of Nuclear Medicine, 49, 713–720.

    PubMed  CAS  Google Scholar 

  103. Malech, H., et al. (1997). Prolonged production of NADPH oxidase-corrected granulocytes after gene therapy of chronic granulomatous disease. Proceedings of the National Academy of Sciences of the USA, 94, 12133–12138.

    PubMed  CAS  Google Scholar 

  104. Malik, P., Krall, W. J., Yu, X. J., Zhou, C., & Kohn, D. B. (1995). Retroviral-mediated gene expression in human myelomonocytic cells: A comparison of hematopoietic cell promoters to viral promoters. Blood, 86, 2993–3005.

    PubMed  CAS  Google Scholar 

  105. Mann, R., Mulligan, R. C., & Baltimore, D. (1983). Construction of a retrovirus packaging mutant and its use to produce helper-free defective retrovirus. Cell, 33, 153–159.

    PubMed  CAS  Google Scholar 

  106. Manning, J. S., Hackett, A. J., & Darby, N. B., Jr. (1971). Effect of polycations on sensitivity of BALD-3T3 cells to murine leukemia and sarcoma virus infectivity. Applied Microbiology, 22, 1162–1163.

    PubMed  CAS  Google Scholar 

  107. Markowitz, D., Goff, S., & Bank, A. (1988). Construction and use of a safe and efficient amphotropic packaging cell line. J of Virology, 167, 400–406.

    CAS  Google Scholar 

  108. Maruggi, G., et al. (2009). Transcriptional enhancers induce insertional gene deregulation independently from the vector type and design. Molecular Therapy, 17, 851–856.

    PubMed  CAS  Google Scholar 

  109. Matzow, T., et al. (2007). Hypoxia-targeted over-expression of carboxylesterase as a means of increasing tumour sensitivity to irinotecan (CPT-11) . The Journal of Gene Medicine, 9, 244–252.

    PubMed  CAS  Google Scholar 

  110. May, C., et al. (2000). Therapeutic haemoglobin synthesis in beta-thalassaemic mice expressing lentivirus-encoded human beta-globin. Nature, 406, 82–86.

    PubMed  CAS  Google Scholar 

  111. Mazzolari, E., et al. (2007). Long-term immune reconstitution and clinical outcome after stem cell transplantation for severe T-cell immunodeficiency. Journal of Allergy and Clinical Immunology, 120, 892–899.

    PubMed  CAS  Google Scholar 

  112. Miller, A. D. (1990). Retrovirus packaging cells. Human Gene Therapy, 1, 5–14.

    PubMed  CAS  Google Scholar 

  113. Miller, A. D., Garcia, J. V., von Suhr, N., Lynch, C. M., Wilson, C., & Eiden, M. V. (1991). Construction and properties of retrovirus packaging cells based on gibbon ape leukemia virus. Journal of Virology, 65, 2220–2224.

    PubMed  CAS  Google Scholar 

  114. Milsom, M. D., & Fairbairn, L. J. (2004). Protection and selection for gene therapy in the hematopoietic system. The Journal of Gene Medicine, 6, 133–146.

    PubMed  CAS  Google Scholar 

  115. Milsom, M. D., & Williams, D. A. (2007). Live and let die: In vivo selection of gene-modified hematopoietic stem cells via MGMT-mediated chemoprotection. DNA Repair (Amst), 6, 1210–1221.

    CAS  Google Scholar 

  116. Milsom, M. D., et al. (2005). Overexpression of HOXB4 confers a myelo-erythroid differentiation delay in vitro. Leukemia, 19, 148–153.

    PubMed  CAS  Google Scholar 

  117. Milsom, M. D., et al. (2008). Reciprocal relationship between O6-methylguanine-DNA methyltransferase P140K expression level and chemoprotection of hematopoietic stem cells. Cancer Research, 68, 6171–6180.

    PubMed  CAS  Google Scholar 

  118. Mitchell, R. S., et al. (2004). Retroviral DNA integration: ASLV, HIV, and MLV show distinct target site preferences. PLoS Biology, 2, E234.

    PubMed  Google Scholar 

  119. Modlich, U., et al. (2005). Leukemias following retroviral transfer of multidrug resistance 1 (MDR1) are driven by combinatorial insertional mutagenesis. Blood, 105, 4235–4246.

    PubMed  CAS  Google Scholar 

  120. Modlich, U., et al. (2006). Cell culture assays reveal the importance of retroviral vector design for insertional genotoxicity. Blood, 108, 2545–2553.

    PubMed  CAS  Google Scholar 

  121. Modlich, U., et al. (2008). Leukemia induction after a single retroviral vector insertion in Evi1 or Prdm16. Leukemia, 22, 1519–1528.

    PubMed  CAS  Google Scholar 

  122. Modlich, U., et al. (2009). Insertional transformation of hematopoietic cells by self-inactivating lentiviral and gammaretroviral vectors. Molecular Therapy, 17, 1919–1928.

    PubMed  CAS  Google Scholar 

  123. Montini, E., et al. (2006). Hematopoietic stem cell gene transfer in a tumor-prone mouse model uncovers low genotoxicity of lentiviral vector integration. Nature Biotechnology, 24, 687–696.

    PubMed  CAS  Google Scholar 

  124. Montini, E., et al. (2009). The genotoxic potential of retroviral vectors is strongly modulated by vector design and integration site selection in a mouse model of HSC gene therapy. The Journal of Clinical Investigation, 119, 964–975.

    PubMed  CAS  Google Scholar 

  125. Moolten, F. L., & Cupples, L. A. (1992). A model for predicting the risk of cancer consequent to retroviral gene therapy. Human Gene Therapy, 3, 479–486.

    PubMed  CAS  Google Scholar 

  126. Moreno-Carranza, B., et al. (2009). Transgene optimization significantly improves SIN vector titers, gp91phox expression and reconstitution of superoxide production in X-CGD cells. Gene Therapy, 16, 111–118.

    PubMed  CAS  Google Scholar 

  127. Moritz, T., Patel, V. P., & Williams, D. A. (1994). Bone marrow extracellular matrix molecules improve gene transfer into human hematopoietic cells via retroviral vectors. The Journal of Clinical Investigation, 93, 1451–1457.

    PubMed  CAS  Google Scholar 

  128. Moser, H. W., Mahmood, A., & Raymond, G. V. (2007). X-linked adrenoleukodystrophy. Nature Clinical Practice, 3, 140–151.

    PubMed  Google Scholar 

  129. Neale, G. A., Rehg, J. E., & Goorha, R. M. (1995). Ectopic expression of rhombotin-2 causes selective expansion of CD4-CD8- lymphocytes in the thymus and T-cell tumors in transgenic mice. Blood, 86, 3060–3071.

    PubMed  CAS  Google Scholar 

  130. Neven, B., et al. (2009). Long-term outcome after hematopoietic stem cell transplantation of a single-center cohort of 90 patients with severe combined immunodeficiency. Blood, 113, 4114–4124.

    PubMed  CAS  Google Scholar 

  131. Nienhuis, A. (2006). Assays to Evaluate the Genotoxicity of Retroviral Vectors. Molecular Therapy, 13, 461–462.

    Google Scholar 

  132. Noguchi, M., et al. (1993). Interleukin-2 receptor gamma chain mutation results in X-linked severe combined immunodeficiency in humans. Cell, 73, 147–157.

    PubMed  CAS  Google Scholar 

  133. Nolta, J. A., Crooks, G. M., Overell, R. W., Williams, D. E., & Kohn, D. B. (1992). Retroviral vector-mediated gene transfer into primitive human hematopoietic progenitor cells: Effects of mast cell growth factor (MGF) combined with other cytokines. Experimental Hematology, 20, 1065–1071.

    PubMed  CAS  Google Scholar 

  134. Notarangelo, L. D., Miao, C. H., & Ochs, H. D. (2008). Wiskott-Aldrich syndrome. Current Opinion in Hematology, 15, 30–36.

    PubMed  CAS  Google Scholar 

  135. Ochs, H. D., Filipovich, A. H., Veys, P., Cowan, M. J., & Kapoor, N. (2008). Wiskott-Aldrich syndrome: Diagnosis, clinical and laboratory manifestations, and treatment. Biology of Blood and Marrow Transplant, 15, 84–90.

    Google Scholar 

  136. Orkin, S. H., & Nathan, D. G. (1981). The molecular genetics of thalassemia. Advances in Human Genetics, 11, 233–280.

    PubMed  CAS  Google Scholar 

  137. Ott, M. G., Seger, R., Stein, S., Siler, U., Hoelzer, D., & Grez, M. (2007). Advances in the treatment of chronic granulomatous disease by gene therapy. Current Gene Therapy, 7, 155–161.

    PubMed  CAS  Google Scholar 

  138. Ott, M. G., et al. (2006). Correction of X-linked chronic granulomatous disease by gene therapy, augmented by insertional activation of MDS1-EVI1, PRDM16 or SETBP1. Nature Medicine, 12, 401–409.

    PubMed  CAS  Google Scholar 

  139. Ott, M. G., et al. (2007). Phase I/II gene therapy study for Chronic Granulomatous Disease: Results, lessons and perspectives. American Society of Hematology, ISI:000251100800504, 503.

    Google Scholar 

  140. Otto, E., et al. (1994). Characterization of a replication-competent retrovirus resulting from recombination of packaging and vector sequences. Human Gene Therapy, 5, 567–575.

    PubMed  CAS  Google Scholar 

  141. Perumbeti, A., et al. (2009). A novel human gamma-globin gene vector for genetic correction of sickle cell anemia in a humanized sickle mouse model: Critical determinants for successful correction. Blood, 114, 1174–1185.

    PubMed  CAS  Google Scholar 

  142. Pike-Overzet, K., et al. (2006). Gene therapy: Is IL2RG oncogenic in T-cell development? Nature, 443, E5, discussion E6–7.

    PubMed  CAS  Google Scholar 

  143. Pike-Overzet, K., et al. (2007). Ectopic retroviral expression of LMO2, but not IL2Rgamma, blocks human T-cell development from CD34+ cells: Implications for leukemogenesis in gene therapy. Leukemia, 21, 754–763.

    PubMed  CAS  Google Scholar 

  144. Pluta, K., Luce, M. J., Bao, L., Agha-Mohammadi, S., & Reiser, J. (2005). Tight control of transgene expression by lentivirus vectors containing second-generation tetracycline-responsive promoters. The Journal of Gene Medicine, 7, 803–817.

    PubMed  CAS  Google Scholar 

  145. Radcliffe, P. A., et al. (2008). Analysis of factor VIII mediated suppression of lentiviral vector titres. Gene Therapy, 15, 289–297.

    PubMed  CAS  Google Scholar 

  146. Raper, S. E., et al. (2003). Fatal systemic inflammatory response syndrome in a ornithine transcarbamylase deficient patient following adenoviral gene transfer. Molecular Genetics and Metabolism, 80, 148–158.

    PubMed  CAS  Google Scholar 

  147. Relander, T., et al. (2005). Gene transfer to repopulating human CD34+ cells using amphotropic-, GALV-, or RD114-pseudotyped HIV-1-based vectors from stable producer cells. Molecular Therapy, 11, 452–459.

    PubMed  CAS  Google Scholar 

  148. Robinson, D., Elliott, J. F., & Chang, L. J. (1995). Retroviral vector with a CMV-IE/HIV-TAR hybrid LTR gives high basal expression levels and is up-regulated by HIV-1 Tat. Gene Therapy, 2, 269–278.

    PubMed  CAS  Google Scholar 

  149. Roussel, M., et al. (1979). Three new types of viral oncogene of cellular origin specific for haematopoietic cell transformation. Nature, 281, 452–455.

    PubMed  CAS  Google Scholar 

  150. Russell, M., et al. (1994). Expression of EVI1 in myelodysplastic syndromes and other hematologic malignancies without 3q26 translocations. Blood, 84, 1243–1248.

    PubMed  CAS  Google Scholar 

  151. Ryu, B. Y., Evans-Galea, M. V., Gray, J. T., Bodine, D. M., Persons, D. A., & Nienhuis, A. W. (2008). An experimental system for the evaluation of retroviral vector design to diminish the risk for proto-oncogene activation. Blood, 111, 1866–1875.

    PubMed  CAS  Google Scholar 

  152. Samakoglu, S., et al. (2006). A genetic strategy to treat sickle cell anemia by coregulating globin transgene expression and RNA interference. Nature Biotechnology, 24, 89–94.

    PubMed  CAS  Google Scholar 

  153. Sandrin, V., et al. (2002). Lentiviral vectors pseudotyped with a modified RD114 envelope glycoprotein show increased stability in sera and augmented transduction of primary lymphocytes and CD34+ cells derived from human and nonhuman primates. Blood, 100, 823–832.

    PubMed  CAS  Google Scholar 

  154. Sankaran, V. G., et al. (2008). Human fetal hemoglobin expression is regulated by the developmental stage-specific repressor BCL11A. Science, 322, 1839–1842.

    PubMed  CAS  Google Scholar 

  155. Sauvageau, G., et al. (1995). Overexpression of HOXB4 in hematopoietic cells causes the selective expansion of more primitive populations in vitro and in vivo. Genes and Development, 9, 1753–1765.

    PubMed  CAS  Google Scholar 

  156. Schambach, A., Galla, M., Maetzig, T., Loew, R., & Baum, C. (2007). Improving transcriptional termination of self-inactivating gamma-retroviral and lentiviral vectors. Molecular Therapy, 15, 1167–1173.

    PubMed  CAS  Google Scholar 

  157. Schambach, A., Swaney, W. P., & van der Loo, J. C. (2009). Design and production of retro- and lentiviral vectors for gene expression in hematopoietic cells. Methods in Molecular Biology, 506, 191–205.

    PubMed  CAS  Google Scholar 

  158. Schambach, A., et al. (2006). Woodchuck hepatitis virus post-transcriptional regulatory element deleted from X protein and promoter sequences enhances retroviral vector titer and expression. Gene Therapy, 13, 641–645.

    PubMed  CAS  Google Scholar 

  159. Schiedlmeier, B., et al. (2003). High-level ectopic HOXB4 expression confers a profound in vivo competitive growth advantage on human cord blood CD34+ cells, but impairs lymphomyeloid differentiation. Blood, 101, 1759–1768.

    PubMed  CAS  Google Scholar 

  160. Schmidt, M., et al. (2001). Detection and direct genomic sequencing of multiple rare unknown flanking DNA in highly complex samples. Human Gene Therapy, 12, 743–749.

    PubMed  CAS  Google Scholar 

  161. Schmidt, M., et al. (2002). Polyclonal long-term repopulating stem cell clones in a primate model. Blood, 100, 2737–2743.

    PubMed  CAS  Google Scholar 

  162. Schmidt, M., et al. (2007). High-resolution insertion-site analysis by linear amplification-mediated PCR (LAM-PCR). Nature Methods, 4, 1051–1057.

    PubMed  CAS  Google Scholar 

  163. Schwarzwaelder, K., et al. (2007). Gammaretrovirus-mediated correction of SCID-X1 is associated with skewed vector integration site distribution in vivo. The Journal of Clinical Investigation, 117, 2241–2249.

    PubMed  CAS  Google Scholar 

  164. Scobie, L., et al. (2009). A novel model of SCID-X1 reconstitution reveals predisposition to retrovirus-induced lymphoma but no evidence of gammaC gene oncogenicity. Molecular Therapy, 17, 1031–1038.

    PubMed  CAS  Google Scholar 

  165. Segal, A. W. (1996). The NADPH oxidase and chronic granulomatous disease. Molecular Medicine Today, 2, 129–135.

    PubMed  CAS  Google Scholar 

  166. Seger, R. A., et al. (2002). Treatment of chronic granulomatous disease with myeloablative conditioning and an unmodified hemopoietic allograft: A survey of the European experience, 1985–2000. Blood, 100, 4344–4350.

    PubMed  CAS  Google Scholar 

  167. Seggewiss, R., et al. (2006). Acute myeloid leukemia is associated with retroviral gene transfer to hematopoietic progenitor cells in a rhesus macaque. Blood, 107, 3865–3867.

    PubMed  CAS  Google Scholar 

  168. Semmler, A., Kohler, W., Jung, H. H., Weller, M., & Linnebank, M. (2008). Therapy of X-linked adrenoleukodystrophy. Expert review of Neurotherapeutics, 8, 1367–1379.

    PubMed  CAS  Google Scholar 

  169. Shou, Y., Ma, Z., Lu, T., & Sorrentino, B. P. (2006). Unique risk factors for insertional mutagenesis in a mouse model of XSCID gene therapy. Proceedings of the National Academy of Sciences of the USA, 103, 11730–11735.

    PubMed  CAS  Google Scholar 

  170. Somia, N., & Verma, I. M. (2000). Gene therapy: Trials and tribulations. Nature Reviews, 1, 91–99.

    PubMed  CAS  Google Scholar 

  171. Stephan, V., et al. (1996). Atypical X-linked severe combined immunodeficiency due to possible spontaneous reversion of the genetic defect in T cells. The New England Journal of Medicine, 335, 1563–1567.

    PubMed  CAS  Google Scholar 

  172. Stolworthy, T. S., et al. (2008). Yeast cytosine deaminase mutants with increased thermostability impart sensitivity to 5-fluorocytosine. Journal of Molecular Biology, 377, 854–869.

    PubMed  CAS  Google Scholar 

  173. Suzuki, T., Minehata, K., Akagi, K., Jenkins, N. A., & Copeland, N. G. (2006). Tumor suppressor gene identification using retroviral insertional mutagenesis in Blm-deficient mice. EMBO Journal, 25, 3422–3431.

    PubMed  CAS  Google Scholar 

  174. Takeshita, T., et al. (1992). Cloning of the gamma chain of the human IL-2 receptor. Science, 257, 379–382.

    PubMed  CAS  Google Scholar 

  175. Tassi, C., Fortuna, A., Bontadini, A., Lemoli, R. M., Gobbi, M., & Tazzari, P. L. (1991). CD34 or S313 positive cells selection by avidin-biotin immunoadsorption. Haematologica, 76(Suppl 1), 41–43.

    PubMed  Google Scholar 

  176. Thornhill, S. I., et al. (2008). Self-inactivating gammaretroviral vectors for gene therapy of X-linked severe combined immunodeficiency. Molecular Therapy, 16, 590–598.

    PubMed  CAS  Google Scholar 

  177. Thrasher, A. J., et al. (2006). Gene therapy: X-SCID transgene leukaemogenicity. Nature, 443, E5–6, discussion E6–7.

    Google Scholar 

  178. Touw, I. P., & Erkeland, S. J. (2007). Retroviral insertion mutagenesis in mice as a comparative oncogenomics tool to identify disease genes in human leukemia. Molecular Therapy, 15, 13–19.

    PubMed  CAS  Google Scholar 

  179. Trasher, A. J., & Gaspar, B. Successful bone marrow gene therapy for SCID. Blood Cells Molecules, and Diseases, 40, 287–288.

    Google Scholar 

  180. Trobridge, G. D., et al. (2006). Foamy virus vector integration sites in normal human cells. Proceedings of the National Academy of Sciences of the USA, 103, 1498–1503.

    PubMed  CAS  Google Scholar 

  181. Uchiyama, T., et al. (2006). Application of HSVtk suicide gene to X-SCID gene therapy: Ganciclovir treatment offsets gene corrected X-SCID B cells. Biochemical and Biophysical Research Communications, 341, 391–398.

    PubMed  CAS  Google Scholar 

  182. Vanin, E. F., Kaloss, M., Broscius, C., & Nienhuis, A. W. (1994). Characterization of replication-competent retroviruses from non-human primates with virus-induced T cell lymphomas and observations regarding the mechanism of oncogenesis. Journal of Virology, 68, 4241–4250.

    PubMed  CAS  Google Scholar 

  183. Wagner, W., et al. (2005). Retroviral integration sites correlate with expressed genes in hematopoietic stem cells. Stem Cells, 23, 1050–1058.

    PubMed  CAS  Google Scholar 

  184. Walsh, C. E., et al. (1994). A functionally active retrovirus vector for gene therapy in Fanconi anemia group C. Blood, 84, 453–459.

    PubMed  CAS  Google Scholar 

  185. Wang, G. P., et al. (2008). DNA bar coding and pyrosequencing to analyze adverse events in therapeutic gene transfer. Nucleic Acids Research, 36, e49.

    PubMed  Google Scholar 

  186. Wang, W. (2007). Emergence of a DNA-damage response network consisting of Fanconi anaemia and BRCA proteins. Nature Reviews, 8, 735–748.

    PubMed  CAS  Google Scholar 

  187. Weatherall, D. J. (1998). Pathophysiology of thalassaemia. Bailliere’s Clinical Haematology, 11, 127–146.

    PubMed  CAS  Google Scholar 

  188. Wicke, D. C., et al. (2008). Ectopic Expression of the Extracellular Domain of Mpl Is Sufficient to Induce a Hematopoietic Population Crisis. American Society of Hematology, ISI:000262104703324, 994.

    Google Scholar 

  189. Williams, D. A., Lemischka, I. R., Nathan, D. G., & Mulligan, R. C. (1984). Introduction of new genetic material into pluripotent haematopoietic stem cells of the mouse. Nature, 310, 476–480.

    PubMed  CAS  Google Scholar 

  190. Winkelstein, J. A., et al. (2000). Chronic granulomatous disease. Report on a national registry of 368 patients. Medicine, 79, 155–169.

    PubMed  CAS  Google Scholar 

  191. Woods, N. B., Bottero, V., Schmidt, M., von Kalle, C., & Verma, I. M. (2006). Gene therapy: Therapeutic gene causing lymphoma. Nature, 440, 1123.

    PubMed  CAS  Google Scholar 

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

    PubMed  CAS  Google Scholar 

  193. Yamada, Y., Warren, A. J., Dobson, C., Forster, A., Pannell, R., & Rabbitts, T. H. (1998). The T cell leukemia LIM protein Lmo2 is necessary for adult mouse hematopoiesis. Proceedings of the National Academy of Sciences of the USA, 95, 3890–3895.

    PubMed  CAS  Google Scholar 

  194. Zhang, F., et al. (2007). Lentiviral vectors containing an enhancer-less ubiquitously acting chromatin opening element (UCOE) provide highly reproducible and stable transgene expression in hematopoietic cells. Blood, 110, 1448–1457.

    PubMed  CAS  Google Scholar 

  195. Zhang, X. B., et al. (2008). High incidence of leukemia in large animals after stem cell gene therapy with a HOXB4-expressing retroviral vector. The Journal of Clinical Investigation, 118, 1502–1510.

    PubMed  CAS  Google Scholar 

  196. Zychlinski, D., et al. (2008). Physiological promoters reduce the genotoxic risk of integrating gene vectors. Molecular Therapy, 16, 718–725.

    PubMed  CAS  Google Scholar 

Download references

Acknowledgments

We thank Megan Smith, Miriam Edlund and Elise Porter for administrative support and members of our laboratory and the Transatlantic Gene Therapy Consortium for helpful discussions. This work is supported by NIH HL081499.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David A. Williams .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Müller, L.U., Milsom, M.D., Williams, D.A. (2011). Insertional Mutagenesis in Hematopoietic Cells: Lessons Learned from Adverse Events in Clinical Gene Therapy Trials. In: Dupuy, A., Largaespada, D. (eds) Insertional Mutagenesis Strategies in Cancer Genetics. Springer, New York, NY. https://doi.org/10.1007/978-1-4419-7656-7_6

Download citation

Publish with us

Policies and ethics