Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-06-30T02:04:15.677Z Has data issue: false hasContentIssue false

27 - Gene transfer: methods and applications

from Part III - Evaluation and treatment

Published online by Cambridge University Press:  01 July 2010

Martin Pulé
Affiliation:
Postdoctoral Research Fellow Center for Cell and Gene Therapy, Baylor College of Medicine, Houston, TX, USA
Malcolm K. Brenner
Affiliation:
Director, Center for Cell and Gene Therapy, Professor, Departments of Medicine and Pediatrics, Baylor College of Medicine, Houston, TX, USA
Ching-Hon Pui
Affiliation:
St. Jude Children's Research Hospital, Memphis
Get access

Summary

Introduction

The concept of using gene transfer techniques to express a new gene in the somatic cells of a patient has stimulated considerable interest, speculation, and hyperbole. The inevitable backlash against promises that so far have not been fulfilled has led to much confusion about the aims and achievements of gene transfer and to a lurking suspicion that the entire field is simply a South Sea bubble waiting to burst. This chapter seeks to provide a balanced account of the current status of gene transfer as applied to leukemia and related disorders, and to review the accomplishments of the field as well as the impediments to progress. Most importantly, it will try to give an idea of the incremental way in which gene transfer technologies will supplement, long before they supplant, current therapeutic approaches to hematologic malignancies.

There are two broad strategies of gene transfer applicable to the treatment of leukemia and lymphoma. First, the tumor cell itself can be genetically modified to “repair” its intrinsic molecular defect. Alternatively, a toxic gene can be introduced to destroy the tumor cell, or it can be transduced to express molecules that trigger an immune response against it. Second, the host's T cells can be redirected, their antitumor activity augmented, or they can be transduced with suicide genes to terminate potentially harmful immune reactions. The drug sensitivity of normal host tissues can be decreased by delivering cytotoxic drug-resistance genes to sensitive tissues, thereby increasing the therapeutic index of chemotherapy.

Type
Chapter
Information
Childhood Leukemias , pp. 661 - 678
Publisher: Cambridge University Press
Print publication year: 2006

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Bender, M. A., Palmer, T. D., Gelinas, R. E., & Miller, A. D.Evidence that the packaging signal of Moloney murine leukemia virus extends into the gag region. J Virol, 1987; 61: 1639–46.Google ScholarPubMed
Brenner, M. K., Rill, D. R., Holladay, M. S.et al.Gene marking to determine whether autologous marrow infusion restores long-term haemopoiesis in cancer patients. Lancet, 1993; 342: 1134–7.CrossRefGoogle ScholarPubMed
Hacein-Bey-Abina, S., Le Deist, F., Carlier, F., et al.Sustained correction of X-linked severe combined immunodeficiency by ex-vivo gene therapy. N Engl J Med, 2002; 346: 1185–93.CrossRefGoogle ScholarPubMed
Hacein-Bey-Abina, S., Kalle, C. von, Schmidt, M., et al.LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science, 2003; 302: 415–19.CrossRefGoogle ScholarPubMed
Amado, R. G. & Chen, I. S.Lentiviral vectors – the promise of gene therapy within reach ?Science, 1999; 285: 674–6.CrossRefGoogle ScholarPubMed
Sutton, R. E., Wu, H. T., Rigg, R., Bohnlein, E., & Brown, P. O.Human immunodeficiency virus type 1 vectors efficiently transduce human hematopoietic stem cells. J Virol, 1998; 72: 5781–8.Google ScholarPubMed
Mascarenhas, L., Stripecke, R., Case, S. S., et al.Gene delivery to human B-precursor acute lymphoblastic leukemia cells. Blood, 1998; 92: 3537–45.Google ScholarPubMed
Case, S. S., Price, M. A., Jordan, C. T., et al.Stable transduction of quiescent CD34(+)CD38(‒) human hematopoietic cells by HIV-1-based lentiviral vectors. Proc Natl Acad Sci U S A, 1999; 96: 2988–93.CrossRefGoogle ScholarPubMed
Buchschacher, G. L. Jr. & Wong-Staal, F.Development of lentiviral vectors for gene therapy for human diseases. Blood, 2000; 95: 2499–504.Google ScholarPubMed
Engelhardt, J. F., Yang, Y., Stratford-Perricaudet, L. D., et al.Direct gene transfer of human CFTR into human bronchial epithelia of xenografts with E1-deleted adenoviruses. Nat Genet, 1993; 4: 27–34.CrossRefGoogle ScholarPubMed
Le Gal, L. S., Robert, J. J., Berrard, S., et al.An adenovirus vector for gene transfer into neurons and glia in the brain. Science, 1993; 259: 988–90.CrossRefGoogle Scholar
Smith, T. A., Mehaffey, M. G., Kayda, D. B., et al.Adenovirus mediated expression of therapeutic plasma levels of human factor Ⅸ in mice. Nat Genet, 1993; 5: 397–402.CrossRefGoogle ScholarPubMed
Amin, R., Wilmott, R., Schwarz, Y., Trapnell, B., & Stark, J.Replication-deficient adenovirus induces expression of interleukin-8 by airway epithelial cells in vitro. Hum Gene Ther, 1995; 6: 145–53.CrossRefGoogle ScholarPubMed
Muzyczka, N.Use of adeno-associated virus as a general transduction vector for mammalian cells. Curr Top Microbiol Immunol, 1992; 158: 97–129.Google ScholarPubMed
Clark, K. R., Voulgaropoulou, F., & Johnson, P. R.A stable cell line carrying adenovirus-inducible rep and cap genes allows for infectivity titration of adeno-associated virus vectors. Gene Ther, 1996; 3: 1124–32.Google ScholarPubMed
Conway, J. E., Rhys, C. M., Zolotukhin, I., et al.High-titer recombinant adeno-associated virus production utilizing a recombinant herpes simplex virus type I vector expressing AAV-2 Rep and Cap. Gene Ther, 1999; 6: 986–93.CrossRefGoogle ScholarPubMed
Flotte, T., Carter, B., Conrad, C., et al.A phase I study of an adeno-associated virus-CFTR gene vector in adult CF patients with mild lung disease. Hum Gene Ther, 1996; 7: 1145–59.CrossRefGoogle ScholarPubMed
Wagner, J. A., Reynolds, T., Moran, M. L., et al.Efficient and persistent gene transfer of AAV-CFTR in maxillary sinus. Lancet, 1998; 351: 1702–3.CrossRefGoogle ScholarPubMed
Wagner, J. A., Moran, M. L., Messner, A. H., et al.A phase I/II study of tgAAV-CF for the treatment of chronic sinusitis in patients with cystic fibrosis. Hum Gene Ther, 1998; 9: 889–909.CrossRefGoogle ScholarPubMed
Wagner, J. A., Nepomuceno, I. B., Messner, A. H., et al.A phase II, double-blind, randomized, placebo-controlled clinical trial of tgAAVCF using maxillary sinus delivery in patients with cystic fibrosis with antrostomies. Hum Gene Ther, 2002; 13: 1349–59.CrossRefGoogle ScholarPubMed
Gool, S. W., Barcy, S., Devos, S., et al.CD80 (B7-1) and CD86 (B7-2): potential targets for immunotherapy ?Res Immunol, 1995; 146: 183–96.CrossRefGoogle ScholarPubMed
Su, H., Chang, J. C., Xu, S. M., & Kan, Y. W.Selective killing of AFP-positive hepatocellular carcinoma cells by adeno-associated virus transfer of the herpes simplex virus thymidine kinase gene. Hum Gene Ther, 1996; 7: 463–70.CrossRefGoogle ScholarPubMed
Nabel, G. J., Nabel, E. G., Yang, Z. Y., et al.Direct gene transfer with DNA-liposome complexes in melanoma: expression, biologic activity, and lack of toxicity in humans. Proc Natl Acad Sci U S A, 1993; 90: 11307–11.CrossRefGoogle ScholarPubMed
Gao, X. & Huang, L.A novel cationic liposome reagent for efficient transfection of mammalian cells. Biochem Biophys Res Commun, 1991; 179: 280–5.CrossRefGoogle ScholarPubMed
Templeton, N. S.Cationic liposome-mediated gene delivery in vivo. Biosci Rep, 2002; 22: 283–95.CrossRefGoogle ScholarPubMed
Trubetskoy, V. S., Torchilin, V. P., Kennel, S., & Huang, L.Cationic liposomes enhance targeted delivery and expression of exogenous DNA mediated by N-terminal modified poly(L-lysine)-antibody conjugate in mouse lung endothelial cells. Biochim Biophys Acta, 1992; 1131: 311–13.CrossRefGoogle ScholarPubMed
Ito, K., Ito, K., Shinohara, N., & Kato, S.DNA immunization via intramuscular and intradermal routes using a gene gun provides different magnitudes and durations on immune response. Mol Immunol, 2003; 39: 847–54.CrossRefGoogle ScholarPubMed
Lin-Hong, L., Biagi, E., Brenner, M., & Fratantoni, J. Electroporation-mediated efficient non-viral gene delivery to CLL-B cells and application in immunotherapy. In 44th Meeting of the American Society of Hematology, December 6–10, 2002 Philadelphia, PA, USA. Meeting abstract 5538.
Yotnda, P., Chen, D. H., Chiu, W., et al.Bilamellar cationic liposomes protect adenovectors from preexisting humoral immune responses. Mol Ther, 2002; 5: 233–41.CrossRefGoogle ScholarPubMed
Rossi, J. J.Therapeutic antisense and ribozymes. Br Med Bull, 1995; 51: 217–25.CrossRefGoogle ScholarPubMed
Zhang, Y., Mukhopadhyay, T., Donehower, L. A., Georges, R. N., & Roth, J. A.Retroviral vector-mediated transduction of K-ras antisense RNA into human lung cancer cells inhibits expression of the malignant phenotype. Hum Gene Ther, 1993; 4: 451–60.CrossRefGoogle ScholarPubMed
Heise, C., Sampson-Johannes, A., Williams, A., et al.ONYX-015, an E1B gene-attenuated adenovirus, causes tumor-specific cytolysis and antitumoral efficacy that can be augmented by standard chemotherapeutic agents. Nat Med, 1997; 3: 639–45.CrossRefGoogle ScholarPubMed
Parr, M. J., Manome, Y., Tanaka, T., et al.Tumor-selective transgene expression in vivo mediated by an E2F-responsive adenoviral vector. Nat Med, 1997; 3: 1145–9.CrossRefGoogle ScholarPubMed
Wang, H., Cai, Q., Zeng, X., et al.Antitumor activity and pharmacokinetics of a mixed-backbone antisense oligonucleotide targeted to the RIalpha subunit of protein kinase A after oral administration. Proc Natl Acad Sci U S A, 1999; 96: 13989–94.CrossRefGoogle ScholarPubMed
Gewirtz, A. M.Myb targeted therapeutics for the treatment of human malignancies. Oncogene, 1999; 18: 3056–62.CrossRefGoogle ScholarPubMed
Bayever, E., Iversen, P. L., Bishop, M. R., et al.Systemic administration of a phosphorothioate oligonucleotide with a sequence complementary to p53 for acute myelogenous leukemia and myelodysplastic syndrome: initial results of a phase I trial. Antisense Res Dev, 1993; 3: 383–90.CrossRefGoogle ScholarPubMed
Waters, J. S., Webb, A., Cunningham, D., et al.Phase I clinical and pharmacokinetic study of bcl-2 antisense oligonucleotide therapy in patients with non-Hodgkin's lymphoma. J Clin Oncol, 2000; 18: 1812–23.CrossRefGoogle ScholarPubMed
Fire, A., Xu, S., Montgomery, M. K., et al.Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature, 1998; 391: 806–11.CrossRefGoogle ScholarPubMed
Elbashir, S. M., Harborth, J., Lendeckel, W., et al.Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells. Nature, 2001; 411: 494–8.CrossRefGoogle ScholarPubMed
Brummelkamp, T. R., Bernards, R., & Agami, R.A system for stable expression of short interfering RNAs in mammalian cells. Science, 2002; 296: 550–3.CrossRefGoogle ScholarPubMed
Beltinger, C., Fulda, S., Kammertoens, T., et al.Herpes simplex virus thymidine kinase/ganciclovir-induced apoptosis involves ligand-independent death receptor aggregation and activation of caspases. Proc Natl Acad Sci U S A, 1999; 96: 8699–704.CrossRefGoogle ScholarPubMed
Hurwitz, M. Y., Marcus, K. T., Chevez-Barrios, P., et al.Suicide gene therapy for treatment of retinoblastoma in a murine model. Hum Gene Ther, 1999; 10: 441–8.CrossRefGoogle Scholar
Hurwitz, A., Finci-Yeheskel, Z., Milwidsky, A., & Mayer, M.Regulation of cyclooxygenase activity and progesterone production in the rat corpus luteum by inducible nitric oxide synthase. Reproduction, 2002; 123: 663–9.CrossRefGoogle ScholarPubMed
Hurwitz, R. L., Brenner, M. K., Poplack, D. G., & Horowitz, M. C.Retinoblastoma treatment. Science, 1999; 285: 663–4.CrossRefGoogle ScholarPubMed
Bartlett, J. S., Kleinschmidt, J., Boucher, R. C., & Samulski, R. J.Targeted adeno-associated virus vector transduction of nonpermissive cells mediated by a bispecific F(ab'gamma)2 antibody. Nat Biotechnol, 1999; 17: 181–6.CrossRefGoogle ScholarPubMed
Girod, A., Ried, M., Wobus, C., et al.Genetic capsid modifications allow efficient re-targeting of adeno- associated virus type 2. Nat Med, 1999; 5: 1438.CrossRefGoogle ScholarPubMed
Russell, S. J. & Cosset, F. L.Modifying the host range properties of retroviral vectors. J Gene Med, 1999; 1: 300–11.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Grill, J., Beusechem, V. W., Valk, P., et al.Combined targeting of adenoviruses to integrins and epidermal growth factor receptors increases gene transfer into primary glioma cells and spheroids. Clin Cancer Res, 2001; 7: 641–50.Google ScholarPubMed
Krasnykh, V., Belousova, N., Korokhov, N., Mikheeva, G., & Curiel, D. T.Genetic targeting of an adenovirus vector via replacement of the fiber protein with the phage T4 fibritin. J Virol, 2001; 75: 4176–83.CrossRefGoogle ScholarPubMed
Osaki, T., Tanio, Y., Tachibana, I., et al.Gene therapy for carcinoembryonic antigen-producing human lung cancer cells by cell type-specific expression of herpes simplex virus thymidine kinase gene. Cancer Res, 1994; 54: 5258–61.Google ScholarPubMed
Barnett, B. G., Tillman, B. W., Curiel, D. T., & Douglas, J. T.Dual targeting of adenoviral vectors at the levels of transduction and transcription enhances the specificity of gene expression in cancer cells. Mol Ther, 2002; 6: 377–85.CrossRefGoogle ScholarPubMed
Curiel, D. T.The development of conditionally replicative adenoviruses for cancer therapy. Clin Cancer Res, 2000; 6: 3395–9.Google ScholarPubMed
Chen, L., Rao, A., & Harrison, S. C.Signal integration by transcription-factor assemblies: interactions of NF-AT1 and AP-1 on the IL-2 promoter. Cold Spring Harb Symp Quant Biol, 1999; 64: 527–31.CrossRefGoogle ScholarPubMed
Yui, M. A., Hernandez-Hoyos, G., & Rothenberg, E. V.A new regulatory region of the IL-2 locus that confers position-independent transgene expression. J Immunol, 2001; 166: 1730–9.CrossRefGoogle ScholarPubMed
Pizzato, M., Marlow, S. A., Blair, E. D., & Takeuchi, Y.Initial binding of murine leukemia virus particles to cells does not require specific Env-receptor interaction. J Virol, 1999; 73: 8599–611.Google Scholar
Tepper, R. I., Pattengale, P. K., & Leder, P.Murine interleukin-4 displays potent anti-tumor activity in vivo. Cell, 1989; 57: 503–12.CrossRefGoogle ScholarPubMed
Golumbek, P. T., Lazenby, A. J., Levitsky, H. I., et al.Treatment of established renal cancer by tumor cells engineered to secrete interleukin-4. Science, 1991; 254: 713–16.CrossRefGoogle ScholarPubMed
Fearon, E. R., Pardoll, D. M., Itaya, T., et al.Interleukin-2 production by tumor cells bypasses T helper function in the generation of an antitumor response. Cell, 1990; 60: 397–403.CrossRefGoogle Scholar
Colombo, M. P., Ferrari, G., Stoppacciaro, A., et al.Granulocyte colony-stimulating factor gene transfer suppresses tumorigenicity of a murine adenocarcinoma in vivo. J Exp Med, 1991; 173: 889–97.CrossRefGoogle ScholarPubMed
Dilloo, D., Bacon, K., Holden, W., et al.Combined chemokine and cytokine gene transfer enhances antitumor immunity. Nat Med, 1996; 2: 1090–5.CrossRefGoogle ScholarPubMed
Soiffer, R., Lynch, T., Mihm, M., et al.Vaccination with irradiated autologous melanoma cells engineered to secrete human granulocyte-macrophage colony-stimulating factor generates potent antitumor immunity in patients with metastatic melanoma. Proc Natl Acad Sci U S A, 1998; 95: 13141–6.CrossRefGoogle ScholarPubMed
Nelson, W. G., Simons, J. W., Mikhak, B., et al.Cancer cells engineered to secrete granulocyte-macrophage colony-stimulating factor using ex vivo gene transfer as vaccines for the treatment of genitourinary malignancies. Cancer Chemother Pharmacol, 2000; 46(Suppl.): S67–72.CrossRefGoogle ScholarPubMed
DeAngelo, D. J., Dranoff, G., Galinsky, I., et al.A Phase I study of vaccination with lethally irradiated, autologous myeloblasts engineered by adenoviral-mediated gene transfer to secrete granulocyte-macrophage colony-stimulating factor [abstract]. Blood, 2001; 98: 463a.Google Scholar
Asher, A. L., Mule, J. J., Kasid, A., et al.Murine tumor cells transduced with the gene for tumor necrosis factor-alpha. Evidence for paracrine immune effects of tumor necrosis factor against tumors. J Immunol, 1991; 146: 3227–34.Google ScholarPubMed
Hock, H., Dorsch, M., Diamantstein, T., & Blankenstein, T.Interleukin 7 induces CD4+ T cell-dependent tumor rejection. J Exp Med, 1991; 174: 1291–8.CrossRefGoogle ScholarPubMed
Tepper, R. I., Pattengale, P. K., & Leder, P.Murine interleukin-4 displays potent anti-tumor activity in vivo. Cell, 1989; 57: 503–12.CrossRefGoogle ScholarPubMed
Watanabe, Y., Kuribayashi, K., Miyatake, S., et al.Exogenous expression of mouse interferon gamma cDNA in mouse neuroblastoma C1300 cells results in reduced tumorigenicity by augmented anti-tumor immunity. Proc Natl Acad Sci U S A, 1989; 86: 9456–60.CrossRefGoogle ScholarPubMed
Leimig, T., Foreman, N., Rill, D., et al.Immunomodulatory effects of human neuroblastoma cells transduced with a retroviral vector encoding interleukin-2. Cancer Gene Ther, 1994; 1: 253–8.Google ScholarPubMed
Kooten, C., & Banchereau, J.CD40-CD40 ligand: a multifunctional receptor-ligand pair. Adv Immunol, 1996; 61: 1–77.CrossRefGoogle ScholarPubMed
Banchereau, J. & Steinman, R. M.Dendritic cells and the control of immunity. Nature, 1998; 392: 245–52.CrossRefGoogle ScholarPubMed
Dilloo, D., Brown, M., Roskrow, M., et al.CD40 ligand induces an antileukemia immune response in vivo. Blood, 1997; 90: 1927–33.Google ScholarPubMed
Fujita, N., Kagamu, H., Yoshizawa, H., et al.CD40 ligand promotes priming of fully potent antitumor CD4(+) T cells in draining lymph nodes in the presence of apoptotic tumor cells. J Immunol, 2001; 167: 5678–88.CrossRefGoogle ScholarPubMed
Kato, K., Cantwell, M. J., Sharma, S., & Kipps, T. J.Gene transfer of CD40-ligand induces autologous immune recognition of chronic lymphocytic leukemia B cells. J Clin Invest, 1998; 101: 1133–41.CrossRefGoogle ScholarPubMed
Guinan, E. C., Gribben, J. G., Boussiotis, V. A., Freeman, G. J., & Nadler, L. M.Pivotal role of the B7:CD28 pathway in transplantation tolerance and tumor immunity. Blood, 1994; 84: 3261–82.Google ScholarPubMed
Ranheim, E. A. & Kipps, T. J.Activated T cells induce expression of B7/BB1 on normal or leukemic B cells through a CD40-dependent signal. J Exp Med, 1993; 177: 925–35CrossRefGoogle ScholarPubMed
Wulf, G. G., Wang, R. Y., Kuehnle, I., et al.A leukemic stem cell with intrinsic drug efflux capacity in acute myeloid leukemia. Blood, 2001; 98: 1166–73.CrossRefGoogle ScholarPubMed
Rousseau, R. & Biagi, E. Treatment of high-risk acute leukemia with an autologous vaccine expressing transgenic IL-2 and CD40L. In 44th Meeting of the American Society of Hematology, December 6–10, 2002, Philadelpia, PA, USA. Meeting abstract 3420.
Biagi, E. & Rousseau, R. Bystander transfer of functional human CD40 ligand from gene-modified fibroblasts to leukemia cells: a method suitable for clinical cancer vaccine production. In 44th Meeting of the American Society of Hematology, December 6–10, 2002, Philadelpia, PA, USA. Meeting abstract 1462.
Luznik, L., Slansky, J. E., Jalla, S., et al.Successful therapy of metastatic cancer using tumor vaccines in mixed allogeneic bone marrow chimeras. Blood, 2003; 101: 1645–52.CrossRefGoogle ScholarPubMed
Raz, E., Carson, D. A., Parker, S. E., et al.Intradermal gene immunization: the possible role of DNA uptake in the induction of cellular immunity to viruses. Proc Natl Acad Sci U S A, 1994; 91: 9519–23.CrossRefGoogle ScholarPubMed
Kim, J. J., Yang, J. S., Lee, D. J.et al.Macrophage colony-stimulating factor can modulate immune responses and attract dendritic cells in vivo. Hum Gene Ther, 2000; 11: 305–21.CrossRefGoogle ScholarPubMed
Ohashi, T., Hanabuchi, S., Kato, H., et al.Prevention of adult T-cell leukemia-like lymphoproliferative disease in rats by adoptively transferred T cells from a donor immunized with human T-cell leukemia virus type 1 Tax-coding DNA vaccine. J Virol, 2000; 74: 9610–6.CrossRefGoogle ScholarPubMed
Syrengelas, A. D., Chen, T. T., & Levy, R.DNA immunization induces protective immunity against B-cell lymphoma. Nat Med, 1996; 2: 1038–41.CrossRefGoogle ScholarPubMed
Krieg, A. M. & Wagner, H.Causing a commotion in the blood: immunotherapy progresses from bacteria to bacterial DNA. Immunol Today, 2000; 21: 521–6.CrossRefGoogle ScholarPubMed
Romieu, R., Baratin, M., Kayibanda, M., et al.Passive but not active CD8+ T cell-based immunotherapy interferes with liver tumor progression in a transgenic mouse model. J Immunol, 1998; 161: 5133–7.Google ScholarPubMed
Granziero, L., Krajewski, S., Farness, P., et al.Adoptive immunotherapy prevents prostate cancer in a transgenic animal model. Eur J Immunol, 1999; 29: 1127–38.3.0.CO;2-X>CrossRefGoogle Scholar
Drobyski, W. R., Roth, M. S., Thibodeau, S. N., & Gottschall, J. L.Molecular remission occurring after donor leukocyte infusions for the treatment of relapsed chronic myelogenous leukemia after allogeneic bone marrow transplantation. Bone Marrow Transplant, 1992; 10: 301–4.Google ScholarPubMed
Rooney, C. M., Smith, C. A., Ng, C. Y., et al.Use of gene-modified virus-specific T lymphocytes to control Epstein–Barr-virus-related lymphoproliferation. Lancet, 1995; 345: 9–13.CrossRefGoogle ScholarPubMed
Brenner, M. K. & Heslop, H. E.Graft-versus-host reactions and bone-marrow transplantation. Curr Opin Immunol, 1991; 3: 752–7.CrossRefGoogle ScholarPubMed
Melief, C. J. & Kast, W. M.Potential immunogenicity of oncogene and tumor suppressor gene products. Curr Opin Immunol, 1993; 5: 709–13.CrossRefGoogle ScholarPubMed
Heslop, H. E. & Rooney, C. M.Adoptive cellular immunotherapy for EBV lymphoproliferative disease. Immunol Rev, 1997; 157: 217–22.CrossRefGoogle ScholarPubMed
Bruggen, P., Traversari, C., Chomez, P., et al.A gene encoding an antigen recognized by cytolytic T lymphocytes on a human melanoma. Science, 1991; 254: 1643–7.CrossRefGoogle ScholarPubMed
Gross, G., Gorochov, G., Waks, T., & Eshhar, Z.Generation of effector T cells expressing chimeric T cell receptor with antibody type-specificity. Transplant Proc, 1989; 21: 127–30.Google ScholarPubMed
Irving, B. A. & Weiss, A.The cytoplasmic domain of the T cell receptor zeta chain is sufficient to couple to receptor-associated signal transduction pathways. Cell, 1991; 64: 891–901.CrossRefGoogle Scholar
Altenschmidt, U., Klundt, E., & Groner, B.Adoptive transfer of in vitro-targeted, activated T lymphocytes results in total tumor regression. J Immunol, 1997; 159: 5509–15.Google ScholarPubMed
Kershaw, M. H., Darcy, P. K., Trapani, J. A., & Smyth, M. J.The use of chimeric human Fc(epsilon) receptor I to redirect cytotoxic T lymphocytes to tumors. J Leukoc Biol, 1996; 60: 721–8.CrossRefGoogle ScholarPubMed
Moritz, D., Wels, W., Mattern, J., & Groner, B.Cytotoxic T lymphocytes with a grafted recognition specificity for ERBB2-expressing tumor cells. Proc Natl Acad Sci U S A, 1994; 91: 4318–22.CrossRefGoogle ScholarPubMed
Walker, R. E., Bechtel, C. M., Natarajan, V., et al.Long-term in vivo survival of receptor-modified syngeneic T cells in patients with human immunodeficiency virus infection. Blood, 2000; 96: 467–74.Google ScholarPubMed
Anderson, S. J., Levin, S. D., & Perlmutter, R. M.Involvement of the protein tyrosine kinase p56lck in T cell signaling and thymocyte development. Adv Immunol, 1994; 56: 151–78.CrossRefGoogle Scholar
Brocker, T. & Karjalainen, K.Signals through T cell receptor-zeta chain alone are insufficient to prime resting T lymphocytes. J Exp Med, 1995; 181: 1653–9.CrossRefGoogle ScholarPubMed
Finney, H. M., Lawson, A. D., Bebbington, C. R., & Weir, A. N.Chimeric receptors providing both primary and costimulatory signaling in T cells from a single gene product. J Immunol, 1998; 161: 2791–7.Google Scholar
Maher, J., Brentjens, R. J., Gunset, G., Riviere, I., & Sadelain, M.Human T-lymphocyte cytotoxicity and proliferation directed by a single chimeric TCRzeta /CD28 receptor. Nat Biotechnol, 2002; 20: 70–5.CrossRefGoogle ScholarPubMed
Geiger, T. L., Nguyen, P., Leitenberg, D., & Flavell, R. A.Integrated src kinase and costimulatory activity enhances signal transduction through single-chain chimeric receptors in T lymphocytes. Blood, 2001; 98: 2364–71.CrossRefGoogle ScholarPubMed
Kershaw, M. H., Westwood, J. A., & Hwu, P.Dual-specific T cells combine proliferation and antitumor activity. Nat Biotechnol, 2002; 20: 1221–7.CrossRefGoogle ScholarPubMed
Rossig, C., Bollard, C. M., Nuchtern, J. G., Rooney, C. M., & Brenner, M. K.Epstein–Barr virus-specific human T lymphocytes expressing antitumor chimeric T-cell receptors: potential for improved immunotherapy. Blood, 2002; 99: 2009–16.CrossRefGoogle Scholar
Rosenberg, S. A., Aebersold, P., Cornetta, K., et al.Gene transfer into humans – immunotherapy of patients with advanced melanoma, using tumor-infiltrating lymphocytes modified by retroviral gene transduction. N Engl J Med, 1990; 323: 570–8.CrossRefGoogle ScholarPubMed
Brodie, S. J., Patterson, B. K., Lewinsohn, D. A., et al.HIV-specific cytotoxic T lymphocytes traffic to lymph nodes and localize at sites of HIV replication and cell death. J Clin Invest, 2000; 105: 1407–17.CrossRefGoogle ScholarPubMed
Rosenberg, S. A., Lotze, M. T., Yang, J. C., et al.Experience with the use of high-dose interleukin-2 in the treatment of 652 cancer patients. Ann Surg, 1989; 210: 474–84.CrossRefGoogle ScholarPubMed
Liu, K. & Rosenberg, S. A.Transduction of an IL-2 gene into human melanoma-reactive lymphocytes results in their continued growth in the absence of exogenous IL-2 and maintenance of specific antitumor activity. J Immunol, 2001; 167: 6356–65.CrossRefGoogle ScholarPubMed
Eaton, D., Gilham, D. E., O'Neill, A., & Hawkins, R. E.Retroviral transduction of human peripheral blood lymphocytes with Bcl- X(L) promotes in vitro lymphocyte survival in pro-apoptotic conditions. Gene Ther, 2002; 9: 527–35.CrossRefGoogle ScholarPubMed
Gorelik, L. & Flavell, R. A.Immune-mediated eradication of tumors through the blockade of transforming growth factor-beta signaling in T cells. Nat Med, 2001; 7: 1118–22.CrossRefGoogle Scholar
Bollard, C. M., Rossig, C., Calonge, M. J., et al.Adapting a transforming growth factor beta-related tumor protection strategy to enhance antitumor immunity. Blood, 2002; 99: 3179–87.CrossRefGoogle ScholarPubMed
Bonini, C., Ferrari, G., Verzeletti, S., et al.HSV-TK gene transfer into donor lymphocytes for control of allogeneic graft-versus-leukemia. Science, 1997; 276: 1719–24.CrossRefGoogle ScholarPubMed
Tiberghien, P., Ferrand, C., Lioure, B., et al.Administration of herpes simplex-thymidine kinase-expressing donor T cells with a T-cell-depleted allogeneic marrow graft. Blood, 2001; 97: 63–72.CrossRefGoogle ScholarPubMed
Riddell, S. R., Elliott, M., Lewinsohn, D. A., et al.T-cell mediated rejection of gene-modified HIV-specific cytotoxic T lymphocytes in HIV-infected patients. Nat Med, 1996; 2: 216–23.CrossRefGoogle ScholarPubMed
Clackson, T., Yang, W., Rozamus, L. W., et al.Redesigning an FKBP-ligand interface to generate chemical dimerizers with novel specificity. Proc Natl Acad Sci U S A, 1998; 95: 10437–42.CrossRefGoogle ScholarPubMed
Thomis, D. C., Marktel, S., Bonini, C., et al.A Fas-based suicide switch in human T cells for the treatment of graft-versus-host disease. Blood, 2001; 97: 1249–57.CrossRefGoogle Scholar
Iuliucci, J. D., Oliver, S. D., Morley, S., et al.Intravenous safety and pharmacokinetics of a novel dimerizer drug, AP1903, in healthy volunteers. J Clin Pharmacol, 2001; 41: 870–9.CrossRefGoogle ScholarPubMed
Introna, M., Barbui, A. M., Bambacioni, F., et al.Genetic modification of human T cells with CD20: a strategy to purify and lyse transduced cells with anti-CD20 antibodies. Hum Gene Ther, 2000; 11: 611–20.CrossRefGoogle Scholar
Klump, H., Schiedlmeier, B., Vogt, B., et al.Retroviral vector-mediated expression of HoxB4 in hematopoietic cells using a novel coexpression strategy. Gene Ther, 2001; 8: 811–17.CrossRefGoogle ScholarPubMed
Murphy, D., Crowther, D., Renninson, J., et al.A randomised dose intensity study in ovarian carcinoma comparing chemotherapy given at four week intervals for six cycles with half dose chemotherapy given for twelve cycles. Ann Oncol, 1993; 4: 377–83.CrossRefGoogle ScholarPubMed
Levin, L. & Hryniuk, W. M.Dose intensity analysis of chemotherapy regimens in ovarian carcinoma. J Clin Oncol, 1987; 5: 756–67.CrossRefGoogle ScholarPubMed
Levin, L., Simon, R., & Hryniuk, W.Importance of multiagent chemotherapy regimens in ovarian carcinoma: dose intensity analysis. J Natl Cancer Inst, 1993; 85: 1732–42.CrossRefGoogle ScholarPubMed
Pastan, I. & Gottesman, M. M.Multidrug resistance. Annu Rev Med, 1991; 42: 277–86.CrossRefGoogle ScholarPubMed
Mickisch, G. H., Licht, T., Merlino, G. T., Gottesman, M. M., & Pastan, I.Chemotherapy and chemosensitization of transgenic mice which express the human multidrug resistance gene in bone marrow: efficacy, potency, and toxicity. Cancer Res, 1991; 51: 5417–24.Google ScholarPubMed
Mickisch, G. H., Merlino, G. T., Galski, H., Gottesman, M. M., & Pastan, I.Transgenic mice that express the human multidrug-resistance gene in bone marrow enable a rapid identification of agents that reverse drug resistance. Proc Natl Acad Sci U S A, 1991; 88: 547–51.CrossRefGoogle ScholarPubMed
Moscow, J. A., Huang, H., Carter, C., et al.Engraftment of MDR1 and NeoR gene-transduced hematopoietic cells after breast cancer chemotherapy. Blood, 1999; 94: 52–61.Google ScholarPubMed
Appelbaum, F. R. & Buckner, C. D.Overview of the clinical relevance of autologous bone marrow transplantation. Clin Haematol, 1986; 15: 1–18.CrossRefGoogle ScholarPubMed
Burnett, A. K., Tansey, P., Watkins, R., et al.Transplantation of unpurged autologous bone-marrow in acute myeloid leukaemia in first remission. Lancet, 1984; 2: 1068–70.CrossRefGoogle ScholarPubMed
Goldstone, A. H., Anderson, C. C., Linch, D. C., et al.Autologous bone marrow transplantation following high dose chemotherapy for the treatment of adult patients with acute myeloid leukaemia. Br J Haematol, 1986; 64: 529–37.CrossRefGoogle ScholarPubMed
Shpall, E. J. & Jones, R. B.Release of tumor cells from bone marrow. Blood, 1994; 83: 623–5.Google ScholarPubMed
Brugger, W., Bross, K. J., Glatt, M., et al.Mobilization of tumor cells and hematopoietic progenitor cells into peripheral blood of patients with solid tumors. Blood, 1994; 83: 636–40.Google ScholarPubMed
Rill, D. R., Santana, V. M., Roberts, W. M., et al.Direct demonstration that autologous bone marrow transplantation for solid tumors can return a multiplicity of tumorigenic cells. Blood, 1994; 84: 380–3.Google ScholarPubMed
De Fabritiis, P., Ferrero, D., Sandrelli, A., et al.Monoclonal antibody purging and autologous bone marrow transplantation in acute myelogenous leukemia in complete remission. Bone Marrow Transplant, 1989; 4: 669–74.Google ScholarPubMed
Gambacorti-Passerini, C., Rivoltini, L., Fizzotti, M., et al.Selective purging by human interleukin-2 activated lymphocytes of bone marrows contaminated with a lymphoma line or autologous leukaemic cells. Br J Haematol, 1991; 78: 197–205.CrossRefGoogle ScholarPubMed
Gorin, N. C., Aegerter, P., Auvert, B., et al.Autologous bone marrow transplantation for acute myelocytic leukemia in first remission: a European survey of the role of marrow purging. Blood, 1990; 75: 1606–14.Google ScholarPubMed
Santos, G. W., Yeager, A. M., & Jones, R. J.Autologous bone marrow transplantation. Annu Rev Med, 1989; 40: 99–112.CrossRefGoogle ScholarPubMed
Gribben, J. G., Freedman, A. S., Neuberg, D., et al.Immunologic purging of marrow assessed by PCR before autologous bone marrow transplantation for B-cell lymphoma. N Engl J Med, 1991; 325: 1525–33.CrossRefGoogle ScholarPubMed
Brenner, M., Krance, R., Heslop, H. E., et al.Assessment of the efficacy of purging by using gene marked autologous marrow transplantation for children with AML in first complete remission. Hum Gene Ther, 1994; 5: 481–99.CrossRefGoogle ScholarPubMed
Cornetta, K., Tricot, G., Broun, E. R., et al.Retroviral-mediated gene transfer of bone marrow cells during autologous bone marrow transplantation for acute leukemia. Hum Gene Ther, 1992; 3: 305–18.CrossRefGoogle ScholarPubMed
Cai, Q., Rubin, J. T., & Lotze, M. T.Genetically marking human cells – results of the first clinical gene transfer studies. Cancer Gene Ther, 1995; 2: 125–36.Google ScholarPubMed
Santana, V. M., Brenner, M. K., Ihle, J., et al.A phase I trial of high-dose carboplatin and etoposide with autologous marrow support for treatment of stage D neuroblastoma in first remission: use of marker genes to investigate the biology of marrow reconstitution and the mechanism of relapse. Hum Gene Ther, 1991; 3: 257–72.CrossRefGoogle Scholar
Deisseroth, A. B., Kantarjian, H., Talpaz, M., et al.Autologous bone marrow transplantation for CML in which retroviral markers are used to discriminate between relapse which arises from systemic disease remaining after preparative therapy versus relapse due to residual leukemia cells in autologous marrow: A pilot trial. Hum Gene Ther, 1991; 2: 359–76.Google Scholar
Rill, D. R. & Smith, S.Long term in vivo fate of human hemopoietic cells transduced by moloney-based retroviral vectors [abstract]. Blood, 2000; 84: 96.Google Scholar
Moritz, T., Mackay, W., & Feng, L. J.Gene transfer of 0-6-methylguanine methyltransferase (MGMT) protects hematopoietic cells (HQ) from nitrosourea (NU) induced toxicity in vitro and in vivo [abstract]. Blood, 1993; 81: 18a.Google Scholar
Dunbar, C. E., Bodine, D. M., Sorrentino, B., et al.Gene transfer into hematopoietic cells. Implications for cancer therapy. Ann N Y Acad Sci, 1994; 716: 216–24.CrossRefGoogle ScholarPubMed
Dunbar, C. E., Cottler-Fox, M., O'Shaughnessy, J. A., et al.Retrovirally marked CD34-enriched peripheral blood and bone marrow cells contribute to long-term engraftment after autologous transplantation. Blood, 1995; 85: 3048–57.Google ScholarPubMed
Brenner, M. K., Rill, D. R., Heslop, H. E., et al.Gene marking after bone marrow transplantation. Eur J Cancer, 1994; 30A: 1171–6.CrossRefGoogle ScholarPubMed
Roskrow, M. A., Rooney, C. M., Heslop, H. E., et al.Administration of neomycin resistance gene marked EBV specific cytotoxic T-lymphocytes to patients with relapsed EBV-positive Hodgkin disease. Hum Gene Ther, 1998; 9: 1237–50.CrossRefGoogle ScholarPubMed
Rooney, C. M., Smith, C. A., Ng, C. Y. C., et al.Infusion of cytotoxic T cells for the prevention and treatment of Epstein–Barr virus-induced lymphoma in allogeneic transplant recipients. Blood, 1998; 92: 1549–55.Google ScholarPubMed
Schmidt, M., Hoffmann, G., Wissler, M., et al.Detection and direct genomic sequencing of multiple rare unknown flanking DNA in highly complex samples. Hum Gene Ther, 2001; 12: 743–9.CrossRefGoogle ScholarPubMed
Vanin, E. F., Kaloss, M., Broscius, C., & Nienhuis, A. W.Characterization of replication-competent retroviruses from nonhuman primates with virus-induced T-cell lymphomas and observations regarding the mechanism of oncogenesis. J Virol, 1994; 68: 4241–50.Google ScholarPubMed
Davé, U. P., Jenkins, N. A., & Copeland, N. G.Gene therapy insertional mutagenesis insights. Science, 2004; 303: 333.CrossRefGoogle ScholarPubMed
McCormack, M. P. & Rabbitts, T. H.Activation of the T-cell oncogene LMO2 after gene therapy for X-linked severe combined immunodeficiency expectations. N Engl J Med, 2004; 350: 913–22.CrossRefGoogle Scholar
Li, Z., Dullmann, J., Schiedlmeier, B., et al.Murine leukemia induced by retroviral gene marking. Science, 2002; 296: 497.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Gene transfer: methods and applications
    • By Martin Pulé, Postdoctoral Research Fellow Center for Cell and Gene Therapy, Baylor College of Medicine, Houston, TX, USA, Malcolm K. Brenner, Director, Center for Cell and Gene Therapy, Professor, Departments of Medicine and Pediatrics, Baylor College of Medicine, Houston, TX, USA
  • Edited by Ching-Hon Pui
  • Book: Childhood Leukemias
  • Online publication: 01 July 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511471001.028
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Gene transfer: methods and applications
    • By Martin Pulé, Postdoctoral Research Fellow Center for Cell and Gene Therapy, Baylor College of Medicine, Houston, TX, USA, Malcolm K. Brenner, Director, Center for Cell and Gene Therapy, Professor, Departments of Medicine and Pediatrics, Baylor College of Medicine, Houston, TX, USA
  • Edited by Ching-Hon Pui
  • Book: Childhood Leukemias
  • Online publication: 01 July 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511471001.028
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Gene transfer: methods and applications
    • By Martin Pulé, Postdoctoral Research Fellow Center for Cell and Gene Therapy, Baylor College of Medicine, Houston, TX, USA, Malcolm K. Brenner, Director, Center for Cell and Gene Therapy, Professor, Departments of Medicine and Pediatrics, Baylor College of Medicine, Houston, TX, USA
  • Edited by Ching-Hon Pui
  • Book: Childhood Leukemias
  • Online publication: 01 July 2010
  • Chapter DOI: https://doi.org/10.1017/CBO9780511471001.028
Available formats
×