Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-27T08:44:42.020Z Has data issue: false hasContentIssue false

11 - Viral delivery of shRNA

Published online by Cambridge University Press:  31 July 2009

Ying Mao
Affiliation:
BD Biosciences Clontech
Chris Mello
Affiliation:
BD Biosciences Clontech
Laurence Lamarcq
Affiliation:
BD Biosciences Clontech
Brad Scherer
Affiliation:
BD Biosciences Clontech
Thomas Quinn
Affiliation:
BD Biosciences Clontech
Patty Wong
Affiliation:
BD Biosciences Clontech
Andrew Farmer
Affiliation:
BD Biosciences Clontech
Krishnarao Appasani
Affiliation:
GeneExpression Systems, Inc., Massachusetts
Andrew Fire
Affiliation:
Stanford University, California
Get access

Summary

Introduction

The completion of the human genome has made available the sequences of thousands of genes (Baltimore, 2001), allowing researchers to switch focus from identifying genes to understanding their function. In broad terms, gene function studies can be classified into two categories: those where the gene of interest is introduced into a system in which it is not expressed, and those in which the gene is disrupted or removed. While over-expression studies are fairly straightforward, methods for gene inactivation have been hampered in higher eukaryotes by the difficulty in manipulating their genetic material. Thus, although it is possible to generate mice lacking genes of interest by homologous recombination (Capecchi, 1989; van der Weyden et al., 2002), such studies remain technically challenging and expensive. Moreover, in some cases, deletion of a gene may be lethal, preventing its analysis (e.g., Lui et al., 1996). Alternatively, the phenotype produced may differ from that expected in humans (Harlow, 1992; Lee et al., 1992). A simple method for effective genetic inactivation in somatic cells in vitro is greatly needed, but has remained elusive (Sedivy and Dutriaux, 1999). Not surprisingly, recent years have seen considerable interest in a novel method for inactivating gene function in somatic cells that exploits the phenomenon of RNA interference (RNAi), first described by Fire et al. (1998). In their seminal study, they showed that double-stranded (ds)RNA homologous to a gene of interest could inhibit its expression. The dsRNA is digested into 21–23 nucleotide small interfering RNAs (siRNAs).

Type
Chapter
Information
RNA Interference Technology
From Basic Science to Drug Development
, pp. 161 - 173
Publisher: Cambridge University Press
Print publication year: 2005

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

Abremski, K. and Hoess, R. (1984). Bacteriophage P1 site-specific recombination. Purification and properties of the Cre recombinase protein. Journal of Biological Chemistry, 259, 1509–1514Google ScholarPubMed
Agami, R. and Bernards, R. (2000). Distinct initiation and maintenance mechanisms cooperate to induce G1 cell cycle arrest in response to DNA damage. Cell, 102, 55–66CrossRefGoogle ScholarPubMed
Armentano, D., Yu, S-F., Kantoff, P. W., Ruden, T., Anderson, W. F., and Gilboa, E. (1987). Effect of internal viral sequences on the utility of retroviral vectors. Journal of Virology, 61, 1647–1650Google ScholarPubMed
Baltimore, D. (2001). Our genome unveiled. Nature, 409, 814–816CrossRefGoogle ScholarPubMed
Barton, G. M., and Medzhitov, R. (2002). Retroviral delivery of small interfering RNA into primary cells. Proceedings of the National Academy of Sciences USA, 99, 14943–14945CrossRefGoogle ScholarPubMed
Brummelkamp, T. R., Bernards, R., and Agami, R. (2002a). A system for stable expression of short interfering RNAs in mammalian cells. Science, 296, 550–553CrossRefGoogle Scholar
Brummelkamp, T. R., Bernards, R., and Agami, R. (2002b). Stable suppression of tumorigenicity by virus-mediated RNA interference. Cancer Cell, 2, 243–247CrossRefGoogle Scholar
Capecchi, M. R. (1989). Altering the genome by homologous recombination. Science, 244, 1288–1292CrossRefGoogle ScholarPubMed
Correll, P. H., Colilla, S., and Karlsson, S. (1994). Retroviral vector design for long-term expression in murine hematopoietic cells in vivo. Blood, 84, 1812–1822Google ScholarPubMed
Devroe, E., and Silver, P. A. (2002). Retrovirus-delivered siRNA. BMC Biotechnology, 2, 15CrossRefGoogle ScholarPubMed
Donze, O., and Picard, D. (2002). RNA interference in mammalian cells using siRNAs synthesized with T7 RNA polymerase. Nucleic Acids Research, 30, e46CrossRefGoogle ScholarPubMed
Elbashir, S. M., Harborth, J., Lendeckel, W., Yalcin, A., Weber, K., and Tuschl, T. (2001). Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells. Nature, 411, 494–498CrossRefGoogle ScholarPubMed
Emerman, M., and Temin, H. M. (1984). Genes with promoters in retrovirus vectors can be independently suppressed by an epigenetic mechanism. Cell, 39, 449–467CrossRefGoogle ScholarPubMed
Fire, A., Xu, S., Montgomery, M. K., Kostas, S. A., Driver, S. E., and Mello, C. C. (1998). Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature, 391, 806–811CrossRefGoogle ScholarPubMed
Graham, F. L., and Prevec, L. (1991). Manipulation of adenovirus vectors. Methods in Molecular Biology, 7, 109–128Google ScholarPubMed
Hacein-Bey-Abina, S., Kalle, C., Schmidt, M., Deist, F., Wulffraat, N., McIntyre, E., Radford, I., Villeval, J. L., Fraser, C. C., Cavazzana-Calvo, M., and Fischer, A. (2003). A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. New England Journal of Medicine, 348, 255–256CrossRefGoogle ScholarPubMed
Hammond, S. M., Caudy, A. A., and Hannon, G. J. (2001). Post-transcriptional gene silencing by double-stranded RNA. Nature Reviews Genetics, 2, 110–119CrossRefGoogle ScholarPubMed
Harlow, E. (1992). Retinoblastoma. For our eyes only. Nature, 359, 270–271CrossRefGoogle ScholarPubMed
Hawley, T. S., Sabourin, L. A., and Hawley, R. G. (1989). Comparative analysis of retroviral vector expression in mouse embryonal carcinoma cells. Plasmid, 22, 120–131CrossRefGoogle ScholarPubMed
Hemann, M. T., Fridman, J. S., Zilfou, J. T., Hernando, E., Paddison, P. J., Cordon-Cardo, C., Hannon, G. J., and Lowe, S. W. (2003). An epi-allelic series of p53 hypomorphs created by stable RNAi produces distinct tumor phenotypes in vivo. Nature Genetics, 33, 396–400CrossRefGoogle ScholarPubMed
Hütvagner, G., and Zamore, P. D. (2002). RNAi: Nature abhors a double-strand. Current Opinion in Genetics and Development, 12, 225–232CrossRefGoogle ScholarPubMed
Ilves, H., Barske, C., Junker, U., Bohnlein, E., and Veres, G. (1996). Retroviral vectors designed for targeted expression of RNA polymerase III-driven transcripts: A comparative study. Gene, 171, 203–208CrossRefGoogle ScholarPubMed
Julius, M. A., Yan, Q., Zheng, Z., and Kitajewski, J. (2000). Q Vectors, Bicistronic retroviral vectors for gene transfer. BioTechniques, 28, 702–707Google ScholarPubMed
Kawasaki, H., and Taira, K. (2003). Short hairpin type of dsRNAs that are controlled by tRNA(Val). promoter significantly induce RNAi-mediated gene silencing in the cytoplasm of human cells. Nucleic Acids Research, 31, 700–707CrossRefGoogle ScholarPubMed
Lee, E. Y., Chang, C. Y., Hu, N., Wang, Y. C., Lai, C. C., Herrup, K., Lee, W. H., and Bradley, A. (1992). Mice deficient for Rb are nonviable and show defects in neurogenesis and haematopoiesis. Nature, 359, 288–294CrossRefGoogle ScholarPubMed
Lee, N. S., Dohjima, T., Bauer, G., Li, H., Li, M-J., Ehsani, A., Salvaterra, P., and Rossi, J. (2002). Expression of small interfering RNAs targeted against HIV-1 rev transcripts in human cells. Nature Biotechnology, 20, 500–505CrossRefGoogle ScholarPubMed
Liu, C. Y., Flesken-Nikitin, A., Li, S., Zeng, Y., and Lee, W. H. (1996). Inactivation of the mouse Brca1 gene leads to failure in the morphogenesis of the egg cylinder in early postimplantation development. Genes & Development, 10, 1835–1843CrossRefGoogle ScholarPubMed
Miller, V. M., Xia, H., Marrs, G. L., Gouvion, C. M., Lee, G., Davidson, B. L., and Paulson, H. L. (2003). Allele-specific silencing of dominant disease genes. Proceedings of the National Academy of Sciences USA, 100, 195–7200CrossRefGoogle ScholarPubMed
Miyagishi, M., and Taira, K. (2002). U6-promoter-driven siRNAs with four uridine 3′ overhangs efficiently suppress targeted gene expression in mammalian cells. Nature Biotechnology, 20, 497–500CrossRefGoogle Scholar
Mizuguchi, H., and Kay, M. A. (1998). Efficient construction of a recombinant adenovirus vector by an improved in vitro ligation method. Human Gene Therapy, 9, 2577–2583CrossRefGoogle ScholarPubMed
Nykanen, A., Haley, B., and Zamore, P. D. (2001). ATP requirements and small interfering RNA structure in the RNA interference pathway. Cell, 107, 309–321CrossRefGoogle ScholarPubMed
Paddison, P. J., Caudy, A. A., Bernstein, E., Hannon, G. J., and Conklin, D. S. (2002). Short hairpin RNAs (shRNAs) induce sequence-specific silencing in mammalian cells. Genes & Development, 16, 948–958CrossRefGoogle ScholarPubMed
Paul, C. P., Good, P. D., Winer, I., and Engelke, D. R. (2002). Effective expression of small interfering RNA in human cells. Nature Biotechnology, 20, 505–508CrossRefGoogle ScholarPubMed
Qin, X. F., An, D. S., Chen, I. S., and Baltimore, D. (2003). Inhibiting HIV-1 infection in human T cells by lentiviral-mediated delivery of small interfering RNA against CCR5. Proceedings of the National Academy of Sciences USA, 100, 183–188CrossRefGoogle Scholar
Reynolds, A., Leake, D., Boese, Q., Scaringe, S., Marshall, W. S. and Khvorova, A. (2004). Rational siRNA design for RNA interference. Nature Biotechnology, 22, 326–330CrossRefGoogle ScholarPubMed
Robertson, E., Bradley, A., Kuehn, M., and Evans, M. (1986). Germ-line transmission of genes introduced into cultured pluripotential cells by retroviral vector. Nature, 323, 445–448CrossRefGoogle ScholarPubMed
Rubinson, D. A., Dillon, C. P., Kwiatkowski, A. V., Sievers, C., Yang, L., Kopinja, J., Zhang, M., McManus, M. T., Gertler, F. B., Scott, M. L., and Parijs, L. (2003). A lentivirus-based system to functionally silence genes in primary mammalian cells, stem cells and transgenic mice by RNA interference. Nature Genetics, 33, 401–406CrossRefGoogle ScholarPubMed
Schroder, A. R., Shinn, P., Chen, H., Berry, C., Ecker, J. R., and Bushman, F. (2002). HIV-1 integration in the human genome favors active genes and local hotspots. Cell, 110, 521–529CrossRefGoogle ScholarPubMed
Sedivy, J. M., and Dutriaux, A. (1999). Gene targeting and somatic cell genetics – a rebirth or a coming of age?Trends in Genetics, 15, 88–90CrossRefGoogle ScholarPubMed
Sharp, P. A. (2001). RNA Interference – 2001. Genes & Development, 15, 485–490CrossRefGoogle ScholarPubMed
Shen, C., Buck, A. K., Liu, X., Winkler, M., and Reske, S. N. (2003). Gene silencing by adenovirus-delivered siRNA. Federation of European Biochemical Society Letters, 539, 111–114CrossRefGoogle ScholarPubMed
Tiscornia, G., Singer, O., Ikawa, M., and Verma, I. M. (2003). A general method for gene knockdown in mice by using lentiviral vectors expressing small interfering RNA. Proceedings of the National Academy of Sciences USA, 100, 1844–1848CrossRefGoogle ScholarPubMed
Yu, J-Y., DeRuiter, S. L., and Turner, D. L. (2002). RNA interference by expression of short-interfering RNAs and hairpin RNAs in mammalian cells. Proceedings of the National Academy of Sciences USA, 99, 6047–6052CrossRefGoogle ScholarPubMed
Yu, S. F., Ruden, T., Kantoff, P. W., Garber, C., Seiberg, M., Ruther, U., Anderson, W. F., Wagner, E. F., and Gilboa, E. (1986). Self-inactivating retroviral vectors designed for transfer of whole genes into mammalian cells. Proceedings of the National Academy of Sciences USA, 83, 3194–3198CrossRefGoogle ScholarPubMed
Weyden, L., Adams, D. J., and Bradley, A. (2002). Tools for targeted manipulation of the mouse genome. Physiological Genomics, 11, 133–164CrossRefGoogle ScholarPubMed
Wu, X., Li, Y., Crise, B., and Burgess, S. M. (2003). Transcription start regions in the human genome are favored targets for MLV integration. Science, 300, 1749–1751CrossRefGoogle ScholarPubMed
Xia, H., Mao, Q., Paulson, H. L., and Davidson, B. L. (2002). siRNA-mediated gene silencing in vitro and in vivo. Nature Biotechnology, 20, 1006–1010CrossRefGoogle 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.

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.

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.

Available formats
×