Zinc finger nuclease

Zinc finger nuclease

Zinc-finger nucleases (ZFNs) are artificial restriction enzymes generated by fusing a zinc finger DNA-binding domain to a DNA-cleavage domain. Zinc finger domains can be engineered to target desired DNA sequences and this enables zinc-finger nucleases to target unique sequences within complex genomes. By taking advantage of endogenous DNA repair machinery, these reagents can be used to precisely alter the genomes of higher organisms.

Contents

DNA-cleavage domain

The non-specific cleavage domain from the type IIs restriction endonuclease FokI is typically used as the cleavage domain in ZFNs.[1] This cleavage domain must dimerize in order to cleave DNA[2] and thus a pair of ZFNs are required to target non-palindromic DNA sites. Standard ZFNs fuse the cleavage domain to the C-terminus of each zinc finger domain. In order to allow the two cleavage domains to dimerize and cleave DNA, the two individual ZFNs must bind opposite strands of DNA with their C-termini a defined distance apart. The most commonly used linker sequences between the zinc finger domain and the cleavage domain requires the 5' edge of each binding site to be separated by 5 to 7 bp.[3]

Several different protein engineering techniques have been employed to improve both the activity and specificity of the nuclease domain used in ZFNs. Directed evolution has been employed to generate a FokI variant with enhanced cleavage activity that the authors dubbed "Sharkey".[4] Structure-based design has also been employed to improve the cleavage specificity of FokI by modifying the dimerization interface so that only the intended heterodimeric species are active.[5][6][7][8]

DNA-binding domain

The DNA-binding domains of individual ZFNs typically contain between three and six individual zinc finger repeats and can each recognize between 9 and 18 basepairs. If the zinc finger domains are perfectly specific for their intended target site then even a pair of 3-finger ZFNs that recognize a total of 18 basepairs can theoretically target a single locus in a mammalian genome.

Various strategies have been developed to engineer Cys2His2 zinc fingers to bind desired sequences.[9] These include both "modular assembly" and selection strategies that employ either phage display or cellular selection systems.

The most straightforward method to generate new zinc-finger arrays is to combine smaller zinc-finger "modules" of known specificity. The most common modular assembly process involves combining three separate zinc fingers that can each recognize a 3 basepair DNA sequence to generate a 3-finger array that can recognize a 9 basepair target site. Other procedures can utilize either 1-finger or 2-finger modules to generate zinc-finger arrays with six or more individual zinc fingers. The main drawback with this procedure is the specificities of individual zinc fingers can overlap and can depend on the context of the surrounding zinc fingers and DNA. Without methods to account for this "context dependence", the standard modular assembly procedure often fails unless it is used to recognize sequences of the form (GNN)N.[10]

Numerous selection methods have been used to generate zinc-finger arrays capable of targeting desired sequences. Initial selection efforts utilized phage display to select proteins that bound a given DNA target from a large pool of partially randomized zinc-finger arrays. More recent efforts have utilized yeast one-hybrid systems, bacterial one-hybrid and two-hybrid systems, and mammalian cells. A promising new method to select novel zinc-finger arrays utilizes a bacterial two-hybrid system and has been dubbed "OPEN" by its creators.[11] This system combines pre-selected pools of individual zinc fingers that were each selected to bind a given triplet and then utilizes a second round of selection to obtain 3-finger arrays capable of binding a desired 9-bp sequence. This system was developed by the Zinc-Finger Consortium as an alternative to commercial sources of engineered zinc-finger arrays.

(see: Zinc finger chimera for more info on zinc finger selection techniques)

Applications

Zinc finger nucleases have become useful reagents for manipulating the genomes of many plants and animals including arabidopsis[12][13], tobacco[14][15], soybean,[16] corn,[17] Drosophila melanogaster,[18] C. elegans,[19] sea urchin,[20] silkworm,[21] zebrafish,[22], frogs,[23] mice,[24] rats,[25] rabbits,[26] pigs,[27] cattle,[28] and various types of mammalian cells.[29] Zinc finger nucleases have also been used in a mouse model of haemophilia[30] and an ongoing clinical trial is evaluating Zinc finger nucleases that disrupt the CCR5 gene in CD4+ human T-cells as a potential treatment for HIV/AIDS. ZFNs are also used for the creation of a new generation of genetic disease models called isogenic human disease models.

Disabling an allele

ZFNs can be used to disable dominant mutations in heterozygous individuals by producing double strand breaks (DSBs) in the DNA (see Genetic recombination) in the mutant allele which will, in the absence of a homologous template, be repaired by non-homologous end-joining (NHEJ). NHEJ repairs DSBs by joining the two ends together and usually produces no mutations, provided that the cut is clean and uncomplicated. In some instances however, the repair will be imperfect, resulting in deletion or insertion of base-pairs, producing frame-shift and preventing the production of the harmful protein.[31] Multiple pairs of ZFNs can also be used to completely remove entire large segments of genomic sequence.[32]

ZFNs have also been used modify disease-causing alleles in triplet repeat disorders. Expanded CAG/CTG repeat tracts are the genetic basis for more than a dozen inherited neurological disorders including Huntington’s disease, myotonic dystrophy, and several spinocerebellar ataxias. It has been demonstrated in human cells that ZFNs can direct double-strand breaks (DSBs) to CAG repeats and shrink the repeat from long pathological lengths to short, less toxic lengths.[33]

Recently, a group of researchers have successfully applied the ZFN technology to genetically modify the gol pigment gene and the ntl gene in zebrafish embryo. Specific zinc-finger motifs were engineered to recognize distinct DNA sequences. The ZFN-encoding mRNA was injected into one-cell embryos and a high percentage of animals carried the desired mutations and phenotypes. Their research work demonstrated that ZFNs can specifically and efficiently create heritable mutant alleles at loci of interest in the germ line, and ZFN-induced alleles can be propagated in subsequent generations.

Similar research of using ZFNs to create specific mutations in zebrafish embryo has also been carried out by other research groups. The kdr gene in zebra fish encodes for the vascular endothelial growth factor-2 receptor. Mutagenic lesions at this target site was induced using ZFN technique by a group of researchers in US. They suggested that the ZFN technique allows straightforward generation of a targeted allelic series of mutants; it does not rely on the existence of species-specific embryonic stem cell lines and is applicable to other vertebrates, especially those whose embryos are easily available; finally, it is also feasible to achieve targeted knock-ins in zebrafish, therefore it is possible to create human disease models that are heretofore inaccessible.

Allele editing

ZFNs are also used to rewrite the sequence of an allele by invoking the homologous recombination (HR) machinery to repair the DSB using the supplied DNA fragment as a template. The HR machinery searches for homology between the damaged chromosome and the extra-chromosomal fragment and copies the sequence of the fragment between the two broken ends of the chromosome, regardless of whether the fragment contains the original sequence. If the subject is homozygous for the target allele, the efficiency of the technique is reduced since the undamaged copy of the allele may be used as a template for repair instead of the supplied fragment.

Gene therapy

The success of gene therapy depends on the efficient insertion of therapeutic genes at the appropriate chromosomal target sites within the human genome, without causing cell injury, oncogenic mutations or an immune response. The construction of plasmid vectors is simple and straightforward. Custom-designed ZFNs that combine the non-specific cleavage domain (N) of FokI endonuclease with zinc-finger proteins (ZFPs) offer a general way to deliver a site-specific DSB to the genome, and stimulate local homologous recombination by several orders of magnitude. This makes targeted gene correction or genome editing a viable option in human cells. Since ZFN-encoded plasmids could be used to transiently express ZFNs to target a DSB to a specific gene locus in human cells, they offer an excellent way for targeted delivery of the therapeutic genes to a pre-selected chromosomal site. The ZFN-encoded plasmid-based approach has the potential to circumvent all the problems associated with the viral delivery of therapeutic genes.[34] The first therapeutic applications of ZFNs are likely to involve ex vivo therapy using a patients own stem cells. After editing the stem cell genome, the cells could be expanded in culture and reinserted into the patient to produce differentiated cells with corrected functions. The initial targets will likely include the causes of monogenic diseases such as the IL2Rγ gene and the b-globin gene for gene correction and CCR5 gene for mutagenesis and disablement.[31]

Potential Problems

Off-target Cleavage

If the zinc finger domains are not specific enough for their target site or they do not target a unique site within the genome of interest, off-target cleavage may occur. Such off-target cleavage may lead to the production of enough double-strand breaks to overwhelm the repair machinery and consequently yield chromosomal rearrangements and/or cell death. Off-target cleavage events may also promote random integration of donor DNA.[31] Despite advances in engineering both more specific zinc finger domains and modified FokI cleavage domains ,[35] ZFN off-target activity is still a significant concern.[36] Two separate methods have been demonstrated to decrease off-target cleavage for 3-finger ZFNs that target two adjacent 9-basepair sites.[37] Other groups use ZFNs with 4, 5 or 6 zinc fingers that target longer and presumably rarer sites and such ZFNs could theoretically yield less off-target activity. A comparison of a pair of 3-finger ZFNs and a pair of 4-finger ZFNs detected off-target cleavage in human cells at 31 loci for the 3-finger ZFNs and at 9 loci for the 4-finger ZFNs.[38] Whole genome sequencing of C. elegans modified with a pair of 5-finger ZFNs found only the intended modification and a deletion at a site "unrelated to the ZFN site" indicating this pair of ZFNs was capable of targeting a unique site in the C. elegans genome.[19]

Immunogenicity

As with many foreign proteins inserted into the human body, there is a risk of an immunological response against the therapeutic agent and the cells in which it is active. Since the protein will only need to be expressed transiently however, the time over which a response may develop is short.[31]

Prospects

The ability to precisely manipulate the genomes of plants, animals and insects has numerous applications in basic research, agriculture, and human therapeutics. Using ZFNs to modify endogenous genes has traditionally been a difficult task due mainly to the challenge of generating zinc finger domains that target the desired sequence with sufficient specificity. Improved methods of engineering zinc finger domains and the availability of ZFNs from a commercial supplier now put this technology in the hands of increasing numbers of researchers. Several groups are also developing other types of engineered nucleases including engineered homing endonucleases [39] [40] and nucleases based on engineered TAL effectors. [41] [42] TAL effector nucleases (TALENs) are particularly interesting because TAL effectors appear to be very simple to engineer [43] [44] and TALENs can be used to target endogenous loci in human cells.[45] But to date no one has reported the isolation of clonal cell lines or transgenic organisms using such reagents. One type of ZFN, known as SB-728-T, has been tested for potential application in the treatment of HIV.[46]

See also

References

  1. ^ Kim, YG; Cha, J., Chandrasegaran, S. (1996). "Hybrid restriction enzymes: zinc finger fusions to Fok I cleavage domain". Proc Natl Acad Sci USA 93 (3): 1156–60. doi:10.1073/pnas.93.3.1156. PMC 40048. PMID 8577732. http://www.pnas.org/content/93/3/1156.abstract. 
  2. ^ Bitinaite, J.; D. A. Wah, Aggarwal, A. K., Schildkraut, I. (1998). "FokI dimerization is required for DNA cleavage". Proc Natl Acad Sci USA 95 (18): 10570–5. doi:10.1073/pnas.95.18.10570. PMC 27935. PMID 9724744. http://www.pnas.org/cgi/content/full/95/18/10570. 
  3. ^ Cathomen T, Joung JK (July 2008). "Zinc-finger nucleases: the next generation emerges". Mol. Ther. 16 (7): 1200–7. doi:10.1038/mt.2008.114. PMID 18545224. http://www.nature.com/mt/journal/v16/n7/abs/mt2008114a.html. 
  4. ^ Guo, J.; Gaj, T.; Barbas Iii, C. F. (2010). "Directed Evolution of an Enhanced and Highly Efficient FokI Cleavage Domain for Zinc Finger Nucleases". Journal of Molecular Biology 400 (1): 96. doi:10.1016/j.jmb.2010.04.060. PMC 2885538. PMID 20447404. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2885538.  edit
  5. ^ Szczepek, M.; Brondani, V.; Büchel, J.; Serrano, L.; Segal, D. J.; Cathomen, T. (2007). "Structure-based redesign of the dimerization interface reduces the toxicity of zinc-finger nucleases". Nature Biotechnology 25 (7): 786. doi:10.1038/nbt1317. PMID 17603476.  edit
  6. ^ Miller, J. C.; Holmes, M. C.; Wang, J.; Guschin, D. Y.; Lee, Y. L.; Rupniewski, I.; Beausejour, C. M.; Waite, A. J. et al. (2007). "An improved zinc-finger nuclease architecture for highly specific genome editing". Nature Biotechnology 25 (7): 778–785. doi:10.1038/nbt1319. PMID 17603475.  edit
  7. ^ Doyon, Y.; Vo, T. D.; Mendel, M. C.; Greenberg, S. G.; Wang, J.; Xia, D. F.; Miller, J. C.; Urnov, F. D. et al. (2010). "Enhancing zinc-finger-nuclease activity with improved obligate heterodimeric architectures". Nature Methods 8 (1): 74. doi:10.1038/nmeth.1539. PMID 21131970.  edit
  8. ^ Ramalingam, S.; Kandavelou, K.; Rajenderan, R.; Chandrasegaran, S. (2011). "Creating Designed Zinc-Finger Nucleases with Minimal Cytotoxicity". Journal of Molecular Biology 405 (3): 630. doi:10.1016/j.jmb.2010.10.043. PMC 3017627. PMID 21094162. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3017627.  edit
  9. ^ C.O. Pabo; E.Peisach; R.A. Grant (2001). "Design and Selection of Novel Cys2His2 Zinc Finger Proteins". Annu. Rev. Biochem. 70: 313–40. doi:10.1146/annurev.biochem.70.1.313. PMID 11395410. http://arjournals.annualreviews.org/doi/abs/10.1146/annurev.biochem.70.1.313?url_ver=Z39.88-2003&rfr_id=ori:rid:crossref.org&rfr_dat=cr_pub%3dncbi.nlm.nih.gov. 
  10. ^ Ramirez CL, Foley JE, Wright DA, et al. (May 2008). "Unexpected failure rates for modular assembly of engineered zinc fingers". Nat. Methods 5 (5): 374–5. doi:10.1038/nmeth0508-374. PMID 18446154. 
  11. ^ Maeder ML, et al.' (September 2008). "Rapid "open-source" engineering of customized zinc-finger nucleases for highly efficient gene modification". Mol. Cell 31 (2): 294–301. doi:10.1016/j.molcel.2008.06.016. PMC 2535758. PMID 18657511. http://linkinghub.elsevier.com/retrieve/pii/S1097-2765(08)00461-9. 
  12. ^ Zhang, F.; Maeder, M. L.; Unger-Wallace, E.; Hoshaw, J. P.; Reyon, D.; Christian, M.; Li, X.; Pierick, C. J. et al. (2010). "High frequency targeted mutagenesis in Arabidopsis thaliana using zinc finger nucleases". Proceedings of the National Academy of Sciences 107 (26): 12028–12033. doi:10.1073/pnas.0914991107.  edit
  13. ^ Osakabe, K.; Osakabe, Y.; Toki, S. (2010). "Site-directed mutagenesis in Arabidopsis using custom-designed zinc finger nucleases". Proceedings of the National Academy of Sciences 107 (26): 12034–12039. doi:10.1073/pnas.1000234107.  edit
  14. ^ Cai, C. Q.; Doyon, Y.; Ainley, W. M.; Miller, J. C.; Dekelver, R. C.; Moehle, E. A.; Rock, J. M.; Lee, Y. L. et al. (2008). "Targeted transgene integration in plant cells using designed zinc finger nucleases". Plant Molecular Biology 69 (6): 699–709. doi:10.1007/s11103-008-9449-7. ISBN 1110300894497. PMID 19112554.  edit
  15. ^ Townsend, J. A.; Wright, D. A.; Winfrey, R. J.; Fu, F.; Maeder, M. L.; Joung, J. K.; Voytas, D. F. (2009). "High-frequency modification of plant genes using engineered zinc-finger nucleases". Nature 459 (7245): 442–445. Bibcode 2009Natur.459..442T. doi:10.1038/nature07845. PMC 2743854. PMID 19404258. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2743854.  edit
  16. ^ Curtin, S. J.; Zhang, F.; Sander, J. D.; Haun, W. J.; Starker, C.; Baltes, N. J.; Reyon, D.; Dahlborg, E. J. et al. (2011). "Targeted Mutagenesis of Duplicated Genes in Soybean with Zinc-Finger Nucleases". Plant Physiology 156 (2): 466–473. doi:10.1104/pp.111.172981. PMID 21464476.  edit
  17. ^ Shukla VK, Doyon Y, Miller JC, et al. (May 2009). "Precise genome modification in the crop species Zea mays using zinc-finger nucleases". Nature 459 (7245): 437–41. Bibcode 2009Natur.459..437S. doi:10.1038/nature07992. PMID 19404259. 
  18. ^ Bibikova, M.; Beumer, K.; Trautman, J.; Carroll, D. (2003). "Enhancing Gene Targeting with Designed Zinc Finger Nucleases". Science 300 (5620): 764. doi:10.1126/science.1079512. PMID 12730594.  edit
  19. ^ a b Wood, A. J.; Lo, T. -W.; Zeitler, B.; Pickle, C. S.; Ralston, E. J.; Lee, A. H.; Amora, R.; Miller, J. C. et al. (2011). "Targeted Genome Editing Across Species Using ZFNs and TALENs". Science. doi:10.1126/science.1207773.  edit
  20. ^ Ochiai H, Fujita K, Suzuki K, et al. (Aug 2010). "Targeted mutagenesis in the sea urchin embryo using zinc-finger nucleases". Genes to Cells 15 (8): 875–85,. doi:10.1111/j.1365-2443.2010.01425.x. PMID 20604805. 
  21. ^ Takasu, Y.; Kobayashi, I.; Beumer, K.; Uchino, K.; Sezutsu, H.; Sajwan, S.; Carroll, D.; Tamura, T. et al. (2010). "Targeted mutagenesis in the silkworm Bombyx mori using zinc finger nuclease mRNA injection". Insect Biochemistry and Molecular Biology 40 (10): 759–765. doi:10.1016/j.ibmb.2010.07.012. PMID 20692340.  edit
  22. ^ S.C. Ekker (2008). "Zinc Finger–Based Knockout Punches for Zebrafish Genes". Zebrafish 5 (2): 1121–3. doi:10.1089/zeb.2008.9988. PMC 2849655. PMID 18554175. http://www.liebertonline.com/doi/abs/10.1089/zeb.2008.9988. 
  23. ^ Young, J. J.; Cherone, J. M.; Doyon, Y.; Ankoudinova, I.; Faraji, F. M.; Lee, A. H.; Ngo, C.; Guschin, D. Y. et al. (2011). "Efficient targeted gene disruption in the soma and germ line of the frog Xenopus tropicalis using engineered zinc-finger nucleases". Proceedings of the National Academy of Sciences 108 (17): 7052–7057. doi:10.1073/pnas.1102030108.  edit
  24. ^ Goldberg, A. D.; Banaszynski, L. A.; Noh, K. M.; Lewis, P. W.; Elsaesser, S. J.; Stadler, S.; Dewell, S.; Law, M. et al. (2010). "Distinct Factors Control Histone Variant H3.3 Localization at Specific Genomic Regions". Cell 140 (5): 678–691. doi:10.1016/j.cell.2010.01.003. PMC 2885838. PMID 20211137. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2885838.  edit
  25. ^ Geurts, A. M.; Cost, G. J.; Freyvert, Y.; Zeitler, B.; Miller, J. C.; Choi, V. M.; Jenkins, S. S.; Wood, A. et al. (2009). "Knockout Rats via Embryo Microinjection of Zinc-Finger Nucleases". Science 325 (5939): 433–433. doi:10.1126/science.1172447. PMC 2831805. PMID 19628861. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2831805.  edit
  26. ^ Flisikowska, T.; Thorey, I. S.; Offner, S.; Ros, F.; Lifke, V.; Zeitler, B.; Rottmann, O.; Vincent, A. et al. (2011). Milstone, David S.. ed. "Efficient Immunoglobulin Gene Disruption and Targeted Replacement in Rabbit Using Zinc Finger Nucleases". PLoS ONE 6 (6): e21045. doi:10.1371/journal.pone.0021045. PMC 3113902. PMID 21695153. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3113902.  edit
  27. ^ Hauschild, J.; Petersen, B.; Santiago, Y.; Queisser, A. -L.; Carnwath, J. W.; Lucas-Hahn, A.; Zhang, L.; Meng, X. et al. (2011). "Efficient generation of a biallelic knockout in pigs using zinc-finger nucleases". Proceedings of the National Academy of Sciences. doi:10.1073/pnas.1106422108.  edit
  28. ^ Yu, S.; Luo, J.; Song, Z.; Ding, F.; Dai, Y.; Li, N. (2011). "Highly efficient modification of beta-lactoglobulin (BLG) gene via zinc-finger nucleases in cattle". Cell Research. doi:10.1038/cr.2011.153.  edit
  29. ^ D. Carroll (2008). "Zinc-finger Nucleases as Gene Therapy Agents". Gene Therapy 15 (22): 1463–1468. doi:10.1038/gt.2008.145. PMC 2747807. PMID 18784746. http://www.nature.com/gt/journal/v15/n22/abs/gt2008145a.html. 
  30. ^ Li, H.; Haurigot, V.; Doyon, Y.; Li, T.; Wong, S. Y.; Bhagwat, A. S.; Malani, N.; Anguela, X. M. et al. (2011). "In vivo genome editing restores haemostasis in a mouse model of haemophilia". Nature 475 (7355). doi:10.1038/nature10177.  edit
  31. ^ a b c d Durai S, Mani M, Kandavelou K, Wu J, Porteus MH, Chandrasegaran S (2005). "Zinc finger nucleases: custom-designed molecular scissors for genome engineering of plant and mammalian cells". Nucleic Acids Res. 33 (18): 5978–90. doi:10.1093/nar/gki912. PMC 1270952. PMID 16251401. http://nar.oxfordjournals.org/cgi/pmidlookup?view=long&pmid=16251401. 
  32. ^ Lee HJ, Kim E, Kim JS (December 2009). "Targeted chromosomal deletions in human cells using zinc finger nucleases". Genome Res. 20 (1): 81–9. doi:10.1101/gr.099747.109. PMC 2798833. PMID 19952142. http://www.genome.org/cgi/pmidlookup?view=long&pmid=19952142. 
  33. ^ Mittelman, D; Moye, C, Morton, J, Sykoudis, K, Lin, Y, Carroll, D, Wilson, JH (2009-06-16). "Zinc-finger directed double-strand breaks within CAG repeat tracts promote repeat instability in human cells". Proceedings of the National Academy of Sciences of the United States of America 106 (24): 9607–12. doi:10.1073/pnas.0902420106. PMC 2701052. PMID 19482946. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2701052. 
  34. ^ Kandavelou K; Chandrasegaran S (2008). "Plasmids for Gene Therapy". Plasmids: Current Research and Future Trends. Caister Academic Press. ISBN 978-1-904455-35-6. 
  35. ^ T. Cathomen; J.K. Joung (2008). "Zinc-finger Nucleases: The Next Generation Emerges". Molecular Therapy 16 (7): 1200–1207. doi:10.1038/mt.2008.114. PMID 18545224. http://www.nature.com/mt/journal/v16/n7/abs/mt2008114a.html. 
  36. ^ Zinc-finger Nuclease-induced Gene Repair With Oligodeoxynucleotides: Wanted and Unwanted Target Locus Modifications Molecular Therapy vol. 18 no.4, 743-753 (2010)
  37. ^ Gupta A, Meng X, Zhu LJ, Lawson ND, Wolfe SA (September 2010). "Zinc finger protein-dependent and -independent contributions to the in vivo off-target activity of zinc finger nucleases". Nucleic Acids Res 39 (1): 381–392. doi:10.1093/nar/gkq787. PMC 3017618. PMID 20843781. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3017618. .
  38. ^ Pattanayak, V.; Ramirez, C. L.; Joung, J. K.; Liu, D. R. (2011). "Revealing Off-Target Cleavage Specificities of Zinc Finger Nucleases by in Vitro Selection". Nature Methods 8 (9): 765–770. doi:10.1038/nmeth.1670. PMC 3164905. PMID 21822273. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3164905.  edit
  39. ^ Grizot S, Smith J, Daboussi F, et al. (September 2009). "Efficient targeting of a SCID gene by an engineered single-chain homing endonuclease". Nucleic Acids Res. 37 (16): 5405–19. doi:10.1093/nar/gkp548. PMC 2760784. PMID 19584299. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2760784. 
  40. ^ Gao H, Smith J, Yang M, et al. (January 2010). "Heritable targeted mutagenesis in maize using a designed endonuclease". Plant J. 61 (1): 176–87. doi:10.1111/j.1365-313X.2009.04041.x. PMID 19811621. 
  41. ^ Christian M, Cermak T, Doyle EL, et al. (July 2010). "Targeting DNA Double-Strand Breaks with TAL Effector Nucleases". Genetics 186 (2): 757–61. doi:10.1534/genetics.110.120717. PMC 2942870. PMID 20660643. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2942870. 
  42. ^ Li T, Huang S, Jiang WZ, et al. (August 2010). "TAL nucleases (TALNs): hybrid proteins composed of TAL effectors and FokI DNA-cleavage domain". Nucleic Acids Res 39 (1): 359–372. doi:10.1093/nar/gkq704. PMC 3017587. PMID 20699274. http://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3017587. 
  43. ^ Moscou MJ, Bogdanove AJ (December 2009). "A simple cipher governs DNA recognition by TAL effectors". Science 326 (5959): 1501. Bibcode 2009Sci...326.1501M. doi:10.1126/science.1178817. PMID 19933106. 
  44. ^ Boch J, Scholze H, Schornack S, et al. (December 2009). "Breaking the code of DNA binding specificity of TAL-type III effectors". Science 326 (5959): 1509–12. doi:10.1126/science.1178811. PMID 19933107. 
  45. ^ Miller, J. C.; Tan, S.; Qiao, G.; Barlow, K. A.; Wang, J.; Xia, D. F.; Meng, X.; Paschon, D. E. et al. (2010). "A TALE nuclease architecture for efficient genome editing". Nature Biotechnology 29 (2): 143. doi:10.1038/nbt.1755. PMID 21179091.  edit
  46. ^ [[cite news |title=How Patient X Could Save Millions of Lives}}

Further reading

External links


Wikimedia Foundation. 2010.

Игры ⚽ Нужно сделать НИР?

Look at other dictionaries:

  • Zinc finger chimera — Zinc finger protein chimera are chimeric proteins composed of a DNA binding zinc finger protein domain and another domain through which the protein exerts its effect. The effector domain may be a transcriptional activator (A) or repressor… …   Wikipedia

  • Zinc finger protein transcription factor — Zinc finger protein transcription factors or ZFP TFs, consisting of activators and repressors are transcription factors composed of a zinc finger protein domain (ZFP) and any of a variety of transcription factor effector domains which exert their …   Wikipedia

  • Zinc finger protein — A zinc finger protein is a DNA binding protein domain comprised of zinc fingers ranging from two in the Drosophila regulator ADR1, the more common three in mammalian Sp1 up to nine in TFIIIA. They occur in nature as the part of transcription… …   Wikipedia

  • Nucléase à doigt de zinc — Les nucléases à doigts de zinc sont des enzymes de restriction artificielles créées par la fusion d un domaine de liaison à l ADN, de type « doigt de zinc », et d un domaine catalytique de coupure de l ADN (nucléase). Ingénierie par… …   Wikipédia en Français

  • Chimeric nuclease — Chimeric nucleases are an example of engineered proteins which must comprise a DNA binding domain to give sequence specificity and a nuclease domain for DNA cleavage. Contents 1 DNA binding domains 2 Nuclease domain 3 See also …   Wikipedia

  • Restriction enzyme — Glossary Restriction …   Wikipedia

  • FokI (biology) — The enzyme FokI, naturally found in Flavobacterium okeanokoites , is a bacterial type IIS restriction endonuclease consisting of an N terminal DNA binding domain and a non specific DNA cleavage domain at the C terminal.cite journal | author =… …   Wikipedia

  • CompoZr — The CompoZr Zinc finger nuclease (ZFN) platform is a technology developed by Sigma Aldrich that allows researchers to target and manipulate the genome of living cells thereby creating cell lines or entire organisms with permanent and heritable… …   Wikipedia

  • ZFN — is an acronym that may refer to:*Tulita Airport, an airport in Canada *Zinc finger nuclease, a nuclease enzyme coupled to a zinc finger based DNA binding domain *The Zero Configuration File Network, an open source network program …   Wikipedia

  • Génie génétique — Grains de blé résistants à une maladie, obtenus à partir d’une enzyme fabriquant naturellement des antibiotiques Le génie génétique (ou ingénierie génétique[1]) est un ensemble de techniques, faisant partie de la biologie moléculaire et ayant p …   Wikipédia en Français

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”