Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-19T18:41:57.169Z Has data issue: false hasContentIssue false

Nucleic acid transfection and transgenesis in parasitic nematodes

Published online by Cambridge University Press:  31 August 2011

JAMES B. LOK*
Affiliation:
Department of Pathobiology, School of Veterinary Medicine, University of Pennsylvania, 3800 Spruce Street, Philadelphia, PA, 19104, USA
*
*Corresponding author: Tel: +1215-898-7892. Fax: +1 215-573-7023. E-mail: jlok@vet.upenn.edu

Summary

Transgenesis is an essential tool for assessing gene function in any organism, and it is especially crucial for parasitic nematodes given the dwindling armamentarium of effective anthelmintics and the consequent need to validate essential molecular targets for new drugs and vaccines. Two of the major routes of gene delivery evaluated to date in parasitic nematodes, bombardment with DNA-coated microparticles and intragonadal microinjection of DNA constructs, draw upon experience with the free-living nematode Caenorhabditis elegans. Bombardment has been used to transiently transfect Ascaris suum, Brugia malayi and Litomosoides sigmodontis with both RNA and DNA. Microinjection has been used to achieve heritable transgenesis in Strongyloides stercoralis, S. ratti and Parastrongyloides trichosuri and for additional transient expression studies in B. malayi. A third route of gene delivery revisits a classic method involving DNA transfer facilitated by calcium-mediated permeabilization of recipient cells in developing B. malayi larvae and results in transgene inheritance through host and vector passage. Assembly of microinjected transgenes into multi-copy episomal arrays likely results in their transcriptional silencing in some parasitic nematodes. Methods such as transposon-mediated transgenesis that favour low-copy number chromosomal integration may remedy this impediment to establishing stable transgenic lines. In the future, stable transgenesis in parasitic nematodes could enable loss-of-function approaches by insertional mutagenesis, in situ expression of inhibitory double-stranded RNA or boosting RNAi susceptibility through heterologous expression of dsRNA processing and transport proteins.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2011

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

REFERENCES

Aboobaker, A. A. and Blaxter, M. L. (2003). Use of RNA interference to investigate gene function in the human filarial nematode parasite Brugia malayi. Molecular and Biochemical Parasitology 129, 4151. doi: 10.1016/S0166-6851(03)00092-6CrossRefGoogle ScholarPubMed
Ashton, F. T., Li, J. and Schad, G. A. (1999). Chemo- and thermosensory neurons: structure and function in animal parasitic nematodes. Veterinary Parasitology 84, 297316.CrossRefGoogle ScholarPubMed
Ashton, F. T., Zhu, X., Boston, R., Lok, J. B. and Schad, G. A. (2007). Strongyloides stercoralis: Amphidial neuron pair ASJ triggers significant resumption of development by infective larvae under host-mimicking in vitro conditions. Experimental Parasitology 115, 9297. doi: 10.1016/j.exppara.2006.08.010CrossRefGoogle ScholarPubMed
Ayuk, M. A., Suttiprapa, S., Rinaldi, G., Mann, V. H., Lee, C. M. and Brindley, P. J. (2011). Schistosoma mansoni U6 gene promoter-driven short hairpin RNA induces RNA interference in human fibrosarcoma cells and schistosomules. International Journal for Parasitology 41, 783789. doi: 10.1016/j.ijpara.2011.02.004.CrossRefGoogle ScholarPubMed
Bailey, M., Chauhan, C., Liu, C. and Unnasch, T. R. (2011). The role of polymorphisms in the spliced leader addition domain in determining promoter activity in Brugia malayi. Molecular and Biochemical Parasitology 176, 3741. doi: 10.1016/j.molbiopara.2010.11.012.CrossRefGoogle ScholarPubMed
Balu, B. and Adams, J. H. (2006). Functional genomics of Plasmodium falciparum through transposon-mediated mutagenesis. Cellular Microbiology 8, 15291536. doi: 10.1111/j.1462-5822.2006.00776.x.CrossRefGoogle ScholarPubMed
Balu, B., Chauhan, C., Maher, S. P., Shoue, D. A., Kissinger, J. C., Fraser, M. J. Jr. and Adams, J. H. (2009). piggyBac is an effective tool for functional analysis of the Plasmodium falciparum genome. BMC Microbiology 9, 83. doi: 10.1186/1471-2180-9-83.CrossRefGoogle ScholarPubMed
Balu, B., Shoue, D. A., Fraser, M. J. Jr. and Adams, J. H. (2005). High-efficiency transformation of Plasmodium falciparum by the lepidopteran transposable element piggyBac. Proceedings of the National Academy of Sciences, USA 102, 1639116396. doi: 10.1073/pnas.0504679102.CrossRefGoogle ScholarPubMed
Bargmann, C. I. (2006). Chemosensation in C. elegans. In WormBook (ed. The C. elegans Research Community) WormBook, doi/10.1895/wormbook.1.123.1, http://www.wormbook.orgGoogle Scholar
Bargmann, C. I. and Avery, L. (1995). Laser killing of cells in Caenorhabditis elegans. Methods in Cell Biology 48, 225250.CrossRefGoogle ScholarPubMed
Berezikov, E., Bargmann, C. I. and Plasterk, R. H. (2004). Homologous gene targeting in Caenorhabditis elegans by biolistic transformation. Nucleic Acids Research 32, e40. doi: 10.1093/nar/gnh033.CrossRefGoogle ScholarPubMed
Bhopale, V. M., Kupprion, E. K., Ashton, F. T., Boston, R. and Schad, G. A. (2001). Ancylostoma caninum: the finger cell neurons mediate thermotactic behavior by infective larvae of the dog hookworm. Experimental Parasitology 97, 7076. doi: 10.1006/expr.2000.4575CrossRefGoogle ScholarPubMed
Brindley, P. J., Laha, T., McManus, D. P. and Loukas, A. (2003). Mobile genetic elements colonizing the genomes of metazoan parasites. Trends in Parasitology 19, 7987. doi: 10.1016/S1471-4922(02)00061-2CrossRefGoogle ScholarPubMed
Castelletto, M. L., Massey, H. C. Jr. and Lok, J. B. (2009). Morphogenesis of Strongyloides stercoralis infective larvae requires the DAF-16 ortholog FKTF-1. PLoS Pathogens 5, e1000370. doi: 10.1371/journal.ppat.1000370.CrossRefGoogle ScholarPubMed
Chelur, D. S. and Chalfie, M. (2007). Targeted cell killing by reconstituted caspases. Proceedings of the National Academy of Sciences, USA 104, 22832288. doi: 10.1073/pnas.0610877104.CrossRefGoogle ScholarPubMed
Cheng, G., Cohen, L., Mikhli, C., Jankowska-Anyszka, M., Stepinski, J., Darzynkiewicz, E. and Davis, R. E. (2007). In vivo translation and stability of trans-spliced mRNAs in nematode embryos. Molecular and Biochemical Parasitology 153, 95106. doi: 10.1016/j.molbiopara.2007.02.003.CrossRefGoogle ScholarPubMed
Cohen, L. S., Mikhli, C., Friedman, C., Jankowska-Anyszka, M., Stepinski, J., Darzynkiewicz, E. and Davis, R. E. (2004). Nematode m7GpppG and m3(2,2,7)GpppG decapping: activities in Ascaris embryos and characterization of C. elegans scavenger DcpS. RNA 10, 16091624. doi: 10.1261/rna.7690504.CrossRefGoogle Scholar
Cohen, S. N., Chang, A. C. and Hsu, L. (1972). Nonchromosomal antibiotic resistance in bacteria: genetic transformation of Escherichia coli by R-factor DNA. Proceedings of the National Academy of Sciences of the United States of America 69(8), 21102114.CrossRefGoogle ScholarPubMed
Dalzell, J. J., McVeigh, P., Warnock, N. D., Mitreva, M., Bird, D. M., Abad, P., Fleming, C. C., Day, T. A., Mousley, A., Marks, N. J. and Maule, A. G. (2011). RNAi Effector Diversity in Nematodes. PLoS Neglected Tropical Diseases 5, e1176. doi: 10.1371/journal.pntd.0001176.CrossRefGoogle ScholarPubMed
Davis, R. E., Parra, A., LoVerde, P. T., Ribeiro, E., Glorioso, G. and Hodgson, S. (1999). Transient expression of DNA and RNA in parasitic helminths by using particle bombardment. Proceedings of the National Academy of Sciences, USA 96, 86878692. doi: 10.1073/pnas.96.15.8687.CrossRefGoogle ScholarPubMed
Dzierszinski, F., Nishi, M., Ouko, L. and Roos, D. S. (2004). Dynamics of Toxoplasma gondii differentiation. Eukaryotic Cell 3, 9921003. doi: 10.1128/EC.3.4.992-1003.2004.CrossRefGoogle ScholarPubMed
Dzierszinski, F., Pepper, M., Stumhofer, J. S., LaRosa, D. F., Wilson, E. H., Turka, L. A., Halonen, S. K., Hunter, C. A. and Roos, D. S. (2007). Presentation of Toxoplasma gondii antigens via the endogenous major histocompatibility complex class I pathway in nonprofessional and professional antigen-presenting cells. Infection and Immunity 75, 52005209. doi: 10.1128/IAI.00954-07.CrossRefGoogle Scholar
Evans, T. C. (2006). Transformation and microinjection. In WormBook (ed. The C. elegans Research Community) WormBook, doi/10.1895/wormbook.1.108.1, http://www.wormbook.org.Google Scholar
Fire, A. (1986). Integrative transformation of Caenorhabditis elegans. EMBO Journal 5, 26732680.CrossRefGoogle ScholarPubMed
Fonager, J., Franke-Fayard, B. M., Adams, J. H., Ramesar, J., Klop, O., Khan, S. M., Janse, C. J. and Waters, A. P. (2011). Development of the piggyBac transposable system for Plasmodium berghei and its application for random mutagenesis in malaria parasites. BMC Genomics 12, 155. doi: 10.1186/1471-2164-12-155.CrossRefGoogle ScholarPubMed
Forbes, W. M., Ashton, F. T., Boston, R., Zhu, X. and Schad, G. A. (2004). Chemoattraction and chemorepulsion of Strongyloides stercoralis infective larvae on a sodium chloride gradient is mediated by amphidial neuron pairs ASE and ASH, respectively. Veterinary Parasitology 120, 189198. doi: 10.1016/j.vetpar.2004.01.005.CrossRefGoogle ScholarPubMed
Ford, L., Guiliano, D. B., Oksov, Y., Debnath, A. K., Liu, J., Williams, S. A., Blaxter, M. L. and Lustigman, S. (2005). Characterization of a novel filarial serine protease inhibitor, Ov-SPI-1, from Onchocerca volvulus, with potential multifunctional roles during development of the parasite. Journal of Biological Chemistry 280(49), 4084540856. doi: 10.1074/jbc.M504434200.CrossRefGoogle ScholarPubMed
Ford, L., Zhang, J., Liu, J., Hashmi, S., Fuhrman, J. A., Oksov, Y. and Lustigman, S. (2009). Functional analysis of the cathepsin-like cysteine protease genes in adult Brugia malayi using RNA interference. PLoS Neglected Tropical Diseases 3, e377. doi: 10.1371/journal.pntd.0000377.CrossRefGoogle ScholarPubMed
Grant, W. N., Skinner, S. J. M., Howes, J. N., Grant, K., Shuttle worth, G., Heath, D. D. and Shoemaker, C. B. (2006 a). Heritable transgenesis of Parastrongyloides trichosuri: a nematode parasite of mammals. International Journal for Parasitology 36, 475483. doi: 10.1016/j.ijpara.2005.12.002.CrossRefGoogle ScholarPubMed
Grant, W. N., Stasiuk, S., Newton-Howes, J., Ralston, M., Bisset, S. A., Heath, D. D. and Shoemaker, C. B. (2006 b). Parastrongyloides trichosuri, a nematode parasite of mammals that is uniquely suited to genetic analysis. International Journal for Parasitology 36, 453466. doi: 10.1016/j.ijpara.2005.11.009.CrossRefGoogle ScholarPubMed
Gray, J. M., Hill, J. J. and Bargmann, C. I. (2005). A circuit for navigation in Caenorhabditis elegans. Proceedings of the National Academy of Sciences, USA 102, 31843191. doi: 10.1073/pnas.0409009101.CrossRefGoogle ScholarPubMed
Harbinder, S., Tavernarakis, N., Herndon, L. A., Kinnell, M., Xu, S. Q., Fire, A. and Driscoll, M. (1997). Genetically targeted cell disruption in Caenorhabditis elegans. Proceedings of the National Academy of Sciences, USA 94, 1312813133.CrossRefGoogle ScholarPubMed
Higazi, T. B., Deoliveira, A., Katholi, C. R., Shu, L., Barchue, J., Lisanby, M. and Unnasch, T. R. (2005). Identification of elements essential for transcription in Brugia malayi promoters. Journal of Molecular Biology 353, 113. doi: 10.1016/j.jmb.2005.08.014CrossRefGoogle ScholarPubMed
Higazi, T. B., Merriweather, A., Shu, L., Davis, R. and Unnasch, T. R. (2002). Brugia malayi: transient transfection by microinjection and particle bombardment. Experimental Parasitology 100, 95102. doi: 10.1016/S0014-4894(02)00004-8.CrossRefGoogle ScholarPubMed
Higazi, T. B., Shu, L. and Unnasch, T. R. (2004). Development and transfection of short-term primary cell cultures from Brugia malayi. Molecular and Biochemical Parasitology 137, 345348. doi: 10.1016/j.molbiopara.2004.06.004CrossRefGoogle ScholarPubMed
Higazi, T. B. and Unnasch, T. R. (2004). Intron encoded sequences necessary for trans splicing in transiently transfected Brugia malayi. Molecular and Biochemical Parasitology 137, 181184. doi: 10.1016/j.molbiopara.2004.04.014.CrossRefGoogle ScholarPubMed
Jackstadt, P., Wilm, T. P., Zahner, H. and Hobom, G. (1999). Transformation of nematodes via ballistic DNA transfer. Molecular and Biochemical Parasitology 103, 261266. doi: 10.1016/S0166-6851(99)00089-4CrossRefGoogle ScholarPubMed
Johnson, N. M., Behm, C. A. and Trowell, S. C. (2005). Heritable and inducible gene knockdown in C. elegans using Wormgate and the ORFeome. Gene 359, 2634. doi: 10.1016/j.gene.2005.05.034CrossRefGoogle Scholar
Joiner, K. A. and Roos, D. S. (2002). Secretory traffic in the eukaryotic parasite Toxoplasma gondii: less is more. The Journal of cell biology 157, 557563. doi: 10.1083/jcb.200112144.CrossRefGoogle ScholarPubMed
Jordan, K. A., Dupont, C. D., Tait, E. D., Liou, H. C. and Hunter, C. A. (2010). Role of the NF-kappaB transcription factor c-Rel in the generation of CD8+ T-cell responses to Toxoplasma gondii. International Immunology 22, 851861. doi: 10.1093/intimm/dxq439.CrossRefGoogle ScholarPubMed
Jordan, K. A., Wilson, E. H., Tait, E. D., Fox, B. A., Roos, D. S., Bzik, D. J., Dzierszinski, F. and Hunter, C. A. (2009). Kinetics and phenotype of vaccine-induced CD8+ T-cell responses to Toxoplasma gondii. Infection and Immunity 77, 38943901. doi: 10.1128/IAI.00024-09.CrossRefGoogle ScholarPubMed
Junio, A. B., Li, X., Massey, H. C. Jr., Nolan, T. J., Todd Lamitina, S., Sundaram, M. V. and Lok, J. B. (2008). Strongyloides stercoralis: cell- and tissue-specific transgene expression and co-transformation with vector constructs incorporating a common multifunctional 3′ UTR. Experimental Parasitology 118, 253265. doi: 10.1016/j.exppara.2007.08.018CrossRefGoogle Scholar
Kelly, W. G., Xu, S., Montgomery, M. K. and Fire, A. (1997). Distinct requirements for somatic and germline expression of a generally expressed Caernorhabditis elegans gene. Genetics 146, 227238.CrossRefGoogle ScholarPubMed
Ketschek, A. R., Joseph, R., Boston, R., Ashton, F. T. and Schad, G. A. (2004). Amphidial neurons ADL and ASH initiate sodium dodecyl sulphate avoidance responses in the infective larva of the dog hookworm Anclyostoma caninum. International Journal for Parasitology 34, 13331336. doi: 10.1016/j.ijpara.2004.08.008.CrossRefGoogle ScholarPubMed
Kines, K. J., Mann, V. H., Morales, M. E., Shelby, B. D., Kalinna, B. H., Gobert, G. N., Chirgwin, S. R. and Brindley, P. J. (2006). Transduction of Schistosoma mansoni by vesicular stomatitis virus glycoprotein-pseudotyped Moloney murine leukemia retrovirus. Experimental Parasitology 112, 209220. doi: 10.1016/j.exppara.2006.02.003CrossRefGoogle ScholarPubMed
Kines, K. J., Morales, M. E., Mann, V. H., Gobert, G. N. and Brindley, P. J. (2008). Integration of reporter transgenes into Schistosoma mansoni chromosomes mediated by pseudotyped murine leukemia virus. FASEB Journal 22, 29362948. doi: 10.1096/fj.08-108308.CrossRefGoogle ScholarPubMed
Knox, D. P., Geldhof, P., Visser, A. and Britton, C. (2007). RNA interference in parasitic nematodes of animals: a reality check? Trends in Parasitology 23, 105107. doi: 10.1016/j.pt.2007.01.007CrossRefGoogle ScholarPubMed
Laha, T., Kewgrai, N., Loukas, A. and Brindley, P. J. (2006). The dingo non-long terminal repeat retrotransposons from the genome of the hookworm, Ancylostoma caninum. Experimental Parasitology 113, 142153. doi: 10.1016/j.exppara.2005.12.018.CrossRefGoogle ScholarPubMed
Laha, T., Loukas, A., Wattanasatitarpa, S., Somprakhon, J., Kewgrai, N., Sithithaworn, P., Kaewkes, S., Mitreva, M. and Brindley, P. J. (2007). The bandit, a new DNA transposon from a hookworm-possible horizontal genetic transfer between host and parasite. PLoS Neglected Tropical Diseases 1, e35. doi: 10.1371/journal.pntd.0000035.CrossRefGoogle ScholarPubMed
Lee, R. Y., Hench, J. and Ruvkun, G. (2001). Regulation of C. elegans DAF-16 and its human ortholog FKHRL1 by the daf-2 insulin like signalling pathway. Current Biology 11, 19501957. doi: 10.1016/S0960(01)00595-4.CrossRefGoogle Scholar
Li, X., Massey, H. C., Nolan, T. J., Schad, G. A., Kraus, K., Sundaram, M. and Lok, J. B. (2006). Successful transgenesis of the parasitic nematode Strongyloides stercoralis requires endogenous non-coding control elements. International Journal for Parasitology 36, 671679. doi: 10.1016/j.ijpara.2005.12.007CrossRefGoogle ScholarPubMed
Li, X., Shao, H., Junio, A. B., Nolan, T. J., Massey, H. C. Jr., Pearce, E. J., Viney, M. E. and Lok, J. B. (2011). Transgenesis in the parasitic nematode Strongyloides ratti. Molecular and Biochemical Parasitology 179, 114119. doi: 10.1016/jmolbiopara.2011.06.0002.CrossRefGoogle ScholarPubMed
Lin, K., Dorman, J. B., Rodan, A. and Kenyon, C. (1997). daf-16: An HNF-3/forkhead family member that can function to double the life-span of Caenorhabditis elegans. Science 278, 13191322. doi: 10.1126/science.278.5341.1319.CrossRefGoogle ScholarPubMed
Liu, C., de Oliveira, A., Higazi, T. B., Ghedin, E., DePasse, J. and Unnasch, T. R. (2007). Sequences necessary for trans-splicing in transiently transfected Brugia malayi. Molecular and Biochemical Parasitology 156, 6273. doi: 10.1016/j.molbiopara.2007.07.012CrossRefGoogle ScholarPubMed
Liu, C., Chauhan, C., Katholi, C. R. and Unnasch, T. R. (2009). The splice leader addition domain represents an essential conserved motif for heterologous gene expression in B. malayi. Molecular and Biochemical Parasitology 166, 1521. doi: 10.1016/j.molbiopara.2009.02.004.CrossRefGoogle ScholarPubMed
Liu, C., Chauhan, C. and Unnasch, T. R. (2010 a). The role of local secondary structure in the function of the trans-splicing motif of Brugia malayi. Molecular and Biochemical Parasitology 169, 115119. doi: 10.1016/j.molbiopara.2009.10.003.CrossRefGoogle ScholarPubMed
Liu, C., Oliveira, A., Chauhan, C., Ghedin, E. and Unnasch, T. R. (2010 b). Functional analysis of putative operons in Brugia malayi. International Journal for Parasitology 40, 6371. doi: 10.1016/j.ijpara.2009.07.001.CrossRefGoogle ScholarPubMed
Lok, J. B. (2009). Transgenesis in parasitic nematodes: building a better array. Trends in Parasitology 25, 345347. doi: 10.1016/j.pt.2009.05.002CrossRefGoogle ScholarPubMed
Lok, J. B. and Artis, D. (2008). Transgenesis and neuronal ablation in parasitic nematodes: revolutionary new tools to dissect host-parasite interactions. Parasite Immunology 30(4), 203214. doi: 10.1111/j.1365-3024.2008.01006.x.CrossRefGoogle ScholarPubMed
Lok, J. B. and Massey, H. C. Jr. (2002). Transgene expression in Strongyloides stercoralis following gonadal microinjection of DNA constructs. Molecular and Biochemical Parasitology 119, 279284. doi: 10.1016/S0166-6851(01)00414-5CrossRefGoogle ScholarPubMed
Lopez, P. M., Boston, R., Ashton, F. T. and Schad, G. A. (2000). The neurons of class ALD mediate thermotaxis in the parasitic nematode, Strongyloides stercoralis. International Journal for Parasitology 30, 11151121. doi: 10.1016/S0020-7519(00)00087-4CrossRefGoogle ScholarPubMed
Lustigman, S., Zhang, J., Liu, J., Oksov, Y. and Hashmi, S. (2004). RNA interference targeting cathepsin L and Z-like cysteine proteases of Onchocerca volvulus confirmed their essential function during L3 molting. Molecular and Biochemical Parasitology 138, 165170. doi: 10.1016/j.molbiopara.2004.08.003CrossRefGoogle ScholarPubMed
Maizels, R. M., Pearce, E. J., Artis, D., Yazdanbakhsh, M. and Wynn, T. A. (2009). Regulation of pathogenesis and immunity in helminth infections. Journal of Experimental Medicine 206, 20592066. doi: 10.1084/jem.20091903.CrossRefGoogle ScholarPubMed
Mandel, M. and Higa, A. (1970). Calcium-dependent bacteriophage DNA infection. Journal of Molecular Biology 53, 159162.CrossRefGoogle ScholarPubMed
Mann, V. H., Morales, M. E., Kines, K. J. and Brindley, P. J. (2007). Transgenesis of schistosomes: approaches employing mobile genetic elements. Parasitology 134, 113. doi: 10.1017/S0031182007003824.Google Scholar
Markstein, M., Pitsouli, C., Villalta, C., Celniker, S. E. and Perrimon, N. (2008). Exploiting position effects and the gypsy retrovirus insulator to engineer precisely expressed transgenes. Nature Genetics 40, 476483. doi: 10.1038/ng.101.CrossRefGoogle ScholarPubMed
Massey, H. C. Jr., Nishi, M., Chaudhary, K., Pakpour, N. and Lok, J. B. (2003). Structure and developmental expression of Strongyloides stercoralis fktf-1, a proposed ortholog of daf-16 in Caenorhabditis elegans. International Journal for Parasitology 33, 15371544. doi: 10.1016/S0020-7519(03)00205-4CrossRefGoogle ScholarPubMed
Mello, C. and Fire, A. (1995). DNA transformation. Methods in Cell Biology 48, 451482.CrossRefGoogle ScholarPubMed
Mello, C., Kramer, J. M., Stinchcomb, D. T. and Ambros, V. (1991). Efficient gene transfer in C. elegans: extrachromosomal maintenance and integration of transforming sequences. The EMBO Journal 10(12), 39593970.CrossRefGoogle Scholar
Morales, M. E., Mann, V. H., Kines, K. J., Gobert, G. N., Fraser, M. J. Jr., Kalinna, B. H., Correnti, J. M., Pearce, E. J. and Brindley, P. J. (2007). piggyBac transposon mediated transgenesis of the human blood fluke, Schistosoma mansoni. FASEB Journal 21, 34793489. doi: 10.1096/fj.07-8726com.CrossRefGoogle ScholarPubMed
Nakazawa, Y., Huye, L. E., Dotti, G., Foster, A. E., Vera, J. F., Manuri, P. R., June, C. H., Rooney, C. M. and Wilson, M. H. (2009). Optimization of the piggyBac transposon system for the sustained genetic modification of human T lymphocytes. Journal of Immunotherapy 32, 826836. doi: 10.1097/CJI.0b013e3181ad762b.CrossRefGoogle ScholarPubMed
Nishi, M., Hu, K., Murray, J. M. and Roos, D. S. (2008). Organellar dynamics during the cell cycle of Toxoplasma gondii. Journal of Cell Science 121, 15591568. doi: 10.1242/jcs.021089.CrossRefGoogle ScholarPubMed
Ogg, S., Paradis, S., Gottlieb, S., Patterson, G. I., Lee, L., Tissenbaum, H. A. and Ruvkun, G. (1997). The fork head transcription factor DAF-16 transduces insulin-like metabolic and longevity signals in C. elegans. Nature 389, 994999. doi: 10.1038/40194.CrossRefGoogle ScholarPubMed
Oliveira, A., Katholi, C. R. and Unnasch, T. R. (2008). Characterization of the promoter of the Brugia malayi 12 kDa small subunit ribosomal protein (RPS12) gene. International Journal for Parasitology 38, 11111119. doi: 10.1016/j.ijpara.2008.02.002.CrossRefGoogle ScholarPubMed
Pepper, M., Dzierszinski, F., Crawford, A., Hunter, C. A. and Roos, D. (2004). Development of a system to study CD4+-T-cell responses to transgenic ovalbumin-expressing Toxoplasma gondii during toxoplasmosis. Infection and Immunity 72, 72407246. doi: 10.1128/IAI.72.12.7240-7246.2004.CrossRefGoogle ScholarPubMed
Pfarr, K., Heider, U. and Hoerauf, A. (2006). RNAi mediated silencing of actin expression in adult Litomosoides sigmodontis is specific, persistent and results in a phenotype. International Journal for Parasitology 36, 661669. doi: 10.1016/j.ijpara.2006.01.010.CrossRefGoogle ScholarPubMed
Plasterk, R. H., Izsvak, Z. and Ivics, Z. (1999). Resident aliens: the Tc1/mariner superfamily of transposable elements. Trends in Genetics 15, 326332. doi: 10.1016/S0168-9525(99)01777-1CrossRefGoogle ScholarPubMed
Praitis, V. (2006). Creation of transgenic lines using microparticle bombardment methods. Methods in Molecular Biology 351, 93107.Google ScholarPubMed
Praitis, V., Casey, E., Collar, D. and Austin, J. (2001). Creation of low-copy integrated transgenic lines in Caenorhabditis elegans. Genetics 157, 12171226.CrossRefGoogle ScholarPubMed
Saenz, S. A., Noti, M. and Artis, D. (2010). Innate immune cell populations function as initiators and effectors in Th2 cytokine responses. Trends in Immunology 31, 407413. doi: 10.1016/j.it.2010.09.001.CrossRefGoogle ScholarPubMed
Samarasinghe, B., Knox, D. P. and Britton, C. (2011). Factors affecting susceptibility to RNA interference in Haemonchus contortus and in vivo silencing of an H11 aminopeptidase gene. International Journal for Parasitology 41, 5159. doi: 10.1016/j.ijpara.2010.07.005.CrossRefGoogle ScholarPubMed
Sambrook, J., Fritsch, E. F. and Maniatis, T. (1989). Molecular Cloning, A Laboratory Manual. pp. 16.3016.40. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.Google Scholar
Schad, G. A. (1989). Morphology and life history of Strongyloides stercoralis. In Strongyloidiasis: a Major Roundworm Infection of Man (ed. Grove, D. I.), pp. 85104. Taylor and Francis Inc., Philadelphia.Google Scholar
Schlager, B., Wang, X., Braach, G. and Sommer, R. J. (2009). Molecular cloning of a dominant roller mutant and establishment of DNA-mediated transformation in the nematode Pristionchus pacificus. Genesis 47, 300304. doi: 10.1002/dvg.20499.CrossRefGoogle ScholarPubMed
Shu, L., Katholi, C. R., Higazi, T. and Unnasch, T. R. (2003). Analysis of the Brugia malayi HSP70 promoter using a homologous transient transfection system. Molecular and Biochemical Parasitology 128, 6775. doi: 10.1016/S0166-6851(03)00052-5CrossRefGoogle ScholarPubMed
Song, C., Gallup, J. M., Day, T. A., Bartholomay, L. C. and Kimber, M. J. (2010). Development of an in vivo RNAi protocol to investigate gene function in the filarial nematode, Brugia malayi. PLoS Pathogens 6, e1001239. doi: 10.1371/journal.ppat.1001239.CrossRefGoogle Scholar
Soragni, E., Herman, D., Dent, S. Y., Gottesfeld, J. M., Wells, R. D. and Napierala, M. (2008). Long intronic GAA*TTC repeats induce epigenetic changes and reporter gene silencing in a molecular model of Friedreich ataxia. Nucleic Acids Research 36, 60566065. doi: 10.1093/nar/gkn604.CrossRefGoogle Scholar
Stinchcomb, D. T., Shaw, J. E., Carr, S. H. and Hirsh, D. (1985). Extrachromosomal DNA transformation of Caenorhabditis elegans. Molecular and Cellular Biology 5, 34843496.Google ScholarPubMed
Tachu, B., Pillai, S., Lucius, R. and Pogonka, T. (2008). Essential role of chitinase in the development of the filarial nematode Acanthocheilonema viteae. Infection and Immunity 76, 221228. doi: 10.1128/IAI.00701-07.CrossRefGoogle ScholarPubMed
Tait, E. D., Jordan, K. A., Dupont, C. D., Harris, T. H., Gregg, B., Wilson, E. H., Pepper, M., Dzierszinski, F., Roos, D. S. and Hunter, C. A. (2010). Virulence of Toxoplasma gondii is associated with distinct dendritic cell responses and reduced numbers of activated CD8+ T cells. Journal of Immunology 185, 15021512. doi: 10.4049/jimmunol.0903450.CrossRefGoogle ScholarPubMed
Tavernarakis, N., Wang, S. L., Dorovkov, M., Ryazanov, A. and Driscoll, M. (2000). Heritable and inducible genetic interference by double-stranded RNA encoded by transgenes. Nature Genetics 24, 180183. doi: 10.1038/72850CrossRefGoogle ScholarPubMed
Tchoubrieva, E. B., Ong, P. C., Pike, R. N., Brindley, P. J. and Kalinna, B. H. (2010). Vector-based RNA interference of cathepsin B1 in Schistosoma mansoni. Cellular and molecular life sciences : CMLS 67, 37393748. doi: 10.1007/s00018-010-0345-3.CrossRefGoogle ScholarPubMed
Thibault, S. T., Singer, M. A., Miyazaki, W. Y., Milash, B., Dompe, N. A., Singh, C. M., Buchholz, R., Demsky, M., Fawcett, R., Francis-Lang, H. L., Ryner, L., Cheung, L. M., Chong, A., Erickson, C., Fisher, W. W., Greer, K., Hartouni, S. R., Howie, E., Jakkula, L., Joo, D., Killpack, K., Laufer, A., Mazzotta, J., Smith, R. D., Stevens, L. M., Stuber, C., Tan, L. R., Ventura, R., Woo, A., Zakrajsek, I., Zhao, L., Chen, F., Swimmer, C., Kopczynski, C., Duyk, G., Winberg, M. L. and Margolis, J. (2004). A complementary transposon tool kit for Drosophila melanogaster using P and piggyBac. Nature Genetics 36, 283287. doi: 10.1038/ng1314.CrossRefGoogle ScholarPubMed
Tzertzinis, G., Egana, A. L., Palli, S. R., Robinson-Rechavi, M., Gissendanner, C. R., Liu, C., Unnasch, T. R. and Maina, C. V. (2010). Molecular evidence for a functional ecdysone signaling system in Brugia malayi. PLoS Neglected Tropical Diseases 4(3), e625. doi: 10.1371/journal.pntd.0000625.CrossRefGoogle ScholarPubMed
Unnasch, T. R., Bradley, J., Beauchamp, J., Tuan, R. and Kennedy, M. W. (1999). Characterization of a putative nuclear receptor from Onchocerca volvulus. Molecular and Biochemical Parasitology 104, 259269. doi: 10.1016/S0166-6851(99)00152-8.CrossRefGoogle ScholarPubMed
Viney, M. E. (1999). Exploiting the life cycle of Strongyloides ratti. Parasitology Today 15, 231235. doi: 10.1016/S0169-4758(99)01452-0.CrossRefGoogle Scholar
Viney, M. E. and Ashford, R. W. (1990). The ultrastructure of the peri-vulval region of Strongyloides cebus. Journal of Helminthology 64, 2328.CrossRefGoogle ScholarPubMed
Viney, M. E., Green, L. D., Brooks, J. A. and Grant, W. N. (2002). Chemical mutagenesis of the parasitic nematode Strongyloides ratti to isolate ivermectin resistant mutants. International Journal for Parasitology 32, 16771682. doi: 10.1016/S0020-7519(02)00157-1.CrossRefGoogle ScholarPubMed
Viney, M. E. and Lok, J. B. (2007). Strongyloides spp. In WormBook (ed. The C. elegans Research Community) WormBook, doi/10.1895/wormbook.1.141.1, http://www.wormbook.org.Google Scholar
Wilm, T., Demel, P., Koop, H. U., Schnabel, H. and Schnabel, R. (1999). Ballistic transformation of Caenorhabditis elegans. Gene 229, 3135.CrossRefGoogle ScholarPubMed
Winston, W. M., Sutherlin, M., Wright, A. J., Feinberg, E. H. and Hunter, C. P. (2007). Caenorhabditis elegans SID-2 is required for environmental RNA interference. Proceedings of the National Academy of Sciences of the United States of America 104, 1056510570. doi: 10.1073/pnas.0611282104.CrossRefGoogle ScholarPubMed
Xu, S., Liu, C., Tzertzinis, G., Ghedin, E., Evans, C. C., Kaplan, R. and Unnasch, T. R. (2011). In vivo transfection of developmentally competent Brugia malayi infective larvae. International Journal for Parasitology 41, 355362. doi: 10.1016/j.ijpara.2010.10.005.CrossRefGoogle ScholarPubMed
Yang, S., Brindley, P. J., Zeng, Q., Li, Y., Zhou, J., Liu, Y., Liu, B., Cai, L., Zeng, T., Wei, Q., Lan, L. and McManus, D. P. (2010). Transduction of Schistosoma japonicum schistosomules with vesicular stomatitis virus glycoprotein pseudotyped murine leukemia retrovirus and expression of reporter human telomerase reverse transcriptase in the transgenic schistosomes. Molecular and Biochemical Parasitology 174, 109116. doi: 10.1016/j.molbiopara.2010.07.007.CrossRefGoogle ScholarPubMed
Zhao, Z. R., Lei, L., Liu, M., Zhu, S. C., Ren, C. P., Wang, X. N. and Shen, J. J. (2008). Schistosoma japonicum: inhibition of Mago nashi gene expression by shRNA-mediated RNA interference. Experimental Parasitology 119, 379384. doi: 10.1016/j.exppara.2008.03.015.CrossRefGoogle ScholarPubMed
Zwaal, R. R., Mendel, J. E., Sternberg, P. W. and Plasterk, R. H. (1997). Two neuronal G proteins are involved in chemosensation of the Caenorhabditis elegans dauer-inducing pheromone. Genetics 145, 715727.CrossRefGoogle ScholarPubMed