„Zinkfingernukleasen“ – Versionsunterschied

aus Wikipedia, der freien Enzyklopädie
Zur Navigation springen Zur Suche springen
[ungesichtete Version][ungesichtete Version]
Inhalt gelöscht Inhalt hinzugefügt
Rjwilmsi (Diskussion | Beiträge)
K →‎Disabling an allele: Journal cites:, added 1 PMC, using AWB (7751)
imported>ScienceGeekling
Updated references
Zeile 1: Zeile 1:
'''Zinc-finger nucleases''' ('''ZFN'''s) 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 which enables zinc-finger nucleases to target a unique sequence within a complex genome. By taking advantage of endogenous DNA repair machinery, these reagents can be used to precisely alter the genomes of higher organisms.
'''Zinc-finger nucleases''' ('''ZFN'''s) 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.


== DNA-cleavage domain ==
== DNA-cleavage domain ==


The non-specific cleavage domain from the type IIs [[restriction endonuclease]] [[FokI]] is typically used as the cleavage domain in ZFNs.
The non-specific cleavage domain from the type IIs [[restriction endonuclease]] [[FokI]] is typically used as the cleavage domain in ZFNs.<ref name="kim1996">{{cite journal| last=Kim| first=YG| coauthors= Cha, J., Chandrasegaran, S.| year=1996| title=Hybrid restriction enzymes: zinc finger fusions to Fok I cleavage domain| journal=Proc Natl Acad Sci USA| volume= 93| issue=3| pages=1156–60 |url=http://www.pnas.org/content/93/3/1156.abstract| doi=10.1073/pnas.93.3.1156| pmid=8577732| pmc=40048}}</ref>
<ref name="kim1996">{{cite journal| last=Kim| first=YG| coauthors= Cha, J., Chandrasegaran, S.| year=1996| title=Hybrid restriction enzymes: zinc finger fusions to Fok I cleavage domain| journal=Proc Natl Acad Sci USA| volume= 93| issue=3| pages=1156–60 |url=http://www.pnas.org/content/93/3/1156.abstract| doi=10.1073/pnas.93.3.1156| pmid=8577732| pmc=40048}}</ref>
This cleavage domain must dimerize in order to cleave DNA<ref name="bitinaite1998">{{cite journal| last=Bitinaite| first=J.| coauthors= D. A. Wah, Aggarwal, A. K., Schildkraut, I.| year=1998| title=FokI dimerization is required for DNA cleavage| journal=Proc Natl Acad Sci USA| volume= 95| issue=18| pages=10570–5 |url=http://www.pnas.org/cgi/content/full/95/18/10570| doi=10.1073/pnas.95.18.10570| pmid=9724744| pmc=27935}}</ref>
This cleavage domain must dimerize in order to cleave DNA
<ref name="bitinaite1998">{{cite journal| last=Bitinaite| first=J.| coauthors= D. A. Wah, Aggarwal, A. K., Schildkraut, I.| year=1998| title=FokI dimerization is required for DNA cleavage| journal=Proc Natl Acad Sci USA| volume= 95| issue=18| pages=10570–5 |url=http://www.pnas.org/cgi/content/full/95/18/10570| doi=10.1073/pnas.95.18.10570| pmid=9724744| pmc=27935}}</ref>
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.<ref>{{cite journal |author=Cathomen T, Joung JK |title=Zinc-finger nucleases: the next generation emerges |journal=Mol. Ther. |volume=16 |issue=7 |pages=1200–7 |year=2008 |month=July |pmid=18545224 |doi=10.1038/mt.2008.114 |url=http://www.nature.com/mt/journal/v16/n7/abs/mt2008114a.html}}
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.<ref>{{cite journal |author=Cathomen T, Joung JK |title=Zinc-finger nucleases: the next generation emerges |journal=Mol. Ther. |volume=16 |issue=7 |pages=1200–7 |year=2008 |month=July |pmid=18545224 |doi=10.1038/mt.2008.114 |url=http://www.nature.com/mt/journal/v16/n7/abs/mt2008114a.html}}
</ref>
</ref>
Zeile 47: Zeile 45:


== Applications ==
== Applications ==
[[Zinc finger nucleases]] have become useful reagents for manipulating genomes of many higher organisms including ''[[Drosophila melanogaster]]'', ''[[Caenorhabditis elegans]]'', [[Sea urchin]],<ref>
[[Zinc finger nucleases]] have become useful reagents for manipulating the genomes of many plants and animals including [[arabidopsis]]<ref name=Zhang2010>{{Cite doi|10.1073/pnas.0914991107}}</ref><ref name=Osakabe2010>{{Cite doi|10.1073/pnas.1000234107}}</ref>, [[tobacco]]<ref name=Cai2008>{{Cite doi|10.1007/s11103-008-9449-7}}</ref><ref name=Townsend2009>{{Cite doi|10.1038/nature07845}}</ref>, [[soybean]],<ref name=Curtin2011>{{Cite doi|10.1104/pp.111.172981}}</ref> [[maize|corn]],<ref>
{{cite journal
|author=Ochiai H, Fujita K, Suzuki K, ''et al.''
|title=Targeted mutagenesis in the sea urchin embryo using zinc-finger nucleases
|journal=Genes to Cells
|volume=15
|issue=8
|pages=875–85,
|year=2010
|month=Aug
|pmid= 20604805
|doi=10.1111/j.1365-2443.2010.01425.x
|unused_data=DUPLICATE DATA: doi=10.1111/j.1365-2443.2010.01425.x
}}</ref> [[tobacco]], [[maize|corn]],<ref>
{{cite journal
{{cite journal
|author=Shukla VK, Doyon Y, Miller JC, ''et al.''
|author=Shukla VK, Doyon Y, Miller JC, ''et al.''
Zeile 74: Zeile 59:
|unused_data=DUPLICATE DATA: doi=10.1038/nature07992
|unused_data=DUPLICATE DATA: doi=10.1038/nature07992
|bibcode=2009Natur.459..437S
|bibcode=2009Natur.459..437S
}}</ref> ''[[Drosophila melanogaster]]'',<ref name=Bibikoca2003>{{Cite doi|10.1126/science.1079512}}</ref> ''[[Caenorhabditis elegans| C. elegans]]'',<ref name=Wood2011>{{Cite doi|10.1126/science.1207773}}</ref> [[Sea urchin]],<ref>
}}</ref>
{{cite journal
|author=Ochiai H, Fujita K, Suzuki K, ''et al.''
|title=Targeted mutagenesis in the sea urchin embryo using zinc-finger nucleases
|journal=Genes to Cells
|volume=15
|issue=8
|pages=875–85,
|year=2010
|month=Aug
|pmid= 20604805
|doi=10.1111/j.1365-2443.2010.01425.x
|unused_data=DUPLICATE DATA: doi=10.1111/j.1365-2443.2010.01425.x
}}</ref>
[[silkworm]],<ref name=Takasua2010>{{Cite doi|10.1016/j.ibmb.2010.07.012}}</ref>
[[zebrafish]],<ref>
[[zebrafish]],<ref>
{{cite journal
{{cite journal
Zeile 88: Zeile 87:
|issue = 2
|issue = 2
|pmc = 2849655
|pmc = 2849655
}}</ref>, [[frog]]s,<ref name=Young2011>{{Cite doi|10.1073/pnas.1102030108}}</ref> [[mice]],<ref name=Goldberg2010>{{Cite doi|10.1016/j.cell.2010.01.003}}</ref> [[rat]]s,<ref name=Geurts2009>{{Cite doi|10.1126/science.1172447}}</ref> [[rabbit]]s,<ref name=Flisikowska2011>{{Cite doi|10.1371/journal.pone.0021045}}</ref> [[pig]]s,<ref name=Hauschild2011>{{Cite doi|10.1073/pnas.1106422108}}</ref> and
}}</ref>
various types of mammalian cells,<ref>
various types of mammalian cells.<ref>
{{cite journal
{{cite journal
|author = D. Carroll
|author = D. Carroll
Zeile 102: Zeile 101:
|issue = 22
|issue = 22
|pmc = 2747807
|pmc = 2747807
}}</ref> [[Zinc finger nucleases]] have also been used in a mouse model of [[haemophilia]]<ref name=Li2011>{{Cite doi|10.1038/nature10177}}</ref> 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]].
}}</ref>
and [[rats]].<ref>
{{cite journal
|author=Geurts AM, Cost GJ, Freyvert Y, ''et al.''
|title=Knockout rats via embryo microinjection of zinc-finger nucleases
|journal=Science
|volume=325
|issue=5939
|pages=433
|year=2009
|month=July
|pmid=19628861
|doi=10.1126/science.1172447
|url=http://www.sciencemag.org/cgi/pmidlookup?view=long&pmid=19628861
|pmc=2831805
}}</ref>
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 ===
=== Disabling an allele ===

Version vom 8. Juli 2011, 07:37 Uhr

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.

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 generated 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] and various types of mammalian cells.[28] Zinc finger nucleases have also been used in a mouse model of haemophilia[29] 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.[30] Multiple pairs of ZFNs can also be used to completely remove entire large segments of genomic sequence.[31]

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.[32]

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.[33] 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.[30]

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.[30] Despite advances in engineering both more specific zinc finger domains and modified FokI cleavage domains ,[34] ZFN off-target activity is still a significant concern.[35] Two separate methods have been demonstrated to decrease off-target cleavage for 3-finger ZFNs that target two adjacent 9-basepair sites. [36] Other groups use ZFNs with 4, 5 or 6 zinc fingers that target longer and presumably rarer sites and such ZFNs may tend to yield less off-target activity, but this has not been conclusively demonstrated.

Immunogenicity

Vorlage:Details

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.[30]

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 [37] [38] and nucleases based on engineered TAL effectors. [39] [40] TAL effector nucleases (TALENs) are particularly interesting because TAL effectors appear to be very simple to engineer [41] [42] and TALENs can be used to target endogenous loci in human cells.[43] But to date no one has reported the isolation of clonal cell lines or transgenic organisms using such reagents.

See also

References

Vorlage:Reflist

Further reading

  • Mandell JG, Barbas CF: Zinc Finger Tools: custom DNA-binding domains for transcription factors and nucleases. In: Nucleic Acids Res. 34. Jahrgang, Web Server issue, Juli 2006, S. W516–23, doi:10.1093/nar/gkl209, PMID 16845061, PMC 1538883 (freier Volltext) – (oxfordjournals.org).
  • Porteus MH, Carroll D: Gene targeting using zinc finger nucleases. In: Nat. Biotechnol. 23. Jahrgang, Nr. 8, August 2005, S. 967–73, doi:10.1038/nbt1125, PMID 16082368.
  • Doyon Y, McCammon JM, Miller JC, et al.: Heritable targeted gene disruption in zebrafish using designed zinc-finger nucleases. In: Nat. Biotechnol. 26. Jahrgang, Nr. 6, Juni 2008, S. 702–8, doi:10.1038/nbt1409, PMID 18500334, PMC 2674762 (freier Volltext).
  • Meng X, Noyes MB, Zhu LJ, Lawson ND, Wolfe SA: Targeted gene inactivation in zebrafish using engineered zinc-finger nucleases. In: Nat. Biotechnol. 26. Jahrgang, Nr. 6, Juni 2008, S. 695–701, doi:10.1038/nbt1398, PMID 18500337, PMC 2502069 (freier Volltext).

External links

  1. YG Kim, Cha, J., Chandrasegaran, S.: Hybrid restriction enzymes: zinc finger fusions to Fok I cleavage domain. In: Proc Natl Acad Sci USA. 93. Jahrgang, Nr. 3, 1996, S. 1156–60, doi:10.1073/pnas.93.3.1156, PMID 8577732, PMC 40048 (freier Volltext) – (pnas.org).
  2. J. Bitinaite, D. A. Wah, Aggarwal, A. K., Schildkraut, I.: FokI dimerization is required for DNA cleavage. In: Proc Natl Acad Sci USA. 95. Jahrgang, Nr. 18, 1998, S. 10570–5, doi:10.1073/pnas.95.18.10570, PMID 9724744, PMC 27935 (freier Volltext) – (pnas.org).
  3. Cathomen T, Joung JK: Zinc-finger nucleases: the next generation emerges. In: Mol. Ther. 16. Jahrgang, Nr. 7, Juli 2008, S. 1200–7, doi:10.1038/mt.2008.114, PMID 18545224 (nature.com).
  4. Vorlage:Cite doi
  5. Vorlage:Cite doi
  6. Vorlage:Cite doi
  7. Vorlage:Cite doi
  8. Vorlage:Cite doi
  9. C.O. Pabo; E.Peisach; R.A. Grant: Design and Selection of Novel Cys2His2 Zinc Finger Proteins. In: Annu. Rev. Biochem. 70. Jahrgang, 2001, S. 313–40, doi:10.1146/annurev.biochem.70.1.313, PMID 11395410 (annualreviews.org).
  10. Ramirez CL, Foley JE, Wright DA, et al.: Unexpected failure rates for modular assembly of engineered zinc fingers. In: Nat. Methods. 5. Jahrgang, Nr. 5, Mai 2008, S. 374–5, doi:10.1038/nmeth0508-374, PMID 18446154.
  11. Maeder ML, et al.': Rapid "Open-Source" Engineering of Customized Zinc-Finger Nucleases for Highly Efficient Gene Modification. In: Mol. Cell. 31. Jahrgang, Nr. 2, September 2008, S. 294–301, doi:10.1016/j.molcel.2008.06.016, PMID 18657511, PMC 2535758 (freier Volltext) – (elsevier.com).
  12. Vorlage:Cite doi
  13. Vorlage:Cite doi
  14. Vorlage:Cite doi
  15. Vorlage:Cite doi
  16. Vorlage:Cite doi
  17. Shukla VK, Doyon Y, Miller JC, et al.: Precise genome modification in the crop species Zea mays using zinc-finger nucleases. In: Nature. 459. Jahrgang, Nr. 7245, Mai 2009, S. 437–41, doi:10.1038/nature07992, PMID 19404259, bibcode:2009Natur.459..437S.
  18. Vorlage:Cite doi
  19. Vorlage:Cite doi
  20. Ochiai H, Fujita K, Suzuki K, et al.: Targeted mutagenesis in the sea urchin embryo using zinc-finger nucleases. In: Genes to Cells. 15. Jahrgang, Nr. 8, August 2010, S. 875–85,, doi:10.1111/j.1365-2443.2010.01425.x, PMID 20604805.
  21. Vorlage:Cite doi
  22. S.C. Ekker: Zinc finger-based knockout punches for zebrafish genes. In: Zebrafish. 5. Jahrgang, Nr. 2, 2008, S. 1121–3, doi:10.1089/zeb.2008.9988, PMID 18554175, PMC 2849655 (freier Volltext) – (liebertonline.com).
  23. Vorlage:Cite doi
  24. Vorlage:Cite doi
  25. Vorlage:Cite doi
  26. Vorlage:Cite doi
  27. Vorlage:Cite doi
  28. D. Carroll: Progress and prospects: Zinc-finger nucleases as gene therapy agents. In: Gene Therapy. 15. Jahrgang, Nr. 22, 2008, S. 1463–1468, doi:10.1038/gt.2008.145, PMID 18784746, PMC 2747807 (freier Volltext) – (nature.com).
  29. Vorlage:Cite doi
  30. a b c d Durai S, Mani M, Kandavelou K, Wu J, Porteus MH, Chandrasegaran S: Zinc finger nucleases: custom-designed molecular scissors for genome engineering of plant and mammalian cells. In: Nucleic Acids Res. 33. Jahrgang, Nr. 18, 2005, S. 5978–90, doi:10.1093/nar/gki912, PMID 16251401, PMC 1270952 (freier Volltext) – (oxfordjournals.org).
  31. Lee HJ, Kim E, Kim JS: Targeted chromosomal deletions in human cells using zinc finger nucleases. In: Genome Res. 20. Jahrgang, Nr. 1, Dezember 2009, S. 81–9, doi:10.1101/gr.099747.109, PMID 19952142, PMC 2798833 (freier Volltext) – (genome.org).
  32. D Mittelman, Moye, C, Morton, J, Sykoudis, K, Lin, Y, Carroll, D, Wilson, JH: Zinc-finger directed double-strand breaks within CAG repeat tracts promote repeat instability in human cells. In: Proceedings of the National Academy of Sciences of the United States of America. 106. Jahrgang, Nr. 24, 16. Juni 2009, S. 9607–12, doi:10.1073/pnas.0902420106, PMID 19482946, PMC 2701052 (freier Volltext).
  33. Kandavelou K; Chandrasegaran S: Plasmids: Current Research and Future Trends. Caister Academic Press, 2008, ISBN 978-1-904455-35-6, Plasmids for Gene Therapy.
  34. T. Cathomen; J.K. Joung: Zinc-finger Nucleases: The Next Generation Emerges. In: Molecular Therapy. 16. Jahrgang, Nr. 7, 2008, S. 1200–1207, doi:10.1038/mt.2008.114, PMID 18545224 (nature.com).
  35. Zinc-finger Nuclease-induced Gene Repair With Oligodeoxynucleotides: Wanted and Unwanted Target Locus Modifications Molecular Therapy vol. 18 no.4, 743-753 (2010)
  36. Gupta A, Meng X, Zhu LJ, Lawson ND, Wolfe SA: Zinc finger protein-dependent and -independent contributions to the in vivo off-target activity of zinc finger nucleases. In: Nucleic Acids Res. 39. Jahrgang, Nr. 1, September 2010, S. 381–392, doi:10.1093/nar/gkq787, PMID 20843781, PMC 3017618 (freier Volltext)..
  37. Grizot S, Smith J, Daboussi F, et al.: Efficient targeting of a SCID gene by an engineered single-chain homing endonuclease. In: Nucleic Acids Res. 37. Jahrgang, Nr. 16, September 2009, S. 5405–19, doi:10.1093/nar/gkp548, PMID 19584299, PMC 2760784 (freier Volltext).
  38. Gao H, Smith J, Yang M, et al.: Heritable targeted mutagenesis in maize using a designed endonuclease. In: Plant J. 61. Jahrgang, Nr. 1, Januar 2010, S. 176–87, doi:10.1111/j.1365-313X.2009.04041.x, PMID 19811621.
  39. Christian M, Cermak T, Doyle EL, et al.: TAL Effector Nucleases Create Targeted DNA Double-strand Breaks. In: Genetics. 186. Jahrgang, Nr. 2, Juli 2010, S. 757–61, doi:10.1534/genetics.110.120717, PMID 20660643, PMC 2942870 (freier Volltext).
  40. Li T, Huang S, Jiang WZ, et al.: TAL nucleases (TALNs): hybrid proteins composed of TAL effectors and FokI DNA-cleavage domain. In: Nucleic Acids Res. 39. Jahrgang, Nr. 1, August 2010, S. 359–372, doi:10.1093/nar/gkq704, PMID 20699274, PMC 3017587 (freier Volltext).
  41. Moscou MJ, Bogdanove AJ: A simple cipher governs DNA recognition by TAL effectors. In: Science. 326. Jahrgang, Nr. 5959, Dezember 2009, S. 1501, doi:10.1126/science.1178817, PMID 19933106, bibcode:2009Sci...326.1501M.
  42. Boch J, Scholze H, Schornack S, et al.: Breaking the code of DNA binding specificity of TAL-type III effectors. In: Science. 326. Jahrgang, Nr. 5959, Dezember 2009, S. 1509–12, doi:10.1126/science.1178811, PMID 19933107.
  43. Vorlage:Cite doi